You are on page 1of 579

Sandra 

Fiori
Paula D. Pratolongo  Editors

The Bahía
Blanca
Estuary
Ecology and Biodiversity
The Bahía Blanca Estuary
Sandra M. Fiori • Paula D. Pratolongo
Editors

The Bahía Blanca Estuary


Ecology and Biodiversity
Editors
Sandra M. Fiori Paula D. Pratolongo
Instituto Argentino de Oceanografía IADO Centro de Recursos Naturales Renovables
(Universidad Nacional del Sur-CONICET) de la Zona Semiárida, CERZOS
Departamento de Biología, Bioquímica y (Universidad Nacional del Sur-CONICET)
Farmacia, Universidad Nacional del Sur Departamento de Biología, Bioquímica y
Bahía Blanca, Argentina Farmacia, Universidad Nacional del Sur
Bahía Blanca, Argentina

ISBN 978-3-030-66485-5    ISBN 978-3-030-66486-2 (eBook)


https://doi.org/10.1007/978-3-030-66486-2

© Springer Nature Switzerland AG 2021


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of
the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation,
broadcasting, reproduction on microfilms or in any other physical way, and transmission or information
storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology
now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors, and the editors are safe to assume that the advice and information in this book
are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or the
editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.

This Springer imprint is published by the registered company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Preface

The Bahía Blanca Estuary is among the most ecologically rich and complex areas
along the Argentine coast. It is composed of 2,300  km2 of intertidal flats, salt
marshes, shallow channels, and emerged islands, arranged to create intricate pat-
terns of land-sea interfaces where terrestrial and marine realms converge. Ecosystems
in the estuary are shaped by complex exchanges of energy, water, nutrients, sedi-
ments, and biota. These natural environments sustain a high concentration of marine
and terrestrial species, including endemic, threatened, and endangered fish, shore-
birds, and marine mammals, whose major resting and reproductive habitats have
been included within the several nature reserves established in the area. Puerto
Cuatreros, in the inner zone of the estuary hosts a permanent station of marine
research, whose records span more than 30 years of biophysical variables and rep-
resent one of the largest time series of ecological data in South America.
Large scale deterioration of coastal ecosystem is a global issue, and the Bahía
Blanca Estuary is no exception. Despite its large ecological relevance, the estuary is
under increasing anthropogenic pressure by large urban settlements, industrial
installations, and port expansion projects, raising the question of how we can bal-
ance both conservation and human development. Overfishing, invasive species that
establish and thrive, and runoff of nutrients and pollutants from cities and land-­
based industries are major environmental issues that arise from conflicts of interest
between the various users of the coastal zone. This book deals primarily with the
biological aspects of ecosystem health, but we recognize the many dimensions of
the problem and offer an insight into some relevant social challenges for the envi-
ronmental management in the region.
This book presents an updated revision of the biology and ecology of the major
groups of living organisms inhabiting the coastal zone of Bahía Blanca, including
new material on current research revealing the presence of previously unrecognized
species and changes in ecosystem functions. The different chapters throughout the
book include additions on how humans adversely impact the natural communities,
but specific chapters have been also incorporated to deal with major anthropogenic
pressures and conservation efforts in the area.

v
vi Preface

The book is organized into 20 chapters written by the local experts on each topic.
The first four chapters introduce readers to the regional context, geographical set-
tings, and physical and chemical processes that affect the distribution and abun-
dance of living organisms. Chapters 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15 and 16
comprise the core of the book, on the composition, organization, management
issues, and conservation status of the relevant biological communities of major hab-
itat types within the estuary and the area of influence. The final four chapters in the
book deal with socio-environmental aspects relevant for the ecosystem health of the
estuary.
At the regional level, we hope this book will help young students and citizens to
appreciate the many valuable ecosystems around us and promote a new generation
to continue efforts towards sensible management and conservation of our natural
environments. From a broader perspective, we expect to offer access to the interna-
tional readership seeking for updated information on the biology and ecology of the
Bahía Blanca Estuary.

Bahía Blanca, Argentina Sandra M. Fiori


Paula D. Pratolongo
Acknowledgement

This book would not be complete without the acknowledgement of the exceptional
job carried out by the authors who contributed their expertise to each chapter. They
are all renowned local specialists in the different topics covered within the biologi-
cal, social, conservation, and education sciences, and their distinctive views of the
Bahía Blanca Estuary are all represented in this book. As editors, we want to express
our gratitude to the authors who produced high-quality material, mostly based on
their own research in the region, and contributed original photographs, maps, and
illustrations. Through their contributions they have made this book a high-quality
education resource for a broad community.
Sandra M. Fiori and Paula D. Pratolongos
Bahía Blanca, July 2020

vii
Contents

1 The Bahía Blanca Estuary in a Regional Context��������������������������������    1


Paula D. Pratolongo and Sandra M. Fiori
2 Geography of the Bahía Blanca Estuary ����������������������������������������������   17
Walter Daniel Melo
3 Physical Oceanography of the Bahía Blanca Estuary��������������������������   31
Gerardo M. E. Perillo and M. Cintia Piccolo
4 Bahía Blanca Estuary: A Chemical Oceanographic Approach ����������   51
Jorge E. Marcovecchio, Ana L. Oliva, Noelia S. La Colla,
Andrés H. Arias, Sandra E. Botté, Pía Simonetti, Analía V. Serra,
Vanesa L. Negrin, Ana C. Ronda, and Claudia E. Domini
5 Plankton Ecology and Biodiversity in the Bahía Blanca Estuary ������   83
Anabela A. Berasategui, M. Sofía Dutto, Celeste López-Abbate,
and Valeria A. Guinder
6 Biology and Ecology of the Benthic Algae ��������������������������������������������  113
M. Emilia Croce, M. Cecilia Gauna, Carolina Fernández,
Ailen M. Poza, and Elisa R. Parodi
7 The Intertidal Meiobenthos of the Bahía Blanca Estuary ������������������  153
Verónica N. Bulnes, Agustín G. Menechella, Kevin A. Rucci,
and Michel Sciberras
8 The Intertidal Soft-Bottom Macrobenthic Invertebrates��������������������  179
M. Cecilia Carcedo, Sabrina Angeletti, Georgina Zapperi,
Eder P. Dos Santos, and Sandra M. Fiori
9 Taxonomic and Functional Assessment
of Subtidal Macrobenthic Communities
in the Bahía Blanca Estuary (Argentina)����������������������������������������������  215
M. Emilia Bravo, M. Cecilia Carcedo, Eder P. Dos Santos,
and Sandra M. Fiori

ix
x Contents

10 Shrimps and Prawns��������������������������������������������������������������������������������  253


Patricia Marta Cervellini and Jorge Omar Pierini
11 Ecology and Biology of Fish Assemblages ��������������������������������������������  275
Juan Manuel Molina, Gabriela Blasina,
and Andrea Lopez Cazorla
12 Bahía Blanca Estuary and the Importance
of Wetlands for the Conservation of Sea Turtles����������������������������������  307
Victoria Massola, Laura Prosdocimi, Cristina Suldrup,
and Juan Facundo Sosa
13 Shorebirds and Seabirds’ Ecology and Conservation��������������������������  327
Natalia S. Martínez-Curci, Germán O. García, Leandro Marbán,
Pía Simonetti, and Sergio M. Zalba
14 Marine Mammals: Is the Bahía Blanca Estuary
and Its Area of Influence Important for Their Conservation?������������  359
Gisela Giardino, Estela M. Luengos Vidal, Victoria Massola,
M. Agustina Mandiola, Joaquín C. M. Gana, Diego Rodríguez,
and Ricardo Bastida
15 Use of Coastal Area Habitats by Land Mammals��������������������������������  397
Estela M. Luengos Vidal, Nicolás Caruso, Sabrina Martinez,
Emma Casanave, and Lucherini Mauro
16 Coastal Wetlands of the Bahía Blanca Estuary:
Landscape Structure and Plant Associations����������������������������������������  435
Paula Daniela Pratolongo, Flavia Funk, María Julia Piovan,
Carla Celleri, and Vanesa L. Negrin
17 Environmental Diagnosis of the Protected Coastal
Areas of the Bahía Blanca Estuary��������������������������������������������������������  469
M. Elizabeth Carbone, María Ángeles Speake,
and Walter Daniel Melo
18 Small-Scale Artisanal Fishers and Socio-­environmental
Conflicts in Estuarine and Coastal Wetlands����������������������������������������  493
Daniela M. Truchet and M. Belén Noceti
19 Estuarine Environmental Monitoring Programs:
Long-Term Studies����������������������������������������������������������������������������������  521
Jorge E. Marcovecchio, Sandra E. Botté, Silvia G. De Marco,
Andrea Lopez Cazorla, Andrés H. Arias, Mónica Baldini,
María Amelia Cubitto, Sandra M. Fiori, Ana L. Oliva,
Noelia La Colla, Gabriela Blasina, Juan Manuel Molina,
Pia Simonetti, Analía V. Serra, Vanesa L. Negrín, Ana C. Ronda,
and Marcelo T. Pereyra
Contents xi

20 Environmental Education: Mud and Salt Classrooms ������������������������  549


Cristina Sanhueza and Paola Germain

Index������������������������������������������������������������������������������������������������������������������  563
About the Editors

Sandra M. Fiori, PhD  is a Professor of Ecology of Communities and Ecosystems


at Universidad Nacional del Sur, Argentina, and also holds a position as Independent
Researcher at CONICET (National Research Council, Argentina). She obtained a
Master in Wild Life Management from Universidad de Córdoba in 1997 and PhD in
Biological Sciences from Universidad de Nacional del Sur in 2003, and completed
postdoctoral training at Instituto Politécnico Nacional – Unidad Mérida (Yucatán,
México).

Paula  D.  Pratolongo, PhD  is a Professor of Marine Ecology at Universidad


Nacional del sur, Argentina, and also holds a position as Independent Researcher at
CONICET (National Research Council, Argentina). She obtained a PhD in
Biological Sciences from Universidad de Buenos Aires in 2005, and completed
postdoctoral training at East Carolina University, North Carolina, under a Fulbright
Award. Dra. Pratolongo received the MaB Young Scientists Award on Ecosystems
and Water, granted by the MaB-UNESCO Program.

xiii
Chapter 1
The Bahía Blanca Estuary in a Regional
Context

Paula D. Pratolongo and Sandra M. Fiori

1.1  Introduction

The Bahía Blanca Estuary is one of the largest coastal systems in Atlantic South
America, shaped by a unique combination of large interannual climatic variations
and a transgressive sea level during the Holocene. This mesotidal estuary is placed
in a sharp transition between humid subtropical and semiarid climates and
encompasses about 2300  km2 of wide intertidal flats, salt marshes, and emerged
islands, arranged to create intricate landscape patterns. If we are to study the
ecosystems and living organisms of the Bahía Blanca Estuary, we need to first
describe the regional context and recognize the physical and sociocultural attributes
that influence species distributions, life histories, and ecosystem’s health. Thus, the
best way to open this book is to provide the readers with an introduction to the
regional context along the Atlantic coast of temperate South America.
Many estuaries and coastal wetland systems occur along the Atlantic coast of
South America, from the temperate southern coasts of Brazil to Tierra del Fuego
Island in Argentina (Fig. 1.1). All these systems share a similar configuration of the
coast, located on a passive margin far back from the seafloor spreading at the mid-­
Atlantic ridge. The east coast of both South and North America is geologically

P. D. Pratolongo ()
Centro de Recursos Naturales Renovables de la Zona Semiárida, CERZOS (Universidad
Nacional del Sur-CONICET), Bahía Blanca, Argentina
Departamento de Biología, Bioquímica y Farmacia, Universidad Nacional del Sur,
Bahía Blanca, Argentina
S. M. Fiori
Departamento de Biología, Bioquímica y Farmacia, Universidad Nacional del Sur,
Bahía Blanca, Argentina
Instituto Argentino de Oceanografía, IADO (Universidad Nacional del Sur-CONICET),
Bahía Blanca, Argentina

© Springer Nature Switzerland AG 2021 1


S. M. Fiori, P. D. Pratolongo (eds.), The Bahía Blanca Estuary,
https://doi.org/10.1007/978-3-030-66486-2_1
2 P. D. Pratolongo and S. M. Fiori

Fig. 1.1  The Atlantic coast of South America from southern Brazil to Tierra del Fuego Island in
Argentina. (Map by Walter D. Melo)

ancient and has been stable for millions of years. During this time, the coastline
eroded, and continental landforms wore down, leaving wide continental margins of
accumulated sediments. Although South and North American Atlantic shorelines
are similar in their geographical configuration, they went through different relative
sea-level histories during the Holocene, which modeled a distinct evolution of
coastal landscapes.
Worldwide, estuaries adapted to the changing sea levels during the Holocene. At
any time, the balance between the regional rates of relative sea-level change and the
local sediment budget determined the evolution of an estuary. Under a rising sea
1  The Bahía Blanca Estuary in a Regional Context 3

level, the evolution of estuaries was mainly controlled by vertical accretion at rates
that kept pace with the sea-level rise. For instance, in the Chesapeake Bay (eastern
North America), shallow estuaries fringed by wetlands formed at locations with
abundant sediment inputs and slow rates of relative sea-level rise over the late
Holocene. Sediment-starved estuaries in the same region, like those on the mainland
edge of the Delmarva Peninsula, moved upland along with their coastal wetlands,
creating a varied succession of wetland habitats. Further north, where the relative
sea level fell and new land emerged constantly (e.g., along the Hudson Bay, Canada),
estuaries migrated seaward and the old coastal features became part of the terrestrial
landscape. So let us first introduce the concept of relative sea level and describe the
different regional evolutions of the relative sea level during the Holocene.

1.2  The Relative Sea Level During the Holocene

The globally averaged (also called eustatic) sea level has been rising from the Last
Glacial Maximum (LGM) to the present. However, the relative sea level (relative
height of the sea with respect to land) can vary from place to place because of
eustatic, isostatic, ocean circulation, tectonic, and local factors (Khan et al. 2015).
After the LGM, the extensive melting of massive ice sheets produced a transfer of
mass from the continents to the ocean. Ice melting increased the ocean volume (thus
augmenting the eustatic sea level), but also uplifted the Earth’s crust, in a process
called glacio-isostatic adjustment (Peltier 2002).
Near-field locations are those regions that were close to the margins of large ice
sheets during the LGM. In these regions, the rate of glacio-isostatic uplift was
higher than the eustatic sea-level rise, and the relative sea level monotonically felt
from the LGM to the present. In these areas, typically high-latitude coastal regions
(e.g., Fennoscandia, Finland, Labrador), along with the falling sea level, new land
emerged, and coastal settings expanded seaward.
Intermediate-field regions extend for several thousand kilometers away from ice
sheet edges (Clark et al. 1978). In these areas, the mantle material flowed off the
forebulge, which is the lump of the lithosphere bordering the ice sheet. This effect
typically produced the glacio-isostatic subsidence that characterizes intermediate-­
field coastal regions (Fleming et al. 1998). In coastal settings along the temperate
Atlantic coasts of North America, for instance, the relative sea level was continually
rising from the LGM to the present due to the combination of isostatic and eustatic
effects. In these and other intermediate-field regions, Holocene estuaries drowned,
and new estuaries formed landward.
Finally, far-field locations are away from large ice sheets. In these regions,
eustatic contributions first exceeded glacio-isostatic effects, and the relative sea
level raised during the early Holocene. Far-field locations commonly present a mid-­
Holocene highstand (e.g., relative sea-level maximum) above the elevation of
present shorelines. After the Holocene highstand, the relative sea level in far-field
regions begun to fall to the current position. In low-latitude regions, the late
4 P. D. Pratolongo and S. M. Fiori

Holocene relative sea-level fall was mainly due to the migration of water away from
the tropics into subsiding forebulges (Mitrovica and Milne 2002). In far-field
locations at higher latitudes, a crustal tilting, upward toward the continent, produced
the observed relative sea-level fall during the late Holocene (Kopp et  al. 2015).
Examples of these far-field locations are the east coast of South America, including
the Bahía Blanca Estuary, West Australia, East China, and South Africa. In these
regions, the coastline initially retreated along with a rising sea level until it reached
a transgressive maximum. After the highstand, the coast prograded as the relative
sea level felt to its present position.

1.3  M
 ajor Coastal Systems Along Temperate Atlantic
South America

The southern coast of Brazil comprises an extensive sequence of barrier islands and
coastal lagoons, which extend to the northern coast of Uruguay. The Merin lagoon,
in northern Uruguay, is a 6000 km2 freshwater body draining into the Dos Patos
lagoon, a 10,360 km2 estuary in southern Brazil, which is connected to the sea by a
single and narrow inlet (Fig. 1.1). The relative sea level in this region reached 2 m
above present at the end of the postglacial marine transgression (about 5000 years
BP) and then started a progressively falling trend to the present position (Angulo
et al. 1999). Barriers along these coastlines comprise a series of four barrier-lagoon
depositional systems, each barrier representing the landward limit of a maximum
transgression event, and with the youngest Holocene barrier located seaward. Along
with the falling sea-level trend after the highstand, vast areas of tidal flats and
marshes developed, which are outstanding elements of the coastal landscape. In the
Dos Patos Estuary, irregularly flooded salt marshes extend over 70 km2. Spartina
alterniflora is the only species in the low marsh, forming dense stands. In the middle
marsh, the southern cordgrass Spartina densiflora is one of the dominant species,
commonly associated with the Cyperaceae Bolboschoenus maritimus. Taller species
like Myrsine parvifolia and Acrostichum danaeifolium gradually displace
S. densiflora at higher elevations (Azevedo 2000). In hypersaline areas, Sarcocornia
ambigua is a common species (Costa and Neves 2006).
The Río de la Plata Estuary is a significant coastal wetland system in northern
coasts of Argentina. The area comprises the Paraná River Delta, located at the head
of the estuary, and extensive coastal plains that formed after the postglacial
transgression (Figs. 1.1 and 1.2). Several authors locate a highstand for this region
about 6000 years BP, when the relative sea level reached around 6 m above present
(Isla 1989; Violante and Parker 2000; Cavallotto et al. 2004). During the regressive
stage, minor deltas developed at the mouths of small rivers, along with beach ridges,
barriers, and spits that began to prograde. Progradation gave place to the formation
of coastal lagoons and tidal flats in the protected areas behind the new formed
barriers and spits. About 3000 years BP, the relative sea level was around 3 m above
1  The Bahía Blanca Estuary in a Regional Context 5

Fig. 1.2  Major coastal features of northern Argentina and geographical location of the Bahía
Blanca Estuary (Map by Walter D. Melo)

present, and a period of drier climate interrupted the progradation of small deltas.
Instead, a sequence of successive estuarine beaches began to form, as sea level
continued to fall. During the final stage of the regression, from about 1600 BP to
present, the increase in the Paraná River discharge induced the formation of the
modern delta and displaced the maximum salinity gradient seaward (Cavallotto
et al. 2005).
A complex mosaic of tidal and nontidal freshwater wetlands occupies the mod-
ern delta, along with the extensive area of former innermost estuarine environments,
including shelly and sandy beach ridges, coastal lagoons, and tidal flats. This
freshwater wetland system covers an area of about 17,500 km2 (Malvárez 1997).
Close to the mouth of the Paraná River, the Lower Delta is a freshwater tidal system
that covers 2700 km2, which is aggrading into Rio de la Plata Estuary at an estimated
rate of 2.64  km2  year−1 (Medina and Codignotto 2013). The Cyperaceae
Schoenoplectus californicus colonizes the new bars, stabilizes sediments, and
establishes monospecific marshes. As these new bars mature, a mixture of forbs like
Hymenachne grumosa, Ludwigia spp., and Senecio bonariensis replace
S. californicus. In a more advanced stage, a surrounding levee begins to develop.
6 P. D. Pratolongo and S. M. Fiori

Mature islands in the frontal zone finally assume the form of a central depression of
marsh, where the Cyperaceae Scirpus giganteus grows in nearly monospecific
stands, and a surrounding levee covered with forest (Kandus and Malvárez 2004).
Tidal influences attenuate upstream, and the pulsing hydrology of the Paraná River
becomes dominant (Neiff 1999). Upstream, nontidal freshwater wetlands
progressively replace tidal freshwater marshes. These wetlands form a large corridor
that extends along 3400  km, from the Paraná River Delta to the Gran Pantanal,
through the Paraná River alluvial valley, and the Paraguay River lowlands (Petean
and Cappato 2005).
The coastal plain southward Río de la Plata developed in the same depositional
environment after the Holocene transgression (Violante and Parker 2000).
Samborombón Bay is an extensive coastal system composed of low-energy
prograding landforms, hosting more than 800  km2 of salt marshes (Isacch et  al.
2006). In the low marsh, S. alterniflora grows in monospecific stands and
S. densiflora marshes establish at higher elevations, along with typical accompanying
species like Apium sellowianum, Limonium brasiliense, and Cortaderia selloana. In
some places, monospecific stands of S. ambigua occur between S. alterniflora and
S. densiflora marshes. Mixed marshes with S. densiflora, S. ambigua, Juncus acutus,
and Distichlis spicata are also common in elevated sites (Cagnoni 1999). Although
there is a general arrangement of communities from the seashore to the uplands, the
plant zonation is not conspicuous. Instead, there is a complex mosaic of different
salt marsh species, intermixed with patches and narrow strips of Celtis tala forest,
associated with shelly ridges parallel to the coast (Vilanova et al. 2006).
Southward, in a gradient of increasing aridity, salt marsh vegetation becomes
sparse. In the Bahía Blanca Estuary, at the northern limit of the Patagonian Region,
most of the intertidal fringe is covered by extensive barren mudflats. Pure stands of
S. alterniflora are commonly restricted to low marshes in the middle reach of the
estuary, but do not appear in the inner zone (Pratolongo et al. 2016). Through the
shallow inner section of the estuary, S. ambigua marshes constrain to elevations
close to the mean high tide level (Pratolongo et al. 2016). Above the elevation of the
mean high tide, irregularly inundated by seawater, land cover is a mosaic of salt
flats, halophytic steppes, and shrubs. Salt flats occupy large areas beyond the limits
of the tidal influence. Sabkha is an Arabic term that describes the extensive, barren,
salt-encrusted, and periodically flooded coastal flats as well as inland salt flats that
form by capillary rise of saline groundwater. Inland and coastal sabkhas (i.e., salt
flats) of a considerable extension also occur further south in San Blas Bay
(Pratolongo et al. 2016). However, despite the vast area that they cover, salt flats in
the region are virtually absent from the literature.
South of the Colorado River extends the Patagonian Region, the largest dry
region of South America. Rivers debouching in the coast of Patagonia typically have
broad alluvial valleys, disproportionate for their little discharge and sediment loads
(Kokot 2004). Cliffed coasts are common along Patagonia and Holocene marine
deposits are widespread, raised at different elevations (Rostami et  al. 2000). San
Sebastián Bay, in Tierra del Fuego Island, is a macrotidal embayment characterized
by a 17 km long gravel spit called Península del Páramo. The spit protects extensive
tidal flats that formed by sedimentation during the regressive phase after the
1  The Bahía Blanca Estuary in a Regional Context 7

Holocene highstand. Chenier ridges and muddy tidal flats developed by rapid
progradation, leaving a sequence that spans about 8 km (Vilas et al. 1999). Vegetation
is virtually absent in the intertidal zone except for the circular mounds of Sarcocornia
magellanica (ex Salicornia ambigua) that colonize the upper portion and scattered
plants of Puccinellia spp. occurring at some locations. At higher elevations, pastures
dominated by Puccinellia magellanica, Puccinellia biflora, and S. magellanica
usually cover the abandoned tidal flats. Lepidophyllum cupressiforme, an endemic
species, appears on the crests of cheniers, forming open scrublands (Collantes and
Faggi 1999).

1.4  T
 he Regional Climate and Vegetation Patterns of Central
East Argentina

The Bahía Blanca Estuary is located in central east Argentina, southern South
America. According to the Köppen-Geiger climate classification (Peel et al. 2007)
(Fig. 1.3), central east Argentina is a transitional zone between temperate oceanic
and cold desert climates and includes zones of humid subtropical and cold semiarid

Fig. 1.3  Climates of central east Argentina according to the Köppen-Geiger climate classification.
(Map by Walter D. Melo)
8 P. D. Pratolongo and S. M. Fiori

climate. This gradient results from the interaction of humid Atlantic air masses from
the northeast with drier and colder masses from the southwest (Burgos and Vidal
1951). Climate in the Southern Hemisphere is controlled by semipermanent pressure
fields of the subtropical South Atlantic and South Pacific anticyclones and temperate
low-pressure cells (Clapperton 1993). Because of the interaction between the
general atmospheric circulation and the Andes Mountain, there are two principal
climatic dominions in southern South America: the Atlantic and the Pacific. A fringe
called the Arid Diagonal crosses southern South America from coastal Peru to
Patagonia and connects zones of arid and semiarid climate. The Arid Diagonal
encompasses several dry lands like Atacama, Monte, and the Patagonian Region,
which isolate the temperate and subtropical forests of Chile and southern Argentina
from other humid regions of South America. In central east Argentina, this climatic
transition is characterized by a strong gradient of decreasing rainfall from the
northeast to the southwest (Bruniard 1982).
Iglesias (1981) proposed a climatic characterization of Argentina. In this classi-
fication, climate in the Pampas (plains of central east Argentina) was typified as
Humid Temperate of the Plains in the east with mean annual rainfall ranging
1200–800 mm, and Subhumid Temperate of the Plains in the west with 800–500 mm.
The mean annual temperature ranges 14–20 °C, there is no dry season, but a decrease
in precipitation occurs at the end of summer and during winter. In northern
Patagonia, Iglesias (1981) recognized the climates Semiarid Temperate of the
Plateau, with mean annual rainfall ranging 200–300 mm, and the Arid Temperate of
the Plateau, with 100–200  mm. Both climates belong to the Arid Diagonal and
define the transition between the Atlantic and the Pacific Dominion. Besides this
well-established pattern in the total annual rainfall, there is a noticeable decreasing
seasonality from northeast to southwest. Throughout the region, mean monthly
precipitation is higher between October and March, and lower values occur between
June and August, but differences between wet and dry months diminish through the
southwest (Celleri et al. 2018).
One of the most distinctive climatic features of central east Argentina is the large
interannual rainfall variability, which has been previously related to El Niño-­
Southern Oscillation (ENSO) (e.g., Rivera and Penalba 2015). ENSO is a naturally
occurring phenomenon that results from interactions between sea surface
temperature anomalies and atmospheric circulation over the tropical Pacific. During
the warm phase of the ENSO cycle (El Niño years), there is a warming in sea surface
temperatures across the central and east-central equatorial Pacific. In central east
Argentina, El Niño years have been associated with positive (above average) rainfall
anomalies. Conversely, La Niña years have been associated with negative (below
average) rainfall anomalies
Celleri et al. (2018) found a significant periodic interannual variability for cen-
tral east Argentina, characterized by an 8-year  cycle, with a strong signal in the
southwestern Pampas. Even though the ENSO cycle explains a large part of this
variability, extremely wet and dry years may also occur under neutral ENSO condi-
tions. These discrepancies may respond to exceptional atmospheric circulations and
unusual water vapor fluxes over the area (Scian et  al. 2006). According to more
1  The Bahía Blanca Estuary in a Regional Context 9

recent studies (Celleri et al. 2018), summer rainfall (November to January) has a
lagged response to the Southern Oscillation Index (SOI), and total values show a
significant negative correlation with the average SOI estimated from August to
October (3-month lag). According to Kayano and Andreoli (2007), ENSO telecon-
nections for the regional rainfall are considerably strong when ENSO and the Pacific
Decadal Oscillation (PDO) are in the same phases, but weak correlations character-
ize those years with oscillations out of phase. While summer rainfall is significantly
associated to El Niño events, the annual totals may not reflect the teleconnection,
and the observed discrepancies may be due to the interaction of ENSO, PDO, and
anomalous circulation patterns.
Vegetation types along central east Argentina are strongly associated with the
regional precipitation gradient, ranging from grasslands in the northeast to
xerophytic shrublands and bushy steppes in the southwest. According to Cabrera
(1978), three different phytogeographic regions can be distinguished within central
east Argentina: Pampas, Espinal, and Monte. The Pampas Region (Fig. 1.4), which
represents the eastern half of the area, is mostly covered by grasslands. Based on its

Fig. 1.4  Phytogeographic regions in central east Argentina and suregions within the  Pampas
Region. (Map by Walter D. Melo)
10 P. D. Pratolongo and S. M. Fiori

microtopography and soil characteristics, this ecoregion can be further divided into
four subregions: Flooding Pampa, Inland Pampa, Southern Pampa, and Rolling
Pampa (Soriano et  al. 1992). Surrounding the Pampas, the Espinal represents a
phytogeographic transition to the Monte Region, and vegetation is composed of a
mixture of grasslands, xerophytic shrublands, and bushy steppes. Finally, the Monte
Region extends through the southwest and is covered by xerophytic shrublands.
Psammophytic and halophytic communities also occur under specific soil conditions
(Cabrera 1978).
Most of the area is dedicated to agriculture and cattle raising. In the Rolling
Pampa, a combination of humid climate, deep and well drained soils allow extensive
agriculture and continuous cropping of soybean, wheat, and maize under rain-fed
regime (INTA 1990). Shallow saline soils, poor drainage, and water erosion impose
limitations to crop production in the Flooding Pampa, and the area is primarily
dedicated to livestock production, mostly sustained by natural pastures (Bilenca
et al. 2012; Frank and Viglizzo 2012). Lands in the rest of central east Argentina are
suitable to cattle and cattle-crop mixed production schemes, including pastures in
rotation with wheat, sunflower, and maize (Hall et al. 1992).

1.5  T
 he Biogeographic Classification
of the Marine Environment

Large Marine Ecosystems (LMEs) are wide areas of ocean space along the Earth’s
continental margins. LMEs represent regional units characterized by distinctive
hydrographic regimens, submarine topography, productivity levels, and populations
related by trophic interactions. Coastal waters off Argentina and Uruguay belong to
the Patagonian Shelf Large Marine Ecosystem (Fig.  1.1), one of the widest
continental shelves in the world, that covers an area of about 1.2  million km2
(Sherman and Duda 1999). Two major surface currents influence the Patagonian
Shelf: the warm, southward-flowing Brazil Current and the Malvinas Current, a
narrow branch of the Antarctic Circumpolar Current that flows northward along the
continental slope of Argentina, and the cold, northward-flowing Malvinas Current
that defines the ecological boundary offshore of this region (Bisbal 1995). The shelf
area is also influenced by low-salinity coastal waters, mainly from the Río de la
Plata Estuary and minor rivers like the Colorado, Negro, and Chubut that also
discharge significant amounts of freshwater and sediments to the shelf
(Heileman 2009).
The Patagonian Shelf is one of the most productive and complex marine systems,
foraging ground for more than 60 species of resident and visiting seabirds (Raya
Rey and Huettmann 2020). The extensive mixing of the Malvinas and the Brazil
Currents results in a conspicuous confluence zone that has many biological, physical,
and meteorological consequences over the entire region. The high productivity of
this marine area is modulated by the presence of several estuarine, tidal, and
1  The Bahía Blanca Estuary in a Regional Context 11

shelf-­break frontal systems that support high concentrations of phytoplankton and


feed trophic webs. Major frontal systems include the Shelf-Break Front, the
Northern Patagonian Front, and the Southern Patagonian Front.
The Shelf-Break Front is a permanent thermohaline frontal system that extends
along the shelf break from the convergence between the Brazil and Malvinas
Currents, south to the Malvinas Islands (Acha et  al. 2004). The Southern
Patagonian Front extends southward from the San Jorge Gulf. Through the south,
the front is characterized by a strong salinity gradient, due to the influence of low-
salinity waters from the Magellan Strait and the Cape Horn Current (Acha et al.
2004). The Northern Patagonian Frontal System includes the Peninsula Valdes
Tidal Front, which forms during spring and summer and represents the boundary
between stratified shelf waters during the warm season and tidally mixed coastal
waters (Acha et al. 2004). Frontal areas commonly have high values of surface
chlorophyll concentration, which can be linked to their primary productivity
(Rivas et al. 2006; Romero et al. 2006). This high productivity supports a diverse
marine biota, with species from warm, temperate, and cold waters. In addition,
coastal areas within the Patagonian Shelf provide reproductive habitats for many
species, some of them of outstanding global ecological, economic, and social
importance, including seabird and marine mammals, as well as fish and inverte-
brates (Miloslavich et al. 2011).
From a zoogeographic perspective, the Patagonian Shelf can be divided into
the Argentine and the Magellanic Provinces (Miloslavich et al. 2011). Provinces
are large areas defined by the presence of particular biotas, with some cohesion
from an evolutionary perspective and endemism, mainly at the level of species
(Spalding et al. 2007). The Argentine Biogeographic Province extends from a
fluctuating northern limit between 30° and 32° S (off Rio Grande do Sul State,
Brazil) to northern Patagonia at 43° S. The Magellanic Province comprises the
complete continental shelf between 55° and 43° S and ascends along the shelf
border up to between 36° and 35° S (Balech and Ehrlich 2008). The transition
between both provinces corresponds to the southernmost extension of the con-
vergence zone between the Malvinas and Brazil Currents. The extreme tem-
perature gradients, intense horizontal and vertical mixing, and changes in
habitat features have a large influence on species distributions and set the tran-
sition between provinces (Boltovskoy 1979; Piola et  al. 2000; Wieters
et al. 2012).
Biogeographic studies considering benthic macroinvertebrate assemblages of
mollusks, echinoderms, and bryozoans over the Argentine Shelf (Bastida
et  al. 1992) confirmed that two major faunal groups can be distinguished, in
agreement with the traditional biogeographic division into Argentine and
Magellanic provinces. Furthermore, these authors suggested that the Magellanic
province could be further divided into the Patagonian and Malvinean districts,
which would cover the warmer inner shelf, and the deeper and colder outer
shelf, respectively.
12 P. D. Pratolongo and S. M. Fiori

1.6  El Rincón Coastal System

Waters of the Bahía Blanca Estuary belong to the Argentine Province, more specifi-
cally to a coastal oceanographic area called El Rincón that extends from 39° S to
41° S, up to a depth of 40 m (Acha et al. 2004). El Rincón is characterized by verti-
cally homogeneous waters due to a shallow bathymetry and strong tidal forcing.
Coastal waters have lower salinities because of the influence of both the Negro and
Colorado Rivers, and a coastal front separates El Rincón from high-salinity shelf
waters (Fig.  1.2). The salinity gradient in the frontal zone is increased by the
influence of high saline waters flowing north from the San Matías Gulf. Bathymetry
and the mean shelf circulation contribute to maintain the front, oriented north-south,
which encloses an area of about 10,000 km2 (Guerrero 1998; Lucas et al. 2005).
El Rincón is regarded as a highly productive area due to the presence of elevated
densities of fish and zooplankton species (Macchi and Acha 1998; Viñas et  al.
2013). It has been ascribed as a very complex oceanographic system, influenced by
variable inputs of continental runoff from the Negro and Colorado Rivers, local
generation of high-salinity cells, and winds that dominate the inner shelf dynamics
(Lucas et al. 2005). The circulation in the area is not yet fully understood, but a
gyre-like pattern has been detected in numerical models, which would favor fish
larval retention and the presence of their predatory fish (Piola and Rivas 1997). The
availability of adequate food, suitable thermal and salinity ranges, and the oceanic
circulation that favors a retention mechanism would facilitate fish larval recruitment
and growth of valuable commercial species (Marrari et  al. 2004; Delgado et  al.
2015). Also, this area is essential for spawning, calving, and breeding of different
emblematic species of the coastal fish complex (Molina and Lopez Cazorla 2011;
Militelli et al. 2013).
Rich oceanic conditions bring nutrients, plankton, and fish into Argentine waters,
but unsustainable practices have led to the collapse of several commercial fisheries.
However, there is an increasing effort directed to improve the conservation status of
the Argentine Sea by creating protected areas and developing strategies for the
ecosystem-based management of fisheries. The National System of Marine
Protected Areas (MPAs) was created in 2014, to protect and conserve marine spaces
that are representative of habitats and ecosystems under environmental policy
objectives. Within this frame, El Rincón Marine National Park and National Marine
Reserve has been proposed as a protected area, along with another five marine
regions, because of their great ecological significance. In the specific case of El
Rincón, the area has been proposed as MPA based on the presence of many
commercially important fish species, as well as feeding resources for top predatory
birds and mammals. The area also entails a large biogeographic relevance, since it
connects two coastal protected areas (Reserva Natural de Uso Múltiple Bahía San
Blas and Reserva Natural de Uso Múltiple Bahía Blanca, Falsa y Verde). The high
mortality of keystone predators by incidental catch and the poor condition of fishing
stocks have been pointed as major threats for this area (Falabella 2014; DNC/
APN 2017).
1  The Bahía Blanca Estuary in a Regional Context 13

References

Acha EM, Mianzan HW, Guerrero R et al (2004) Marine fronts at the continental shelves of austral
South America. J Mar Syst 44:83–105
Angulo RJ, Giannini PC, Suguio K et al (1999) Relative sea-level changes in the last 5500 years
in southern Brazil (Laguna–Imbituba region, Santa Catarina State) based on vermetid 14C ages.
Mar Geol 159:323–339
Azevedo AMG (2000) Hábitats, associações vegetais e fenologia das plantas das marismas da
Ilha da Pólvora, Estuário da Lagoa dos Patos (RS, Brasil). Doctoral dissertation, Fundação
Universidade Federal do Rio Grande
Balech E, Ehrlich MD (2008) Esquema biogeográfico del mar Argentino. Rev Investig Desarro
Pesq 19:45–75
Bastida R, Roux A, Martínez DE (1992) Benthic communities of the argentine continental-shelf.
Oceanol Acta 15:687–698
Bilenca D, Codesido M, González Fischer C et  al (2012) Impactos de la transformación agro-
pecuaria sobre la biodiversidad en la provincia de Buenos Aires. Rev Mus Argent Cienc Nat
14:189–198
Bisbal GA (1995) The Southeast South American shelf large marine ecosystem: evolution and
components. Mar Policy 19:21–38
Boltovskoy D (1979) Zooplankton of the South-western Atlantic. S Afr J Sci 75:541–544
Bruniard E (1982) La Diagonal Árida Argentina: un límite climático real. Rev Geogr 95:5–20
Burgos JJ, Vidal AL (1951) The climates of the Argentine Republic according to the new
Thornthwaite classification. Ann Assoc Am Geogr 41:237–263
Cabrera AL (1978) Flora de la provincia de Jujuy, República Argentina: Compositae. INTA,
Buenos Aires
Cagnoni M (1999) Espartillares de la costa bonaerense de la República Argentina. Un caso de
humedales costeros. In Malvárez AI (ed) Tópicos sobre Humedeles Subtropicales y templados
de Sudamérica. MAB-UNESCO, Montevideo
Cavallotto JL, Violante RA, Parker G (2004) Sea-level fluctuations during the last 8600 years in the
de la Plata river (Argentina). Quat Int 114:155–165
Cavallotto JL, Violante RA, Colombo F (2005) Evolución y cambios ambientales de la llanura
costera de la cabecera del río de la Plata. Rev Asoc Geol Argent 60:353–367
Celleri C, Zapperi G, González Trilla G et al (2018) Spatial and temporal patterns of rainfall vari-
ability and its relationship with land surface phenology in central east Argentina. Int J Climatol
38:3963–3975
Clapperton CM (1993) Quaternary geology and geomorphology of South America. Elsevier,
Amsterdam
Clark JA, Farrell WE, Peltier WR (1978) Global changes in postglacial sea level: a numerical
calculation. Q Rev 9:265–287
Collantes MB, Faggi AM (1999) Los humedales del sur de Sudamérica. In: Malvárez AI (ed)
Tópicos sobre humedales subtropicales y templados de Sudamérica. NESCO, Montevideo,
pp 15–25
Costa CSB, Neves LSD (2006) Respostas cromáticas de Salicornia gaudichaudiana Mog.
(Chenopodiaceae) a diferentes níveis de radiação UV-B e salinidade. Atlantica 28:25–31
Delgado AL, Loisel H, Jamet C et al (2015) Seasonal and inter-annual analysis of chlorophyll-a
and inherent optical properties from satellite observations in the inner and mid-shelves of the
south of Buenos Aires Province (Argentina). Remote Sens 7:11821–11847
DNC/APN (Dirección Nacional de Conservación/Administración de Parques Nacionales) (2017)
Sistema Nacional de Áreas Marinas Protegidas (Ley N° 27.037): Justificación técnica para la
creación de seis áreas Marinas Protegidas. Dirección Nacional de Conservación, Ministerio de
Ambiente y Desarrollo Sustentable, p 52
14 P. D. Pratolongo and S. M. Fiori

Falabella V (2014) Identificación de áreas de alto valor de conservación como potenciales


áreas marinas protegidas. Informe elaborado durante la fase preparatoria del Proyecto GEF
5112-FAO-Secretaria de Ambiente y Desarrollo Sustentable, p 97
Fleming K, Johnston P, Zwartz D et al (1998) Refining the eustatic sea-level curve since the last
glacial maximum using far and intermediate-field sites. Earth Planet Sci Lett 163:327–342
Frank FC, Viglizzo EF (2012) Water use in rain-fed farming at different scales in the Pampas of
Argentina. Agric Syst 109:35–42
Guerrero RA (1998) Oceanografía física del estuario del Río de la Plata y el sistema costero de
El Rincón. Noviembre, 1994. In: Lasta CA (ed) Resultados de una campaña de evaluación de
recursos demersales costeros de la provincia de Buenos Aires y del litoral uruguayo, vol 21.
INIDEP Informe Técnico, pp 29–54
Hall AJ, Rebella CM, Ghersa C et al (1992) Field-crop systems of the Pampas. In: Pearson CJ
(ed) Field crop ecosystems, Serie: Ecosystems of the World. Elsevier, Amsterdam, pp 413–450
Heileman S (2009) XVI-55 Patagonian Shelf LME in the UNEP large marine ecosystem report:
a perspective on changing conditions in LMES of the world’s regional seas. In: Sherman K,
Hempel G (eds) UNEP regional seas report and studies N°182. United Nations Environment
Programme, Nairobi, pp 735–746
Iglesias de Cuello A (1981) Tipos de clima. In: Chiozza E, Figueira E (eds) Atlas Total la República
Argentina. Centro Editor de América Latina, Buenos Aires, pp 193–200
INTA (1990) Atlas de suelos de la República Argentina. Ediciones INTA, Buenos Aires
Isla FI (1989) Holocene sea-level fluctuation in the southern hemisphere. Quat Sci Rev 8:359–368
Isacch JP, Costa CSB, Rodriguez-Gallego L, Conde D, et al (2006) Distribution of saltmarsh plant
communities associated with environmental factors along a latitudinal gradient on the south-­
west Atlantic coast. J Biogeogr 33:888–900
Kandus P, Malvárez AI (2004) Vegetation patterns and change analysis in the lower delta islands
of the Paraná River (Argentina). Wetlands, 24:620–632
Kayano MT, Andreoli RV (2007) Relations of South American summer rainfall interannual varia-
tions with the Pacific decadal oscillation. Int J Climatol 27:531–540
Khan NS, Vane CH, Horton BP (2015) Stable carbon isotope and C/N geochemistry of coastal
wetland sediments as a sea-level indicator. In Shennan I, Long AJ, Horton BP (eds) Handb
Sea-Level Res. John Wiley & Sons, Ltd, Oxford, UK
Kokot RR (2004) Erosión en la costa patagónica por cambio climático. Rev Asoc Geol Argent
59:715–726
Kopp RE, Horton BP, Kemp AC et al (2015) Past and future sea-level rise along the coast of North
Carolina, USA. Clim Chang 132:693–707
Lucas A, Guerrero RA, Mianzán H et al (2005) Coastal oceanographic regimes of the Northern
Argentine Continental Shelf (34–43°S). Estuar Coast Shelf Sci 65:405–420
Macchi GJ, Acha ME (1998) Aspectos reproductivos de las principales especies de peces en la
Zona Común de Pesca Argentino-Uruguaya y en El Rincón. In: Lasta CA (ed) Resultados de
una campaña de evaluación de recursos demersales costeros de la provincia de Buenos Aires y
del litoral uruguayo. Noviembre 1994, vol 21. INIDEP Inf. Téc, pp 67–89
Malvárez AI (1997) Las comunidades vegetales del Delta del río Paraná. Su relación con facto-
res ambientales y patrones del paisaje. PhD thesis. Facultad de Ciencias Exactas y Naturales,
Universidad de Buenos Aires
Marrari M, Viñas MD, Martos P et al (2004) Spatial patterns of mesozooplankton distribution in
the Southwestern Atlantic Ocean (34°–41°S) during austral spring: relationship with the hydro-
graphic conditions. J Mar Sci 61:667–679
Medina R, Codignotto J (2013) Evolución del delta del río Paraná y su posible vinculación con el
calentamiento global. Rev Mus Argent Cienc Nat 15:191–200
Militelli MI, Macchi GJ, Rodrigues KA (2013) Comparative reproductive biology of Sciaenidae
family species in the Río de la Plata and Buenos Aires Coastal Zone, Argentina. J Mar Biol
Assoc UK 93:413–423
1  The Bahía Blanca Estuary in a Regional Context 15

Miloslavich P, Klein E, Díaz JM et al (2011) Marine biodiversity in the Atlantic and Pacific coasts
of South America: knowledge and gaps. PLoS One 6(1):e14631. https://doi.org/10.1371/jour-
nal.pone.0014631
Mitrovica JX, Milne GA (2002) On the origin of late Holocene sea-level highstands within equato-
rial ocean basins. Quat Sci Rev 21:2179–2190
Molina JM, Lopez Cazorla A (2011) Trophic ecology of Mustelus schmitti in a nursery area of
northern Patagonia. J Sea Res 65:381–389
Neiff JJ (1999) El régimen de pulsos en ríos y grandes humedales de Sudamérica. In Malvarez AI,
Kandus P (eds) Tópicos sobre Humedales Sudamericanos ORCYT-MAB (UNESCO), Buenos
Aires, Argentina
Peel MC, Finlayson BL, McMahon TA (2007) Updated world map of the Köppen-Geiger climate
classification. Hydrol Earth Syst Sci 11:1633–1644
Peltier WR, (2002) On eustatic sea level history: Last Glacial Maximum to Holocene. Quat Sci
Rev 21:377–396
Peteán J, Cappato J (2005) Humedales fluviales de América del Sur. Hacia un manejo Sustentable.
Imprenta Lux SA, Santa Fé, Argentina
Piola AR, Rivas AL (1997) Corrientes en la plataforma continental. In: Boschi EE (ed) El Mar
Argentino y Sus Recursos Pesqueros 1: Antecedentes históricos de las exploraciones en el mar
y las características ambientales. Instituto Nacional de Investigación y Desarrollo Pesquero,
Mar del Plata, pp 119–132
Piola AR, Campos EJD, Möller OO Jr et  al (2000) Subtropical shelf front off eastern South
America. J Geophys Res Oceans 105(C3):6565–6578
Pratolongo PD, Piovan MJ, Cuadrado DG (2016) Coastal environments in the Bahía Blanca
estuary, Argentina. In: Khan M, Boër B, Ȫzturk M et al (eds) Sabkha ecosystems, Tasks for
Vegetation Science, vol 48. Springer, Cham, pp 205–224
Raya Rey A, Huettmann F (2020) Telecoupling analysis of the Patagonian Shelf: a new approach
to study global seabird-fisheries interactions to achieve sustainability. J Nat Conserv 53:125748
Rivas AL, Dogliotti AI, Gagliardini DA (2006) Seasonal variability in satellite-measured surface
chlorophyll in the Patagonian Shelf. Cont Shelf Res 26:703–720
Rivera J, Penalba O (2015) El Niño/La Niña events as a tool for regional drought monitoring
in Southern South America. In: Andreu J, Solera A, Paredes-Arquiola J et al (eds) Drought:
research and science-policy interfacing. CRC Press, Valencia, pp 293–299
Romero SI, Piola AR, Charo M et  al (2006) Chlorophyll-a variability off Patagonia based on
SeaWiFS data. J Geophys Res 111:C05021
Rostami K, Peltier WR, Mangini A (2000) Quaternary marine terraces, sea-level changes and
uplift history of Patagonia, Argentina: comparisons with predictions of the ICE-4G (VM2)
model of the global process of glacial isostatic adjustment. Quat Sci Rev 19:1495–1525
Scian B, Labraga JC, Reimers W et al (2006) Characteristics of large-scale atmospheric circula-
tion related to extreme monthly rainfall anomalies in the Pampa region, Argentina, under non-­
ENSO conditions. Theor Appl Climatol 85:89–106
Sherman K, Duda AM (1999) Large marine ecosystems: an emerging paradigm for fishery sustain-
ability. Fisheries 24:15–26
Soriano A, León RJC, Sala OE et  al (1992) Río de la Plata grasslands. In: Coupland RT (ed)
Natural grasslands: introduction and western hemisphere, Serie: Ecosystems of the world.
Elsevier, New York
Spalding MD, Fox HE, Allen GR et al (2007) Marine ecoregions of the world: a bioregionalization
of coastal and shelf areas. Bioscience 57:573–583
Viñas MD, Marrari M, Di Mauro RP et al (2013) El zooplancton del hábitat reproductivo de la
población bonaerense de anchoíta (Engraulis anchoita), con especial énfasis en crustáceos.
Rev Investig Desarr Pesq 23:125–144
Violante RA, Parker G (2000) El Holoceno en las regiones marinas y costeras del nordeste de
Buenos Aires. Rev Asoc Geol Argent 55:337–351
16 P. D. Pratolongo and S. M. Fiori

Vilanova I, Prieto AR, Stutz S (2006) Historia de la vegetación en relación con la evolución
geomorfológica de las llanuras costeras del este de la provincia de Buenos Aires durante el
Holoceno. Ameghiniana 43:147–159
Vilas F, Arche A, Ferrero M, Isla F (1999) Subantarctic macrotidal flats, cheniers and beaches in
San Sebastián Bay, Tierra del Fuego, Argentina. Mar Geol 160:301–326
Wieters EA, McQuaid C, Palomo G et al (2012) Biogeographical boundaries, functional group
structure and diversity of Rocky Shore communities along the Argentinean coast. PLoS One
7(11):e49725. https://doi.org/10.1371/journal.pone.0049725
Chapter 2
Geography of the Bahía Blanca Estuary

Walter Daniel Melo

2.1  Introduction

From the physiographic point of view, the Bahía Blanca Estuary and associated
wetlands represent one of the most remarkable features within the environmental
diversity that characterizes the Argentine coast. The system encompasses a wide
variety of landscape elements like sandy beaches, coastal dunes, tidal plains,
marshes, islands, tidal channels, and extensive salt flats, inherited from the trans-
gressive past.
Bahía Blanca is located at an inflection point in the coast of the Buenos Aires
Province, central east Argentina, where the coastline turns from the predominant
north-south direction to the east-west (Fig. 1.1; Chap. 1). The coastal system is a
2,500 km2 embayment that extends from the cliffs at “Baliza Monte Hermoso” (an
old beacon northwestward) to Laberinto Point on the southern shore of Brightman
Creek (Fig. 2.1). The estuarine system comprises three major first-order channels,
almost parallelly oriented northwest to southeast: Principal Channel, Bahía Falsa,
and Bahía Verde. The Principal Channel, the main navigation channel, extends
along 120 km, naturally reaching around 10 m depth before dredging. Besides main
channels, the estuarine system is composed of more than 128,000 ha of tidal mud-
flats and salt marshes, dissected by countless tidal channels of smaller order and
varying sizes (Perillo and Piccolo 1999a, b). The southern limit of the system is
defined by the Colorado River, the only permanent watercourse with a substantial
discharge (Melo and Limbozzi 2008). Within the Bahía Blanca Estuary, continental
runoff restricts to the northern shore of Principal Channel (Píccolo and Perillo

W. D. Melo ()
Instituto Argentino de Oceanografía, IADO (Universidad Nacional del Sur-CONICET),
Bahía Blanca, Argentina
Departamento de Geografía y Turismo, Universidad Nacional del Sur,
Bahía Blanca, Argentina
e-mail: wdmelo@iado-conicet.gob.ar

© Springer Nature Switzerland AG 2021 17


S. M. Fiori, P. D. Pratolongo (eds.), The Bahía Blanca Estuary,
https://doi.org/10.1007/978-3-030-66486-2_2
18 W. D. Melo

Fig. 2.1  Bahía Blanca Estuary. District, cities, and main coastal environments. (Map by Walter
D. Melo)

1990). Two permanent small rivers represent the main freshwater input to the estu-
ary (Fig. 2.1). The Sauce Chico River has a hydrographic basin that covers 1450 km2
and an average flow of 1,807 m3/s. The Napostá Grande Stream has a hydrographic
basin of 1100 km2 and an average flow rate of 0.8 m3/s (Piccolo et al. 2008). Besides
these two permanent sources, there are intermittent streams that drain rainfall excess
during wet periods. Main intermittent streams are those in the group named
Saladillo, which encompasses three different streams (Saladillo or Dulce, Saladillo
de Lazaga, and Saladillo de García) and the Maldonado Channel, built in 1950 to
manage excess rainfall on the Bahía Blanca City and overflows of the Napostá
Grande River (Fig. 2.2).
Large salt flats inherited from the transgressive past extend through the north-
west of Principal Channel, becoming dominant landscape features beyond tidal
influences. These gently sloped plains, currently unconnected to tides, performed as
intertidal mudflats under a higher relative sea level during the Holocene. Nowadays,
2  Geography of the Bahía Blanca Estuary 19

Fig. 2.2  Coastal uses and environments in Bahía Blanca District. (Map by Walter D. Melo)

these extended salt flats function as transitional areas between the marine and con-
tinental domains. “La Vidriera” is the largest salt flat in the area (Fig. 2.1), extending
along 8 km through the northwest, following the axis Chasicó Lagoon-La Vidriera-­
Principal Channel. This alignment defines a morphologic transition that separates
North and South Continental Regions. The northern shore of Principal Channel,
eastward from La Vidriera Salt Flat, hosts a large environmental variability, includ-
ing minor estuarine areas at the mouth of the Sauce Chico River, Napostá Grande
Stream, and small stream Saladillo de García. Through the mouth and after a salt
flat area, the small village Villa Del Mar, Puerto Belgrano, and the Puerto Rosales
are located in the transition between tide- and wave-dominated systems (Melo and
Carbone 2012) (Fig.  2.3). East from Puerto Rosales, fine-grained mudflats give
place to sandy beaches that develop at the edge of coastal dunes, like those in Tejada
Point. Sandy beaches extend eastward up to the cliffs at Baliza Monte Hermoso.
The total area floodable by extraordinary tides in storm situations totals approxi-
mately 128,000  ha, of which 28,000  ha would correspond to high tidal flats, the
regular tidal flats cover about 90,000  ha, and the marsh areas about 18,000  ha
(Fig. 2.1).
Regarding island morphology, two different zones can be distinguished. North of
the Bahía Falsa Channel, tidal flats of considerable extension cover most of the
estuarine area, and plentiful isles of variable sizes form disaggregated clusters
(Melo et al. 2003). Despite its name, Trinidad Island is actually a group of 46 islands
covering 4,300 ha of emerged land (Fig. 2.1). The surface area of the largest island
20 W. D. Melo

Fig. 2.3  Main coastal uses and environments in Coronel Rosales District. (Map  by Walter
D. Melo)

is 1,340 ha, but smaller isles cover less than half a hectare. Intertidal environments
cover most of Trinidad Islands, including about 15,000 ha of tidal flats, 3,300 ha of
high tidal flats, and 1000  ha of marshes. Through the southeast, under a wave-­
dominated regime, sand flats develop instead of tidal flats (Carbone and Melo 2016).
Similarly, the group Bermejo Island (2,200 ha of emerged land) comprises 52 sepa-
rated isles, the largest island covering about 470  ha, surrounded by more than
8,900 ha of tidal flats and 700 ha of marshes. Finally, the group Embudo Islands is
composed of about 40 islands covering 675 ha of emerged land. About 250 addi-
tional islands and islets form smaller groups, covering 6,800 ha along with small
channels and tidal flats. The emerged land of these islands presents highly saline
soils, and vegetation restricts to halophytic steppes and shrubs.
South of the Bahía Falsa Channel, single larger isles conform geomorphic units
instead of groups. The higher elevations and sandy soils allow the formation of sand
dunes, and sand flats commonly replace mudflats in the intertidal zone. Major isles
2  Geography of the Bahía Blanca Estuary 21

are Wood (also called Monte) and Ariadna, both aligned between the Bahía Falsa
and Bahía Verde Channels, and Verde Peninsula, formerly an island that got attached
to the continent once the relative sea level dropped after the Holocene highstand.
Verde Peninsula covers about 9,000 ha, and extensive dune fields develop near the
eastern coast.
This complex geomorphic setting was shaped by the recent geological history of
the Bahía Blanca Estuary, during the late Pleistocene and Holocene (Melo 2004).
There is a paleomorphologic relationship between the estuary and the adjacent con-
tinental environment. Bahía Blanca Estuary is the reception basin of a drainage
basin having a surface of the order of 3000 km2. However, in the current drainages
of its hydrographic basin, no large watercourses are observed that are directly
related to the geomorphological succession that determined the origin of the estu-
ary. The existence of geomorphologic units such as the Colorado River, dune fields,
sand strandflats, and a series of the valleys and axes of depressions areas in the SW
of La Pampa Province (Fig. 1.2; Chap. 1) they are connected since the Pleistocene
to the genesis and evolution of the estuary. The latter are aligned and related to pre-
vious faulting which may be related directly to the origin of the estuary. Part of these
alignments are below mean sea level (−20 m) and formed by canyons and valleys
containing dunes of varied shapes. These valleys have been arranged in groups. The
first group starts on the NW of Lihué Calel Hills and formed two alignments having
a NW-SE and E-W trends. Both follow the Chasicó-Salinas Chicas depression, con-
tinuing into the La Vidriera Salt Flat and the Principal Channel. The second group
is 120 km in length formed by the alignment passing through the Blanca Grande and
Callaquéo lagoons, ending in the Chasicó-Salinas Chicas depression (Fig. 1.2;
Chap. 1). The alignments in the third group are mostly trending in the E-W direction
starting at the confluence between the Colorado and Salado Rivers ending in the
Chasicó-Salinas Chicas depression. This group is independent of the other two, and
its alignments are related to the Colorado River directed toward the southern portion
of the estuary (González Uriarte 1984). The most significant axis has a length of
140 km and relates the Anzoátegui and La Gotera Salt Lagoons (Fig. 1.2; Chap. 1)
with Bahía Falsa and Bahía Verde. The fluvial discharge along these groups pro-
vided the hydric connection between the continent and the littoral environments
developing the initial stages of the formation of the Bahía Blanca Estuary.
The area presently occupied by the estuary was subaerial. Later on, the climate
changed to warmer and more humid conditions, and the mean sea level increased to
approximately the present conditions (Aguirre 1995). The drain introduced a large
sediment load coming from the three first groups, developing a delta. Major channels
formed from old continental drains that followed the three different drainage axes,
ending in the present area of the Bahía Blanca Estuary. Evidence of these former
drainages persists in the major estuarine channels aligned with continental depres-
sions. According to Aguirre (1995), a sharp change in climate occurred during the
early Holocene, establishing a period of warmer and more humid conditions, before
the transgressive maximum. Additional evidences of humid climate during the trans-
gressive phase were found in the coastal zone that extends up to 100 km west from
Bahía Blanca. In this area, continental lagoons showed a rise in water levels and
22 W. D. Melo

higher abundances of freshwater diatoms, fishes, and vascular plants (Gutiérrez-Téllez


and Schillizzi 2002). Under humid conditions during the early Holocene, rivers may
have introduced large amounts of fine sediments (mostly loessic pampean silts) to the
estuarine area, using these previous topographic depressions.
According to Spalletti and Isla (2003), during the mid-Holocene transgressive
stage, the deltaic deposits were partially covered with medium sand and biogenic
remains, as well as compacted silty clay sediments in shallower zones. The advance
of marine sediments as well as the rework and redistribution of fluvial-estuarine
sediments filled paleochannels, until the establishment of a large estuarine-marine
environment. Because of the higher sea level, the delta was fully covered by the sea
reaching the continent through the La Vidriera Salt Flat (Melo 2004). Humid condi-
tions may have continued until about 5000  years BP, when the establishment of
psammophytic and halophytic communities indicates a climatic change toward sub-
humid to dry conditions (Prieto 1996). The increase of the mean sea level reached
values of 5–12 m above present conditions (González Uriarte 1984). These values
are related to similar studies made by Cavallotto et al. (1999) for the Río de la Plata
and Aliotta et al. (2000) and Spagnuolo et al. (2000) for Arroyo Parejas. Toward
3000 years BP, climate changed to temperate-dry, similar to the present conditions
(Aguirre 1995), which determined the disappearance of the rivers being replaced by
Aeolic environments. Consequently, the Colorado River progressively migrated
southward. In this migration, the river occupied and later released the axis
Anzoátegui Salt Lake-La Gotera-Bahía Falsa.
Under a falling sea level and diminished river discharge, the present islands may
have emerged, and large continental drainages disappeared, or transformed in the
tidal channels Bahía Verde, Bahía Falsa, and Principal Channel. About 2000 years
BP, the connection between the continental and transitional environments (La
Vidriera) was transformed in a coastal lagoon.
As sea level continued to fall, low elevation isles connected and reshaped to form
extensive mudflats. Coastal lagoons got progressively isolated from the marine
influence and transformed in salt flats, and the increased aridity enhanced wind
transport of fine sediments, allowing the formation of large dune fields that pres-
ently cover southern isles. Wind action transported eastward the original sand
deposits from southeast of La Pampa Province, forming extensive dune fields. This
material also covered the Verde, Ariadna, and Monte Island, located to the south of
the estuary. González and Weiler (1983) and Spalletti and Isla (2003) found at the
mouth of the Colorado Nuevo River, to the south of the indicated islands, the pres-
ence of a lobular prograding shelf delta originated by marine reworking.

2.2  Occupation Processes and Land Use in Bahía Blanca

In 1520 the european explorers first entered at the Principal Channel, and Brightman
Cove and the “Cabeza de Buey” Channel were used as a refuge by the navigators.
However, studies on the survey of canals and islands would begin almost 300 years
2  Geography of the Bahía Blanca Estuary 23

later (Melo and Piccolo 2006). The complex environmental situations articulated
with historical processes led to a late survey of the coast, of the islands, and of the
plains being in 1804 when the first geographical recognitions are made. The contri-
bution of fresh water and the Principal Channel presence were the main reasons that
allowed Bahía Blanca City to be founded in 1828, which, as a small military garri-
son and a smaller civilian population, survived almost alone until 1880 with the
effective occupation of Patagonia. Until those years, its port lacked relevant infra-
structure, and in 1884 with the presence of the railroad at “Ingeniero White” town,
the port works began to be significant (Fig. 2.2). Toward the end of that century, in
1898, the creation of the Naval Base “Puerto Belgrano” resulted in the establish-
ment of a small population and which were the origins of the “Punta Alta”  city
(Fig. 2.3). At the beginning of the twentieth century, it already established the con-
ditions for the consolidation of the railway activity and the establishment of local
industries, creating Puerto Galván (1902), Puerto Cuatreros (1903), and Puerto
Rosales (1912) (Coleman 1948). Land uses such as way urban, industrial and rail-
way occupation was carried out only along the north coast of the Principal Channel,
between Puerto Cuatreros and Puerto Rosales. In this strip of the channel, the per-
manent watercourses formed former alluvial cones where its elevated terrain did not
give rise to the formation of extensive tidal plains and thus facilitates the continental
proximity to the greater depths of the Principal Channel, which allowed port filing.
The territorial organization of the Bahía Blanca Estuary area and its wetlands on
the coast, from the political and administrative aspect, three different districts inte-
grate the area: (i) Municipality of Villarino; (ii) Municipality of Bahía Blanca,
including the localities of General Daniel Cerri and Ingeniero White (hereinafter
referred to as Cerri and White, respectively); and (iii) the Municipality of Coronel
de Marina Leonardo Rosales which includes Villa del Mar and Pehuen Co.

2.2.1  Villarino District

This district has a coastline extension of approximately 250 km (Fig. 2.1), of which


about 200 km is of jurisdiction on the estuarial coastal front, which develops from
north to south between the 1st Arm of the Sauce Chico River and Laberinto Point,
at south of Brightman Cove (Figs. 2.1 and 2.2). Although this district has the great-
est extension over the wetlands of the Bahía Blanca, it does not have any localities
or minor settlements on the coast and the activity is rural only. Only one mention,
“Villarino Viejo” Place, which is about 2 km from the head of the Principal Channel
is a small sector of farms next to the aforementioned arm and that initially func-
tioned as the district’s header (Fig. 2.2). However, its activity of coastal use is by
only sports anglres that to arrive at the coast are used neighborhood roads. In pio-
neer times, there were smaller ports or berths such as “Villarino Viejo,” “Puerto
Coloma,” “San Miguel” Farm, and “El Algarrobo” farm were used by rural estab-
lishments on the continent. Today, they are disjointed, and they are used as sport
fishing activity points for private use. The Villarino District dominance reaches
24 W. D. Melo

Verde Peninsula. In its origins of productive occupation, the “Isla Verde” Farm of
9000 ha worked. Since 1947, the occupation belongs to the Argentine Navy, where
a limited sector is for military use, but its use remains predominantly rural (Melo
et  al. 2005). In turn, the isthmus of the peninsula belongs to a private farm. In
“Bahía Verde” Channel and Brightman Cove, the drainage channels of the “Corfo
Rio Colorado” corporation area have their mouth at the time of irrigation (Fig. 2.1).

2.2.2  Bahía Blanca District

The Bahía Blanca jurisdiction covers the entire coastal front of the district, from the
1st Arm of the Sauce Chico River to the cadastral boundary with the Coronel
Rosales District, occupying about 35  km from the north veril of the Principal
Channel (Fig.  2.1). Basically, three sectors are identified on the Bahía Blanca
District coastal front. To the west of said front, the space of 5000 m that covers from
the 1st Arm of the “Sauce Chico” River to the “General Cerri” Town the are lands
in which spaces for rural use are interspersed with spaces without uses. In this sec-
tor, the presence of Puerto Cuatreros stands out, next to the 2nd Arm of the Sauce
Chico River. It is a great breakwater, which is reached by a consolidated road on the
salt flats the tidal plains. This functioned as an export point for the former
“Sansinena” Industrial Meat Plant (then CAP Corporation, today abandoned)
(CGPBB 2020) and which determined the growth of the town of “General Cerri.”
This port today functions as a recreation, environmental observation, and fishing
club (Fig. 2.2). In this west sector, the smaller sector rural use, 400 ha, is interrupted
by the installation of the urban effluent discharge treatment plant and one waste
treatment plant of the meat industry. Between the “Maldonado” Spa and the mouth
of the Napostá Grande Stream, the greatest urban and industrial pressure develops
over the entire coastal area (Melo and Limbozzi 2008) (Fig. 2.4). The mentioned
spa is the only coastal summer recreation center in the district and operates on the
east veril of the “Maldonado” Urban Channel. To the southeast of the same, at the
estuary in the east border of Maldonado Channel, there is a coastal front project that
contains the Municipal Natural Reserve and that the “Belisario Roldán” Municipal
Dump was formerly operating. Neighboring this reserve is the most important port
complex on the Argentine coast that covers 9  km between Puerto Galván and
Ingeniero White Town (Fig. 2.4).

2.2.2.1  Puerto Galván

It was founded as a cereal terminal in 1905 by the company “Buenos Aires al


Pacífico” (BAP), a former British railroad company (CGPBB 2020). At present, it
has diversified its operational activity toward oil and gas activity. Among its original
facilities is the specialized terminal for the management of cereals and by-products
by the companies “Oleaginosa Moreno Hnos. S.A.” and the incorporation of the
2  Geography of the Bahía Blanca Estuary 25

Fig. 2.4  Coastal uses and environments in main pressure area of Bahía Blanca District. (Map by
Walter D. Melo)

companies “Los Grobo Inversora S.A.” and “Louis Dreyfus Commodities (LDC).”
At the west end of Puerto Galván is the Flammables dock (CGPBB 2020) (Fig. 2.4).
It is composed of two berthing sites, site No. 1, destined to the operation of liquid
fuels by oil companies and caustic soda produced by the firm “Indupa S.A.,” and
site No. 2, assigned to the operation with gaseous and petrochemical products by the
companies of the Bahía Blanca Petrochemical Pole and “Transportadora de Gas del
Sur” (TGS). Also, a floating petrochemical plant (“Polisur S.A.”) dedicated to the
production of high- and low-density polyethylene is located. To the west of the main
pier is a sector of 6 hectares dedicated to general cargoes. In that same sector, the
“Puerto Galván” Yacht Club (CNPG) has its facilities (Fig. 2.4).

2.2.2.2  Petrochemical Pole

Between Puerto Galván and the Port of Ingeniero White, the largest petrochemical
complex in South America is located, which together with the Industrial Park occu-
pies some 600 hectares (Figs. 2.2 and 2.4). To the north of the sector is the “Ricardo
26 W. D. Melo

Eliçabe” Oil Refinery, founded in 1922 was the first plant of the industrial complex
whose main activity is the refining of hydrocarbons. The Petrochemical Industrial
Pole represents more than a half of the industrial production of Buenos Aires
Province (London 2017), it was inaugurated in 1986 (Ferrera 2003) and constitutes
an integrated production chain, basically generating three types of industries:
Oil industry: with an installed capacity of 4 million tons per year. Products: eth-
ane, gasoline, LPG, fuel oil, gas oil, asphalt, kerosene.
Petrochemical industry: with an installed capacity of 3.4  million per year.
Products: ethylene, VCM, PVC, polyethylene, urea, pure ammonia.
Chemical industry: with an installed capacity of 350 thousand tons per year.
Products: chlorine, caustic soda, oxygen, nitrogen. In the north of Puerto Galván are
plants of the AXION Energy, YPF, and Shell oil companies and a TGS gas plant.
Since the founding of the Petrochemical Pole of Bahía Blanca, international
business groups have a high dynamism about changes in their composition, group-
ing, or ownership of plants. At present, the most important companies that make up
the Bahía Blanca Petrochemical Pole are “PBB-Polisur,” “Unipar-Indupa,” “Air
Liquide,” “Mega” Company, and “Profertil.”
Between Puerto Galván and Port of Ingeniero White, the Company “Mega” and
“Profertil” plants are located on the coastal front and also belong to the group of
port companies. “Mega” is dedicated to the production of natural gas to get ethane,
propane, butane, and gasoline, while “Profertil,” being the largest fertilizer com-
pany in South America, produces urea and ammonia (Fig. 2.4).

2.2.2.3  Port of Ingeniero White

The Port of Ingeniero White, originally called “Puerto Commercial Bahía Blanca,”
was founded in 1884 as the railroad head of the “Ferrocarril del Sud,” a company
with capitals of British origin. The port consisted of a series of iron docks, of a
structure called transparent, that crossed the tidal plains to the edge of the Principal
Channel (CGPBB 2020). In its origins, the so-called fruits of the country were
exported, consisting of cereals, leather, and wool. Subsequently, fruits and fishing
products were added. Since the end of the nineteenth century and throughout the
twentieth century, artisanal fishing was carried out in the estuary canals. The large
number of dock workers, railways workers, and fishermen also determined the ori-
gin of the Ingeniero White Town (Fig. 2.4). Between 1948 and 1993, the National
State administered the port through the “Administración General de Puertos” (AGP).
In 1993, the “Consorcio General de Puertos de Bahía Blanca” (CGPBB), an admin-
istrative entity composed of representatives of private port companies, the provin-
cial and local state, replaced this administration.
Its access navigation channel, the Principal Channel, has 40 feet of the draft that
makes it the most important deepwater port in Argentina. Cereals are exported
through the “Cargill” and “Terminal Bahía Blanca” companies. Containers enter the
general cargo dock and tie up boats of various uses, such as tourism, defense, and
deep-sea fishing. The boat dock dedicated to artisanal fishing is still maintained in a
2  Geography of the Bahía Blanca Estuary 27

small sector. In the western sector are the facilities of the “Bahía Blanca” Yacht
Club (Fig. 2.4).
To the east and bordering on Coronel Rosales District, a salt area, abandoned
tidal plains and bush land coverage that total some 2,500 hectares of vacant land are
linked to the future port, industrial, and services expansion. This sector is also
crossed by an area of urban water discharges from Grünbeim Town.
The area occupied by the “Great Bahía Blanca,” which involves the city of Bahía
Blanca, General Cerri, and Ingeniero White Towns, is 150 km2. The total population
of the districts is 305,000 inhabitants, discriminated for Bahía Blanca 285,000 habi-
tants, Ingeniero White 12,489, and General Cerri 6,745 habitants (INDEC 2010).

2.2.3  Coronel Rosales District

This district totals a coastal front of about 78 km, of which about 60 km of extension
are on coastal of the Bahía Blanca Estuary. On the border with the Bahía Blanca
District, the vacant of salt areas, abandoned tidal plains, and shrub coverage contin-
ues, where in Coronel Rosales District it covers about 2000 hectares. “Villa del
Mar” town stands out as the only residential urban coastal center of the Villarino,
Bahía Blanca, and Coronel Rosales District (Fig. 2.1). Born from lots as a seaside
resort in 1944 (Cinti 2017), today and with 327 inhabitants, it behaves like a neigh-
borhood center, the result of the urban expansion of Punta Alta city (Fig. 2.3).
Neighboring to “Villa del Mar” is the Naval Base “Puerto Belgrano,” which with
about 8 km of development and about 1,200 hectares is imposed with a strong pres-
ence on the coastal front (Fig.  2.3). Being the main military settlement of the
Argentine Navy, it determined the foundation and growth of the town of Punta Alta
(Chalier 2019), which has about 58,000 inhabitants (INDEC 2010). However, this
presence also prevented the direct geographical linking of this town with the coastal
area, which was only achieved with its urban expansion toward “Villa del Mar”
(Fig. 2.3).
The Puerto Rosales area is presented as a complex sector (Fig. 2.3). It is a large
dock and breakwater of 700 m built as a poorly planned railroad project in 1912
(Melo et al. 1997). In the western sector of the breakwater, the “Arroyo Parejas”
Spa, the main summer recreation site immediately to the city of Punta Alta, oper-
ates. While in the east sector this operates a minor port linked to fishing activities
and the maintenance and service of a system of boats that serve the maintenance of
oil monobuoys of “Oiltanking” company. The access channel of this port is the
“Arroyo Parejas” Stream, which also functions as a discharge channel for the urban
effluents of Punta Alta. To the east of the access channel to dock of Puerto Rosales,
in the spikes and coastal dunes sector called “Cantarelli Island,” there is a place
dedicated to the future Free Trade Zone of Coronel Rosales District and next to it is
the oil pumping and shipping plant of “Oiltanking” company (Fig. 2.3).
In the last 35  km of coastal development, up to the cliffs of “Beacon Monte
Hermoso,” the sandy beaches are under the total control of the Marine Base
28 W. D. Melo

“Baterias” and its access is restricted. In the east sector of said base, in the area near
the mentioned cliffs, the “Manuel Belgrano” Space Center is built whose function
will be to be a launching platform for national satellites.

2.2.4  The Islands

In the Buenos Aires Province, the islands are provincial property, where the districts
have their municipal jurisdiction. Since the end of the nineteenth century, the
“Bermejo,” “Trinidad,” and “Wood” Islands were given in concession for produc-
tive purposes, where they worked from factories to process sharks and sea lions as
well as sheep, swine, and bovine livestock exploitation (Cinti 2017). Since 1992, all
the islands were declared Natural Reserves and their commercial exploitation
ceased. Only “Wood” Island is maintained with livestock activity, while “Ariadne”
Island, like the northern islands, has no commercial use (Fig. 2.1).

References

Aguirre ML (1995) Cambios ambientales en la región costera bonaerense durante el Cuaternario


tardío. Evidencias Malacológicas. IV Jornadas Geológicas y Geofísicas Bonaerenses. Actas
I: 35–40
Aliotta S, Schnack EJ, IslaF et al (2000) Desarrollo secuencial de formas de fondo en un régimen
macromareal. Latin American Journal of Sedimentology and Basin Analysis 7(12):95–107
Carbone ME, Melo WD (2016) Aplicación de geotecnologias en el análisis de procesos mor-
fodinámicos del sector costero externo del estuario de Bahía Blanca. 3ras Jornadas de Nuevas
Tecnologías de la Información Geográfica del Sur Argentino. Departamento de Geografía y
Turismo UNS. Bahía Blanca
Cavallotto JL, Violante R, Parker G (1999) Historia Evolutiva del Río de la Plata durante el
Holoceno. Actas XIV Congreso Geológico Argentino, I: 508–511
CGPBB (2020) La Bahía… Razones para una Identidad. Consorcio General de Puerto de Bahía
Blanca. Disponible en: https://puertobahiablanca.com
Chalier G (2019) Algunas consideraciones acerca de Puerto Belgrano como área patrimonial: con-
cientización y revalorización para el fortalecimiento del sentido de pertenencia. RES GESTA,
nro 55, Instituto de Historia – Fac. Der. y Cs. Ss. del Rosario – UCA. Rosario – Argentina.
ISSN: 2525-0884
Cinti S (2017) Las Islas de la bahía Blanca/Los forjadores de su historia, 1ra Edición. Vacasagrada
Ediciones, Bahía Blanca
Coleman AH (1948) Mi vida de ferroviario inglés en la Argentina: 1887–1948. Ediciones del
autor, Bahía Blanca
Ferrera IM (2003) El polo petroquímico de Bahía Blanca como eje y estrategia del desarrollo en el
corredor bioceánico Sur. Revista Tiempo y Espacio. Temuco, Chile. 10(13):63–80
González Uriarte M (1984) Características geomorfológicas de la porción continental que rodea la
Bahía Blanca, Provincia de Buenos Aires. IX Congreso Geológico. Actas III. 556–576
González MA, Weiler N (1983) Ciclicidad de niveles marinos holocénicos en Bahía Blanca en
delta del río Colorado (Provincia de Buenos Aires), en base a edades Carbono-14. Actas de la
2  Geography of the Bahía Blanca Estuary 29

Oscilación del Nivel Medio del Mar durante el último Hemiciclo De glacial de la Argentina.
69–88. Mar del Plata
Gutiérrez Téllez B, Schillizzi R (2002) Asociaciones de Diatomeas en Paleoambientes Cuaternarios
de la Costa Sur de la Provincia de Buenos Aires. Argentina Pesquisas em Geociências 29:59e70
INDEC (2010) Censo Nacional de Población, Hogares y Viviendas 2010 – Datos provistos por la
Agencia de Inversión y Comercio Exterior del Municipio de Bahía Blanca
London SA (2017) Methodological Note about the Analysis of the Economic Growth and
Environment Ecology & Environmental Sciences vol. 2
Melo W (2004) Origen del estuario de Bahía Blanca. En: Piccolo MC, Hoffmeyer MS (eds)
Ecosistema del Estuario de Bahía Blanca. Instituto Argentino de Oceanografía. Sapienza Ed,
pp 21–27
Melo WD, Carbone ME (2012) Regionalización ambiental del Estuario de Bahía Blanca. IX
Jornadas Nacionales de Geografía Física de Bahía Blanca, Bahía Blanca
Melo WD, Limbozzi F (2008) Geomorphology, hidrological systems and land use of Bahía Blanca
Estuary region. In: Neves R, Baretta J, Mateus M (eds) Perspectives on integrated coastal zone
management in South America. IST Press. Scientific Publishers, Lisboa, pp 317–331
Melo WD, Piccolo MC (2006) Cartografía del estuario de Bahía Blanca. XXIII Reunión Científica
Asociación Argentina de Geofísicos y Geodesias
Melo WD, Cuadrado D, Gómez E (1997) Influencia de las condiciones ambientales en la planifi-
cación territorial de Puerto Rosales. Iras Jornadas Nacionales de Geografía Física 1996
Melo WD, Schillizzi R, Perillo GME et al (2003) Influencia del área continental pampeana en la
evolución morfológica del estuario de Bahía Blanca. AAS Revista 10(1):39–52
Melo WD, Piccolo MC, Perillo GME (2005) Propuesta de desarrollo turístico en playas del estu-
ario de Bahía Blanca: Faro El Rincón y Baliza Monte Hermoso. VII Jornadas Nacionales y
Primer Simposio Internacional de Investigación-Acción en Turismo
Perillo GM, Piccolo MC (1990) Physical characteristics of the Bahía Blanca Estuary (Argentina).
Estuar Coast Shelf Sci 34:303–347
Perillo GME, Piccolo MC (1999a) Geomorphological and physical characteristics of the Bahía
Blanca Estuary, Argentina. In: Piccolo MC, Pino-Quivira M (eds) Perillo GME. Estuaries of
South America. Environmental Science. Springer, Berlin/Heidelberg, pp 195–216
Perillo GM, Piccolo MC (1999b) Geomorphology and physical characteristics of the Bahía Blanca
Estuary (Argentina). In: Piccolo MC, Pino-Quivira M (eds) Perillo GME. Estuaries of South
America. Environmental Science. Springer, Berlin/Heidelberg, 223 p
Piccolo MC, Perillo GME, Melo WD (2008) The Bahía Blanca Estuary: an integrated overview
of its geomorphology and dynamics. In: Neves R, Baretta J, Mateus M (eds) Perspectives on
integrated coastal management in South America. IST Press, Lisboa, pp 231–240
Prieto AR (1996) Late Quaternary Vegetational and Climatic Changes in the Pampa Grassland of
Argentina. Quat Res 45:73–88
Spagnuolo J, Farinati E, Aliotta S (2000) Tafonomía de Brachidontes rodriguezi (Bivalvia) de
cordones costeros holocenos de Argentina. Profil. Institut für Geologie und Paläontologie,
Universität Stuttgart, Band 18, CD- ROOM (76)
Spalletti LA, Isla FI (2003) Características y Evolución del delta del río Colorado (“Colú-Leuvú”),
Provincia de Buenos Aires, República Argentina. Rev Asoc Argent Sedimentol 10(1):23–37
Chapter 3
Physical Oceanography of the Bahía
Blanca Estuary

Gerardo M. E. Perillo and M. Cintia Piccolo

3.1  Introduction

Through history, estuaries have played a significant role in the evolution of coastal
civilizations as they have fulfilled many essential roles that accelerated the estab-
lishment and economic advance of the ports and cities located on their shores
(Perillo 1995a). Most estuaries are associated with rivers that allow a connection
with the hinterland, especially those that have navigational capability. A low per-
centage of the major estuaries in the world does not have significant rivers providing
freshwater and sediment input. The Bahía Blanca Estuary (Fig. 2.1; Chap. 2) is an
exception as the present-day freshwater input is low and concentrated mostly at the
head and middle reach of the Canal Principal.
The Bahía Blanca Estuary has a very complex evolution during the Late
Pleistocene-Early Holocene (for detailed descriptions, see Chapters 1 and 2). The
estuary today was the northern portion of a very large (over 300  km wide) delta
formed by the combination of the Colorado and Negro Rivers (Melo et al. 2003,
2013), but mainly dominated by the former. The formation of the delta corresponds
to a significant input of sediments from the Colorado River through several distribu-
taries as well as other rivers that are no longer active. The excess sediment provided
during about 10,000 years generated a delta which its actual extension into the inner
shelf is today a matter of discussion since there is no clear geomorphologic and
sedimentological indication of its outer border within the El Rincón area (Fig. 1.2;
Chap. 1). Upon the diminishing discharges of the Colorado River, including the
disappearance of many of its distributaries (Melo 2004), sediment input reduced
sensibly, being presently basically negligible.

G. M. E. Perillo () · M. C. Piccolo


Instituto Argentino de Oceanografía, IADO (Universidad Nacional del Sur-CONICET),
Bahía Blanca, Argentina
Departamento de Geología, Universidad Nacional del Sur, Bahía Blanca, Argentina

© Springer Nature Switzerland AG 2021 31


S. M. Fiori, P. D. Pratolongo (eds.), The Bahía Blanca Estuary,
https://doi.org/10.1007/978-3-030-66486-2_3
32 G. M. E. Perillo and M. C. Piccolo

On the other hand, mean sea level (msl) rose at extraordinary rates (about 9 mm/
year; Gómez and Perillo 1995) from 8000 to 6000 years ago, reaching the highest
level so far measured in the world of 7 m above present. Therefore, what is today the
estuary was a shallow sea which penetrated along the Vidriera Salt Flat up to the
Chasicó Lake (Melo 2004; Melo et al. 2013). Sea level dropped at a rate of the order
of 4  mm/year (Gómez and Perillo 1995) until about the Little Ice Age
(1400–1700 year AC) where msl bounced back and started to rise again. Nevertheless,
Gómez et al. (2006) have clear proofs (correlated with similar data from Brazil) that
about 2400 year BP msl dropped about 3 m below present. Still, the actual story of
the origin of the estuary needs to unfold further. In the last 1000 years, the condi-
tions have changed dramatically. Since there is no further sediment input, added to
the advance of the msl of the order of 1.6 mm/year (Lanfredi et al. 1988, 1998), the
Bahía Blanca Estuary has become a marine-dominated environment being in a sig-
nificant erosional stage (Perillo 1995b).
To understand the dynamic processes that occur in the estuary is essential to have
a clear image of the geomorphology of the system since all processes are controlled
by the shape of the channels and the distribution of the intertidal and island areas.
Therefore, the aim of this chapter is to provide an integral review of the geomorpho-
logic, climatic, and dynamic processes acting on the estuary.

3.2  Geomorphologic Features of the Bahía Blanca Estuary

The geomorphology of the Bahía Blanca Estuary has been described many times
already (i.e., Perillo and Piccolo 1991, 1999; Perillo et  al. 2001; Melo 2004).
Therefore, for the present case, only its general structure will be discussed with the
addition of specific features that appear as the most relevant, but also those that have
the higher impact on the dynamic processes. The Bahía Blanca Estuary is a complex
network of tidal channels, with a set of major ones (Principal, Bermejo, Bahía Falsa,
Bahía Verde, and Caleta Brightman) which have a general NW-SE orientation. All
these major channels are remnants of the original distributaries of the Colorado River.
Although at first sight, the Bahía Blanca Estuary appears with little or no changes
along the time, when analyzed in detail, there are significant modifications in the
migration of major channels (Ginsberg and Perillo 1999; Ginsberg et al. 2009), as
well as significant changes that occur among the various marshes and tidal flats.
After the msl retreated about 1500 years ago, the intertidal and supratidal areas were
dominated by Sarcocornia marshes. However, the erosional processes undermined
these marshes lowering the level of the surface making them more prone to have a
longer hydroperiod (meaning passing from inundations about 40–70 times per year
to about 700 times per year). Most of the Sarcocornia marshes were replaced by
tidal flats which themselves are, in part, populated by Spartina alterniflora marshes,
plants that are more adaptable to longer periods of inundation (Pratolongo
et al. 2013).
3  Physical Oceanography of the Bahía Blanca Estuary 33

Nevertheless, as one travels along the tidal channels of the estuary, the original
strata of the delta deposits are outcropping along the flanks. Although, there are
clear signs of erosion in the inner channels and along the island coasts. It is common
to find parts of the islands with their typical continental vegetation deposited, after
being eroded, along the banks of the channels being latter redistributed by the tidal
currents.
The sediment eroded, mostly from the inner portion of the estuary, is transported
toward the mouth and deposited on the lower tidal flats. Therefore, there is a sedi-
ment compaction gradient from the head to the mouth of the estuary. At the head,
where most of the material is still the original from the delta, the level of compac-
tion is higher. However, toward the mouth, most of the sediment that was eroded
and was not transported outside the estuary by the ebb currents is deposited on the
low flats. These are generating layers of unconsolidated sediments, which, if
affected by strong currents or waves, may be easily resuspended. The places where
these sediments are deposited tend to be natural traps, meaning that the hydrody-
namic conditions make them adequate to concentrate the material.
One of the most interesting things observed in the Bahía Blanca Estuary, espe-
cially in the outer portions of the estuary, is that the Spartina marshes develop along
the tidal channels (Fig. 3.1a) whereas tidal flats are located farther away from the
channel (Perillo 2019). The normal distribution in all coastal wetlands of the world
is precisely the opposite (Allen 2000; Hopkinson et  al. 2018; Pratolongo et  al.
2019). Typically, one expects a standard ramp from the channel inland where the
lower levels are dominated by unvegetated (considering vascular plants only) areas.
Along the channel, usually the bare deposits of the tidal flat appear and, as we move
further up, the stability of the sediments and the slightly lower hydroperiod allows
for the settlement of the pioneer plants and further development of the marsh struc-
ture. The presence of plants (specifically their leaves and stems) increases frictional
drag, thus reducing the turbulence generated by waves and currents and helping the
deposition of the sediments being transported by those processes (Leonard and
Luther 1995; Pratolongo et al. 2019).
In the case of the outer portions of the Bahía Blanca Estuary, the mechanism is
inverse likely due to the possible presence of remnants of older delta channel levees.
These levees, being slightly higher than the adjacent flats, are better suited to allow
the settlement of plants right at the border of the flat. Once the pioneer plants settle,
they start working in reducing the turbulence at the moment water flows over the
bankfull levels inducing sediment deposition. Together with the sediment, nutrients
are also deposited, which are employed by the plants for further growth, reducing
still more the sediment transported farther inland. Therefore, it becomes a virtuous
circle, higher plants reduce more the turbulence, generating more sedimentation and
nutrient concentration, which then helps increase the size of the plants. Spartina
plants grow from rhizomes in a radial structure (Fig. 3.1b). As these circles grow,
they coalesce and then form a large patch that spreads along the border of the chan-
nel (Perillo 2019).
Tidal flats and marshes of the Bahía Blanca Estuary are dominated by a complex
network of tidal courses (sensu Perillo 2009) and depressions (Revollo Sarmiento
34 G. M. E. Perillo and M. C. Piccolo

Fig. 3.1  Examples of the


characteristics of the
typical wetlands and tidal
courses found in the Bahía
Blanca Estuary. (a)
Development of Spartina
marsh along the border of
a tidal channel while the
tidal flats are farther away,
(b) complex tidal course
pattern on a tidal flat being
colonized by circular
patches of Spartina, (c)
example of a tidal creek
with higher sinuosity near
the mouth plus some
rectangular meanders
probably due to the
influence of the underlying
deltaic sediments. (Photos
by Gerardo M. Perillo)

et al. 2016, 2020; Perillo 2019). Courses vary from incipient rills and groves along
the flanks of larger courses to extensive creeks and channels. They were described
in detail by Perillo and Cuadrado (1991), Angeles et al. (2004), Ginsberg and Perillo
(2004), among others; however, Perillo (2009, 2019) provided an analysis of the
various types of drainage systems as well as the particular shape of the courses. The
3  Physical Oceanography of the Bahía Blanca Estuary 35

latter requires further analysis, though, because most of the courses change their
meandering pattern from the head to the mouth.
Most commonly, as also occurs in rivers, courses tend to have higher sinuosity in
the low relief sectors of the basin. In the case of the Bahía Blanca Estuary occurs the
contrary. The low relief part of the tidal flats presents creeks with low sinuosity
meanders, but as they approach their mouth on the steeper side of larger creeks or
channels, the degree of sinuosity increases significantly (Fig.  3.1c). A possible
explanation is that meandering in these creeks does not develop, as usually occur in
a river system, as a mechanism to lengthen the channel to retain more water under
the bankfull level, but a process known as instability propagation (Perillo 2019).
The initial meander occurs at the mouth of the creek due to the formation of a bar or
spit due to the sediment transported along the flank of the main channel. This devia-
tion of the mouth creates a damming of the ebbing water which is forced to erode
the course upstream generating a new meander. In some cases, these meanders are
activated by the formation of erosional cusps (Ginsberg and Perillo 1990; Zhao
et al. 2019). Each new meander generates a modification of the circulation pattern
within the creek that propagates headward.
On the other hand, the Bahía Blanca Estuary is worldwide unique due to the
number and diversity of tidal depressions (Perillo 2019) found mostly on the tidal
flats (Fig. 3.2). They range from ponds to pools with circular, elliptic, or complex
shapes. Based on the analysis of Google Earth images of 14 sites selected from the
estuary, the pond density varies from 6 per 100  m2 to 25 per 100  m2 (Revollo
Sarmiento et al. 2020), whereas the size range varies from 0.6 to 173 m2. Depressions
fulfill a significant role in both the ecology and geomorphology of the tidal flats and
marshes. In the former, due to the level of water retention, they provide the condi-
tions for certain species of worms and algae to establish colonies. In the latter case,
depressions are many times related to the process of formation of tidal courses.
The Bahía Blanca Estuary is also characterized by the presence of scour holes
that only were previously reported for North Bay (USA) (Kjerfve et al. 1979). The
most significant scour holes are located at the intersection of tributary channels like

Fig. 3.2  Examples of the extensive distributions of tidal depressions, some of them connected by
small grooves and creeks. Their interconnection may develop into a tidal creek. (Photos by Gerardo
M. Perillo)
36 G. M. E. Perillo and M. C. Piccolo

those found at the Tres Brazas, Cabeza de Buey (Ginsberg and Perillo 1999), or
Maldonado channels (Pierini et al. 2005). Nevertheless, there is a particular scour
hole that is unique for the Bahía Blanca Estuary found at the center of a meander
(Mavo Manstretta et al. 2018; Perillo 2019) which is developed by the formation of
a particular eddy due to interaction of the tidal currents with the specific geomor-
phology of the channel.
Large bedforms in the deeper portions of the Principal Channel are another
unique feature of the Bahía Blanca Estuary. Aliotta and Perillo (1987) described the
first-ever large dunes (up to 6 m high) that generate flow separation at the crest due
to the high slope of the lee side (in some dunes they are 30° (average 11°) near the
angle of repose of the sediment). The particular characteristic of these dunes is that
they are solitary bedforms as defined by Perillo and Ludwick (1984) since they are
separated by a flat bed of fine sediment (silt/clay). The average migration rate was
calculated in 33  m/year which was later confirmed by a newer study (Minor
Salvatierra et al. 2015) of the same bedform field reaching an average migration of
43 m/year. Although this is the major dune field, there are others found both on the
middle reach (Cuadrado et al. 2010) and in the outer portion of the estuary (Gómez
and Perillo 1992).
Even though the extensive area that the Bahía Blanca Estuary (2300 km2) has,
most of it has not been studied from both its geomorphology and dynamics. On the
other hand, about 90% of the publications on these subjects are concentrated in
Canal Principal’s issues. The Principal Channel (Fig. 2.1; Chap. 2) has a total length
of 60 km and varies in width from about 3–4 km at the mouth (22 m depth) to 200 m
at the head (3 m depth); both depth and width increase almost an order of magnitude
from head to mouth (Perillo and Piccolo 1991). Like other major channels (bays)
that flow toward the inner shelf, the Canal Principal is partly closed by a modified
ebb delta (Cuadrado and Perillo 1997). The channel cross section is steep on the
sides, with a U-shaped bottom having a small asymmetry to the right. Upstream of
Puerto Galvan (Fig. 2.2; Chap. 2), the channel narrows and becomes more V-shaped
with the asymmetry following the meandering pattern headward (Gómez et al. 1996).
At the confluence with the Principal Channel, the funnel-shaped mouth of tribu-
tary channels is turned seaward, due to the ebb dominance in the Principal Channel.
Most ebb deltas at the mouth of major channels have changed (Gómez and Perillo
1992; Cuadrado and Perillo 1997a, b); however, the original delta shape and south-
ward orientation of the ebb channel and associated shoals in the Principal Channel
are still preserved despite strong marine dynamics. This is due to its stable connec-
tion to the northern shore and its position on top of a sill (Chasicó Formation),
which has reduced sediment transport and served as an anchor for the ebb delta
(Cuadrado and Perillo 1997a, b). The mobility of delta shoals depends on the
approach angle of tidal currents (Perillo and Cuadrado 1991), while 3D dunes in the
Canal Principal and tributary channels form due to high current velocities and geo-
morphologic “traps” which favor sedimentation (Aliotta and Perillo 1987; Gómez
et al. 1996).
Except for few bedforms in flood-dominated channels, all 3D dunes and shoals
in the Bahía Blanca Estuary have ebb dominance. Since sediment runoff from rivers
3  Physical Oceanography of the Bahía Blanca Estuary 37

is virtually absent and ebb delta characteristics impede sediment input from the
shelf, the high concentration of suspended sediments in the estuary is due to erosion
of tidal flats and island shores (Ginsberg and Perillo 1990). The southern coast of
the Principal Channel across the port of Ingeniero White has retreated up to 50 m
between 1980 and 1986 and 1.5 million m3 were exported from an 8 km stretch in
the mid-reach of the channel (Perillo and Sequeira 1989). However, the dredging of
the port and navigation channel to a nominal depth of 45′ (13.5 m) in 1989–1990
affected most of the tidal flats located along large portions of the southern coast of
the Principal Channel. The tidal flats have become Spartina marshes in about 3
years after the dredging. At present, there are indications that the Spartina marshes
are now in an incipient erosional stage, probably due to lack of sediment input.
Because most sediments in the estuary are silts and clays (Gelós et al. 1988),
strong currents and short slack water intervals impede their deposition in channels
and on the tidal flats, while short, locally generated waves erode old sediments and
prevent any accumulation of new ones. These conditions explain the erosional stage
of most of the estuary and the prevalence of sediments from the deltaic deposition
period. Furthermore, the biological and physical interactions in the system are
rather crucial in the development of tidal creeks (Perillo and Iribarne 2003) and play
a role in the erosional processes. For instance, Minkoff et al. (2005) and Minkoff
(2007) have demonstrated that crabs and plants acting together were responsible for
eroding over 2200 m3 of sediments of a 270 ha high marsh in 45 years.

3.3  Climatology

The Bahía Blanca Estuary is located in a temperate climate zone in the southeast of
Buenos Aires Province. Mid-latitude westerly winds and the influence of the
Subtropical South Atlantic High dominate the typical weather pattern of the region.
Climate variability and air-water interaction processes determine the main charac-
teristics of the estuarine water parameters and dynamics. The area is windy with
prevailing winds direction from NNW, NW, and N with a mean velocity of 24 km/h
during 40% of the year and gusts over 100 km/h. The wind is high during the spring
and summer and diminishes in the fall. The highest average speed occurred in the
winter of 1986 with 82 km/h. The highest mean record was 70.3 km/h in 1980 and
the lowest 28.2 km/h in 1964. A study of 50 years (1960–2014) of meteorological
variables on the estuary coast shows the significance of the climate variability in
different decades. The maximum average wind speeds were observed in 1970–1980
with mean values (62 km/h) that exceeded the average (50.3 km/h). The 1980–1990
decade had an average of 51.4 km/h. After that decade, the average speed decreased
registering speeds of 45.5 km/h in 1990–2000 and 47.2 km/h in 2000–2010. From
2003 to the present, this parameter evidenced a decrease in the annual maximum
average speeds (Ferrelli 2016; Ferrelli et al. 2019).
The mean annual temperature varies from 14 to 20 °C, with an average annual
temperature of 15.5 °C. The lowest value corresponded to 1964 with 14 °C and the
38 G. M. E. Perillo and M. C. Piccolo

highest to 2014 with 16.6 °C. Periods in which the air temperature was lower than
the average (15.3 °C) occurred in 1960–1966, 1971–1979, and 2005–2007, while
the warm ones were 1980–1985, 1993–2001, 2008–2009, and 2011–2014. Due to
the interaction processes between water and the atmosphere, an increase in the aver-
age air temperature causes an increase in the water temperature. The occurrence of
El Niño events approximately coincided with air and water temperatures above the
mean values, while the manifestation of La Niña events in the Bahía Blanca Estuary
caused air and water temperatures below the mean values. During the last 50 years
(1960–2014), the air temperature of the area increased 0.9 °C, the maximum wind
speed decreased 7.5 km/h, and the relative humidity presented two different cycles,
marking wet and dry periods. The annual mean rainfall is 654 mm.
Like temperature, precipitation in the estuary showed notable differences on a
seasonal scale. Summer is the rainiest season with an average of 206.2 mm, present-
ing a maximum of 405 mm in 1985 and a minimum of 56.5 mm in 1972. The spring
average is 198.4 mm. The maximum was recorded in 1976 with 496 mm and the
minimum in 1991 with 83.7 mm. Autumn presents an average value of 138.1 mm,
ranging between 364.2 and 30.7  mm in 1982 and 1988, respectively. The winter
average value is 96.3 mm with a maximum of 238.5 mm in 1989 and a minimum of
11.5 mm in 1995. The dry and wet years were related to the Oceanic Niño Index
(ONI) since they occurred together with weak and moderate El Niño and La Niña
years, while the normal ones generally coincided with weak Niño or Niña periods.
The climate of the region presents a marked seasonality, and the climate variability
has an essential influence of the estuary dynamics, because wind, among others,
affects tides, waves, and water circulation in general.

3.4  Physical Oceanography of the Estuary

As indicated, the Bahía Blanca Estuary receives very little input of freshwater, but
it is enough to generate a thermohaline circulation (Piccolo and Perillo 1990). The
two major tributaries enter the estuary from the northern shore. The Sauce Chico
River (a drainage area of 1600  km2) discharges into the Principal Channel about
3 km downstream from the head of the estuary, and the Napostá Grande Creek (a
drainage area of 1240 km2) reaches the estuary about 1 km downstream of the port
of Ingeniero White (Fig. 2.2; Chap. 2). Both tributaries behave similarly in spring
and summer during maximum mean rainfall, but they are out of phase in autumn
when the Sauce Chico River has a secondary mode (Piccolo and Perillo 1990).
Although mean annual runoff flows of the Sauce Chico River and the Napostá
Grande Creek are low (1.5–1.9 and 0.5–0.9  m3/s, respectively), runoff from the
Sauce Chico River may peak between 10 and 50 m3/s, with a recorded maximum of
106  m3/s in 1977, although other unpublished sources indicate a maximum of
200 m3/s. Besides these significant freshwater inputs, the inflows from other, smaller
tributaries into the estuary are intermittent and only significant during periods of
high local precipitation. However, the most substantial input of freshwater,
3  Physical Oceanography of the Bahía Blanca Estuary 39

nutrients, and contaminants is provided by the sewage discharges from Bahía


Blanca, Punta Alta, and Ingeniero White cities. These sewage discharges are located
at the Maldonado Channel and Napostá Creek (for Bahía Blanca City), Puerto
Rosales (Punta Alta City) (Figs. 2.2 and 2.4; Chap. 2). Only in the last year, all three
discharges have treatment plants, otherwise, before the sewage had only some minor
primary treatment which implied a high level of pollution mostly around the dis-
charge places (see Chap. 4, “Bahía Blanca Estuary: A Chemical Oceanographic
Approach”). However, the high level of turbulence and mixing due to the tidal
dynamics allows a dramatic dilution of the pollutants within a few kilometers from
the discharges.
When a flash flood in either of both tributaries occurs, the plume of freshwater
only remains within a few tens of kilometers from the mouth of the rivers. In those
conditions, the water column shows a significant stratification that in places may
reach a salt wedge structure. However, rarely are more than partly mixed conditions.
These conditions are exceptional since most of the time, the estuary behaves as
vertically homogeneous (Perillo et al. 2001).
Mean annual (13 °C), summer (21.6 °C), and winter (8.5 °C) surface water tem-
peratures (Fig. 3.3a) in the Principal Channel are always slightly higher at the head

Fig. 3.3  Average historical distribution of (a) temperature and (b) salinity and density along the
Principal Channel of the Bahía Blanca Estuary. Position 0 km corresponds to the mouth of the
Sauce Chico River. (Modified from Piccolo and Perillo 1990)
40 G. M. E. Perillo and M. C. Piccolo

of the estuary (Piccolo et al. 1987; Piccolo and Perillo 1990; Perillo et al. 2001),
while mean surface salinity increases exponentially from the head to mid-reaches of
the estuary (Fig. 3.3b). Longitudinal temperature distributions vary between rainy
periods in spring/summer and low runoff in winter, when the vertical thermal struc-
ture of the estuary is homogeneous, and there are longitudinal variations. The data
provided in Fig. 3.4 correspond to average values taken from historical measure-
ments made between 1967 and 1986 along the Principal Channel (Perillo et  al.
1987). However, recent monitoring made along the middle reach of the channel has
shown similar temperature values but marked an increase in the salinity data vary-
ing from 33.8 in winter to 37.1 in summer (IADO 2016), but, in some cases, values
reached up to 45.3 in the fall of 2018 basically along the whole channel (IADO 2018).
Because of its intricate geomorphology, the dynamic processes acting in the estuary
are also complex. The main driver is the tide, but the wind plays a significant role that

Fig. 3.4  Examples of cross-section velocity profiles measured at different places of the inner
reach of the Principal Channel showing the effects of the tributaries, adjacent tidal flats, as well as
the dynamics along a large channel meander. (a) Location of the cross section, (b) velocity vectors
along the cross section, (c) amplitude and direction of cross-section velocity vectors during flood
conditions, (d) amplitude and direction of cross-section velocity vectors during ebb conditions.
Note the changes in speed across the channel due to the geomorphology. (Modified from López
Gregori et al. 2017)
3  Physical Oceanography of the Bahía Blanca Estuary 41

cannot be discounted in the circulation of the waters. In the present section, we propose
to provide a simplified view of the main processes and the interaction among them.
The estuary is dominated by a quasi-stationary tidal wave that cooscillates with
the tidal wave that propagates from south to north along the Argentinian continental
shelf (Palma et al. 2004). Due to how the wave propagates, the water penetrates first
through the Bahía Falsa and Bahía Verde channels and later through the Principal
Channel (Perillo and Piccolo 1991). The ebbing condition is reverse, moving out
first through the southern channels and ending with the outflow from the Principal
Channel. The complexity of this situation arises because all these main channels are
interconnected by the complex network of channels and creeks that transfer water
from one primary channel to another. Furthermore, when the tide level covers the
flats and marshes, except for the islands, all the estuary is covered by water with a
mean depth of the order of 0.5–1 m (Perillo et al. 2001). Therefore, water that may
have initially entered the estuary through one of the southern bays, it may ebb by
one or more of the northern channels. In several channels that connect two major
channels (i.e., El Embudo), the direction of the flow may change up to four times
during the same tidal cycle depending on which major channel is generating the
most considerable influence in the circulation.
The complexity of this circulation has a significant correlate because it makes
impossible to establish with adequate precision the residence time of the water, but
still worse, of any possible substance that is introduced in the estuary. Numerical
models of the estuary so far (Pierini et al. 2008a, b) failed in defining this situation
because we still lack detailed geomorphology of the estuary which is the essential
element to have a well-behaved model.
The Bahía Blanca Estuary is a mesotidal, semidiurnal system with tidal ranges
that were measured only along the Principal Channel with mean values ranging
from 2.2 m at the Oceanographic Tower (OT) (Fig. 2.4; Chap. 2) to 3.5 m in Puerto
Belgrano, 3.8 m in the Ingeniero White port reaching a maximum of 4 m at the head
in Villarino Viejo (Perillo and Piccolo 1991). Although along all the estuary the tide
behaves as pure semidiurnal, at the OT the tide is mixed semidiurnal (Perillo and
Piccolo 1991). The relationship between the convergence of the channel and the
friction effect of the geomorphology results in a hypersynchronous estuary (based
on the classification by Le Floch 1961) as the tidal range and the intensity of the
tidal currents increase headward.
The comparison of the amplitude of the main harmonics at the three most impor-
tant tidal stations between the data back in 1991 with information gathered from 2008
to 2017 shows basically no changes. However, when the phases are compared, they
show variations of up to 40° (Table  3.1). Although there is no specific correlation
between the changes in geomorphology and the possible shift in the phases of the tide,
as a first approximation we assume that the changes in this variable may have occurred
due to the dredging of the Principal Channel as well as the significant deposition of the
dredged material on top of the tidal flats and marshes along the Principal Channel.
As indicated, the estuary is fully dominated by the tides and behaves as a lower,
marine estuary as all of it is affected by marine processes. The middle and upper
estuaries are concentrated within the last 5 km of the Sauce Chico River, at the head
of the estuary (Piccolo and Perillo 1990; Perillo et al. 2001).
42 G. M. E. Perillo and M. C. Piccolo

Table 3.1  Changes in the amplitude and phase of the main harmonic components between the
data informed by Perillo and Piccolo (1991) (91) and the values estimated in 2017 (17) by Blanco
Monroy (in preparation)
Station Harmonic IW PB OT
Zo 2.32 2.43 2.93
M2 A(91) 1.63 1.46 1.09
∅(91) 271 266 233
A(17) 1.56 1.41 1.05
∅(17) 244 236 213
S2 A(91) 0.22 0.18 0.17
∅(91) 48 44 356
A(17) 0.2 0.20 0.17
∅(17) 5 356 334
O1 A(91) 0.13 0.14 0.15
∅(91) 43 51 19
A(17) 0.14 0.16 0.15
∅(17) 30 25 12
K1 A(91) 0.19 0.17 0.20
∅(91) 110 81 85
A(17) 0.18 0.19 0.18
∅(17) 87 81 73
N2 A(91) 0.24 0.12 0.17
∅(91) 172 149 159
A(17) 0.2 0.17 0.12
∅(17) 163 154 134
IW Ingeniero White, PB Puerto Belgrano, OT Oceanographic Tower

Winds in the Bahía Blanca Estuary play a significant role in all aspects of the estua-
rine dynamics. As indicated, wind speeds can be very high blowing from rather par-
ticular directions, specifically N and NW, SE and SW.  All these directions have
immediate repercussions regarding their relationship with the orientation of the larg-
est channels (mostly NW-SE). Winds generate waves, storm surges, and subtidal sea-
level variations in the estuary. Wind waves (about 5–10  cm high and 1–3  m long)
occur in channels and on tidal flats when covered by water. The incoming tide, together
with N and NW winds, forms interaction waves which are steep and up to 1.5 m high
and 10–20 m long. However, the main effect of the predominant NW and N winds is
advancing the time of low water, delaying the time of high water, and modifying pre-
dicted water levels. Significant deviations between predicted astronomical tides and
actual tide levels occur at the Ingeniero White port (−4.01 and 2.39  m) and the
Oceanographic Tower (−1.51 and 1.87 m; Perillo and Piccolo 1991). The maximum
negative values coincide with NW winds and maximum positive values with winds
from the SW. At the same stations, the low-frequency sea-­level response to wind indi-
cates that time scales of more than 10 days prevail while short-scale energy peaks
correspond to approximately 3-day periods (Piccolo and Perillo 1989).
Perillo and Sequeira (1989) proposed three types of waves for the middle reach
of the Bahía Blanca Estuary (although they can be observed in any of the other
3  Physical Oceanography of the Bahía Blanca Estuary 43

indicated environments): (1) formed by wind action, (2) formed by the interaction
of wind and tidal currents, and (3) formed by ship wakes. Wind waves are origi-
nated by the continuous action of the wind over any fetch large enough to develop
them. If the wind blows parallel to channels, short, very steep waves are the most
common ones. These low period waves have little effect on the channel bottom,
but they can be very effective in resuspending sediments on tidal flats and eroding
channel banks. It is common to observe along channel borders a turbid, relatively
narrow band of water derived by this wave activity. The continuous impinging of
these very steep waves upon the channel flanks may also be a decisive factor in the
sediment fatigue and resuspension of any material deposited in the tidal flats dur-
ing high tide slack which may result in a coarse sediment but thin laminae within
the sediment structure of the flats (Ginsberg and Perillo 1990). In places where
strong winds blow parallel to large channels, as is common along the Principal
Channel of the Bahía Blanca Estuary, very steep waves, up to 1.5 m high and with
wavelengths of the order of 10–30 m, are formed by the interaction of the incom-
ing tide and the wind.
The presence of these waves sometimes makes the navigation of small crafts and
fishing vessels, which are active in the area, very difficult. Once generated, these
waves propagate against the flanks, contributing to their mechanical erosion. These
waves disappear as soon as the wind or tide changes directions. Waves are not only
a natural phenomenon in estuaries with active harbors, wake action against the
channel coast and machine vibration may also induce fatigue in the interparticle
bonding (Ginsberg and Perillo 1990), thus increasing the erosion potential of the
material especially if the channels are relatively narrow.
Tidal currents in the estuary, in synchronicity with the geomorphology, are also
rather complicated because of the interaction of the channels with the adjacent tidal
flats and marshes. During the tidal stages in which the water level is below the bank-
full, the water flow is similar to what may be observed in a regular river with the dif-
ference of its reversibility. However, when the water level floods the adjacent flats and
marshes, there is a tendency to reduce the velocity while the water spreads out over
the large flat areas. This is also because the tide must work against the slope, although
small, still large enough to extend the time to complete the inundation. As the condi-
tions reverse, the return flow is accelerated by two reasons: first is the geomorphologic
slope in the direction of the flow, but also because there is a marked increase in the
energy slope between the water on the flats and that in the main channel.
An example of the tidal circulation in the Principal Channel that probes the pre-
vious ideas is the study made by López Gregori et al. (2017) at the inner portion of
the channel. Along two successive tidal cycles (over 27 h) measured twice in spring
and neap conditions, in 7 cross sections a total of 430 transects (there is no previous
reference of so many transects elsewhere). The idea was to analyze both the circula-
tion along the channel but also to define the meandering effects as well as the influ-
ence of tributary channels upon the Principal Channel. As expected, ebb currents
were significantly stronger than their flood counterparts, even though the duration
of the flood was about 1.5 h longer than the ebb. The current direction was, for the
most part, parallel to the flanks; however, in transects adjacent to tributary channels,
we observed a higher level of variability (Fig. 3.4c). Employing the velocity profiles
44 G. M. E. Perillo and M. C. Piccolo

in the thalweg, they estimated the average direction vector for the flood and ebb
currents with a difference between them of 176°.
We detected a progressive variation of the velocity along the width of the tran-
sect, being this phenomenon more pronounced during the ebb (Fig. 3.5a). In the first
case, the depth-averaged velocity analysis showed a lateral reduction of this param-
eter from the southern flank to the north flank, estimated at an average value of the
order of 0.60 m/s. In contrast, we observed an inversion for transect II in the loca-
tion of the zones of maximum and minimum velocity with respect to the former.
The mean difference was 0.52 m/s, and the core of higher velocities was character-
ized by a larger lateral extension. During the flood, the lateral variations of the

Fig. 3.5  Examples of cross-section velocity profiles measured at different places along a channel
meander in the inner reach of the Principal Channel showing the effects of the adjacent tidal flats
as well as the dynamics along a large channel meander: (a) ebb conditions, (b) flood conditions.
(Modified from López Gregori et al. 2017)
3  Physical Oceanography of the Bahía Blanca Estuary 45

velocity for both transects were diffusive, being possible to observe a field of speeds
defined by a higher intensity in the vicinity of the southern slope (Fig. 3.5b).
During the whole tidal cycle, on the northern portion of the transects, we observed
significant variability in the magnitude and direction of the current. For instance,
mouthward the Bahía del Pejerrey Channel, they detected a strong velocity gradient.
While in the center of the channel, maximum depth-averaged velocities were between
0.80 and 0.90 m/s southward, and between 0.15 and 0.37 m/s northward, respectively
(Fig. 3.5). This effect was more pronounced during mid ebb, disappearing as the cycle
approached low-water slack. Similar results were observed by Payares Peña et  al.
(2018) along a pronounced meander where a major scour hole was observed.
Velocity profiles clearly show the effects of acceleration and deceleration (Sassi
2009) at the surface levels (Fig. 3.6). Nevertheless, wind influence plays an essential
role in the velocity profile by actually generating two boundary layers: (1) a bottom
boundary layer principally characterized by the logarithmic velocity law, and (2) at
the surface, the wind shear stress produces a boundary layer (Fig. 3.7). The momen-
tum transfer to the lower levels results in a logarithmic defect velocity profile (Sassi
and Perillo 2006).

3.5  Summary

The Bahía Blanca Estuary is one of the most complex estuaries in the world and,
besides more than 50  years of intensive studies, one of the less known from the
oceanographic point of view. The complex geomorphology, the logistics required to
obtain adequate data from areas south of the Principal Channel, makes the task
rather complicated and particularly expensive both in time and budget. Therefore,

Fig. 3.6  Examples of velocity profiles along a tidal cycle of the longitudinal and transversal com-
ponents of the flow on the Bahía del Pejerrey Channel. These profiles are purely tidal with no wind
influence. In all cases (a–d), the effects of acceleration and deceleration of the flow are observed
in particular during the times around slack water. (Modified from Sassi and Perillo 2006;
Sassi 2009)
46 G. M. E. Perillo and M. C. Piccolo

Fig. 3.7  Examples of velocity profiles along a tidal cycle for the longitudinal and transversal
components of the flow in the Principal Channel near the mouth of the Maldonado Channel but
strongly affected by wind. In all cases (a–d), the effects of acceleration and deceleration of the flow
are observed in particular during the times around slack water. (Modified from Sassi and Perillo
2006; Sassi 2009)

most researchers concentrated their effort on the Principal Channel and the areas
adjacent to use a pilot zone that may represent the rest of the estuary.
However, this may not be true. The Principal Channel has both geomorphologic
and dynamic conditions that are unique within the context of the whole estuary.
None of the other channels has the same structure neither they are being effected by
anthropic activities. Rather the contrary, channels like Bermejo or Bahía Falsa and
Bahía Verde are quite pristine and require to be approached in a different way as we
have done for the Principal Channel.

References

Aliotta S, Perillo GME (1987) A sand wave field in the entrance to the Bahía Blanca Estuary,
Argentina. Mar Geol 76(1):1–14
Allen J (2000) Morphodynamics of Holocene salt marshes: a review sketch from the Atlantic and
Southern North Sea coasts of Europe. Quat Sci Rev 19:1155–1231
3  Physical Oceanography of the Bahía Blanca Estuary 47

Angeles GR, Perillo GME, Piccolo MC, Pierini JO (2004) Fractal analysis of tidal channels in the
Bahía Blanca Estuary (Argentina). Geomorphology 57:263–274
Cuadrado DG, Perillo GME (1997) Principal component analysis applied to geomorphologic evo-
lution. Est Coast Shelf Sci 44(4):411–419
Cuadrado DG, Perillo GME (1997a) Migration of intertidal sand banks at the entrance of the Bahia
Blanca Estuary, Argentina. J Coast Res 13:139–147
Cuadrado DG, Perillo GME (1997b) Principal component analysis applied to geomorphologic
evolution. Est Coast Shelf Sci 44:411–419
Cuadrado DG, Gómez EA, Pierini JO (2010) Sand transport on an estuarine submarine dune field.
Geomorphology 121:257–265
Ferrelli F (2016) Análisis del clima local y microlocal de la ciudad de Bahía Blanca. PhD disserta-
tion (unpublished). Universidad Nacional del Sur
Ferrelli F, Brendel AS, Aliaga VS et al (2019) Climate regionalization and trends based on daily
temperature and precipitation extremes in the south of the Pampas (Argentina). Cuad Investn
Geogr 45:393–416
Gelós EM, Spagnuolo JO, Lizasoain GO (1988) Mineralogía y caracterización granulométrica de
sedimentos actuales de la plataforma Argentina entre los paralelos de 39° y 43° de latitud sur y
del Golfo San Matías. Rev Asoc Geol Arg 43:63–79
Ginsberg SS, Perillo GME (1990) Channel bank recession in the Bahía Blanca Estuary, Argentina.
J Coast Res 6(4):999–1010
Ginsberg SS, Perillo GME (1999) Deep scour holes at the confluence of tidal channels in the Bahia
Blanca Estuary, Argentina. Mar Geol 160:171–182
Ginsberg SS, Perillo GME (2004) Characteristics of tidal channels in a mesotidal estuary of
Argentina. J Coast Res 20:489–497
Ginsberg SS, Aliotta S, Lizasoain GO (2009) Morphodynamics and seismostratigraphy of a deep
hole at tidal channel confluence. Geomorphology 104:53–261
Gómez EA, Perillo GME (1992) Largo Bank: a shoreface-connected linear shoal at the Bahía
Blanca Estuary entrance, Argentina. Mar Geol 104:193–204
Gómez EA, Perillo GME (1995) Sediment outcrops underneath shoreface-connected sand ridges,
outer Bahía Blanca estuary, Argentina. Quat South Am Antar Pen 9(3):27–42
Gómez EA, Ginsberg SS, Perillo GME (1996) Geomorfología y sedimentología de la zona interior
del Canal Principal del Estuario de Bahía Blanca. Rev Asoc Arg Sediment 3(2):55–61
Gómez EA, Martinez DE, Borel CM et al (2006) Negative sea-level oscillation in Bahía Blanca
Estuary related to a global climatic change around 2,650 yr. B.P. J of Coast Res SI 39
(Proceedings of the 8th International Coastal Symposium), 181–185
Hopkinson CS, Wolanski E, Brinson MM, Cahoon DR, Perillo GME (2018) Coastal wetlands:
a synthesis. In: GME P, Wolanski E, Cahoon DR, Hopkinson CS (eds) Coastal wetlands: an
integrated ecosystem approach. Elsevier, Amsterdam, pp 1–75
IADO (2016) Programa de monitoreo de la calidad ambiental de la zona interior del estuario de
Bahía Blanca. Informe Final, Instituto Argentino de Oceanografía, 228 pp. Available in: http://
www.bahiablanca.gov.ar/cte/index.html
IADO (2018) Programa de monitoreo de la calidad ambiental de la zona interior del estuario de
Bahía Blanca. Informe Final, Instituto Argentino de Oceanografía, 364 pp. Available in: http://
www.bahiablanca.gov.ar/cte/index.html
Kjerfve B, Shao C-C, Stapor FW (1979) Formation of deep scour holes at the junction of tidal
creeks: a hypothesis. Mar Geol 33:M9–M14
Lanfredi NW, D’Onofrio EE, Mazio CC (1988) Variations of the mean sea level in the Southwest
Atlantic Ocean. Cont Shelf Res 8:1211–1220
Lanfredi NW, Pouza JL, D’Onofrio EE (1998) Sea-level rise and related potential hazards on the
Argentine coast. J Coast Res 14:47–60
Le Floch J (1961) Propagation de la marée dans l’estuaire de la Seine et en Seine Maritime. PhD
dissertation, Université de Paris, 507 pp (unpublished)
48 G. M. E. Perillo and M. C. Piccolo

Leonard LA, Luther ME (1995) Flow hydrodynamics in tidal marsh canopies. Limnol Oceanogr
40:1474–1484
López Gregori M, Alvarez LI, Perillo GME (2017) Circulation pattern in a tidal channel influenced
by tributary channels and adjacent tidal flats. Rio Accoustics 2017. https://ieeexplore.ieee.org/
document/8349706
Mavo Manstretta GM, Payares Peña NB, Vitale AJ et al (2018) Banco sedimentario como agente
modelador de corrientes y transporte de sedimentos en un canal de marea del estuario de Bahía
Blanca, Argentina. XVI Reunión Argentina de Sedimentología, Gral. Roca (abstract)
Melo WD, Schillizzi R, Perillo GME et al (2003) Influencia del área continental pampeana sobre el
origen y la morfología del estuario de Bahía Blanca. Rev Asoc Arg Sediment 10:65–72
Melo WD (2004) Cuenca de Recepción del Estuario de Bahía Blanca. PhD. Dissertation (unpub-
lished) Departamento de Geografía, Universidad Nacional del Sur
Melo WD, Perillo GME, Perillo MM et al (2013) Late Pleistocene-Holocene deltas in the southern
Buenos Aires Province, Argentina. In: Young G, GME P (eds) Deltas: landforms, ecosystems
and human activities, IAHS Publication 358. IAHS Press, Wallingford, pp 187–195
Minkoff DR (2007) Geomorfología y dinámica de canales de mareas en ambientes intermareales.
PhD dissertation Departamento de Ingeniería, Universidad Nacional del Sur (unpublished)
Minkoff DR, Escapa CM, Ferramola FE et al (2005) Erosive processes due to physical - biologi-
cal interactions based in a cellular automata model. Lat Am J Sediment Basin Anal 12:25–34
Minor Salvatierra M, Aliotta S, Ginsberg SS (2015) Morphology and dynamics of large subtidal
dunes in Bahia Blanca estuary, Argentina. Geomorphology 246:168–177
Palma ED, Matano RP, Piola AR (2004) A numerical study of the southwestern Atlantic shelf
circulation: Barotropic response to tidal and wind forcing. J Geophys Res 109:C08014. https://
doi.org/10.1029/2004JC002315
Payares Peña NB, Mavo Manstretta GM, Arena M et al (2018) Características hidrodinámicas de
un canal de marea del estuario de Bahía Blanca, Argentina. XXVIII Congreso Latinoamericano
de Hidráulica. Buenos Aires (extended abstract)
Perillo GME (1995a) Geomorphology and sedimentology of estuaries: an introduction. In: GME P
(ed) Geomorphology and sedimentology of estuaries. Elsevier, Amsterdam, pp 1–16
Perillo GME (1995b) Definition and geomorphologic classifications of estuaries. In: Perillo GME
(ed) Geomorphology and sedimentology of estuaries. Elsevier, Amsterdam, pp 17–47
Perillo GME (2009) Tidal courses: classification, origin and functionality. En: Perillo GME,
Wolanski E, Cahoon DR, y Brinson MM (eds) Coastal wetlands: an integrated ecosystem
approach. Elsevier, Amsterdam. 185–210
Perillo GME (2019) Geomorphology of tidal courses and depressions. In: Perillo GME, Wolanski
E, Cahoon DR, y Hopkinson CS (eds) Coastal wetlands: an integrated ecosystem approach.
Elsevier, Amsterdam. 221–261
Perillo GME, Cuadrado DG (1991) Geomorphologic evolution of El Toro Channel, Bahía Blanca
Estuary (Argentina) prior its dredging. Mar Geol 97(3/4):405–412
Perillo GME, Iribarne OO (2003) New mechanisms studied for creek formation in tidal flats: from
crabs to tidal channels. EOS Am Geophys Union Trans 84:1–5
Perillo GME, Ludwick JC (1984) Geomorphology of a sand wave in lower Chesapeake Bay,
Virginia, USA. Geo-Mar Lett 4(2):110–114
Perillo GME, Piccolo MC (1991) Tidal response in the Bahía Blanca Estuary. J Coast Res
7:437–449
Perillo GME, Piccolo MC (1999) Geomorphologic and physical characteristics of the Bahía
Blanca Estuary. Argentina. En: Perillo GME, Piccolo MC, y Pino Quivira M (eds) Estuaries of
South America: their geomorphology and dynamics. Environmental Science Series, Springer-­
Verlag, Berlín 195–216
Perillo GME, Sequeira ME (1989) Geomorphologic and sediment transport characteristics of the
middle reach of the Bahía Blanca Estuary, Argentina. Journal of Geophysical Research-Oceans
94(C10):14351–14362
Perillo GME, Arango JM, Piccolo MC (1987) Parámetros físicos del estuario de Bahía Blanca.
Período 1967–1986. Instituto Argentino de Oceanografia Contribución Técnica, 250 pp
3  Physical Oceanography of the Bahía Blanca Estuary 49

Perillo GME, Piccolo MC, Parodi E et al (2001) The Bahía Blanca estuary, Argentina. In: Seeliger
U, Kjerfve B (eds) Coastal marine ecosystems of Latin America. Springler, Berlín, pp 205–217
Piccolo MC, Perillo GME (1989) Subtidal sea level response to atmospheric forcing in Bahía.
Proceedings of the 3rd International Congress on Southern Hemisphere Meteorology and
Oceanography, Buenos Aires, pp 323–324
Piccolo MC, Perillo GME, Arango JM (1987) Hidrografía del estuario de Bahía Blanca, Argentina.
Revista Geofisica 26:75–89
Piccolo MC, Perillo GME (1990) Physical characteristics of the Bahía Blanca estuary (Argentina).
Est Coast Shelf Sci 31:303–317
Pierini JO, Perillo GME, Carbone ME et al (2005) Residual flow structure at a scour hole in Bahía
Blanca Estuary, Argentina. J Coast Res 21:784–796
Pierini JO, Marcovecchio J, Campuzano F, Perillo GME (2008a) Evolution of salinity and temper-
ature in Bahía Blanca Estuary, Argentina. En: Neves R, Baretta J, Mateus M (eds) Perspectives
on integrated coastal zone management in South America. IST Press, Lisbon. 501–509
Pierini JO, Campuzano F, Marcovecchio J, Perillo GME (2008b) The application of MOHID to
assess the potential effect of sewage discharge system at Bahía Blanca Estuary (Argentina). In:
Neves R, Baretta J, Mateus M (eds) Perspectives on integrated coastal zone management in
South America. IST Press, Lisbon. 511–518
Pratolongo P, Mazzon C, Zapperi G et al (2013) Land cover changes in tidal salt marshes of the
Bahía Blanca estuary (Argentina) during the past 40 years. Est Coast Shelf Sci 113:23–31
Pratolongo P, Leonardi N, Kirby JR et al (2019) Temperate coastal wetlands: morphology, sedi-
ment processes, and plant communities. In: Perillo GME, Wolanski E, Cahoon DR et al (eds)
Coastal wetlands: an integrated ecosystem approach. Elsevier, Amsterdam, pp 105–152
Revollo Sarmiento GN, Cipolletti MP, Perillo MM et al (2016) Methodology for classification of
geographical features with remote sensing images: application to tidal flats. Geomorphology
257:10–22
Revollo Sarmiento GN, Perillo GME, Delrieux CA (2020) Morphological characterization of
ponds and tidal courses in coastal wetlands using rjr. Est Coast Shelf Sci (accepted)
Sassi MG (2009) Circulación y transporte de sedimentos en estuarios. PhD dissertation.
Departamento de Física, Universidad del Centro de la Provincia de Buenos Aires (unpublished)
Sassi M, Perillo GME (2006) Influencia del viento en el perfil de velocidades de un canal de mar-
eas. VI Jornadas de Ciencias del Mar, Puerto Madryn (abstract)
Zhao K, Gong Z, Xu F et al (2019) The role of collapsed bank soil on tidal channel evolution:
a process-based model involving bank collapse and sediment dynamics. Water Resour Res.
https://doi.org/10.1029/2019WR025514
Chapter 4
Bahía Blanca Estuary: A Chemical
Oceanographic Approach

Jorge E. Marcovecchio, Ana L. Oliva, Noelia S. La Colla, Andrés H. Arias,


Sandra E. Botté, Pía Simonetti, Analía V. Serra, Vanesa L. Negrin,
Ana C. Ronda, and Claudia E. Domini

J. E. Marcovecchio ()
Instituto Argentino de Oceanografía IADO (Universidad Nacional del Sur-CONICET),
Bahía Blanca, Argentina
Facultad Regional Bahía Blanca (UTN-BHI), Universidad Tecnológica Nacional,
Bahía Blanca, Argentina
Facultad de Ingeniería, Universidad FASTA (FI-UFASTA), Mar del Plata, Argentina
Academia Nacional de Ciencias Exactas, Físicas y Naturales (ANCEFN),
Ciudad Autónoma de Buenos Aires, Argentina
e-mail: jorgemar@iado-conicet.gob.ar
A. L. Oliva · N. S. La Colla · P. Simonetti · A. V. Serra
Instituto Argentino de Oceanografía IADO (Universidad Nacional del Sur-CONICET),
Bahía Blanca, Argentina
A. H. Arias
Instituto Argentino de Oceanografía IADO (Universidad Nacional del Sur-CONICET),
Bahía Blanca, Argentina
Departamento de Química, Universidad Nacional del Sur (DQ-UNS),
Bahía Blanca, Argentina
S. E. Botté · V. L. Negrin · A. C. Ronda
Instituto Argentino de Oceanografía IADO (Universidad Nacional del Sur-CONICET),
Bahía Blanca, Argentina
Departamento de Biología, Bioquímica y Farmacia, Universidad Nacional del Sur (DBByF-­
UNS), Bahía Blanca, Argentina
C. E. Domini
Departamento de Química, Universidad Nacional del Sur (DQ-UNS),
Bahía Blanca, Argentina

© Springer Nature Switzerland AG 2021 51


S. M. Fiori, P. D. Pratolongo (eds.), The Bahía Blanca Estuary,
https://doi.org/10.1007/978-3-030-66486-2_4
52 J. E. Marcovecchio et al.

4.1  The Chemical Scenario Within the Estuary

Aquatic environments are fully characterized by a set of parameters and chemical


processes that make up the natural scenario where organisms can properly develop
(Odum 2014). The mentioned parameters include temperature, salinity, pH,
dissolved oxygen, inorganic nutrients, organic matter, other dissolved gases, and
photosynthetic pigments, among others (Roy et  al. 2011). In addition, these
parameters can show numerous processes that take place within the corresponding
biogeochemical cycles (i.e., seasonal variations, adsorption/desorption, alternative
circulation between water, sediments, organisms, and atmosphere driven by
physicochemical control) (Yakushev and Newton 2013). The interaction of these
parameters and processes determines the scenario where the whole biological
processes will occur, also including population movements along the different
spatial gradients due to physiological problems (Hester and Harrison 2007).
Moreover, the chemical scenario results in an essential framework to carry out the
biological production within the system, including the most transcendental
biological processes for the environment such as photosynthesis and respiration
(Yosim and Fry 2015), and the consequent transference of energy to higher trophic
levels (Xu et al. 2011). Finally, it is important to remark that this chemical scenario
fully and continuously interact with other components within the environment (i.e.,
hydrological, geomorphological, or climatic systems) to allow the development of
biodiversity over time (Grimm et al. 2013).
Two sets of chemical and/or physicochemical parameters could be mentioned to
build up the abovementioned scenario: (i) structural parameters and (ii) eco-­
physiological parameters (Marcovecchio and Freije 2013). The first group provides
the conditions under which biological processes develop at the stage considered and
includes parameters such as temperature, salinity, and pH/alkalinity, among others.
Meanwhile, the second group indicates the production capacity of the system under
study and includes inorganic nutrients, photosynthetic pigments, and organic matter,
among others. The integration of both groups provides an image of the health of the
ecosystem as well as the functioning of its biological components (Balvanera
et al. 2006).
Bahía Blanca Estuary has been the object of a large amount of environmental
studies along the last 40  years, including most of the scientific disciplines (e.g.,
chemistry, biology, geology, physics, hydrology, meteorology, among others).
Within this framework, a large time-series database on physicochemical parameters
has been developed at the inner part of the estuary (Puerto Cuatreros and Ingeniero
White Port; Fig.  4.1) which allows to bring a nice picture of the condition and
functioning outline of this area. It is also a very useful tool to diagnose the
environmental condition of the system considering that the mentioned inner zone of
the estuary includes not only the discharges of the main rivers from the region but
also of the largest anthropic activities such as cities, industries, and harbor locations,
with their impacts on the estuary. In addition, much information has also been
4  Bahía Blanca Estuary: A Chemical Oceanographic Approach 53

Fig. 4.1  Location of sampling stations for chemical studies within Bahía Blanca Estuary. (Map by
Walter D. Melo)

recorded through cruises along the Principal Channel of the estuary, as well as on
the external areas within the estuary (Fig. 4.1).
The present chapter has two main goals: (1) to summarize the available informa-
tion on the physicochemical conditions of the Bahía Blanca Estuary, including the
corresponding range of values of the analyzed parameters (temperature, salinity,
turbidity, dissolved oxygen, inorganic nutrients (of N, P, and Si), particulate organic
matter, and photosynthetic pigments (chlorophyll a and phaeopigments), identify-
ing seasonal variations, as well as the influence of external sources to the estuarine
balance, and (2) to present a brief summary of the state of contamination within the
estuary, including records of trace metals, hydrocarbons, pesticides, and microplas-
tics in both abiotic (estuarine water, sediments) and biological (fish species)
compartments.
54 J. E. Marcovecchio et al.

4.1.1  Some Aspects to Consider About the Estuary

Bahía Blanca Estuary (Fig. 4.1), located in the southeast coast of Argentina, is a


mesotidal system, with an elongated shape NW-SE oriented, and includes a main
navigation channel of 60 km long (Piccolo 2008). The estuary is formed by a series
of tidal channels separated by extensive tidal flats and saltmarsh patches covering an
area of ~2300 km2 (Piccolo et al. 2008). The system has been regarded as a turbid
one (Andrade et  al. 2000) with an annual mean particulate suspended matter
concentration of 78 mg.L−1 toward the inner zone (Guinder et al. 2009). The Bahía
Blanca Estuary is characterized by high salinities, usually varying between 33 and
40 PSU, even though values as low as 10 PSU have opportunely been reported (La
Colla et al. 2015).
Freshwater discharges (~241,000 m3.day−1) inflows on its northern coast, mainly
from two contributors: the Sauce Chico River and the Napostá Grande Stream
(Limbozzi and Leitào 2008). Other tributaries comprise small volumes of water,
whereas some channels discharge water only during periods of rainfall and behave
as tidal channels during the rest of the time. The Bahía Blanca Estuary also receives
groundwater contributions (overall estimated inflow of 2000 m3.day−1 according to
CTE 2003). Furthermore, the estuary receives wastewater discharges from Bahía
Blanca city, the main urban settlement located on its side, which supports ~300,000
inhabitants (INDEC 2010), with a discharge of ~70,000 m3.day−1 (Cifuentes et al.
2011). Additional freshwater inputs (i.e., Punta Alta city ~19,000  m3.day−1 of
wastewater (CTE 2003) or the industrial nucleus including petrochemical refineries
which discharges ~106,000  m3.day−1 (Cifuentes et  al. 2011)) are continuously
dumped into the Bahía Blanca Estuary.
The cities and industries enclosing the Bahía Blanca Estuary are in continuous
development (expansion and production). By 2002, the industrial area surrounding
the petrochemical center embraced only 9 industries, while in 2012, it included
more than 135 (Sznaiberg 2012). The harbor area modifies the coastal environment
through the corresponding maintenance dredging activities, the infill material from
the dredged areas, and the modification of the coastlines. So, this set of anthropogenic
stressors result in the deterioration of the Bahía Blanca estuarine environment.

4.2  M
 ain Chemical Features Within the Bahía
Blanca Estuary

A huge amount of information on the chemical aspects of the Bahía Blanca


Estuary has been collected, processed, and published in the last five decades, and
it has allowed identifying both spatial and temporal variations and distribution
trends, as well as the main biogeochemical cycles which drive the observed
performance.
4  Bahía Blanca Estuary: A Chemical Oceanographic Approach 55

4.2.1  Temperature

Estuarine water temperature was one of the usually measured parameters within the
Bahía Blanca Estuary, and the analysis of the obtained data (achieved between 1974
and 2019) has shown a very stable behavior along the Principal Channel, from the
head of the estuary and up to its mouth (Fig. 4.2) (Marcovecchio and Freije 2004).
The corresponding statistical analysis has indicated that non-significant differences
have been observed between the considered areas of the estuary (inner, middle, and
outer), allowing to sustain that estuarine temperature is governed by both the ocean
water temperature as well as the air one, which regulate its variations within the
estuary (Marcovecchio et  al. 2010a). Unlike this, freshwater discharges have not
shown to be a significant regulator of the temperature even on the area close to the
outlets (Marcovecchio 2001). Furthermore, water temperature measured in the ports
of Puerto Cuatreros and Ingeniero White (at the inner estuarine area) have
demonstrated to be strongly regulated by the air temperature from the nearby region,
and their corresponding curves of measured values have shown quite similar trend
in their distribution of values (Fig. 4.3) (Freije and Marcovecchio 2004).
During the herein analyzed period (1974–2019), the range of temperature values
recorded within the estuary water column has oscillated between 4.1 °C (on June
1983) and 27.8 °C (on January 2011). The distribution of temperature values has
followed a sinusoidal curve, which indicates the occurrence of a thermic cycle
opportunely indicated as characteristic of the estuarine conditions (James et  al.
2013; Ralston et al. 2014). This trend in coastal water temperatures is also reflected
in estuarine conditions as well as in the biological processes occurring (Barletta and
Valença Dantas 2016).
In addition, it has been observed a smooth but permanent rise in water tempera-
ture within the estuary along the last 40 years (Table 4.1). This fact fully agrees with
a general trend described for numerous coastal and estuarine systems all over the
world (i.e., Nguyen et al. 2011; Sunda and Cai 2012; Wetz and Yoskowitz 2013),
which indicates the global increase that temperature is showing within different
environments from the Earth (Hansen et al. 2010; IPCC 2018).

4.2.2  Salinity

The Bahía Blanca Estuary has been characterized as a particular estuarine environ-
ment, considering it is not associated to the outflow of big river (Piccolo and Perillo
1999). Consequently, the salinity distribution pattern does not present a sharp gradi-
ent along the estuary, which is one of the main characteristics usually applied to
describe traditional estuaries (Nguyen et  al. 2012; Cloern et  al. 2017; Bowman
2018). Nevertheless, a clear variation in salinity values could be observed at the
inner estuary, with ranges from 17.9 PSU and up to 41.3 PSU opportunely recorded
56 J. E. Marcovecchio et al.

Fig. 4.2  Distribution of physico-chemical parameters values along the Principal Channel within
Bahía Blanca Estuary. Temperature. Salinity. Dissolved oxygen and % of saturation

(Freije and Marcovecchio 2004), and allowing to characterize the Bahía Blanca
Estuary as an estuarine environment (Perillo et al. 2001).
An important fact is that the estuary behaves as “hypersaline” within its inner
area (Marcovecchio and Freije 2004) at almost every summer transforms it in a
4  Bahía Blanca Estuary: A Chemical Oceanographic Approach 57

Fig. 4.3  Annual cycle of temperature in two fixed sampling stations (Ingeniero White and Puerto
Cuatreros) located in the inner area within the Bahía Blanca Estuary

Table 4.1  Decadal variation of physico-chemical parameters’ average concentrations within


Bahía Blanca estuarine water
Decadal averages
Analyzed parameters 1974–1984 1984–1994 1994–2004 2004–2014
Temperature (°C) 13.8 ± 5.8 14.2 ± 5.5 14.4 ± 5.5 15.7 ± 5.5
Salinity (psu) 31.79 ± 3.66 33.58 ± 3.50 32.65 ± 3.60 27.12 ± 12.06
Dissolved oxygen (mg.L−1) 7.7 ± 1.5 – 7.9 ± 1.8 7.4 ± 1.6
pH – – 8,2 ± 0.48 8.2 ± 0.32
Mean value ± standard deviation

negative one during that period – in the terms as defined by de Silva Samarasinghe
and Lennon 1987, Nunez Vaz et  al. 1990, or Nunez Vaz 2012, among others,
allowing an inward flux. Furthermore, it should be highlighted that salinity has
shown a very stable behavior along the Principal Channel of the Bahía Blanca
Estuary, with its maximum variations in the inner area within the estuary (Fig. 4.2).
Finally, the observed distribution of salinity values fully agrees with previous
reports by other authors on different estuarine systems (i.e., Telesh and Khlebovich
2010; Whitfield et al. 2012; Telesh et al. 2013).

4.2.3  pH

pH measurements have presented a consistent distribution trend of values along the


estuary, with nearly constant levels within the different sampling stations. In
addition, it has been observed that the range of measured values has shown
differences, usually linked to seasonal changes and related to biological processes
(Feely et  al. 2010). In this way, highest pH values have been recorded after the
occurrence of the large phytoplankton blooms (winter and summer), reaching up to
58 J. E. Marcovecchio et al.

levels of ~9 (Fig.  4.4a) (Popovich and Marcovecchio 2008; Marcovecchio et  al.
2010a). In the analysis of the long-time series at Puerto Cuatreros and Ingeniero
White Ports, it can be observed that, even though the distribution trend is the same
at both sites, the measured values at Puerto Cuatreros Port seemed to be slightly
higher than those from Ingeniero White Port. The maximum values at both sites
have been identified on summer (January) and winter (August), at the same time the

Fig. 4.4  Distribution and variation of pH values within Bahía Blanca Estuary along the studied
period. (a) pH temporal variation for the 2004–2006 period, (b) pH distribution values along the
1996–2009 period
4  Bahía Blanca Estuary: A Chemical Oceanographic Approach 59

main phytoplankton blooms took place (Fig. 4.4a). This kind of distribution pattern
as well as the importance of this parameter values on the framework of biological
processes development have been opportunely highlighted by several authors on
different estuarine systems (Nixon et al. 2015; Villafañe et al. 2015).
Moreover, it must be highlighted that along the assessed period (1974–2019) the
interannual variability of pH was quite stable, even though the results show that
during 2001 and 2002 the occurrence of the climatic event “El Niño Southern
Oscillation” produced the highest rainfall of the decade in line with the higher pH
values. In particular, during those years and taking into account the previous
condition of acid-base balance, there was a significant change in the acidification of
water toward its alkalization. In fact, during that period, the pH increased
significantly more than 0.5 units, doubling the average standard deviation in less
than a year (Fig. 4.4b) (Arias et al. 2012). Thereafter, pH values gradually returned
to general acidification trend.

4.2.4  Turbidity

Turbidity concentrations have demonstrated to decrease from the inner area of the
estuary seaward, considering that near the head it ranges between 50 and 300 ntu,
while seaward it decreases to less than 200 ntu in the middle area of the estuary, and
at the open ocean observed values close to Bahía Blanca’s mouth are lower than
30 ntu. In this sense, two consistent processes adding suspended sediments into the
estuarine water column must be considered: (i) the main land sources are located at
the inner area (i.e., streams, rivers, sewage outfalls, harbors), and (ii) the increasing
depth from the head to the mouth of the estuary which generates higher sediment
resuspension (Cuadrado et al. 1994; Perillo et al. 2005). In addition, the analysis of
Puerto Cuatreros and Ingeniero White Port long-time data series has shown a
slightly lower mean turbidity values at Ingeniero White Port compared to Puerto
Cuatreros, even in both cases the maximum values have been recorded during
winter. This distribution trend agreed with the reports from other authors regarding
different estuaries at other latitudes, and not only from field work (Yang et al. 2014;
Watts et al. 2017) but also from remote sensing analyses (Chen et al. 2007; Garaba
and Zielinski 2015).

4.2.5  Dissolved Oxygen

The distribution of dissolved oxygen (DO) within the Bahía Blanca Estuary has
shown enough high values so as to support a valuable biological production, with
average levels close to 7 mg.L−1 and reaching up to approximately 13 mg.L−1 during
the highest productive periods (winter and late summer) (Fig.  4.2). Furthermore,
and even considering that the highest concentrations of dissolved oxygen were
60 J. E. Marcovecchio et al.

always recorded at the inner area of the estuary, the spatial distribution trend of this
parameter has shown very stable values along the whole estuary (Fig. 4.2), presenting
not any level associated to hypoxia or anoxia (Gillanders et al. 2015; Breitburg et al.
2015). This fact, together with the recognized higher productivity ability of the
system within the inner area, allows to sustain that both the mixture processes
occurring inside the estuary due to different drivers (tide, currents, waves, winds) as
well as the input of highly oxygenated seawater incoming from the continental shelf
tend to homogenize the DO levels within the system (Borja et al. 2015; Harris et al.
2015). This observation has also agreed with the corresponding distribution of the
oxygen saturation percentage, which has been homogeneous along the estuary, and
presented the highest values at the inner region, with an average value close to 85%,
and the lowest at the region where the sewage outfall from Bahía Blanca city
discharges into the estuary (Fig. 4.2). Finally, it can be mentioned that this percentage
occasionally surpasses the maximum theoretical value (100%) during the
phytoplankton blooms at the inner zone (Popovich and Marcovecchio 2008).

4.2.6  E
 cophysiological Parameters: Inorganic Nutrients–
Organic Matter–Chlorophyll a

The Bahía Blanca Estuary has been recognized as a nutrient-enriched environment,


which usually maintains high levels of these inorganic compounds along most of
the year (Freije and Marcovecchio 2004). The available studies developed since
1974 have consistently shown a general nutrients distribution trend, with their
maximum concentrations within the inner area of the estuary and slightly decreasing
toward the mouth, showing different concentrations but a similar distribution model
with no complete depletions anywhere in the studied environment (Fig. 4.5). Thus,
the oxidized nitrogen species have behaved as follows: the mean levels of nitrate
and nitrite have varied from ~10μM and ~ 2μM, respectively, at the inner area down
to ~0.1μM for both close to the estuary’s mouth. It is important to highlight that the
levels of nitrate at the inner zone are usually no less than 4–5μM, reaching up to
10–12μM, with the exception of the phytoplankton blooms’ periods (late winter and
early spring: August–September), when the concentration of this nutrient can be
close to depletion (Popovich et al. 2008). On the other hand, a very high stock of
ammonium is usually available within the system, mainly also in the inner area,
with mean values of ~20–25μM and reaching up to peaks of ~100μM (Fig.  4.5)
(Popovich et  al. 2008). This is a quite important aspect, considering that this
nitrogenated compound has never been completely depleted, and so it represents a
permanent potential stock of nitrogen for the estuary’s biological processes.
In addition, it should be highlighted that – within the above-described nitrogen
compounds’ distribution trend – these nutrients are not acting as limiting elements
of biological production in the terms defined opportunely by Moss et al. (2013) and
Kennish (2019).
4  Bahía Blanca Estuary: A Chemical Oceanographic Approach 61

Fig. 4.5  Distribution of inorganic nutrients values along the Principal Channel within Bahía
Blanca Estuary
62 J. E. Marcovecchio et al.

Unlike the nitrogenated nutrients, dissolved reactive phosphate distribution has


shown a slight decrease from the head to the mouth of the estuary, with quite stable
concentrations and variations which are usually not fully depleted (Fig.  4.5).
Nevertheless, at the inner area of the estuary where the largest phytoplankton
blooms used to occur, phosphate concentrations occasionally decreased down to ~0,
but these situations have been quickly surpassed, and the corresponding levels of
this nutrient strongly recovered up to its usual concentrations (Popovich et al. 2008;
Popovich and Marcovecchio 2008). Despite this, phosphate could eventually act as
a limiting element of the biological production at the inner area of the estuary
although for brief periods, but its recovery up to baseline levels has proven to be
very fast (Freije et al. 2008).
Finally, silicate has shown a strong decreasing trend from the head of the estuary
(mean value ~120μM) down to the mouth (mean value ~8μM) (Fig. 4.5). Nevertheless,
these values of silicate have demonstrated to be fully adequated as to support the
development of phytoplankton blooms within the estuary, most of them dominated by
diatoms which have been recognized as the most important group of Si consumers
within this environment (Popovich and Gayoso 1999; Guinder et al. 2012, 2013).
The above-described nutrients scenario together with the corresponding physico-­
chemical one have allowed to support a significant biological production which
deserves to be remarked within coastal ecosystems all over the world (Cloern and
Jassby 2010). This fact was pointed out for the Bahía Blanca Estuary since long
time ago, characterizing it as a highly productive one (i.e., Gayoso 1983, 1998a, b,
1999; Freije and Gayoso 1988; Popovich and Marcovecchio 2008; López Abbate
et  al. 2015; Guinder et  al. 2009, 2013, 2015, among others). So, the distribution
trend of photosynthesizer pigments (i.e., herein represented by chlorophyll a mean
concentrations) has also been evaluated for this system and has presented a very
consistent tendency (Fig.  4.6). Thus, a clear decreasing trend of chlorophyll a
concentration from the inner area (mean value ~10μg.L−1 with values reaching up to

Fig. 4.6  Distribution of chlorophyll a and particulate organic matter values along the Main
Navigation Channel within Bahía Blanca Estuary
4  Bahía Blanca Estuary: A Chemical Oceanographic Approach 63

~20μg.L−1) down to the outer one (mean value ~2μg.L−1) has been observed
(Fig. 4.6). However, it is important to highlight that the chlorophyll a levels within
the estuary have never been null, which indicates that the system is a permanently
productive estuary, and its lower values are similar to those usually recorded on
coastal marine waters from the Argentine Sea even during bloom periods (Rivas
et al. 2006; Romero et al. 2006; Olguín and Alder 2011; Paparazzo et al. 2017). In
addition, during the three decades from the 1970s to 1990s, chlorophyll a values of
approximately 55μg.L−1 have been recorded at different years (Popovich and
Marcovecchio 2008); this situation was modified after the big event El Niño from
2001, when structural conditions of the system changed (Arias et al. 2012). As a
consequence of the new environmental scenario, the previously described trend of
biological production was preserved both in time and in characteristics, but the
usual dominant species on the periodical blooms (the diatom Thalassiosira
curviseriata after Gayoso 1998 or Popovich and Gayoso 1999) was replaced by
Thalassiosira minima, which is significantly smaller and consequently includes less
chlorophyll molecules than the first one (Guinder et al. 2012, 2017); so, from 2001
and up to the present, maximum chlorophyll values as recorded within the estuary
have not surpassed ~12μg.L−1 (Guinder et al. 2017).
Simultaneously, with the above-described pigments distribution and concentra-
tions, the values of net primary productivity (NPP) determined at the inner area of
the estuary have reached up to ~300 mgC.m−3.h−1 (Fernández et al. 2014), which
could be mentioned between the highest records reviewed in the international litera-
ture (Sheaves et al. 2015). These high productivity levels determine the occurrence
of high concentrations of organic matter (OM) within the system (mean values
~2000 mg C.m−3all along the Principal Channel of the estuary, and with top levels
reaching up to 3680  mg C.m−3; Fig.  4.6); furthermore, the maximum OM levels
agree with the peaks of chlorophyll a, as well as with the depletion of nitrate, nitrite,
silicate, and dissolved oxygen, indicating that most of the determined particulate
organic matter (POM) is originated through biological production. In fact, the anal-
ysis of stable isotopes signature has demonstrated that Bahía Blanca Estuary phyto-
plankton was the main contributor of organic matter to the sediments, followed by
sewage and microbial mats within the system (La Colla et al. 2014). In addition,
different OM sources occur within the system, including sewage outfall discharges,
rivers, and streams, among others, which could significantly modify the available
OM stock for the system (Marcovecchio et al. 2008).

4.3  S
 tudies on Pollutants Occurrence, Levels,
and Distribution Within Bahía Blanca Estuary

Bahía Blanca Estuary is an excellent study case of potentially toxic substances


occurrence within the environment as well as related pollution processes, because it
is a large transitional environment with a great human activity within its inner area,
64 J. E. Marcovecchio et al.

including non-adequate use of soils (including unplanned urbanization), untreated


domestic or industrial sewage discharge, harbor activities, and/or incorrect solid
waste disposal, between others (Marcovecchio et al. 2008). This environment has
been particularly studied since the 1970s, including its water physico-chemical
parameters, associated biological processes, and pollutants occurrence. These
studies allow to characterize the potential effects on the system, as well as to
recognize its response ability. The information included in the next paragraphs
presents a brief overview on the occurrence, levels, and distribution trends of the
main pollutants determined within Bahía Blanca Estuary (i.e., heavy metals,
hydrocarbons, pesticides, or microplastics) in both the abiotic and the biological
compartments. The identified trends are analyzed within a historical viewpoint,
which allows pointing out evolutive processes on the estuarine environmental
quality. This information is largely useful to carry out concrete monitoring and
management programs within the estuary, and the main information recorded along
the last decades is synthesized in the next paragraphs.

4.3.1  Heavy Metals

Studies on heavy metals at Bahía Blanca Estuary (i.e., Cd, Cr, Cu, Fe, Pb, Zn)
started on the early 1980s and included data on estuarine water, sediments,
suspended particulate matter (SPM), and biota. This information has been obtained
applying internationally standardized methodologies (i.e., samples’ wet acid
mineralization followed by atomic absorption spectroscopy (AAS) or induced
coupled plasma with optical resolution (ICP-OES)) opportunely described and
compiled in different previous papers (i.e., Marcovecchio and Ferrer 2005; De
Marco et al. 2006; Botté et al. 2010a, b; Marcovecchio et al. 2014). Most of these
results were obtained within the inner area of the estuary, even additional information
exists on other regions along the system (Fig. 4.1). Analytical quality (AQ) of the
developed analysis was checked against internationally certified reference materials,
provided by the National Institute for Environmental Studies (NIES) from Tsukuba
(Japan) (mussel and pepperbush tissues, marine and estuarine sediments), as well as
by the National Institute of Standards and Technology (NIST) from Boulder,
Colorado (USA) (mussel tissue, marine sediments). Statistical comparisons were
developed using analysis of variance (ANOVA), mean values assessment (Tukey’s
test), correlation analysis, and single linear regression analysis (Sokal and
Rohlf 1995).
The main obtained results are summarized in Table 4.2, which shows interesting
trends within their distribution. So, dissolved metals have presented a spread range
of values from non-detectable ones (meaning lower than the applied analytical
method detection limit) which have been more of the analyzed samples along the
considered four decades, and up to relatively high concentrations (i.e., 6.8μg Cd.L−1,
48.6μg Cr.L−1, or 19.6μg Pb.L−1) which have been eventually recorded within the
Bahía Blanca Estuary (Table  4.2). It is important to highlight that the dissolved
4  Bahía Blanca Estuary: A Chemical Oceanographic Approach 65

fraction of trace metals indicates the recent entry of these elements into the system
taking into account their ephemeral persistency within the aquatic environment
(Kalnejais et al. 2010; Kent and Vikesland 2016). In this sense, the occurrence of
eventual inputs of dissolved metals into the estuary could be supported, although it
must be exhaustively clarified that these incomes should not be continuous and that
the possibility of the existence of metals released from those deposited in sediments
due to changes in the physico-chemical conditions that govern them must also be
considered (Namieśnik and Rabajczyk 2010; de Souza Machado et  al. 2016).
However, everything seems to indicate that the occurrence of the first mentioned
situation is much more likely than the second one (Gautam et al. 2014).

Table 4.2  Trace metals (Cd, Cu, Cr, Fe, Hg, Ni, Pb, Zn) in estuarine water (μg.L−1), surface
sediments (μg.g−1, dry weight) and fish muscle (μg.g−1, wet weight) from Bahía Blanca Estuary
Range of values (maximum–minimum)
Analyzed metals 1980–1989 1990–1999 2000–2009 2010–2019
Cadmium (Cd) Dissolved (μg.L−1) n.d.–1.86 n.d.–1.79 n.d.–1.21 n.d.–6.8
Sediments (μg.g−1, d.w.) n.d.–2.36 n.d.–3.17 n.d.–2.20 n.d.–0.94
Fish tissues (μg.g−1, w.w.) n.d.–0.46 n.d.–0.54 n.d.–0.66 n.d.–0.02
Copper (Cu) Dissolved (μg.L−1) – – 0.54–16.1 0.31–9.7
Sediments (μg.g−1, d.w.) 3.86–29.9 4.77–31.3 2.46–25.9 8.7–34.1
Fish tissues (μg.g−1, w.w.) n.d.–2.66 n.d.–2.16 n.d.–4.47 n.d.–1.70
Chromium (Cr) Dissolved (μg.L−1) – 2.04–29.4 0.75–21.1 n.d.–48.6
Sediments (μg.g−1, d.w.) 2.9–16.3 3.5–18.8 1.27–21.06 7.2–14.7
Fish tissues (μg.g−1, w.w.) n.d.–3.39 n.d.–3.84 n.d.–6.67 n.d.–2.10
Iron (Fe)a
Dissolved (μg.L−1) – – 0.01–38.6 n.d.–62.5
Sediments (mg.g−1, d.w.) 5.9–51.5 6.1–43.3 4.7–36.6 5.9–82.2
Fish tissues (μg.g−1, w.w.) – – – –
Mercury (Hg) Dissolved (ng.L−1)
b
n.d. n.d.–1.11 n.d.–2.07 n.d.–1.96
Sediments (ng.g−1, d.w.) 10–1670 10–720 n.d.–200 n.d.–220
Fish tissues (ng.g−1, w.w.) 220–3970 200–640 90–120 n.d.–100
Nickel (Ni) Dissolved (μg.L−1) – – n.d.–9.5 n.d.–1.35
Sediments (μg.g−1, d.w.) – – 1.75–26.1 4.5–14.4
Fish tissues (μg.g−1, w.w.) – – n.d.–2.08 n.d.–0.66
Lead (Pb) Dissolved (μg.L−1) n.d.–10.4 n.d.–6.55 n.d.–15.7 n.d.–19.6
Sediments (μg.g−1, d.w.) 4.76–31.05 3.87–26.65 n.d.–42.71 4.3–19.3
Fish tissues (μg.g−1, w.w.) n.d.–3.02 n.d.–3.83 n.d.–5.94 n.d.–1.70
Zinc (Zn) Dissolved (μg.L−1) n.d.–50.6 n.d.–73.6 n.d.–71.3 n.d.–54.7
Sediments (μg.g−1, d.w.) 19.6–103.7 21.75–706.1 14.2–98.5 21.4–82
Fish tissues (μg.g−1, w.w.) n.d.–11.2 0.2–7.13 n.d.–8.09 0.9–30.1
After Marcovecchio (1988), Marcovecchio et al. (1986, 1988a, b, c), Botté et al. (2008, 2010a, b),
IADO (1997, 1999, 2002, 2006, 2008, 2009, 2010, 2012, 2014, 2016, 2018), Marcovecchio and
Ferrer (2005), Marcovecchio et al. (2010b, 2016), Simonetti et al. (2017), Buzzi and Marcovecchio
(2018) and La Colla et al. (2018a, b, 2019)
Range of values (minimum–maximum) along the 1980–2019 period
n.d.: values below the detection limit of the applied analytical method
a
Values in (mg.g−1); bValues in (ng.g−1)
66 J. E. Marcovecchio et al.

A totally different trend was observed in the distribution of trace metals in sedi-
ments of the estuary. In this case, most of the metals studied were permanently
detected in that compartment, and only a few exceptions (i.e., Cd) showed values
below the corresponding detection limits (Table  4.2). Although the range of
concentrations of metals measured in sediments of the estuary has been very wide
and varied, the corresponding average values can be classified as intermediate ones
(Cairns Jr. 2009), indicating the existence of anthropogenic sources that input them
into the system but without reaching critical levels for organisms living there
(Sindermann 2006). These sediments’ contamination status due to their trace metal
content was checked against opportunely reported Geo-accumulation Index (Igeo)
and Enrichment Factor (EF) within Bahía Blanca Estuary (Fernández Severini et al.
2018, following Förstner et al. 1990) as well as comparison with different Sediment
Quality Guidelines (i.e., Canadian, European Union, FAO, among others)(Roach
2005; Hübner et  al. 2009). In this latter sense, it must be considered that these
Quality Guides must be generated for particular environments or regions, and
therefore are not universalizable (Kwok et al. 2013). Thus, the most convenient is to
generate own guides for each environment or region evaluated (Förstner et al. 1990).
Finally, when trace metals distribution within fishes from Bahía Blanca Estuary
was considered, a variable trend has been observed but showing that most of studied
metals seemed to be present within the analyzed tissues (Table 4.2). In all cases the
obtained results include numerous non-detectable values, which means that the
corresponding environmental exposure to this element is far to critical levels
(Jezierska et al. 2009; Govind and Madhuri 2014). In any case, the levels of metals
determined in the tissues of the analyzed fish demonstrate the existence of exposure
to these compounds (eventual or periodic), as well as the corresponding biological
accumulation processes opportunely described for this environment (Marcovecchio
et al. 1986, 1988a, b; De Marco et al. 2006; La Colla et al. 2018b, 2019; among
others). An interesting point to be considered has been the decrease in Hg
concentrations in fish tissues that occurred from the 1980s to the present, and which
meant a significant environmental recovery of these species (Marcovecchio et al.
2001; De Marco et al. 2006).

4.3.2  H
 ydrocarbons, Including Polycyclic Aromatic
Hydrocarbons (PAHs)

PAHs can be originated from three possible sources: petrogenic, pyrolytic and natu-
ral. Although PAHs can be originated naturally, anthropogenic activities are gener-
ally considered the major source of PAHs release into the environment. There are
important differences in the chemical composition of PAHs mixtures depending on
the sources of emission. In general, pyrolytic PAHs present dominance of high
molecular mass compounds (corresponding to four to six rings compounds), while
petrogenic PAHs present dominance of two/three rings compounds. Moreover,
4  Bahía Blanca Estuary: A Chemical Oceanographic Approach 67

pyrolysis process produces PAHs associated to soot carbon. A quite stronger


association and persists during is deposited in aquatic system affecting partitioning
and bioavailability of PAHs.
The Bahía Blanca Estuary has a large history of addressing sources and origin of
this type of hydrocarbons, since Arias et al. (2011), followed by Arias et al. (2009)
and Oliva et  al. (2015), between others. PAHs can be introduced into the Bahía
Blanca Estuary by different ways: for instance, spillage of fossil fuels, ship traffic,
atmospheric depositions, urban runoff, and municipal and industrial wastewater
discharge. From these, the atmospheric transport is the most important pathway for
their environmental distribution. Particle-bound PAHs can be transported long
distances, from the emission source as far as remote areas where they are removed
from the atmosphere through precipitation and dry deposition. Recently, Orazi et al.
tracked the sources and distribution of atmospheric PAHs around the Bahía Blanca
Estuary (Orazi et al. 2020). Results showed a range from 27.97 to 1052.99 ng. m−3
and from 52.40 to 2118.34 ng. g−1 d.w. for air and soil samples, respectively. The
highest air-PAHs levels were registered in Bahía Blanca city (1052.99  ng. m−3,
d.w.), confirming the hypothesis of a significant PAHs source impacting Bahía
Blanca Estuary through the air. Atmospheric PAHs sources are dominated by
pyrolytic origin: high density of traffic, coal combustion, residential emissions, and
industrial activities were tagged as the principal diffuse sources. Finally, PAHs have
been shown to bioaccumulate in the BBE biota, i.e., mussels and pelagic fish (Arias
et al. 2010; Oliva et al. 2015). Figure 4.7 shows the source apportionment of PAHs
in biota at the Bahía Blanca Estuary.

4.3.3  Pesticides

When organochlorine pesticides (OCPs) enter the ocean, they tend to accumulate in
biota and bottom sediments. Smaller organisms incorporate them primarily through
their respiratory surfaces while larger animals do so through food intake (UNEP
2002). Due to their fat solubility and high persistence to biological degradation,
they are biomagnified along the trophic web (UNEP 2002; Roche et  al. 2009;
Lailson-Brito et al. 2010); so the higher trophic levels, such as marine mammals,
top predatory fish or seabirds, usually show the highest OCPs concentrations. The
biota could be affected to toxic effects caused by short- or long-term exposure to
these pesticides. These effects include reproductive damage (Gross et  al. 2002),
endocrine disruption, immune suppression, and cancer, among others (Bergman
et al. 2012; Menzies et al. 2013).
In South America, most of the OCPs were intensively used and produced,
between 1950 and 1990, except Endosulfan, Dicofol, and Methoxychlor, which
were used until recently. During their use, a large part of the OCPs entered the
continental environment and subsequently into the sea, and although all the South
American countries banned or restricted their use and production, detectable and
even harmful environmental concentrations are likely to remain present in the seas
68 J. E. Marcovecchio et al.

Fig. 4.7  Source origin plot indicating main sources at the Bahía Blanca Estuary biota. (a) Mussel’s
IP/276, BaA/228, and An/178 source ratio plot vs Fl/202 ratio, showing different origin zones. (b)
PCA score plot of biota samples. (After Arias et al. 2010)
4  Bahía Blanca Estuary: A Chemical Oceanographic Approach 69

and oceans for several years due to the intrinsic characteristics and environmental
dynamics of the OCPs (UNEP 2002; Iwata et al. 1993; Wania et al. 2006).
The Bahía Blanca Estuary is located in one of the most important agriculture
regions of Argentina. Large amounts of organochlorine compounds (OCs) are
known to be used in areas located close to the estuary basin (Renella and Quirós
2000), and the abundant water and sediments loads of Sauce Chico River, Maldonado
River, Napostá Grande Stream, and other tributaries may play an important role on
the transport and distribution of OCs. In spite of this, in the history of the Bahía
Blanca Estuary, apart from Sericano and Pucci (1984), there was no OC pesticides
information or previous studies on this topic until 2010 (Arias et al. 2010). From
then, frequent monitoring studies could set average values and OCs entries to the
estuary. It has been demonstrated that the estuary sediments are functioning as a
sink for persistent OC pesticides (or their metabolites) that are either in use or have
been recently banned for agricultural purposes, pointing out an extensive use of
OCs in the catchments of the Bahía Blanca Estuary in the past and a present use of
some, also revealing a declining trend in the environmental burden of persistent
pesticides. The main OC access way to the Bahía Blanca Estuary suggested to
function by land runoff and subsequent transport to the estuary by draining tributaries
and/or evaporation and atmospheric wet deposition (Arias et  al. 2010; Tombesi
et al. 2018; Gironés et al. 2020).
DDT and HCH residues are present, but in the low range of levels in comparison
to other worldwide locations as well as to the world coastal sediment concentrations
(Fig. 4.8). Unlike this, ∑OCs average is in the medium range, indicating that the
Bahía Blanca Estuary is not grossly polluted by HCHs or DDTs and that a significant
proportion of OCs pollution is provided by other pesticides (i.e., Endosulfan).

4.3.4  Microplastics

Microplastics (MPs) have passed from being considered emerging pollutants to be


recognized as an emerged threat, with the urgent need to better assess their
distribution in the marine environment, as well as the ecotoxicological and ecological
risks that they have (Avio et  al. 2017). Although MPs have been considered as
vectors for other pollutants (Rochman 2015; Hartmann et al. 2017), these particles
have also been determined themselves as Persistent Organic Pollutants (POP) in the
Convention on POPs (UNEP 2001) due to their characteristics of persistence,
bioaccumulation, long-range transport, and adverse effects. The study of the
presence and distribution of MPs in marine environments has exponentially
increased in recent years; however, there are very few studies in the Bahía Blanca
Estuary. The first evidence was recently published in the gastrointestinal tract of the
marine commercial fish Micropogonias furnieri (Arias et al. 2019). The investigations
of this work focused on two areas of the Bahía Blanca Estuary, finding more particles
in the specimens from the internal zone than in the middle zone of the estuary.
Although the abundance of ingested MPs could reflect variations in the quantity and
70 J. E. Marcovecchio et al.

Fig. 4.8 (a) Example of DDT and degradative isomers spatial distribution at the Bahía Blanca
Estuary. (b) HCH’s spatial distribution over the sampled stations. Schematically, main sewage,
industrial, and freshwater inputs are indicated

type of food consumed between individuals of the same species and between diverse
species, the fact that the same species and length class was analyzed pointed out a
different human anthropic pressure along the Bahía Blanca Estuary with respect to
this pollutant. The second work published for this area was in the same year and
demonstrated, for the first time in the Bahía Blanca Estuary, the existence of MPs in
bottom sediments and surface seawater, and it was determined at the outermost area
of the estuary (Ronda et  al. 2019). Consistently with previous wide world
environmental studies, MPs levels detected in Bahía Blanca Estuary were
comparable to other marine environments, consisted mainly by fibers and the most
abundant size was less than 1 mm (Rochman 2015; Gago et al. 2018). The fact that
4  Bahía Blanca Estuary: A Chemical Oceanographic Approach 71

it has been evidenced MPs contamination in several matrices along the Bahía Blanca
Estuary (organisms, sediment, and seawater) argues the need to continue conducting
research that not only evaluates the presence and distribution of these particles but
also the ecotoxicological and ecological effects that they could have on the Bahía
Blanca Estuary ecosystem.

4.4  B
 rief Comment on the Chemical Functioning
of the Estuary

Bahía Blanca Estuary is a very large system, whose functioning is clearly character-
ized by several processes that on the whole determine the success of the biological
productivity occurring there.
The distribution of the structural parameters within the system is very stable,
mainly in terms of temperature, pH, and turbidity. Salinity presents a relative
stability, even though significant variations occur at the inner area, which could
alternatively be increased or decreased according to the season.
The estuary is usually highly nutrient enriched, and levels of compounds of
nitrogen (basically ammonium) are always available, even though the concentrations
of oxidized nitrogen compounds (NO2− and NO3−) and phosphorus are eventually
used down to close-depletion during the periodical phytoplankton blooms.
Nevertheless, the recovery of these consumed nutrients through mineralization of
organic matter is extremely fast. In addition, a large stock of silicate is usually
available in the estuary, mainly at the inner area, which is completely adequate to
support the biological demand within the system.
Concerning the biological primary production, the most important annual period
is the late winter  – early spring  – when the highest phytoplankton blooms have
historically occurred. It is developed because during this time the nutrients (N, P,
and Si) are largely available, and both the temperature and light intensity are
sufficiently low (~5 to 7 °C, and 400–700μE m−2 s−1, respectively; after Popovich
et al. 2008) as required by the dominant diatom species responsible for the mentioned
bloom. Thus, very high levels of chlorophyll a were detected during this phenomenon
(with values reaching up to 55 mg.m−3), representing densities of ~13 × 106 cells L−1
or net primary productivities of ~300 mgC.m−3.h−1.
The very high amounts of organic matter generated by these biological processes
ensure the regenerated nutrients production, through mineralization processes
occurring within the estuary (Freije et  al. 2008). The obtained results seemed to
indicate that a predominant liberation of ammonium was observed within the
estuarine sediments at the inner estuary, even significant amounts of oxidized
nitrogen compounds (NO2− and NO3−) were eventually also produced (Popovich
et al. 2008). In addition, these are the first nitrogenated nutrients to be consumed
during the phytoplankton bloom, and just when both NO2− and NO3− were depleted,
the NH4+ started to be consumed (Popovich et al. 2008).
72 J. E. Marcovecchio et al.

On the other hand, different kind of pollutants have eventually been recorded
within Bahía Blanca Estuary, mainly in its inner area. Trace metals have been
determined in water, sediments, and fish tissues within the estuary, even their
concentrations were not within recognized critical values. Several of them have
been recorded in the dissolved phase meaning that presumably sources of these
elements are linked to the estuary, being able to discharge them both continuously
or eventually. Sediments within Bahía Blanca Estuary have demonstrated to function
as sink of trace metals, accumulating them at different rates but along the whole
system. This fact, joint to corresponding physico-chemical conditions, fully governs
the probabilities of trace metals input into Bahía Blanca Estuary biota.
As expected to a heavily anthropized environment, a widespread state of PAHs
pollution has been shown to occur at the Bahía Blanca Estuary. Levels range from
negligible (rural) to high in several hotspots, including harbor/industrial facilities
(sediments) and dense urban environments (air).
DDT and HCH residues are present, but in the low range of levels in comparison
to other worldwide locations as well as to the world coastal sediment concentrations.
On the opposite, ∑OCs average is in the medium range, indicating that the Bahía
Blanca Estuary is not grossly polluted by HCHs or DDTs and that a significant
proportion of OCs pollution is provided by other pesticides (i.e., Endosulfan).
Microplastic pollution has been assessed in several matrices along the Bahía
Blanca Estuary (organisms, sediment, and seawater) arguing the need to perform
further research evaluating their ecotoxicological and ecological effects.
Bahía Blanca Estuary’s productive cycle, regulated through bio-geochemical
joint processes, has well functioned for a long time (at least during the last 40 years,
when these studies started at the earlier 1970s). During early 2000 a strong
modification on the composition of the phytoplankton bloom which characterizes
Bahía Blanca Estuary was identified; nevertheless, the system continued to
harmoniously work without showing significant changes in its bio-geochemical
condition. Even though this environment is not strongly stressed by pollutants,
different kinds of potentially toxic compounds have been repeatedly determined,
even at low concentrations. Consequently, this is a nice scenario to control the
evolution and progress of the mentioned estuarine chemical processes, as well as to
monitor the potential occurrence of changes within the identified trends.

References

Andrade S, Pucci AE, Marcovecchio JE (2000) Cadmium concentrations in the Bahía Blanca estu-
ary (Argentina). Potential effects of dissolved cadmium on the diatom Thalassiosira curvise-
riata. Oceanologia 42:505–520
Arias AH (2008) Comportamiento de los hidrocarburos aromáticos policíclicos (PAHs) en aguas,
sedimentos y organismos de la zona interna del estuario de Bahía Blanca. Doctoral Thesis,
Universidad Nacional del Sur (UNS). Bahía Blanca (Argentina), 178 pp
Arias AH, Spetter CV, Freije RH et al (2009) Polycyclic aromatic hydrocarbons in water, mus-
sels (Brachidontes sp., Tagelus sp.) and fish (Odontesthes sp.) from Bahía Blanca Estuary,
Argentina. Estuar Coast Mar Sci 85(1):67–81. https://doi.org/10.1016/j.ecss.2009.06.008
4  Bahía Blanca Estuary: A Chemical Oceanographic Approach 73

Arias AH, Vázquez-Botello A, Tombesi N et al (2010) Presence, distribution, and origins of poly-
cyclic aromatic hydrocarbons (PAHs) in sediments from Bahía Blanca estuary, Argentina.
Environ Monit Assess 160(1–4):301–314. https://doi.org/10.1007/s10661-­008-­0696-­5
Arias AH, Pereyra MT, Marcovecchio JE (2011) Multi-year monitoring of estuarine sediments
as ultimate sink for DDT, HCH, and other organochlorinated pesticides in Argentina. Environ
Monit Assess 172(1–4):17–32. https://doi.org/10.1007/s10661-­010-­1315-­9
Arias AH, Piccolo MC, Spetter CV et al (2012) Lessons from multi-decadal oceanographic moni-
toring at an estuarine ecosystem in Argentina. Int J Environ Res 6(1):219–234. https://doi.
org/10.22059/IJER.2011.488
Arias AH, Ronda A, Oliva AL et al (2019) Evidence of microplastic ingestion by fish from the
Bahía Blanca estuary in Argentina, South America. B Environ Contam Tox 102(6):750–756.
https://doi.org/10.1007/s00128-­019-­02604-­2
Avio CG, Gorbi S, Regoli F (2017) Plastics and microplastics in the oceans: from emerg-
ing pollutants to emerged threat. Mar Environ Res 128:2–11. https://doi.org/10.1016/j.
marenvres.2016.05.012
Balvanera P, Pfisterer AB, Buchmann N et  al (2006) Quantifying the evidence for biodiver-
sity effects on ecosystem functioning and services. Ecol Lett 9:1146–1156. https://doi.
org/10.1111/j.1461-­0248.2006.00963.x
Barletta M, Valença Dantas D (2016) Environmental gradients. In: Kennish MJ (ed) Encyclopedia
of estuaries. Springer, Heidelberg, pp 237–241. 778 pp. ISBN:978-94-017-8800-7
Bergman A, Heindel JJ, Jobling S et  al (eds) (2012) State of the science of endocrine disrupt-
ing chemicals. UNEP (United Nations Environment Programme)-WHO (World Health
Organization), Geneve. 213 pp. ISBN:978-92-807-3274-0
Borja A, Dauer DM, Elliott M et  al (2015) Medium- and long-term recovery of estuarine and
coastal ecosystems: patterns, rates and restoration effectiveness. Estuaries Coast 33:1249–1260.
https://doi.org/10.1007/s12237-­010-­9347-­5
Botté SE, Freije RH, Marcovecchio JE (2008) Dissolved heavy metal (Cd, Pb, Cr, Ni) concentra-
tions in surface water and porewater from Bahia Blanca estuary tidal flats. B Environ Contam
Tox 79(4):415–421. https://doi.org/10.1007/s00128-­007-­9231-­6
Botté SE, Delucchi F, Freije RH et al (2010a) Chapter 11: Cadmium and organotin pollution in
an estuarine environment from Argentina: and overview. In: El Nemr A (ed) Impact, monitor-
ing and management of environmental pollution. Nova Publishers, New  York, pp  263–283.
ISBN:978-1-60876-487-7
Botté SE, Freije RH, Marcovecchio JE (2010b) Distribution of several heavy metals in tidal flats
sediments within Bahía Blanca Estuary (Argentina). Water Air Soil Pollut 210:371–388.
https://doi.org/10.1007/s11270-­009-­0260-­0
Bowman MJ (2018) Chapter 7: Estuarine fronts. In: Kjerfve B (ed) Hydrodynamics of estuaries,
Vol.1: Estuarine physics. CRC Press  – Taylor and Francis Group, Boca Ratón, pp  85–131.
227 pp. ISBN:0-8493-4370-4
Breitburg DL, Hondorp D, Audemard C et al (2015) Landscape-level variation in disease suscep-
tibility related to shallow-water hypoxia. PLoS One 10(2):e0116223. https://doi.org/10.1371/
journal.pone.0116223
Buzzi NS, Marcovecchio JE (2018) Heavy metal concentrations in sediments and in mussels from
Argentinean coastal environments, South America. Environ Earth Sci 77:321–333. https://doi.
org/10.1007/s12665-­018-­7496-­1
Cairns J Jr (2009) Chapter 1: The influence of contaminated sediments on sustainable use of
the planet. In: Kassim TA, Barceló D (eds) Contaminated sediments. Springer, Heidelberg,
pp 1–19. 193 pp. ISBN: 978-3-540-88013-4
Chen Z, Hu C, Muller-Karger F (2007) Monitoring turbidity in Tampa Bay using MODIS/Aqua
250-m imagery. Remote Sens Environ 109:207–220. https://doi.org/10.1016/j.rse.2006.12.019
Cifuentes O, Escudero D, Medus S et al (2011) Estudio de la dinámica (espacial y temporal) de
los efluentes líquidos industriales y urbanos en la zona del Polo Petroquímico y área portu-
aria de Bahía Blanca. En: Puliefito E (ed) Contaminación atmosférica e hídrica en Argentina.
(EdUTecNe), Buenos Aires, pp 539–545
74 J. E. Marcovecchio et al.

Cloern JE, Jassby AD (2010) Patterns and scales of phytoplankton variability in estuarine-coastal
ecosystems. Estuar Coasts 33:230–241. https://doi.org/10.1007/s12237-­009-­9195-­3
Cloern JE, Jassby AD, Tara S et al (2017) Ecosystem variability along the estuarine salinity gra-
dient: examples from long-term study of San Francisco Bay. Limnol Oceanogr 62:272–291.
https://doi.org/10.1002/lno.10537
CTE (Comité Técnico Ejecutivo – Municipalidad de Bahía Blanca) (2003) Programa integral de
monitoreo Bahía Blanca. http://www.bahiablanca.gov.ar/areas-­de-­gobierno/medioambiente/
comite-­tecnicoejecutivo/
Cuadrado D, Perillo GME, Marcos A (1994) Análisis preliminar del transporte del sedimento en
suspension en Puerto Rosales. V Reunión Argentina de Sedimentología:229–234
De Marco SD, Botté SE, Marcovecchio JE (2006) Mercury distribution in abiotic and biologi-
cal compartments within several estuarine systems from Argentina: 1980–2005 period.
Chemosphere 65(2):213–233. https://doi.org/10.1016/j.chemosphere.2006.02.059
de Silva Samarasinghe JR, Lennon GW (1987) Hypersalinity, flushing and transient salt-­
wedges in a tidal gulf: an inverse estuary. Estuar Coast Shelf Sci 24:483–498. https://doi.
org/10.1016/0272-­7714(87)90129-­6
de Souza Machado AA, Spencer K, Kloas W et al (2016) Metal fate and effects in estuaries: a review
and conceptual model for better understanding of toxicity. Sci Total Environ 541:268–281.
https://doi.org/10.1016/j.scitotenv.2015.09.045
Feely RA, Alin SR, Newton J et al (2010) The combined effects of ocean acidification, mixing,
and respiration on pH and carbonate saturation in an urbanized estuary. Estuar Coast Shelf Sci
88:442–449. https://doi.org/10.1016/j.ecss.2010.05.004
Fernández Severini MD, Carbone ME, Villagran DM et al (2018) Toxic metals in a highly urban-
ized industry-impacted estuary (Bahía Blanca Estuary, Argentina): spatio-temporal analysis
based on GIS. Environ Earth Sci 77(10):393–416. https://doi.org/10.1007/s12665-­018-­7565-­5
Fernández EM, Garzón CJE, Martínez AM et al (2014) Estimación de la productividad primaria
del fitoplancton en la zona interna del estuario de Bahía Blanca. In: Marcovecchio JE, Botté SE,
Freije RH (eds) Procesos Geoquímicos Superficiales en Iberoamérica. Sociedad Iberoamericana
de Física y Química Ambiental (SIFyQA), Salamanca, pp 261–278. ISBN:978-84-937437-6-5
Förstner U, Ahlf W, Calmano W et  al (1990) Part II, Chapter 18: Sediments criteria develop-
ment. In: Heling D, Rothe P, Förstner U et al (eds) Sediments and environmental geochemistry.
Springer, Berlin, pp 311–338. 381 pp. ISBN-13:978-3-642-75099-1
Freije RH, Gayoso AM (1988) Producción primaria del estuario de Bahía Blanca. Informes
UNESCO, Ciencias del Mar 47:112–114
Freije RH, Marcovecchio JE (2004) Chapter 8: Oceanografía química del estuario de Bahía
Blanca. En: Piccolo MC, Hoffmeyer MS (eds) El ecosistema del estuario de Bahía Blanca.
Instituto Argentino de Oceanografía (IADO  – CONICET/UNS), Bahía Blanca, 69–78.
ISBN:987-9281-96-9
Freije RH, Spetter CV, Marcovecchio JE et al (2008) Chapter 23: Water chemistry and nutrients of
the Bahía Blanca Estuary. In: Neves R, Baretta J, Mateus M (eds) Perspectives on Integrated
Coastal Zone Management in South America. Part B: From shallow water to the deep fjord: the
study sites. IST Scientific Publishers, Lisbon, pp 243–256. ISBN:978-972-8469-74-0
Gago J, Carretero O, Filgueiras AV, Viñas L (2018) Synthetic microfibers in the marine envi-
ronment: A review on their occurrence in seawater and sediments. Marine pollution bulletin,
127:365–376
Garaba SP, Zielinski O (2015) An assessment of water quality monitoring tools in an estuarine
system. Remote Sens Environ 109(2):207–220. https://doi.org/10.1016/j.rse.2006.12.019
Gautam RK, Sharma SK, Suresh Mahiya S et al (2014) Chapter 1: Contamination of heavy met-
als in aquatic media: transport, toxicity and technologies for remediation. In: Sharma SK
(ed) Heavy metals in water – presence, removal and safety. The Royal Society of Chemistry,
Cambridge, pp 1–24. 380 pp. ISBN:978-1-84973-885-9
Gayoso AM (1983) Estudio de Fitoplancton del Estuario de Bahía Blanco:(Pcia. de Buenos Aires,
Argentina). Zona interna. Puerto Cuatreros. Studia oecológica (2):73–88
4  Bahía Blanca Estuary: A Chemical Oceanographic Approach 75

Gayoso AM (1998) Long-term phytoplankton studies in the Bahías Blanca estuary, Argentina.
ICES Journal of Marine Sciences 55:655–660
Gayoso AM (1998a) Estudio de Fitoplancton del Estuario de Bahía Blanco (Pcia. de Buenos Aires,
Argentina). Zona interna: Puerto Cuatreros. Stud Oecol 2(2):73–88. ISSN:0211-4623
Gayoso AM (1998b) Long-term phytoplankton studies in the Bahía Blanca Estuary, Argentina.
ICES J Mar Sci 55:655–660. https://doi.org/10.1006/jmsc.1998.0375
Gayoso AM (1999) Seasonal succession patterns of phytoplankton in the Bahía Blanca Estuary
(Argentina). Bot Mar 42(4):367–375. https://doi.org/10.1515/bot.1999.042
Gillanders BM, Elsdon TS, Halliday IA et  al (2015) Potential effects of climate change on
Australian estuaries and fish utilising estuaries: a review. Mar Freshw Res 62:1115–1131.
https://doi.org/10.1071/MF11047
Gironés L, Oliva AL, Marcovecchio JE et al (2020) Spatial distribution and ecological risk assess-
ment of residual organochlorine pesticides (OCPs) in South American marine environments.
Curr Environ Health Rep:1–14. https://doi.org/10.1007/s40572-­020-­00272-­7
Govind P, Madhuri S (2014) Heavy metals causing toxicity in animals and fishes. Res J Anim Vet
Fish Sci 2(2):17–23. ISSN:2320-6535
Grimm NB, Stuart Chapin FS, Bierwagen B et al (2013) The impacts of climate change on ecosys-
tem structure and function. Front Ecol Environ 11(9):474–482. https://doi.org/10.1890/120282
Gross TS, Arnold BS, Sepúlveda MS et al (2002) Chapter 39: Endocrine disrupting chemicals and
endocrine active agents. In: Hoffman DJ, Rattner B, Burton GAJ et al (eds) Handbook of eco-
toxicology. CRC Press, Boca Raton, pp 1033–1098. 1315 pp. ISBN:1-56670-546-0
Guinder VA, Popovich CA, Perillo GME (2009) Particulate suspended matter concentrations in the
Bahía Blanca estuary, Argentina: implication for the development of phytoplankton blooms.
Estuar Coast Shelf Sci 85(1):157–165. https://doi.org/10.1016/j.ecss.2009.05.022
Guinder VA, Molinero JC, Popovich CA et  al (2012) Dominance of the planktonic diatom
Thalassiosira minima in recent summers in the Bahía Blanca Estuary, Argentina. J Plankton
Res 34(11):995–1000. https://doi.org/10.1093/plankt/fbs060
Guinder VA, Popovich CA, Molinero JC et al (2013) Phytoplankton summer bloom dynamics in
the Bahía Blanca Estuary in relation to changing environmental conditions. Cont Shelf Res
52:150–158. https://doi.org/10.1016/j.csr.2012.11.010
Guinder VA, López-Abbate MC, Berasategui AA et al (2015) Influence of the winter phytoplank-
ton bloom on the settled material in a temperate shallow estuary. Oceanologia 57:50–60.
https://doi.org/10.1016/j.oceano.2014.10.002
Guinder VA, Molinero JC, López Abbate MC et  al (2017) Phenological changes of blooming
diatoms promoted by compound bottom-up and top-down controls. Estuar Coast 40:95–104.
https://doi.org/10.1007/s12237-­016-­0134-­9
Hansen J, Ruedy R, Sato M et  al (2010) Global surface temperature change. Rev Geophys
48:RG4004. https://doi.org/10.1029/2010RG000345
Harris LA, Hodgkins CLS, Day MC et  al (2015) Optimizing recovery of eutrophic estuaries:
impact of estratification and re-aeration on nutrient and dissolved oxygen dynamics. Ecol Eng
75:470–483. https://doi.org/10.1016/j.ecoleng.2014.11.028
Hartmann NB, Rist S, Bodin J et al (2017) Microplastics as vectors for environmental contami-
nants: exploring sorption, desorption, and transfer to biota. Integr Environ Asses 13(3):488–493.
https://doi.org/10.1002/ieam.1904
Hester RE, Harrison RM (eds) (2007) Biodiversity under threat. Royal Society of Chemistry,
Cambridge. 291 pp. ISBN:978-0-85404-251-7
Hübner R, Astin KB, Herbert RJH (2009) Comparison of sediment quality guidelines (SQGs) for
the assessment of metal contamination in marine and estuarine environments. J Environ Monit
11:713–722. https://doi.org/10.1039/b818593j
IADO (1999) Programa de monitoreo de la calidad ambiental de la zona interior del estuario de
Bahía Blanca. Informe Final, Instituto Argentino de Oceanografía, 25 + 68 pp. Available in:
http://www.bahiablanca.gov.ar/cte/index.html
IADO (2002) Programa de monitoreo de la calidad ambiental de la zona interior del estuario de
Bahía Blanca. Informe Final, Instituto Argentino de Oceanografía: 76 pp. Available in.: http://
www.bahiablanca.gov.ar/cte/index.html
76 J. E. Marcovecchio et al.

IADO (2006) Programa de monitoreo de la calidad ambiental de la zona interior del estuario de
Bahía Blanca. Informe Final, Instituto Argentino de Oceanografía, 82 pp. Available in: http://
www.bahiablanca.gov.ar/cte/index.html
IADO (2008) Programa de monitoreo de la calidad ambiental de la zona interior del estuario de
Bahía Blanca. Informe Final, Instituto Argentino de Oceanografía, 107 pp. Available in: http://
www.bahiablanca.gov.ar/cte/index.html
IADO (2009) Programa de monitoreo de la calidad ambiental de la zona interior del estuario de
Bahía Blanca. Informe Final, Instituto Argentino de Oceanografía, 103 pp. Available in: http://
www.bahiablanca.gov.ar/cte/index.html
IADO (2010) Programa de monitoreo de la calidad ambiental de la zona interior del estuario
de Bahía Blanca. Adenda al Informe Final 2009, Instituto Argentino de Oceanografía, 92 pp.
Available in: http://www.bahiablanca.gov.ar/cte/index.html
IADO (2012) Programa de monitoreo de la calidad ambiental de la zona interior del estuario de
Bahía Blanca. Informe Final, Instituto Argentino de Oceanografía, 122 pp. Available in: http://
www.bahiablanca.gov.ar/cte/index.html
IADO (2014) Programa de monitoreo de la calidad ambiental de la zona interior del estuario de
Bahía Blanca. Informe Final, Instituto Argentino de Oceanografía, 242 pp. Available in: http://
www.bahiablanca.gov.ar/cte/index.html
IADO (2016) Programa de monitoreo de la calidad ambiental de la zona interior del estuario de
Bahía Blanca. Informe Final, Instituto Argentino de Oceanografía, 228 pp. Available in: http://
www.bahiablanca.gov.ar/cte/index.html
IADO (2018) Programa de monitoreo de la calidad ambiental de la zona interior del estuario de
Bahía Blanca. Informe Final, Instituto Argentino de Oceanografía, 364 pp. Available in: http://
www.bahiablanca.gov.ar/cte/index.html
IADO (Instituto Argentino de Oceanografía, CONICET/UNS) (1997) Programa de monitoreo de
la calidad ambiental de la zona interior del estuario de Bahía Blanca. Informe Final, Instituto
Argentino de Oceanografía, 65 pp. Available in: http://www.bahiablanca.gov.ar/cte/index.html
INDEC (2010) Instituto Nacional de Estadística y Censos, Argentina. http://www.indec.gov.ar
IPCC (Intergovernmental Panel on Climate Change) (2018) Global warming of 1.5°C: an IPCC
special report on the impacts of global warming of 1.5°C above pre-industrial levels and related
global greenhouse gas emission pathways, in the context of strengthening the global response
to the threat of climate change, sustainable development, and efforts to eradicate poverty.
Masson-Delmotte V, Zhai P, Pörtner H-O et al (eds) (in press)
Iwata H, Tanabe S, Sakai N et al (1993) Distribution of persistent organochlorines in the oceanic
air and surface seawater and the role of ocean on their global transport and fate. Environ Sci
Technol 27:1080–1098. https://doi.org/10.1021/es00043a007
James NC, van Niekerk L, Whitfield AK, Potts WM, Götz A, Paterson AW (2013) Effects of cli-
mate change on South African estuaries and associated fish species. Clim Res 57:233–248.
https://doi.org/10.3354/cr01178
Jezierska B, Ługowska K, Witeska M (2009) The effects of heavy metals on embryonic devel-
opment of fish (a review). Fish Physiol Biochem 35:625–640. https://doi.org/10.1007/
s10695-­008-­9284-­4
Kalnejais LH, Martin WR, Bothner MH (2010) The release of dissolved nutrients and metals from
coastal sediments due to resuspension. Mar Chem 121(1–4):224–235. https://doi.org/10.1016/j.
marchem.2010.05.002
Kennish MJ (ed) (2019) Ecology of estuaries. Vol 2: Biological aspects. CRC Press – Taylor &
Francis Publ.Co, Boca Raton. 276 pp. ISBN:0-8493-5892-2
Kent RD, Vikesland PJ (2016) Dissolution and persistence of copper-based nanomaterials in under-
saturated solutions with respect to cupric solid phases. Environ Sci Technol 50:6772–6781.
https://doi.org/10.1021/acs.est.5bo4719
Kwok KWH, Batley GE, Wenning RJ et al (2013) Sediment quality guidelines: challenges and
opportunities for improving sediment management. Environ Sci Pollut R 21(1):17–27. https://
doi.org/10.1007/s11356-­013-­1778-­7
4  Bahía Blanca Estuary: A Chemical Oceanographic Approach 77

La Colla N, Franco M, Serra A et al (2014) Saltmarshes: an approach to organic matter contribu-
tion in sediments through isotopic signature. In: III Reunión Argentina de Geoquímica de la
Superficie (III RAGSU), Mar del Plata (Pcia. de Buenos Aires, Argentina), 02-05/DIC/2014.
ISBN:978-987-544-598-7
La Colla NS, Negrin VL, Marcovecchio JE et al (2015) Dissolved and particulate metals dynamics
in a human impacted estuary from the SW Atlantic. Estuar Coast Shelf Sci 166:45–55. https://
doi.org/10.1016/j.ecss.2015.05.009
La Colla NS, Botté SE, Negrin VL et al (2018a) Influence of human-induced pressures on dis-
solved and particulate Pb, Cr, Zn and Ni concentrations in the Bahía Blanca estuary, South
America. Environ Monit Assess 190:532–547. https://doi.org/10.1007/s10661-­018-­6930-­x
La Colla N, Botté SE, Marcovecchio JE (2018b) Metals in coastal zones impacted with urban
and industrial wastes: insights on the metal accumulation pattern in fish species. J Mar Syst
181:53–62. https://doi.org/10.1016/j.jmarsys.2018.01.012
La Colla NS, Botté SE, Marcovecchio JE (2019) Mercury cycling and bioaccumulation in a chang-
ing coastal system: from water to aquatic organisms. Mar Pollut Bull 140:40–50. https://doi.
org/10.1016/j.marpolbul.2018.12.051
Lailson-Brito J, Dorneles PR, Azevedo-Silva CE et al (2010) High organochlorine accumulation
in blubber of Guiana dolphin, Sotalia guianensis, from Brazilian coast and its use to estab-
lish geographical differences among populations. Environ Pollut 158:1800–1808. https://doi.
org/10.1016/j.envpol.2009.11.002
Limbozzi F, Leitào TE (2008) Chapter 31: Characterization of Bahía Blanca main existing pres-
sures and their effects on the state indicators for surface and groundwater quality. In: Neves
R, Baretta J, Mateus M (eds) Perspectives on integrated coastal zone management in South
America. IST Scientific Publishers, Lisbon, pp 333–350. ISBN:978-972-8469-74-0
López Abbate MC, Molinero JC, Guinder VA et al (2015) Microplankton dynamics under heavy
anthropogenic pressure. The case of the Bahía Blanca Estuary, southwestern Atlantic Ocean.
Mar Pollut Bull 95(1):305–314. https://doi.org/10.1016/j.marpolbul.2015.03.026
Marcovecchio JE (1988) Estudio comparativo de la distribución de los metales traza mercurio,
cadmio y zinc en organismos de dos zonas estuariales de Argentina: Bahía Blanca y desembo-
cadura del Rio de la Plata. Doctoral thesis, Mar del Plata National University, 212 pp
Marcovecchio JE (2001) Impact of freshwater on coastal estuaries: South America Atlantic lit-
toral. In: Buddemeier RW, Crossland C, Maxwell B et  al (eds) LOICZ/UNEP Regional
Shyntesis Programme: Australasia-Asia, The Americas, Africa-Europe. Summary Report and
Compendium. LOICZ Reports & Studies N°22, 77 pp. ISSN:1383-4304
Marcovecchio JE, Ferrer LD (2005) Distribution and geochemical partitioning of heavy metals
in sediments of the Bahía Blanca estuary, Argentina. J Coast Res 21(4):826–834. https://doi.
org/10.2112/014-­NIS.1
Marcovecchio JE, Freije RH (2004) Efectos de la intervención antrópica sobre sistemas marinos
costeros: el estuario de Bahía Blanca, vol 56. Anales de la Academia Nacional de Ciencias
Exactas, Físicas y Naturales (ANCEFN), Buenos Aires, pp 115–132. ISSN:0365-1185
Marcovecchio JE, Freije RH (eds) (2013) Procesos Químicos en Estuarios, Editorial
de la Univ Tecnológica Nacional (EdUTecNe), Buenos Aires (Argentina). E-book.
ISBN:978-987-1896-16-5
Marcovecchio JE, Moreno VJ, Pérez A (1986) Bio-magnification of total mercury in Bahia Blanca
shark. Mar Pollut Bull 17(6):276–278. https://doi.org/10.1016/0025-­326X(86)90064-­0
Marcovecchio JE, Moreno VJ, Pérez A (1988a) The sole, Paralichthys sp. as an indicator spe-
cies of heavy metal pollution in the Bahía Blanca estuary, Argentina. Sci Total Environ
75(2–3):191–200. https://doi.org/10.1016/0048-­9697(88)90032-­0
Marcovecchio JE, Moreno VJ, Pérez A (1988b) Determination of some heavy metal baselines
in the biota of Bahía Blanca, Argentina. Sci Total Environ 75(2–3):181–190. https://doi.
org/10.1016/0048-­9697(88)90031-­9
Marcovecchio JE, Moreno VJ, Pérez A (1988c) Total mercury contents in marine organisms of the
Bahía Blanca estuary trophic web. In: Seeliger U, Lacerda LD, Patchineelam SR (eds) Metals
in coastal environments of Latin America. Springer-Verlag, Heidelberg, pp 122–129
78 J. E. Marcovecchio et al.

Marcovecchio JE, Andrade JS, Ferrer LD et al (2001) Mercury distribution in estuarine environ-
ments from Argentina: the detoxification and recovery of salt-marshes after 15 years. Wetland
Ecol Manag 9(4):317–322. https://doi.org/10.1023/A:1011860618461
Marcovecchio JE, Botté SE, Delucchi F et  al (2008) Chapter 28: Pollution processes in Bahía
Blanca estuarine environment. In: Neves R, Baretta J, Mateus M (eds) Perspectives on Integrated
Coastal Zone Management in South America. Part B: From shallow water to the deep fjord: the
study sites. IST Scientific Publishers, Lisbon, pp 303–316. ISBN:978-972-8469-74-0
Marcovecchio JE, Freije RH, Popovich CA, Botté SE et al (2010a) Long-term observational sys-
tem for oceanographic studies within Bahía Blanca estuary (Argentina): state of the art and
perspectives. AQUASHIFT: Life in Warming Waters, Christian Albrecht’s University, Kiel
Marcovecchio JE, Botté SE, Fernández Severini MD et  al (2010b) Geochemical control of
heavy metals concentrations and distribution within Bahía Blanca estuary (Argentina). Aquat
Geochem 16(2):251–266. https://doi.org/10.1007/s10498-­009-­9076-­1
Marcovecchio JE, Botté SE, Domini CE et al (2014) Chapter 15: Heavy metals, major metals, trace
elements. In: Nollet L, de Geelder LSP (eds) Handbook of water analysis, 3rd edn. CRC Press,
Taylor & Francis Group LCC, Boca Ratón, pp 385–433. 979 pp. ISBN:978-1-4398-8966-4
Marcovecchio JE, Botté SE, Fernández Severini MD (2016) Distribution and behavior of zinc in
estuarine environments: an overview on Bahía Blanca estuary (Argentina). Environ Earth Sci
75:1168–1184. https://doi.org/10.1007/s12665-­016-­5942-­5
Menzies R, Quinete NS, Gardinali P et al (2013) Baseline occurrence of organochlorine pesticides
and other xenobiotics in the marine environment: Caribbean and Pacific collections. Mar Pollut
Bull 70:289–295. https://doi.org/10.1016/j.marpolbul.2013.03.003
Moss B, Jeppesen E, Søndergaard M et  al (2013) Nitrogen, macrophytes, shallow lakes and
nutrient limitation: resolution of a current controversy? Hydrobiologia 710:3–21. https://doi.
org/10.1007/s10750-­012-­1033-­0
Namieśnik J, Rabajczyk A (2010) The speciation and physicochemical forms of metals in sur-
face waters and sediments. Chem Spec Bioavailab 22(1):1–24. https://doi.org/10.318
4/095422910X12632119406391
Nguyen KDT, Morley SA, Lai C-H et al (2011) Upper temperature limits of tropical marine ecto-
therms: global warming implications. PLoS One 6(12):e29340. https://doi.org/10.1371/jour-
nal.pone.0029340
Nguyen DH, Umeyama M, ShintaniT (2012) Importance of geometric characteristics for salin-
ity distribution in convergent estuaries. J Hydrol 448–449:1–13. https://doi.org/10.1016/j.
jhydrol.2011.10.044
Nixon SW, Oczkowski AJ, Pilson MEQ et al (2015) On the response of pH to inorganic nutri-
ent enrichment in well-mixed coastal marine waters. Estuar Coast 38:232–241. https://doi.
org/10.1007/s12237-­014-­9805-­6
Nunes Vaz RA (2012) The salinity response of an inverse estuary to climate change and desali-
nation. Estuar Coast Mar Sci 98:49–59. https://doi.org/10.1016/j.ecss.2011.11.023
Nunes Vaz RA, Lennon GW, Bowers DG (1990) Physical behaviour of a large, negative or inverse
estuary. Cont Shelf Res 10(3):277–304. https://doi.org/10.1016/0278-­4343(90)90023-­F
Odum EP (2014) The strategy of ecosystem development. In: Ndubisi FO (ed) The eco-
logical design and planning reader. Island Press, Washington, DC, pp  203–216. 625  pp.
ISBN:978-1-61091-489-5
Olguín HF, Alder VA (2011) Species composition and biogeography of diatoms in Antarctic and
Subantarctic (Argentine shelf) waters (37–76°S). Deep-Sea Res II 58:139–152. https://doi.
org/10.1016/j.dsr2.2010.09.031
Oliva AL, Quintas PY, La Colla NS et al (2015) Distribution, sources, and potential ecotoxico-
logical risk of polycyclic aromatic hydrocarbons in surface sediments from Bahía Blanca
Estuary, Argentina. Arch Environ Contam Tox 69(2):163–172. https://doi.org/10.1007/
s00244-­015-­0169-­0
Orazi MM, Arias AH, Oliva AL et  al (2020) Characterization of atmospheric and soil poly-
cyclic aromatic hydrocarbons and evaluation of air-soil relationship in the Southwest of
4  Bahía Blanca Estuary: A Chemical Oceanographic Approach 79

Buenos Aires province (Argentina). Chemosphere 240:124847. https://doi.org/10.1016/j.


chemosphere.2019.124847
Paparazzo FE, Williams GN, Pisoni JP et al (2017) Linking phytoplankton nitrogen uptake, macro-
nutrients and chlorophyll-a in SW Atlantic waters: the case of the Gulf of San Jorge, Argentina.
J Mar Syst 172:43–50. https://doi.org/10.1016/j.jmarsys.2017.02.007
Perillo GME, Piccolo MC, Parodi E et al (2001) Chapter 14: The Bahía Blanca Estuary, Argentina.
In: Seeliger U, Kjerfve B (eds) Coastal Marine Ecosystems of Latin America, Ecological stud-
ies, vol 144. Springer, Heidelberg, pp 205–217. https://doi.org/10.1007/978-­3-­662-­04482-­7_15
Perillo GME, Pierini JO, Pérez DE, Piccolo MC (2005) Suspended sediment fluxes in the middle
reach of the Bahia Blanca Estuary, Argentina. In: DM FG, Knight J (eds) High resolution
morphodynamics and sedimentary evolution of estuaries. Springer, Heidelberg, pp 101–114.
ISBN:13:978-1-4020-3296-7
Piccolo MC (2008) Chapter 23: Climatological features of the Bahía Blanca Estuary. In: Neves
R, Baretta J, Mateus M (eds) Perspectives on integrated coastal zone management in South
America. IST Scientific Publishers, Lisbon, pp 233–242. ISBN:978-972-8469-74-0
Piccolo MC, Perillo GME (1999) The Argentina estuaries: a review. In: GME P, Piccolo MC, Pino-­
Quivira M (eds) Estuaries of South America: their geomorphology and dynamics, vol 235.
Springer, Heidelberg, pp 101–132. ISBN:13 978-3-642-64269-2
Piccolo MC, Perillo GME, Melo WD (2008) Chapter 22: The Bahía Blanca estuary: an inte-
grated overview of its geomorphology and dynamics. In: Neves R, Baretta J, Mateus M
(eds) Perspectives on integrated coastal zone management in South America. IST Scientific
Publishers, Lisbon, pp 221–232. ISBN:978-972-8469-74-0
Popovich CA, Gayoso AM (1999) Effect of irradiance and temperature on the growth rate of
Thalassiosira curviseriata Takano (Bacillariophyceae), a diatom bloom in Bahía Blanca estu-
ary (Argentina). J Plankton Res 21(6):1101–1110. https://doi.org/10.1093/plankt/21.6.1101
Popovich CA, Marcovecchio JE (2008) Spatial variability of phytoplankton and environmental
factors in a temperate estuary of South América (Atlantic Coast, Argentina). Cont Shelf Res
28:236–244. https://doi.org/10.1016/j.csr.2007.08.001
Popovich CA, Spetter CV, Marcovecchio JE et al (2008) Dissolved nutrients availability during
winter diatom bloom in a turbid and shallow estuary (Bahía Blanca, Argentina). J Coast Res
24(1):95–102. https://doi.org/10.2112/06-­0656.1
Ralston DK, Keafer BA, Brosnahan ML et  al (2014) Temperature dependence of an estuarine
harmful algal bloom: resolving interannual variability in bloom dynamics using a degree-day
approach. Limnol Oceanogr 59(4):1112–1126. https://doi.org/10.4319/lo.2014.59.4.1112
Renella A, Quirós R (2000) Relevamiento del uso de pesticidas agrícolas en la alta cuenca del
río Salado (Provincia de Buenos Aires). Proyecto UBA-Municipio de Junín (CS 2826/99).
Área de Sistemas de Producción Acuática, Departamento de Producción Animal, Facultad de
Agronomía, Universidad de Buenos Aires, 106 pp
Rivas AL, Dogliotti AI, Gagliardini DA (2006) Seasonal variability in satellite-measured surface
chlorophyll in the Patagonian Shelf. Cont Shelf Res 26:703–720. https://doi.org/10.1016/j.
csr.2006.01.013
Roach AC (2005) Assessment of metals in sediments from Lake Macquarie, New South Wales,
Australia, using normalisation models and sediment quality guidelines. Mar Environ Res
59:453–472. https://doi.org/10.1016/j.marenvres.2004.07.002
Roche H, Vollaire Y, Persic A et  al (2009) Organochlorines in the Vaccare’s lagoon trophic
web (Biosphere Reserve of Camargue, France). Environ Pollut 157:2493–2506. https://doi.
org/10.1016/j.envpol.2009.03.016
Rochman CM (2015) Part II, Chapter 5: The complex mixture, fate and toxicity of chemi-
cals associated with plastic debris in the marine environment. In: Bergmann M, Gutow L,
Klages M (eds) Marine anthropogenic litter. Springer, Heidelberg, pp  117–140. 456  pp.
ISBN:978-3-319-16509-7
Romero SI, Piola AR, Charo M et  al (2006) Chlorophyll-a variability off Patagonia based on
SeaWiFS data. J Geophys Res 111:C05021. https://doi.org/10.1029/2005JC003244
80 J. E. Marcovecchio et al.

Ronda AC, Arias AH, Oliva AL et al (2019) Synthetic microfibers in marine sediments and surface
seawater from the Argentinean continental shelf and a marine protected area. Mar Pollut Bull
149:110618. https://doi.org/10.1016/j.marpolbul.2019.110618
Roy S, Llewellyn C, Skarstad Egeland E et al (eds) (2011) Phytoplankton pigments: character-
ization, chemotaxonomy and applications in oceanography. Cambridge University Press,
Cambridge. 892 pp. ISBN:978-1-107-00066-7
Sericano J, Pucci AE (1984) Chlorinated hydrocarbons in the seawater and surface sedi-
ments of Blanca Bay, Argentina. Estuar Coast Shelf Sci 19(1):27–51. https://doi.
org/10.1016/0272-­7714(84)90051-­9
Sheaves M, Baker R, Nagelkerken I et  al (2015) True value of estuarine and coastal nurser-
ies for fish: incorporating complexity and dynamics. Estuar Coast 38:401–414. https://doi.
org/10.1007/s12237-­014-­9846-­x
Simonetti P, Botté SE, Marcovecchio JE (2017) Occurrence and spatial distribution of metals in
intertidal sediments of a temperate estuarine system (Bahía Blanca, Argentina). Environ Earth
Sci 76:636–647. https://doi.org/10.1007/s12665-­017-­6975-­0
Sindermann CJ (ed) (2006) Coastal pollution – effects on living resources and humans. CRC Press,
Taylor & Francis Group LCC, Boca Ratón. 309 pp. ISBN:10:0-8493-9677-8
Sokal RR, Rohlf FJ (eds) (1995) Biometry: the principles and practice of statistics in biological
research. WH Freeman & Co, New York. 487 pp. ISBN:13:978-0716724117
Sunda WG, Cai W-J (2012) Eutrophication induced CO2-acidification of subsurface coastal
waters: interactive effects of temperature, salinity, and atmospheric pCO2. Environ Sci Technol
46:10651–10659. https://doi.org/10.1021/es300626f
Sznaiberg L (2012) Parques Industriales: Luz verde para producir futuro. Revista Informe
Industrial N° 233. http://www.informeindustrial.com.ar/
Telesh IV, Khlebovich VV (2010) Principal processes within the estuarine salinity gradient: a
review. Mar Pollut Bull 61:149–155. https://doi.org/10.1016/j.ecss.2013.10.013
Telesh IV, Schubert H, Skarlato S (2013) Life in the salinity gradient: discovering mechanisms
behind a new biodiversity pattern. Estuar Coast Mar Sci 135:317–327. https://doi.org/10.1016/j.
ecss.2013.10.013
Tombesi N, Pozo K, Arias A et al (2018) Records of organochlorine pesticides in soils and sediments
on the southwest of Buenos Aires Province, Argentina. Environ Earth Sci 77(11):403–414.
https://doi.org/10.1007/s12665-­018-­7582-­4
UNEP (2001) GEO América Latina y el Caribe. Perspectivas del medio ambiente. PNUMA, San
José (Costa Rica), 144 pp
UNEP (United Nations Environment Programme) (2002) Regionally based assessment of persis-
tent toxic substances – Eastern and Western South America. UNEP Regional Report. http://
www.chem.unep.ch/pts
Villafañe VE, Valiñas MS, Cabrerizo MJ et al (2015) Physio-ecological responses of Patagonian
coastal marine phytoplankton in a scenario of global change: role of acidification, nutrients and
solar UVR. Mar Chem 177:411–420. https://doi.org/10.1016/j.marchem.2015.02.01
Wania F, Breivik K, Persson NJ et al (2006) CoZMo-POP 2 – a fugacity-based dynamic multi-­
compartmental mass balance model of the fate of persistent organic pollutants. Environ Model
Softw 21:868–884. https://doi.org/10.1016/j.envsoft.2005.04.003
Watts MJ, Mitra S, Marriott AL et  al (2017) Source, distribution and ecotoxicological assess-
ment of multielements in superficial sediments of a tropical turbid estuarine environment:
a multivariate approach. Mar Pollut Bull 115(1–2):130–140. https://doi.org/10.1016/j.
marpolbul.2016.11.057
Wetz MS, Yoskowitz DW (2013) An “extreme” future for estuaries? Effects of extreme cli-
matic events on estuarine water quality and ecology. Mar Pollut Bull 69:7–18. https://doi.
org/10.1016/j.marpolbul.2013.01.020
Whitfield AK, Elliott M, Basset A et  al (2012) Paradigms in estuarine ecology: a review of
the Remane diagram with a suggested revised model for estuaries. Estuar Coast Shelf Sci
97:78–90. https://doi.org/10.1016/j.ecss.2011.11.026
4  Bahía Blanca Estuary: A Chemical Oceanographic Approach 81

Xu S, Chen Z, Li S et al (2011) Modeling trophic structure and energy flows in a coastal arti-
ficial ecosystem using mass-balance Ecopath model. Estuar Coast 34:351–363. https://doi.
org/10.1007/s12237-­010-­9323-­0
Yakushev EV, Newton A (2013) Chapter 1: Introduction: redox interfaces in marine waters. In:
Yakushev EV (ed) Chemical structure of pelagic redox interfaces: observation and modelling.
Springer, Heidelberg, pp 1–12. 292 pp. ISBN:978-3-642-32124-5
Yang Y-P, Li Y, Sun Z-H et al (2014) Suspended sediment load in the turbidity maximum zone
at the Yangtze River Estuary: the trends and causes. J Geogr Sci 24(1):129–142. https://doi.
org/10.1007/s11442-­014-­1077-­3
Yosim AE, Fry RC (2015) Chapter 1: Systems biology in toxicology and environmental health.
In: Fry RC (ed) Systems biology in toxicology and environmental health. Elsevier Sci. Publ,
Amsterdam, pp 1–10. 273 pp. ISBN:978-0-12-801564-3
Chapter 5
Plankton Ecology and Biodiversity
in the Bahía Blanca Estuary

Anabela A. Berasategui, M. Sofía Dutto, Celeste López-Abbate,


and Valeria A. Guinder

5.1  Introduction

Estuaries support high plankton biomass that often exceeds those found over the
adjacent continental shelves due to the input of organic carbon and nutrients from
land. The species that comprise this high biomass vary according to the geographi-
cal location of the estuary and typically show minimum diversity at high latitudes,
which gradually increases until maximum values at 15–20° and decreases at the
equator. This productivity is one of the key reasons that has attracted humans to
populate estuarine shorelines throughout history. In addition to nutrient and organic
matter availability, estuaries provide a great number of habitat types for plankton
communities mainly defined by physical gradients and geomorphology. Spatial gra-
dients in estuaries pose unique short-term (tidal cycles, freshwater input) and long-­
term (water movement, chemical cycling, and physical structure) environmental
variability that forces plankton to display wide adaptive responses in order to sur-
vive (Day et al. 2013).
Adaptive responses of planktonic communities in estuaries consist on several
behavioral and physiological trade-offs to cope with environmental changes, mainly
caused by freshwater input and the tide dynamic. In this sense, the intertidal region
is subjected to extreme temperatures, water level fluctuations, drying conditions,
and salinity changes (Elliott et al. 2015; Smyth and Elliott 2016). Salinity is largely
the key physical factor that regulates the spatial distribution and structure of estua-
rine phytoplankton and zooplankton. In this context, estuarine species are classified
according to their tolerance to salinity as stenohaline and euryhaline. The first one
encompasses those species which can tolerate only a narrow range of salinity and
are generally found near the mouth of the estuary. The euryhaline species are those
that tolerate salinity fluctuations within a wider range and can thus penetrate further

A. A. Berasategui () · M. S. Dutto · C. López-Abbate · V. A. Guinder


Instituto Argentino de Oceanografía (IADO-CONICET/UNS), Consejo Nacional de
Investigaciones Científicas y Técnicas (CONICET), Bahía Blanca, Argentina

© Springer Nature Switzerland AG 2021 83


S. M. Fiori, P. D. Pratolongo (eds.), The Bahía Blanca Estuary,
https://doi.org/10.1007/978-3-030-66486-2_5
84 A. A. Berasategui et al.

up the estuary. Protists are adapted to either low or high salinity values and display
species-specific distribution patterns based on salinity gradients (Lancelot and
Muylaert 2011). Similarly, most estuarine metazoan zooplankton are osmocon-
formers, and each taxon is associated with a specific range of salinity that is within
their physiological tolerance. The pattern of faunal distribution along estuarine gra-
dients usually displays the highest number of species in the more saline waters near
the mouth, and hence richness tends to decrease with decreasing salinity toward the
riverine area (Whitfield et al. 2012). In this sense, the planktonic species from the
Bahía Blanca Estuary have been adapted for generations to high salinity values
above those in adjacent open sea, which is an uncommon feature in estuaries. For
instance, the average salinity values for the inner and middle zones of the Bahía
Blanca Estuary are around 33, and display hypersalinity (values between 41 and 43)
during hot and dry summers. High evaporation rates and restricted water circulation
in the shallower inner reach can lead to an inverse salinity distribution along the
estuary showing decreasing salinity from the head toward the mouth. Moreover,
hypersalinity is a consequence of the low mean annual runoff of the rivers associ-
ated with the Bahía Blanca Estuary and the residual circulation pattern in the inner
zone. This particularity implies that other environmental factors than salinity gradi-
ents come into play in the distribution of species, such as tides, wind, nutrient load-
ing, and human disturbance.
In general, species diversity increases from highly disturbed habitats such as the
Principal Channel of navigation, toward less impacted secondary channels at the
mouth of the estuary. Tides and wind are the main energy source that contribute to
concentrate higher plankton abundance toward the head of the Bahía Blanca Estuary
during the ebb (Guinder et al. 2009; Menéndez et al. 2012; Chazarreta et al. 2015;
López-Abbate et al. 2019b). The effect of the tidal cycle on short-term variations in
the mesozooplankton community has been reported in this estuary by Menéndez
et al. (2012) and Chazarreta et al. (2015). These authors reported that the abundance
of key mesozooplankton species was greater near the bottom during most of the
tidal cycles and also suggested a lateral movement of key copepod species to areas
of decreased flushing, such as channel margins. These authors also pointed that this
vertical abundance pattern was not observed at low tide, possibly because of the
shallowness of the ecosystem. In fact, at ebb tide, when fast currents occur, higher
abundances of zooplankton were observed near the bottom, which can be inter-
preted as a retention mechanism of organisms within the estuary.
In relation to the continental inputs to the Bahía Blanca Estuary, the largest fresh-
water and nutrient inputs are those provided by the sewage discharges of Bahía
Blanca, Punta Alta, and Ingeniero White cities. There is a subtle balance between
beneficial nutrient enrichment and overfertilization, which can seriously accelerate
primary production, bacterial respiration, and eutrophication. Hence, if elevated
nitrogenous and/or phosphorous loadings occur as a result of anthropic perturba-
tions, the supply rate of silicate may usually become limiting, leading to changes in
the phytoplankton community composition (Day et  al. 2013). In this sense, the
Bahía Blanca Estuary is usually nutrient-enriched showing a large natural stock of
silicate and a high level of nitrogenous compounds, which are adequate to support
5  Plankton Ecology and Biodiversity in the Bahía Blanca Estuary 85

phytoplankton demand of the system, dominated by diatoms (Popovich et al. 2008;


Guinder et  al. 2010). Nitrogen-rich effluents, however, produce a decrease on
microzooplankton probably due to the toxic effect of ammonium (López-Abbate
et al. 2019a).
Another factor that drives plankton distribution is the light availability, which
plays a key role in the control of biomass-specific productivity in turbid and nutrient-­
rich coastal systems like the Bahía Blanca Estuary (Popovich et al. 2008; Guinder
et al. 2009). The reduction in the underwater light availability affects primary pro-
duction by phytoplankton, and in the inner shallow area of the estuary, light penetra-
tion depends upon the wind intensity and direction that erodes the margins of the
tidal flats and saltmarshes and releases soft sediments into the water column (López-­
Abbate et al. 2019b). In agreement, the decrease of suspended particulate matter
concentration in winter with a concomitant increase in the penetration of solar radi-
ation seems to be one of the main drivers for the development of the phytoplankton
winter bloom in the Bahía Blanca Estuary, ascribed as the main biomass event of the
annual cycle (Guinder et al. 2009). When production is reduced by turbidity, highly
nutritive phytoplankton may be replaced by detrital material in the plankton food
web of the estuary (Diodato and Hoffmeyer 2008; Dutto et al. 2014).
It is worth noting that organic detritus plays a key role in this turbid ecosystem
where saltmarshes are crucial contributors to the organic fraction of suspended par-
ticulate matter (Negrin et al. 2011, 2013). Phytoplankton, microzooplankton, and
detritus have an important role in the natural diet of mesozooplankton. For example,
experimental data revealed that key copepods consumers exert higher filtration rates
on microzooplankton than on phytoplankton; however in late spring detritus is a
greater contributor to their natural diet (Diodato and Hoffmeyer 2008). In fact, the
seasonal fatty acid pattern of the mesozooplankton indicates different feeding strat-
egies over the year, suggesting an active feeding mode on protistan plankton during
summer and a more terrestrially derived compounds diet in winter (Dutto et  al.
2014). High environmental tolerance along with feeding plasticity allows plankton
communities to survive under the wide-ranging conditions of the Bahía Blanca
Estuary. Growing human disturbance, however, is pushing plankton to their adap-
tive edge, and profound ecological responses are beginning to become evident.

5.2  P
 hytoplankton and Microzooplankton Biodiversity
and Seasonality

The phytoplankton in the Bahía Blanca Estuary has been studied on a biweekly
basis over more than 30 years (1978–2010), with a monthly or more sporadic sam-
pling in the last decade (2011–2019). The high frequency monitoring allowed the
characterization of the phenology and biodiversity of the phytoplankton blooms in
relation with changing environmental conditions (Gayoso 1981, 1989, 1998;
Popovich and Gayoso 1999; Popovich et al. 2008; Guinder et al. 2010, 2013, 2016).
86 A. A. Berasategui et al.

Similarly, the microzooplankton community (phagotrophic protists with cell size


between 20 and 200 μm, mainly tintinnids and oligotrichs) have been extensively
studied by light microscopy in the estuary during discontinuous periods from 1986
to 2011 (Pettigrosso 2003; Pettigrosso and Popovich 2009; Barría de Cao et  al.
2005, 2011; López-Abbate et al. 2015, 2019a, b). It is worth noting that collection
and processing of samples have been consistent throughout the study period, includ-
ing the identification and quantification of species >5 μm in cell size under light
microscopy. This high resolution plankton survey has been performed by the same
group of specialists, ensuring that changes in the community structure and/or com-
position are due to environmental dynamics and not attributable to analytical pro-
cesses. Together with the sustained observations of protistan plankton, in situ sea
surface temperature, salinity, chlorophyll a and dissolved inorganic nutrients (sili-
cates, DIN, and phosphate) have been measured.
Overall, the long-term monitoring raised key aspects of the phytoplankton phe-
nology and biodiversity in the Bahía Blanca Estuary: (1) dominance of diatoms all
year-round, followed by dinoflagellates and nanoflagellates, (2) recurrence of a late
winter-early spring bloom, regarded as the main yearly biomass event, and (3)
occurrence of a summer bloom of lower magnitude and duration and with higher
interannual variability. Dominance of diatoms is a common feature of eutrophic,
turbid, and vertically mixed estuaries (e.g., Cloern and Dufford 2005; Guinder et al.
2009) and adjacent coastal areas (Garibotti et  al. 2011; Guinder et  al. 2018).
Together with dinoflagellates, both phytoplankton groups are rich in long-chain
essential fatty acids, what confer them high nutritional quality to support secondary
production (Winder et  al. 2017). Concerning microzooplankton, tintinnids domi-
nate the biomass of phagotrophic protists during most of the annual cycle, but espe-
cially during summer, while oligotrichs do not show a clear recurrent pattern. The
ability of oligotrichs to retain plastids from prey confers them with certain indepen-
dence from environmental conditions. As a result, oligotrichs may take advantage of
eventual favorable conditions, regardless of seasons and prey abundance. Overall,
the total biomass of microzooplankton shows maximum values during summer,
with the dominance of tintinnids and oligotrichs, and minimum during winter, when
the community shifts toward the dominance of rotifers, phagotrophic dinoflagel-
lates, and mixotrophic oligotrichs (Barría de Cao et  al. 2011; Pettigrosso and
Popovich 2009; Pettigrosso et  al. 2016). This pattern discloses the seasonal cou-
pling and the tight trophic interconnections between both phototrophic and phago-
trophic protistan communities. Accordingly, microzooplankton are often the
preferred prey of mesozooplankton, in spite of their lower concentration compared
to that of phytoplankton, due to specific characteristics such as body size, swim-
ming mode, and nutritional quality (Stoecker and Capuzzo 1990). In fact, tintinnids
represent the main energy resource of the dominant copepod in the estuary, Acartia
tonsa Dana, during its productive season (summer) (Diodato and Hoffmeyer 2008).
The high carbon transfer efficiency from phytoplankton and microzooplankton to
higher trophic levels in the Bahía Blanca Estuary highlights the key role of protistan
plankton in producing harvestable fish and in providing foraging and nursery
5  Plankton Ecology and Biodiversity in the Bahía Blanca Estuary 87

services for several permanent and migratory animal species (López Cazorla 2007;
Hoffmeyer et al. 2009b; Marrari et al. 2013; Fiori et al. 2016).

5.2.1  Observed Changes in Protistan Plankton

The more than 40 years of data records revealed a decrease in the mean annual lev-
els of chlorophyll a, attributed to a gradual shift in the seasonal patterns of phyto-
plankton biomass. The annual biomass cycle changed from unimodal to bimodal
due to a decrease in the winter-early spring bloom and an intensification of the
summer bloom (Winder and Cloern 2010; López-Abbate et  al. 2017) (Fig.  5.1).
These phenological changes have been driven by compound effects of multiple
biotic and abiotic factors, i.e., warming, changes in wind patterns, increase in water
turbidity, precipitation and nutrient inputs, invasion of species, and shifts in grazing
pressure (Guinder et al. 2010, 2012, 2016; López-Abbate et al. 2017). The regimen
shift in the plankton realm has become evident since the early 2000s, when the typi-
cal blooming species began to decrease and to be replaced, likely causing irrevers-
ible changes in the composition of the phytoplankton blooms (Fig. 5.1).
The onset of the late winter-early spring bloom is around June and commonly
lasts until October. Chlorophyll a over the 1978–present period for the winter-early
spring bloom reached maxima of 54 μg l−1 in 1980, 44 μg l−1 in 2002, and 25 μg l−1
in 2007, displaying a long-term decline at an yearly rate of 1% (López-Abbate et al.
2017). For more than two decades, until the early 2000s, the most abundant and
diverse taxa of the winter-spring bloom were the diatom genera Thalassiosira and
Chaetoceros, with T. curviseriata Takano as the key component of this biomass
event. Over more than 20 years, T. curviseriata has been the dominant blooming
diatom in winter-early spring, reaching up to 80–90% of the total phytoplankton
abundance, representing almost monospecific winter blooms in some years, like in
1991–1993 when it reached maximal abundances of 2.8 × 106 and 12.7 × 106 cells
l−1 (Popovich and Gayoso 1999). T. curviseriata was a key prey for the copepod
Eurytemora americana Williams which was introduced in the estuary in the late
1980s via ballast waters (Hoffmeyer 2004; Berasategui et al. 2009; Guinder et al.
2016). T. curviseriata has decreased toward recent years, with maxima not surpass-
ing the 1.4 × 105 cells l−1 in the late 2000s, and not registered in the samples in the
last winter-spring periods. Other co-dominant blooming Thalassiosira species were
T. anguste-lineata Fryxell and Hasle, T. pacifica Gran and Angst, T. rotula Meunier,
T. eccentrica Ehrenberg Cleve, T. hibernalis Gayoso, T. hendeyi Hasle and Fryxell,
and T. minima Gaarder. Among the most frequent Chaetoceros species were C. debi-
lis Cleve, C. diadema (Ehrenberg) Gran, C. similis Cleve, and a small Chaetoceros
sp. (ca. 8 μm in diameter) characterized by delicate setae. Other frequent species
during the winter bloom but also commonly present all-year round were the diatoms
Skeletonema costatum (Greville) Cleve, Asterionellopsis glacialis (Castracane)
Round, Ditylum brightwellii (West) Grunow, Leptocylindrus minimus Gran,
Guinardia delicatula (Cleve) Hasle, Cerataulina pelagica (Cleve) Hendey,
88 A. A. Berasategui et al.

Fig. 5.1  Schematic representation of the regimen shift in phytoplankton in relation to the multiple
interactive effects of environmental and biological drivers. During the time series (1978–present)
a positive trend of SST, turbidity, nitrogen, and phosphorus was recorded, along with major envi-
ronmental impacts which are enumerated in the lower gray bar. As a consequence, the annual cycle
of phytoplankton biomass changed from unimodal (winter bloom) to bimodal (winter and summer
blooms), while in recent years the winter bloom occurred ca. 4  weeks earlier (dashed lines).
Dominant blooming species also shifted in recent years, denoting profound community-level reor-
ganization. Parallel to phytoplankton changes, a shift in the abundance and composition of con-
sumer’s communities was also observed and is illustrated in the upper panels
5  Plankton Ecology and Biodiversity in the Bahía Blanca Estuary 89

Thalassiosira hendeyi and Paralia sulcata (Ehrenberg) Cleve. These species display
seasonal and interannual variability and tend to bloom during summertime
(December–February). Small phytoflagellates (10–20  μm) are present along the
year with maxima during summer. The dinoflagellates Scrippsiella trochoidea
(Stein) Loeblich III are important in late-spring and early summer, as well as the
heterotrophic dinoflagellates of the genera Protoperidinium Bergh. During summer,
unidentified species of gymnodinians and Cryptophyceae are conspicuous. The
Xantophyceae Ophiocytium sp. is an important component of the phytoplankton in
the Bahía Blanca Estuary, increasing its abundance in the last decade, together with
the large diatom Rhizosolenia sp. in the spring season. In particular, the abundance
of the small diatom Thalassiosira minima (5–15 μm) has risen over the last sum-
mers, likely related to the resuspension of bottom sediments driven by winds and
dredging operations in the estuary, and the consequent release and germination of
resting cysts of this planktonic species in the water column (Guinder et al. 2012).
Tintinnids in the Bahía Blanca Estuary are broadly dominated by species with
agglutinated lorica, consistent with the occurrence of this group within neritic areas
(Dolan 2012). Tintinnidium balechi Barría de Cao represents the dominant species
during most of the annual cycle, followed by several species from the genus
Tintinnopsis (T. brasiliensis Kofoid and Campbell, T. gracilis Kofoid and Campbell,
T. parva Merkle, T. baltica Brandt, T. beroidea Stein, T. levigata Kofoid and
Campbell) and Codonellopsis lusitanica Jörgensen (Barría de Cao et  al. 2005;
López-Abbate et al. 2019b). While tintinnids were in most cases identified to the
species level, oligotrichs were in some cases identified to the genus level and in
other cases were counted as a whole given that species identification requires com-
plex staining and in vivo observation. In spite of these limitations, several species
have been identified as recurrent within the estuary, such as Strombidium capitatum
(Leegaard) Kahl, S. emergens (Leegaard) Kahl, S. acutum Leegaard, S. dalum Lynn,
Montagnes and Small, Strombidinopsis elongata Song and Bradbury, Strobilidium
epacrum Lynn and Montagnes, Lohmanniella oviformis Leegaard, Laboea strobila
Lohmann, and Tontonia appendiculariformis Fauré-Fremiet.
In the last decades (1986–2011), tintinnids have revealed an interannual trend
characterized by a yearly rate of decline of 2.8% (López-Abbate et al. 2019b). In the
same way, oligotrichs declined at a yearly rate of 1.6% in the period 1994–2011, but
with no significant trend. Although no clear species replacement was evident in both
tintinnids and oligotrichs, observations suggest a trend toward the dominance of
tintinnid species with wider oral diameter. Oral diameter is proportional to prey
encounter rate, denoting a competitive advantage against species with narrower oral
area constrained to a smaller prey size spectrum. These profound changes have been
ascribed to the multiple effects of climate and human activities and represent the
tipping point of long-term impacts on mesozooplankton and planktivorous fish
dynamics.
90 A. A. Berasategui et al.

5.2.2  E
 nvironmental Drivers of Long-Term Changes
in Phytoplankton and Microzooplankton

The Bahía Blanca Estuary is characterized by the intense interaction between land
and sea given that half of the estuarine area consists of extensive low-slope tidal
flats densely fragmented by tidal courses (Perillo 2009). The bottom of the estuarine
basin is composed by a massive deposit of fine sediments which relates to an ancient
mouth of a river that was partially buried with marine and tidal plain sediments dur-
ing the Holocene marine transgression (Pratolongo et al. 2017). The grain size in the
lower mudflats influenced by tides ranges between very coarse to fine, with a modal
value of 32 μm. The continuous influence of the mesotidal regime and the effect of
high energy wind waves promote the mobilization of the softer sediments from the
bottom toward the water column, generating high values of suspended sediments
(>50 mg l−1) within a size range between 1 and 50 μm (mode of 10 μm) (Cuadrado
et al. 2005; Guinder et al. 2015; Zapperi et al. 2017).
Water turbidity is an important factor that determines the ecology of suspended
populations in estuaries. Suspended sediments interfere with phytoplankton light
harvesting by reducing light penetration and potentially excluding light-limited spe-
cies. In addition, a high concentration of inedible suspended particles within the
prey size range of phagotrophic protists complicates the prey uptake ultimately
reducing population growth (Boenigk and Novarino 2004). In spite of the limita-
tions posed by turbidity, plankton species in the Bahía Blanca Estuary have devel-
oped a series of adaptations to prosper under turbid conditions. For instance,
T. curviseriata, the keystone species dominating the winter phytoplankton bloom
until the early 2000s, and T. minima dominating in recent summers are well adapted
to thrive light under low intensities (Popovich and Gayoso 1999; Guinder et  al.
2012). Similarly, phagotrophic protists, broadly dominated by tintinnids with agglu-
tinated lorica, on one hand take advantage on high sediments availability to build
their lorica and get some protection against predation and on the other hand, avoid
mechanical cell damage produced by suspended minerals (Dolan 2012).
The annual variability of water turbidity is mainly affected by precipitation
which enhances sediment runoff of continental origin and also by the biological
disturbance of tidal flats. During summer, turbidity reaches maximum values due to
the intense activity of crabs that excavate large burrows and remove up to 5 kg m−2
d−1 of sediments (Zapperi et al. 2017). In contrast, water turbidity reaches minimum
values in winter when benthic communities are less active and precipitation is low
(Guinder et al. 2009). In the last three decades, however, water turbidity has revealed
a significant trend beyond natural variability in the inner area of the estuary. This
trend has been attributed to several factors. One of the primary factors is related to
the shift of wind patterns toward low intensity but highly persistent NW winds in the
last 25 years (López-Abbate et al. 2017). NW winds run parallel to the estuarine
main channels and produce high energy wind waves that enhance the erosion of the
extensive tidal flats (Perillo and Sequeira 1989) and produce the mobilization of soft
sediments toward the water column. Shift on wind patterns has been documented in
5  Plankton Ecology and Biodiversity in the Bahía Blanca Estuary 91

the region as a result of the southward displacement of the South Atlantic subtropi-
cal anticyclone and the persistent anomalies of the Southern Annular Mode (Dragani
et al. 2010). Dredging is an additional factor that produces shear stress and destabi-
lizes bottom sediments. Commercial harbor activities in the inner estuary have
grown gradually in the last decades, which implied the intensification of sediment
removal to allow the navigation of large ships. Deepening and straightening of
waterway began in the late 1950s, while an important deepening occurred in
1999–2000 and produced an 11-fold increase of monthly sediment extraction to
maintain navigation conditions (López-Abbate et  al. 2017). Shear stress of tidal
flats is also produced by the present rate of sea level rise (Lanfredi et  al. 1988).
Lateral erosion of estuarine margins produced a 33% loss of saltmarsh area since
1967 (Pratolongo et al. 2013). Marsh boundary erosion likely stimulated the con-
tinuous transport of mud deposits to the adjacent water column and contributed with
the increase on water turbidity.
The long-term trend of water turbidity implied important consequences on
planktonic communities in the Bahía Blanca Estuary. Although estuarine plankton
is well adapted to turbid conditions, natural populations follow a typical response
curve with a threshold value beyond their tolerance range. When this threshold is
exceeded and individuals are given no enough time to adapt, they can be locally
excluded (Boyd and Brown 2015). Indeed, the concentration of chlorophyll a (as a
proxy of phytoplankton), oligotrichs, tintinnids, and the copepod A. tonsa experi-
enced a significant decrease over the last 30  years (López-Abbate et  al. 2019b).
Phytoplankton and tintinnids were more vulnerable to the increase on water turbid-
ity. Long-term patterns of phytoplankton showed a drastic erosion of seasonal peaks
and a shift from the typical unimodal productivity pattern, i.e., the dominance of a
winter bloom, toward a bimodal regime with moderate productivity events in sum-
mer and winter (Fig. 5.1). The weakening of the winter bloom produced a loss of
synchrony between tintinnids and primary productivity. In addition to the indirect
effect of turbidity, tintinnids were directly affected by the higher concentration of
suspended sediments since the size range of such sediments overlaps precisely with
the prey size spectrum of most tintinnids in the Bahía Blanca Estuary. This suggests
that the decreasing ratio between sediments and phytoplankton forced tintinnids to
an additional energy expenditure to sort inedible particles. Oligotrichs, with their
ability to act both as phototrophs and phagotrophs (i.e., mixotrophs), were less sen-
sitive to water turbidity probably as a result of the exclusion of light-limited phyto-
plankton and filter feeders (López-Abbate et  al. 2019b). The negative trend of
tintinnids impacted on the populations of their main predator, Acartia tonsa. This
copepod dominates mesozooplankton during most of the annual cycle and reaches
maximum abundances in summer. During this period, adults prey preferentially
upon tintinnids, which represents nearly 78% of total particle’s filtration (Diodato
and Hoffmeyer 2008). Although many other factors may have contributed to the
long-term trend of A. tonsa (e.g., competitive exclusion due to the introduction of
invasive species, vulnerability to heavy metal, and nutrient pollution), sustained
observations revealed that the drop on tintinnid’s concentration produced a negative
impact on the numerical abundance of this species (López-Abbate et al. 2019b).
92 A. A. Berasategui et al.

The long-term decline of chlorophyll a was opposed to the positive trend docu-
mented in the continental shelf (Marrari et al. 2017). In the Bahía Blanca Estuary,
the trend was attributed to an ensemble of global and local-scale factors with a time-­
varying dominance that corresponded to the changing estuarine conditions (Guinder
et  al. 2010; López-Abbate et  al. 2017). At the beginning of the time series
(1978–1993), phytoplankton was driven by nutrients concentration and water tem-
perature. However, after 1993, the environmental influence shifted toward the domi-
nance of turbidity and wind velocity, whereas climate signals (SAM and ENSO)
boosted indirect effects through its influence on precipitation, wind, water tempera-
ture, and turbidity. In recent years, the notable dominance of larvae of Magallana
gigas Thunberg in the plankton of the estuary has sharply contributed to the phyto-
plankton clearance, related to the fast growth and voracious feeding of this invasive
species (Fiori et al. 2016). This oyster was introduced in the Argentine coast (40°S)
in 1981 for aquaculture, and documented for the first time in the Bahía Blanca
Estuary in 2010 (Dos Santos and Fiori 2010).
Long-term changes in the Bahía Blanca Estuary denote the prominent role of
wind patterns on water turbidity within estuaries with internal sources of sediments.
Hence, wind modifications driven by climate change, along with other erosive pro-
cesses such as sea level rise, have the potential to profoundly impact the growth,
phenology, and synchronization of estuarine plankton. Moreover, the introduction
and expansion of invasive species, a growing threat of global warming, lead to sig-
nificant impacts on the ecosystem structure, functioning, and biodiversity. Under
the future scenario of climate change and urban development, natural plankton
communities will thereby face unprecedented vulnerability, which urges us to
develop effective tools to mitigate ecological impacts.
Among the most widespread effects of human settlements in coastal areas is the
shift on nutrient balance. The Bahía Blanca Estuary receives excess nitrogen and
phosphorus from point-source sewage effluents. On average, the effluents transport
four times the concentration of nitrogen and three times the concentration of phos-
phorus in estuarine waters. In particular, the concentration of ammonium (NH4+)
within the sewage plume (ca. 1700 m) frequently exceeds the tolerance threshold of
plankton (ca. 100 μM). Sustained observations revealed that when exposed to severe
eutrophication, both phytoplankton and microzooplankton are negatively affected
by a significant reduction on population biomass (López-Abbate et al. 2015). This
reduction on plankton biomass may be partly explained by the dilution of suspended
populations by the continuous inflow of sewage. In fact, sewage effluent contributes
with significant amounts of freshwater into the estuary, which can represent up to
23% of the total freshwater input. In spite of the direct effect on plankton biomass,
excess NH4+ also inhibits the growth of phytoplankton and the uptake of NO3−
(Glibert et  al. 2014). Although nutrient loading does not produced the expected
stimulation on phytoplankton growth, growth inhibition does not occur in response
to excess NH4+ in the Bahía Blanca Estuary (López-Abbate et  al. 2016; López-­
Abbate et al. 2019a). Instead, nutrient addition promoted a shift on species compo-
sition by the increased dominance of nanoflagellates. Severe eutrophication reduced
the grazing rate of microzooplankton on nanoplankton and of nanoflagellates on
5  Plankton Ecology and Biodiversity in the Bahía Blanca Estuary 93

bacteria (López-Abbate et al. 2019a). The lower reactivity potential to control newly
produced biomass under severe eutrophication restricted the ability of microzoo-
plankton to control phytoplankton blooms (López-Abbate et al. 2016). From a bio-
geochemical perspective, an inefficient use of primary producer’s biomass decreases
the overall carbon transfer across the food web (Schmoker et al. 2016) and produces
a positive feedback with sewage effluents by an extra organic matter accumulation.
Environmental quality degradation associated with the dumping of urban effluents
can thereby undermine the compensatory capacity of microbial communities and
reduce their potential to offset nutrient imbalance (Box 5.1).

Box 5.1: Parasitic Outbreak Alleviates Competition and Temporarily


Benefits Thalassiosira curviseriata Population
Changes at the phytoplankton community level emerge from multiple interac-
tive effects of biotic and abiotic stressors at the population level. The species-­
specific responses to compound environmental drivers are non-linear and
depend on individual life history and interspecific connections (Litchman and
Klausmeier 2008). It is well known that the high functional diversity of micro-
bial plankton confers plasticity to their response against natural variability
(Worden et al. 2015). Still, sustained impacts such as warming and coastal
eutrophication gradually alter the phytoplankton yield (López-Abate et  al.
2017). Likewise, sporadic impacts such as extreme weather events and para-
sitic outbreaks drive significant shifts in the community structure and species
composition (Guinder et al. 2017). Eventually, these changes at the base of
the pelagic food web can trigger substantial alterations in the ecosystem
functioning.
In the Bahía Blanca Estuary, the long-term phytoplankton monitoring
allowed the identification of the key blooming species Thalassiosira curvise-
riata Takano as a case study to assess the compound effects of multiple envi-
ronmental drivers on the phenology of blooms. T. curviseriata is a small
centric diatom (6–21  μm in diameter) arranged in curved chains of several
cells. This species is eurythermal (5–20 °C) and euryhaline (20–40) and well
adapted to growth at relatively low underwater light intensity (Guinder et al.
2009). Specimens isolated from the turbid inner zone of the estuary showed
inhibited growth at ∼150 μmol m   −2 s  −1 (Popovich and Gayoso 1999). As
described in previous sections, T. curviseriata has been the dominant bloom-
ing diatom in winter-early spring in the Bahía Blanca Estuary for more than
20  years (1978–2003), but eventually its abundance has notably decreased
over the last years along with changes in its phenology (Guinder et al. 2016).
The winter phytoplankton bloom portrays the main biomass event in the
annual cycle, commonly composed by the same cluster of species from the
genera Thalassiosira and Chaetoceros. Nevertheless, in the last decade the
winter bloom has exhibited a species turnover toward the co-occurrence of
94 A. A. Berasategui et al.

diatoms in detriment of the formerly dominant T. curviseriata. The abundance


decline of T. curviseriata was responsible for the conspicuous biomass drop
of the phytoplankton bloom in recent years, concurrently with an earlier peak
of ca. 4 weeks (Guinder et al. 2010; López-Abbate et al. 2017). The relevance
of this species within the phytoplankton community allowed to compare dom-
inant environmental factors acting at community level (indirectly estimated
through chlorophyll a) versus the species level. The main drivers of the spe-
cies’ decreasing trend were identified from the characterization of its ecologi-
cal niche and from structural equation modelling (Fig.  5.2): precipitation,
water temperature, salinity, turbidity, dissolved inorganic nutrients, and graz-
ing pressure exerted by the invasive copepod Eurytemora americana Williams
(Guinder et al. 2016). The multiple interplay of these drivers has direct and
indirect effects on its population, affecting the competition for resources and
the predator-prey interactions. The rising temperature and drier conditions in
the region over the former years (1990–2008) permeated the pelagic environ-
ment leading to warmer and saltier conditions in recent winters, which
together with changes in nutrient ratios (i.e., rise in nitrite, nitrate, and phos-
phate) (Guinder et  al. 2010; López-Abbate et  al. 2015, 2016) impaired the
growth of T. curviseriata. Moreover, the increase in the population size of the
copepod E. americana since its introduction in the estuary in the late 1980s
(Hoffmeyer 2004; Berasategui et al. 2009, 2013) has caused a temporal niche
differentiation of these two species. The small diatom has shifted its seasonal
distribution toward higher annual ranges of temperature and salinity to avoid
seasonal overlapping with its consumer (Guinder et al. 2016).
Interestingly, in winter 2012 during routinary field research, a severe and
ephemeral parasitic infection on co-dominant blooming diatoms was
observed, concurrent with the re-emergence of T. curviseriata in the plankton
(Guinder et al. 2017). The outbreak was caused by the parasitic nanoflagellate
Pirsonia sp., with a high host-specificity on Thalassiosira pacifica and
Chaetoceros diadema (more than 40% of the cells were infected). The infec-
tion resulted in a severe reduction of their populations and allowed the devel-
opment of the fast-growing opportunistic species T. curviseriata, able to
exploit open niches. The changes in species composition comprised a restruc-
turing of the community by cell sizes and shapes (Guinder et al. 2017). The
ephemeral parasitic outbreak took place during a period of extreme precipita-
tion in the area (fourfold higher than in the same period in the previous decade
2000–2011) that dropped the salinity from 36.5 to 30.0 in 2 weeks. The abrupt
change in salinity was likely responsible of the observed osmotic collapse of
blooming diatoms, manifested by the breakout of cell’s cytoplasm. Eventually,
these extreme changes might overcome the physiological plasticity of phyto-
plankton cells and affect their osmotic regulation. In addition, the sudden
extreme rains might result in unbalanced nutrient ratios and increased turbid-
ity. These changes in the pelagic environment and the massive parasitism on
5  Plankton Ecology and Biodiversity in the Bahía Blanca Estuary 95

blooming diatoms seemed to release species competition and allowed the


development of the confined population of T. curviseriata, as a compensatory
response to the severe reduction in the bloom biomass, by more than 80%
(Guinder et al. 2017).
Commonly, the ephemeral nature of parasitic episodes and the rapidly re-­
establishment of the phytoplankton community after the infection make their
detection difficult. Nevertheless, the high frequency, species-specific survey
in the Bahía Blanca Estuary allowed the discrimination of changes in the
community structure due to environmental stress from natural succession of
phytoplankton species. Protists’ parasitism appears to be a prominent vector
of phytoplankton bloom collapse and species turnover, besides compound
effects of hydroclimatic drivers, nutrients, and grazers. Other studies have
documented that mass host mortalities caused by specific parasitism have
enhanced species diversity and trophic chain complexity (Salomon et  al.
2009; Dunne et al. 2013; Gsell et al. 2013). Overall, these findings highlight
the importance of addressing the proximal ecology of the dominant species to
understand structural modifications in the microbial plankton and their reper-
cussion in higher trophic levels.

Fig. 5.2  Main biotic and abiotic drivers and their compound direct and/or indirect effects on the
abundance and phenology of the diatom Thalassiosira curviseriata assessed in the Bahía Blanca
Estuary, based on high frequency, long-term field observations. Scale bar in the photos of T. cur-
viseriata chains: 10 μm
96 A. A. Berasategui et al.

5.3  A
 n Overview of Metazoan Zooplankton Community
from the Inner Zone of the Bahía Blanca Estuary

Zooplankton in aquatic ecosystems are the functional link between primary produc-
ers and higher trophic levels, mainly the size fraction between 0.2 and 20  mm
(mesozooplankton) (Turner 2004; Sieburth et  al. 1978). These organisms rapidly
respond to changes in the environment and, therefore, are valuable bioindicators of
ecosystem changes (Silva et al. 2004; Chang et al. 2009; Uriarte and Villate 2004).
Zooplankton composition and dynamics in the Bahía Blanca Estuary have been
studied for more than four decades (reviewed by Hoffmeyer 2007; Hoffmeyer and
Mianzan 2007; Hoffmeyer and Cervellini 2007; Berasategui et  al. 2019). Neritic
zooplankton is well adapted to large seasonal variability in temperature, salinity,
turbidity, and nutrient inputs. It is known that there is a positive response of the
mesozooplankton abundance from the Bahía Blanca Estuary to salinity, which indi-
cates that this community appears to be well acclimated to the commonly high
salinity values in the estuary (average annual salinity around 33, with maximum
values of 43) (Berasategui et  al. 2016; Berasategui et  al. 2019). The study of
Guerrero et al. (1976) was pioneer on mesozooplankton from the inner zone of the
Bahía Blanca Estuary, and reported a maximal abundance of 8.105 ind. m−3.
Subsequently, the meso- and macrozooplankton (>20 mm, according to the classifi-
cation by Sieburth et  al. 1978) were extensively studied by Hoffmeyer and col-
leagues (Piccolo and Hoffmeyer 2007), where high seasonal variability in the
abundance of the meso- and macrozooplankton was documented. These authors
recorded 34 taxa in the mesozooplankton fraction. The holoplankton was mostly
represented by the calanoid copepods Acartia tonsa Dana, Paracalanus parvus
Claus, Labidocera fluviatilis Dahl F, and Calanoides carinatus Krøyer, which were
the most frequent and abundant in the study area (Hoffmeyer 2007).
Acartia tonsa is the best-known species among small marine copepods. It has a
cosmopolitan distribution along temperate coastal areas and plays a pivotal role in
the trophic web in the Bahía Blanca Estuary (Lopez Cazorla 2007; Lopez Cazorla
et al. 2011). This copepod is found in the water column throughout the year, with
maximal abundance during the summer season and minimal abundance during the
winter season (Sabatini 1989; Hoffmeyer 1994, 2004; Berasategui et al. 2016). It
shares the temporal ecological niche with the exotic copepod Eurytemora ameri-
cana Williams, during the austral winter-spring (Hoffmeyer 2004; Hoffmeyer
et al. 2009a).
Within the holoplankton, Hoffmeyer (2007) also reported the presence of harpac-
ticoid copepods (i.e., Euterpina acutifrons Dana, Tisbe varians Scott T., Robertsonia
propinqua Scott T., and Heterolaophonte sp. Lang). This group is dominated by
species adapted to live in close association with bottom layers; however, intense
resuspension processes lead to a constant transfer of these organisms from the bot-
tom toward the water column. Several species from gelatinous zooplankton (i.e.,
Mnemiopsis leidyi Mayer, Turritopsis nutricula McCrady, and Obelia sp. Péron and
Lesueur) and the mysids Neomysis americana Smith are also common inhabitants
5  Plankton Ecology and Biodiversity in the Bahía Blanca Estuary 97

within estuarine waters, especially during the warm season (Hoffmeyer and Minazan
2007). Regarding meroplankton, the presence of cirripedia, mollusca, bryozoa, pol-
ichaeta, ascidiacea, cnidaria, and decapods larvae were reported, being the decap-
oda larvae the most abundant group during the warm season with 31 reported taxa
(Hoffmeyer and Cervellini 2007). The highest abundances in this group were attrib-
uted to the zoeas of the crab Neohelice granulata Dana and shrimp larvae of Peisos
petrunkevitchi Burkenroad and Artemesia longinaris Spence Bate.
Subsequent surveys during the years 2009–2010 showed a restructuring of the
mesozooplankton community toward the dominance of typical estuarine species
and a decrease in total mesozooplankton abundance, in comparison to those recorded
by Hoffmeyer (2007) during the years 1990–1991 (Berasategui et  al. 2019). The
mesozooplankton community in 2009–2010 was composed mainly of A. tonsa
(annual mean 695.48 ind m−3), N. granulata (annual mean of 81.97 ind m−3),
B. glandula (annual mean of 28.89 ind m−3), A. amphitrite (annual mean of 13.35
ind m−3), P. parvus (annual mean of 9.63 ind m−3), E. americana (annual mean of
3.91 ind m−3), E. acutifrons (annual mean of 3.62 ind m−3), and Spionidae spp.
(annual mean of 1.15 ind m−3). Berasategui et al. (2019) also documented a decrease
in the abundance of copepod species typical from the adjacent shelf area, such as
L. fluviatilis and C. carinatus (Boltovskoy 1981; Bradford Grieve et al. 1999), and
a shift in the distribution of decapod larvae of Pachycheles laevidactylus Ortmann
1892, Alpheus puapeba Christoffersen 1979, and Cyrtograpsus altimanus Rathbun
1914, probably related to the joint effects of dredging activities in the main channel
and the tidal dynamics of the ecosystem.
In more recent years (2014–2018), mesozooplankton denoted similar annual
dynamics (Table 5.1) to that reported by Hoffmeyer (2007) and Berasategui et al.
(2019) in previous years. Overall, mesozooplankton showed the greatest abun-
dances during the warm months which ranged between 11,704.67 and 557.84 ind.
m−3. During the warm season A. tonsa and the decapod zoeas of Grapsidae domi-
nated (Fig. 5.3a), whereas in winter, a decrease in the abundance of E. americana
(Fig.  5.3b) was notorious in relation to population densities reported in previous
years (Berasategui et al. 2009, 2012, 2019). These changes in the mesozooplankton
community in the estuary are closely interconnected with the long-term modifica-
tions in the microbial plankton phenology, and composition, mentioned in Sects. 5.1
and 5.2 (Fig. 5.1).
The inner zone of the Bahía Blanca Estuary is highly susceptible to biological
invasions due to a poor seawater exchange and its exposure to intense maritime traf-
fic and periodic dredging. These characteristics have probably favored the effective
introduction (via ballast water) and dispersion of exotic species already mentioned
by Hoffmeyer et  al. (2004), such as B. glandula and E. americana, and more
recently, the oyster Magallana gigas (Fig. 5.3d) Thunberg (Dos Santos and Fiori
2010; Chazarreta et al. 2015) as well as the fiddler crab Leptuca (=Uca) uruguayen-
sis Nobili (Truchet et  al. 2019). Indeed the planktonic larvae of both species are
increasingly frequent in zooplankton samples.
Regarding gelatinous fauna, the Bahía Blanca Estuary seems to encompass suit-
able environmental conditions for the development of this kind of organisms (Dutto
98 A. A. Berasategui et al.

Table 5.1  Zooplankton taxa in the Bahía Blanca Estuary


Functional groups Species Abund.
Hydrozoa Eucheilota ventricularis (McCrady, 1859) *
Hydrozoa Gossea brachymera (Bigelow, 1909) *
Hydrozoa Liriope tetraphylla (Chamisso & Eysenhardt, 1821) **
Hydrozoa Turritopsis nutricula (McCrady, 1857) *
Hydrozoa Obelia spp. (Péron & Lesueur, 1810) **
Hydrozoa Olindias sambaquiensis Müller (1861) **
Hydrozoa Actinula (L) *
Scyphozoa Chrysaora lactea (Eschscholtz, 1829) *
Ctenophora Mnemiopsis leidyi (A. Agassiz, 1865) ***
Ctenophora Beroe ovata (Bruguière, 1789) *
Ctenophora Pleurobrachia pileus (O. F. Müller, 1776) *
Mollusca Gastropoda and Bilvalvia undet.(LV) *
Mollusca Crassostrea gigas (Thunberg, 1793) ***
Polychaeta Spionidae (Grube, 1850) **
Polychaeta Syllidae (Grube, 1850) **
Polychaeta Polynoidae (Kinberg, 1856) **
Polychaeta Onuphidae (Kinberg, 1865) **
Polychaeta Aphroditidae (Malmgren, 1867) *
Chaetognatha Parasagitta friderici (Ritter-Záhony, 1911) *
Cladocera Ceriodaphnia (Dana, 1853) **
Cladocera Bosmina longirostris (O.F. Müller, 1785) *
Cladocera Daphnia (O.F. Müller, 1785) *
Mysida Neomysis americana (S.I. Smith, 1873) **
Mysida Arthromysis magellanica (Cunningham, 1871) *
Amphipoda Corophium sp. (Latreille, 1806) *
Amphipoda Corophium, Latreille, 1806 Caprella, Lamarck 1801 *
Decapoda Cyrtograpsus angulatus (Dana, 1851) **
Decapoda Cyrtograpsus altimanus (Rathbun, 1914) **
Decapoda Neohelice granulata (Dana, 1851) ***
Decapoda Pagurixus sp. (Melin, 1939) *
Decapoda Alpheus puapeba (Christoffersen, 1979) *
Decapoda Corystoides chilensis (Lucas in H. Milne Edwards & Lucas, 1844) *
Decapoda Artemesia longinaris (Spence Bate, 1888) **
Decapoda Pinnixa patagoniensis (Rathbun, 1918) *
Decapoda Pachycheles laevidactylus (Ortmann, 1892) *
Decapoda Leptuca uruguayensis (Nobili, 1901) *
Decapoda Peisos petrunkevitchi (Burkenroad, 1945) ***
Sessilia Balanus glandula (Darwin, 1854) ***
Sessilia Amphibalanus amphitrite (Darwin, 1854) ***
Copepoda Acartia tonsa (Dana, 1849) ****
Copepoda Calanoides carinatus (Krøyer, 1849) *
Copepoda Euterpina acutifrons (Dana, 1847) ***
(continued)
5  Plankton Ecology and Biodiversity in the Bahía Blanca Estuary 99

Table 5.1 (continued)
Copepoda Labidocera fluviatilis (Dahl F., 1894) **
Copepoda Monstrilla helgolandica (Claus, 1863) *
Copepoda Eurytemora americana (Williams, 1906) **
Copepoda Monstrilla sp. (Dana, 1849) *
Copepoda Oithona nana (Giesbrecht, 1893) *
Copepoda Paracalanus parvus (Claus, 1863) **
Copepoda Delavalia aff. palustris (Brady, 1868) *
Copepoda Microarthridion aff. littorale (Poppe, 1881) **
Copepoda Halicyclops affi. crassicornis (Herbst, 1955) *
Copepoda Nannopus aff. palustris (Brady, 1880) *
Copepoda Dactylopusia tisboides (Claus, 1863) *
Copepoda Tisbe sp. (Lilljeborg, 1853) *
Copepoda Longipedia (Claus, 1862) **
Copepoda Ectinosomatidae (Sars G.O., 1903) *
Crustacea Isopoda undefined *
Crustacea Stomatopoda undefined *
Bryozoa Amathia sp. (Lamouroux, 1812) *
Bryozoa Cyphonauta (L) *
Anthozoa Cerinula (L) *
Echinodermata Ophiuroidea (L) *
Tunicata Ascidiacea ***
Ichthyoplankton Brevoortia aurea (Spix & Agassiz, 1829) **
Ichthyoplankton Ramnogaster arcuata (Jenyns, 1842) *
Ichthyoplankton Sciaenidae (Cuvier, 1829) *
Ichthyoplankton Atherinidae (Risso, 1827) *
Chordata Oikopleuridae (Lohmann, 1915) *
Abundance data are from 2014 to 2018 (Puerto Cuatreros and Puerto Rosales). Categories of mean
abundance: * less than 10 ind. m−3, ** 10–100 ind. m−3, *** 100–1000 ind. m−3, and **** more
than 1000 ind. m−3

et  al. 2017). Gelatinous zooplankton is a polyphyletic assemblage of disparate


organisms with a gelatin-like consistency (Haddock, 2004). The most common
gelatinous species are those included into the phyla Cnidaria and Ctenophora (i.e.,
medusae and ctenophores, commonly grouped as “jellyfish”). These organisms are
important modulators of marine ecosystems due to their ability to consume large
quantities of zooplanktonic prey including fish larvae and eggs and their “bloom-
ing” nature that potentiates their ecological roles (Boero 2013).
In the temperate Southwestern Atlantic Ocean, investigations on jellyfish have
been intensified in recent years. Studies on the abundance, species composition,
biodiversity patterns, invasions, and ecological traits of medusae and ctenophores
have been carried out at different sites along the Argentine Sea and revealed the
biological relevance of these species within pelagic ecosystems (e.g., Genzano et al.
2006, 2008a, b, 2009a, b; Rodriguez et al. 2007, 2012, 2014, 2017; Schiariti et al.
2008, 2012, 2014; Diaz Briz et al. 2012; Dutto et al. 2019a). As already mentioned,
100 A. A. Berasategui et al.

Fig. 5.3  Zooplankton key species from the Bahía Blanca Estuary. (a) Late-spring mesozooplank-
ton with the copepod Acartia tonsa, Grapsidae zoea, and nauplii of Amphibalanus amphitrite. (b)
Mesozooplankton in winter dominated by the copepod Eurytemora americana, (c) hydroid of
Corymorpha januarii, (d) Magallana gigas, (e) Olindias sambaquiensis, and (f) Liriope tetrap-
hylla. (Photos by Anabela Berasategui (a, b, and d) and Sofía Dutto (c, e, and f))
5  Plankton Ecology and Biodiversity in the Bahía Blanca Estuary 101

one of the geographic zones that seems to enclose suitable environmental conditions
for the development of gelatinous zooplankton communities is the Bahía Blanca
Estuary and the adjacent inner shelf (Dutto et al. 2017, 2019a). Species composition
and ecological studies on medusae and ctenophores in this area concentrated in the
1980s (Mianzan and Sabatini 1985; Zamponi and Mianzan 1985; Mianzan 1986a,
b, 1989a, b, c; Hoffmeyer 1990), but they were discontinued. Currently, the study of
gelatinous zooplankton in the estuarine region has recommenced. Twenty-three
hydromedusae, 3 scyphomedusae, and 3 ctenophores species have been reported for
the area, including some related to economic and sanitary concerns, such as the
cases of Chrysaora lactea Eschscholtz (Scyphozoa), the hydromedusae Olindias
sambaquiensis Müller (Fig. 5.3e), and Liriope tetraphylla Chamisso (Fig. 5.3f) and
Eysenhardt (Dutto et al. 2017). The recent records of high abundances of L. tetrap-
hylla in the region of the Bahía Blanca Estuary (Dutto et al. 2019a) deserve further
investigation and monitoring in order to assess its population dynamic and ecologi-
cal role in this coastal ecosystem. The proliferation of L. tetraphylla may be boosted
by the high productivity of the ecosystem and also promoted by the traits and plas-
ticity of this holoplanktonic species. Likewise, hydroids are also commonly found
in this area (Genzano et  al. 2009a, b). Recently, a population of the solitary and
seasonal polyp of Corymorpha januarii was described from a small tidal channel in
the inner area of the Bahía Blanca Estuary, and one of the first approaches to their
feeding ecology by gut content analysis was provided (Dutto et al. 2019b). Polyps
of C. januarii show a variable diet composed mainly of organic matter and zoo-
planktonic prey. The stomach content analysis reveals that these polyps probably
select copepods, mysids, and other zooplanktonic prey of lower swimming capacity
(e.g., barnacles larvae and benthic invertebrates, and fish eggs) (Boxes 5.2 and 5.3).

Box 5.2: Eurytemora americana as Bioindicator of Water Quality


Contaminants associated with anthropic activities such as pesticides, heavy
metals, and hydrocarbon derivatives produce negative effects on the metabo-
lism of various marine organisms, generating oxidative damage and stress, as
well as a reduction on reproduction and survival (Regoli et al. 2002; Suderman
and Marcus 2002; Hack et al. 2008; Tsangaris et al. 2010). In this sense, cope-
pods respond fast to changes in the environment driven by compound effects
of global scale hydro-climate variability and anthropogenic pressure at local/
regional scale. Their fast response to environmental conditions along with
their short life cycles and fast rate of reproduction constitutes valuable proper-
ties that motivate their use as bioindicators (Parmar et al. 2016; Battuello et al.
2017). These crustaceans have a crucial role in the trophic transfer of con-
taminants in the marine food web, since they tend to accumulate pollutants
from their prey (Wang 2002; Parmar et al. 2016). It is known that in copepods,
pesticides, aromatic polycyclic hydrocarbons, and heavy metals lead to dis-
ruptive effects at the endocrine level and affect growth and reproductive rates
(Marcus 2004).
102 A. A. Berasategui et al.

The Bahía Blanca Estuary is subject to high anthropogenic impact derived


from industrial, port, and urban development, which affects the dynamic and
composition of plankton and seston (Marcovecchio et  al. 2008; Hoffmeyer
et al. 2008). It is also a highly productive ecosystem and constitutes an impor-
tant nursery area for birds, crustaceans, and fish of commercial interest, such
as Micropogonias furnieri (Desmarest, 1823), Cynoscion guatucupa (Cuvier,
1830), Odontesthes argentinensis (Valenciennes, 1835) and Engraulis
anchoita (Hubbs and Marini, 1935) (Lopez Cazorla 2007; Marrari et  al.
2013). Key estuarine copepods (E. americana and A. tonsa) seem well adapted
to severe eutrophication (Biancalana and Torres 2011; Biancalana et al. 2012;
Dutto et al. 2014); however the apparent adaptation of these species to highly
eutrophic conditions may be threatened by rising effluent and pollutant dis-
posal derived from intense human activities (Fernández Severini et al. 2017).
The exotic calanoid copepod E. americana (Fig. 5.2b) was introduced in
the ecosystem via ballast water translocation through commercial ships com-
ing from the Northern Hemisphere (Hoffmeyer 1994), and has successfully
established in the planktonic food web, with maximum abundance from win-
ter to spring (Hoffmeyer et al. 2009b; Berasategui et al. 2019). This species
displays two distinct types of reproductive behavior depending on the prevail-
ing environmental conditions (Berasategui et  al. 2009, 2012). During the
growing season, females produce subitaneous eggs, which hatch as popula-
tion continues to rise in the plankton. When the population peak begins to
decrease after spring, diapause eggs are produced. Thereafter, the species dis-
appears from the pelagic habitat and recruits in banks of diapause eggs in the
bottom sediments, until optimal environmental conditions of subsequent
years trigger their hatching (Berasategui et al. 2013).
In order to assess the adaptive capacity of copepods exposed to sewage and
industrial pollution, Berasategui et al. (2018) evaluated the reproductive per-
formance of E. americana through laboratory experiments under different
levels of contamination. Experimental data revealed that egg production,
number of nauplii, number of fecal pellets, survival, and fertility state of
females responded negatively to increasing concentrations of heavy metals
(cadmium, lead, copper, zinc, and chrome), dissolved ammonium and phos-
phate, and turbidity. The results showed that bioavailable contaminants from
the dissolved phase of sewage effluent reduced E. americana fertility in a
53.09% (egg production 12.4 ± 2.9 egg/female.clutch), while the water from
the bottom of the sewage discharge site was undoubtedly lethal for this spe-
cies (Fig. 5.4a). When exposed to the particulate phase of sewage effluents,
only 40% of the females showed regeneration of their gonads for a second egg
laying (Fig. 5.4b), while the dissolved phase effluent lead to null gonad regen-
eration for a second egg laying. The current rate of environmental change in
the Bahía Blanca Estuary highlights that empirical studies on the physiologi-
cal responses of copepods to anthropogenic impacts are of urgent and critical
concern for planning resource and ecosystem management.
5  Plankton Ecology and Biodiversity in the Bahía Blanca Estuary 103

Box 5.3: Olindias sambaquiensis, a Stinging Jellyfish from the


Southwestern Atlantic
Olindias sambaquiensis (Limnomedusa: Olindiidae, see Fig. 5.3e) is a con-
spicuous hydromedusan species endemic to the coastal waters of southern
Brazil, Uruguay, and Argentina. Its distribution ranges approximately from
23°S to 42°S, where this jellyfish displays outbreaks of variable magnitude
and duration (Chiaverano et al. 2004; Resgalla Jr. et al. 2019). This species
has a meroplanktonic life cycle, which includes small and primitive colonial
polyps (asexual phase), only known through laboratory cultivation and never
found in nature. Its free-living medusa (sexual phase) may reach 10  cm in
diameter and shows diel vertical migration habits staying at the bottom during
the day being, presumably, nocturnal (Chiaverano 2001; Ale et al. 2007). This
species is responsible for a large number of stings to bathers in Argentina and
Uruguay during summer and also during winter and autumn in southern Brazil
(Mianzan and Ramírez 1996; Resgalla Jr. et al. 2011). Olindias sambaquien-
sis presents three different types of tentacles, and, as an adult, it may have
more than 400 tentacles, all of them full of stinging cells or cnidocytes. Its
whole body may also be covered by cnidocytes. The venom of this jellyfish is
very complex and may include 29 putative toxins similar to venom proteins
from diverse phyla, and also two novel cytolysins (Weston et al. 2013; Haddad
Jr. et al. 2014). It can produce persistent skin pain and envenomation resulting
in erythema, severe dermatitis, and systemic symptoms (Kokelj et al. 1993;
Mianzan and Ramírez 1996; Mianzan et al. 2001; Mosovich and Young 2012).
The poisoning by this jellyfish is characterized as moderate to severe, and can
lead to cardiovascular complications including cardiopulmonary arrest
(Haddad Jr. et  al. 2014). Despite its sanitary importance, the population
dynamic in those regions where this jellyfish is common and problematic is
far from being understood. Its life cycle is poorly known because of the small
size of the polyp that makes it difficult to find and, therefore, to study in
nature, and the difficulty in generating and maintaining the polyps under labo-
ratory conditions (Zamponi and Facal 1987; Resgalla Jr. et al. 2019). However,
some congeners have been successfully cultured shedding light on the devel-
opment of Olindias spp. and the environmental factors influencing their popu-
lation dynamics (Patry et  al. 2014; Toshino et  al. 2019). Considering the
results of these investigations, the dynamic of O. sambaquiensis is presum-
ably tightly modulated by water temperature. Winter temperatures may have
an important role in the maintenance and reproduction of polyps, whereas the
liberation and growth of the medusa phase may be linked to rising tempera-
tures (Patry et  al. 2014; Resgalla Jr. et  al. 2019). Although this hypothesis
needs to be confirmed, it is clear that the environment, particularly the tem-
perature, has a crucial role in the development of the life cycle phases of
O. sambaquiensis, and therefore, in the occurrence and the magnitude of
bloom events. Water temperature would be acting at life history level, whereas
104 A. A. Berasategui et al.

other factors, such as winds, seem to have a physical role favoring the pres-
ence or the absence at coastal waters of those jellyfish already liberated. In
southern Brazil, for instance, southerly winds, which may promote the con-
vergence of waters toward the coast accumulating the jellyfish in the beach
zone, are associated with high number of stinging events (Resgalla Jr. et al.
2005, 2011). In the Bahía Blanca region, in contrast, offshore winds (north-
erly winds) trigger coastal upwellings, which produce the emergence of the
bottom layer, favoring the concentration of Olindias jellyfish at coastal waters
(Mianzan and Zamponi 1988; Brendel et al. 2017). In fact, the absence of the
jellyfish in the beach zone of the Bahía Blanca region in recent years would
be explained, in part, by the significant decrease of offshore wind speed
(Brendel et al. 2017). The orientation of the coast, along with other features,
such as the depth and bottom geomorphology, may explain the distinct pat-
terns observed in the Southwestern Atlantic waters in relation to the coastal
occurrence of O. sambaquiensis. Although progress has been made in the
knowledge of O. sambaquiensis (see Resgalla Jr. et al. 2019), much remains
to be understood, particularly, in relation to the life history of this species and
the role of environmental factors in the modulation of the population dynamic.
The comprehension of the life cycle and the life history of O. sambaquiensis
is essential to answer how the environment can modulate the timing and the
magnitude of blooms. To optimize the cultivation and the maintenance of the
species, which will enable the experimentation, is therefore essential.

Fig. 5.4  Sewage effluent effects on fecundity in Eurytemora americana. Experimental response
in females’ culture. (a) Specimen exposed to water from the bottom of the sewage discharge, (b)
specimen exposed to the particulate phase of the sewage effluents. (Photos by Anabela Berasategui)
5  Plankton Ecology and Biodiversity in the Bahía Blanca Estuary 105

References

Ale E, Imazu M, Oliveira OMP et  al (2007) Ocorrência de Olindias sambaquiensis (cnidaria:
hydrozoa) em hábito demersal ao redor da ilha de são sebastião (sp, Brasil). XII Congresso
Latino-Americano de Ciências do Mar – XII COLACMAR Florianópolis
Barría de Cao MS, Beigt D, Piccolo MC (2005) Temporal variability of diversity and biomass
of tintinnids (Ciliophora) in a southwestern Atlantic temperate estuary. J Plankton Res
27:1103–1111
Barría de Cao MS, Piccolo MC, Perillo GME (2011) Biomasss and microzooplankton sea-
sonal assemblages in the Bahía Blanca estuary, Argentinean coast. J Mar Biol Assoc UK
91(5):953–959
Battuello M, Mussat Sartora R, Briziob P et al (2017) The influence of feeding strategies on trace
element bioaccumulation in copepods (Calanoida). Ecol Indic 74:311–320
Berasategui AA, Hoffmeyer MS, Biancalana F et al (2009) Temporal variations in abundance and
fecundity of the invading copepod Eurytemora americana in Bahía Blanca estuary during an
unusual year. Estuar Coast Shelf Sci 85:82–88
Berasategui AA, Hoffmeyer MS, Dutto MS et al (2012) Seasonal variation in the egg morphol-
ogy of the copepod Eurytemora americana and its relationship to reproductive strategy in a
­temperate estuary in Argentina. ICES J Mar Sci 69(3):380–388
Berasategui AA, Dutto MS, Chazarreta (2013) Seasonal occurrence and hatching success of
benthic eggs of calanoid copepods in the Bahía Blanca Estuary, Argentina. Mar Biol Res
9(10):1018–1028
Berasategui AA, Fernandez-Severini MD, Menendez MC (2016) Reproductive trade-off of the
copepod Acartia tonsa in a hypersaline estuary of Southwestern Atlantic. Temporal variations
in the morphology of eggs. Mar Biol Res 12(8):817–829
Berasategui AA, Biancalana F, Fricke A et al (2018) The impact of sewage effluents on the fecun-
dity and survival of Eurytemora americana in a eutrophic estuary of Argentina. Estuar Coast
Shelf Sci 211:208–216
Berasategui AA, López Abbate MC, D’Agostino V et al (2019) Mesozooplankton structure and
seasonal dynamics in three coastal systems of Argentina: Bahía Blanca Estuary, Pirámide Bay
and Ushuaia Bay. In: Hoffmeyer M, Sabatini M, Brandini FP et al (eds) Plankton ecology of the
Southwestern Atlantic from the subtropical to the Subantarctic Realm. Springer, Cham, p 327
Biancalana F, Torres AI (2011) Variation of mesozooplankton composition in a eutrophicated semi-­
enclosed system (Encerrada Bay, Tierra del fuego, Argentina). Braz J Oceangr 59(2):195–199
Biancalana F, Menendez MC, Berasategui AA et al (2012) Sewage pollution effects on mesozoo-
plankton structure in a shallow temperate estuary. Environ Monit Assess 184:3901–3913
Boenigk J, Novarino G (2004) Effect of suspended clay on the feeding and growth of bacterivorous
flagellates and ciliates. Aquat Microb Ecol 34:181–192
Boero F (2013) Review of jellyfish blooms in the Mediterranean and Black Sea. General Fisheries
Commission for the Mediterranean FAO, Rome
Boltovskoy D (Ed) (1981) Atlas del zooplancton del Atlántico Sudoccidental y métodos de trabajo
con el zooplancton Marino. Publicacion especial INIDEP, Mar del Plata, Argentina, pp 938
Boyd PW, Brown CJ (2015) Modes of interactions between environmental drivers and marine
biota. Front Mar Sci 2
Bradford-Grieve JM, Markhaseva EL, Rocha CEF et al (1999) Copepoda. In Boltovskoy D (Ed).
South Atlantic Zooplankton. Backhuy publishers, Leinden, The Netherlands pp 869–1098
Brendel AS, Dutto MS, Menéndez MC et al (2017) Wind pattern change along a period of coastal
occurrence variation of a stinging medusa on a SW Atlantic beach. Anuário do Instituto de
Geociencias UFRJ 40:303–315
Chang KH, Doi H, Nishibe Y et al (2009) Spatial and temporal distribution of zooplankton com-
munities of coastal marine waters receiving different human activities (fish and pearl oyster
farmings). The Open Mar Biol J. https://doi.org/10.2174/18744508090301008
106 A. A. Berasategui et al.

Chazarreta CJ, Hoffmeyer MS, Cuadrado DG (2015) Tidal effects on short-term mesozooplank-
ton distribution in small channels of a temperate-turbid estuary, Southwestern Atlantic. Braz J
Oceanogr 63(2):83–92
Chiaverano L (2001) Historia de vida de Olindias sambaquiensis (Limnomedusae, Olindiidae)
durante su fase sexual en la zona de El Rincón (Buenos Aires, Argentina): Estructura de tallas,
crecimiento, desarrollo e influencia ambiental en sus agregaciones. PhD thesis, Universidad
Nacional de Mar del Plata, Argentina, p 70
Chiaverano L, Mianzan H, Ramírez F (2004) Gonad development and somatic growth patterns of
Olindias sambaquiensis (Limnomedusae, Olindiidae). Hydrobiology 530(531):373–381
Cloern JE, Dufford R (2005) Phytoplankton community ecology: principles applied in San
Francisco Bay. Mar Ecol Progr Ser 285:11–28
Cuadrado DG, Gómez EA, Ginsberg SS (2005) Tidal and longshore sediment transport associated
to a coastal structure. Estuar Coast Shelf S 62:291–300
Day JW, Crump BC, Kemp WM et  al (eds) (2013) Estuarine ecology. Wiley-Blackwell.
Wiley, Hoboken
Diaz Briz L, Martorelli S, Genzano G et  al (2012) Parasitism (Trematoda, Digenea) in medu-
sae from the southwestern Atlantic Ocean: medusa hosts, parasite prevalences, and ecological
implications. Hydrobiologia 690:215–226
Diodato SL, Hoffmeyer MS (2008) Contribution of planktonic and detritic fractions to the natural
diet of mesozooplankton in Bahía Blanca Estuary. Hydrobiologia 614:83–90
Dolan JR (2012) Introduction to tintinnids. In: Dolan JR, Montagnes DJ, Agatha S, Coats DW,
Stoecker D (eds) The biology and ecology of tintinnid ciliates. Wiley, p 16
Dos Santos EP, Fiori SM (2010) Primer registro sobre la presencia de Crassostrea gigas en el
estuario de Bahía Blanca (Argentina). Commun Soc Malacol Urug 9:245–252
Dragani WC, Martin PBC, Simionato G et al (2010) Are wind wave heights increasing in south-­
eastern south American continental shelf between 32°S and 40°S? Cont Shelf Res 30:481–490
Dunne JA, Lafferty KD, Dobson AP et  al (2013) Parasites affect food web structure primarily
through increased diversity and complexity. PLoS Biol. https://doi.org/10.1371/JOURNAL.
PBIO.1001579
Dutto MS, Kopprio GA, Hoffmeyer MS et al (2014) Planktonic trophic interactions in a human
impacted estuary of Argentina: A fatty acid marker approach. J Plankton Res. https://doi.
org/10.1093/plankt/fbu012
Dutto MS, Genzano GN, Schiariti A et al (2017) Medusae and ctenophores from the Bahía Blanca
Estuary and neighboring inner shelf (Southwest Atlantic Ocean, Argentina). Mar Biodivers
Rec 2017:10–14
Dutto MS, Chazarreta CJ, Rodriguez CS et al (2019a) Macroscale abundance patterns of hydrome-
dusae in the temperate Southwestern Atlantic (27°–56°S). PLoS One 14(6):e0217628
Dutto MS, Carcedo MC, Nahuelhual EG et al (2019b) Trophic ecology of a corymorphid hydroid
population in the Bahia Blanca Estuary, Southwestern Atlantic. Reg Stud Mar Sci 31:100746
Elliott M, Borja Á, McQuatters-Gollop A et al (2015) Force majeure: will climate change affect our
ability to attain Good Environmental Status for marine biodiversity? Mar Pollut Bull 95:7–27
Fernández Severini DM, Villagran D, Biancalana F et al (2017) Heavy metal concentrations found
in seston and microplankton from an impacted temperate shallow estuary along the southwest-
ern Atlantic Ocean. J Coast Res 33(5):1196–1209
Fiori SM, Pratolongo PD, Zalba SM et  al (2016) Spatially explicit risk assessment for coastal
invaders under different management scenarios. Mar Biol 163:245. https://doi.org/10.1007/
s00227-­016-­3017-­5
Garibotti IA, Ferrario ME, Almandoz GO et al (2011) Seasonal diatom cycle in Anegada Bay, El
Rincón estuarine system, Argentina. Diatom Res 26(2):227–241
Gayoso AM (1981) Estudio de las diatomeas del estuario de Bahía Blanca. Doctoral dissertation,
Universidad Nacional de La Plata, Argentina, p 100
Gayoso AM (1989) Species of the diatom genus Thalassiosira from the coastal zone of the South
Atlantic (Argentina). Bot Mar 32:331–337
5  Plankton Ecology and Biodiversity in the Bahía Blanca Estuary 107

Gayoso AM (1998) Long-term phytoplankton studies in the Bahía Blanca Estuary, Argentina.
ICES J Mar Sci 55:655–660
Genzano GN, Mianzan HW, Acha EM et al (2006) First record of the invasive medusa Blackfordia
virginica (Hydrozoa: Leptomedusae) in the Río de la Plata estuary, Argentina-Uruguay. Rev
Chil Hist Nat 79:257–261
Genzano GN, Mianzan H, Bouillon J (2008a) Hydromedusae (Cnidaria: Hydrozoa) from the tem-
perate southwestern Atlantic Ocean: a review. Zootaxa 1750:1–18
Genzano GN, Mianzan HW, Diaz BL et al (2008b) On the occurrence of Obelia medusa bloom and
empirical evidence of an unusual Obelia and Amphisbetia hydroids shoreline massive accumu-
lations. Lat Am J Aquat Res 36:301–307
Genzano GN, Giberto D, Schejter L et al (2009a) Hydroid assemblages from the Southwestern
Atlantic Ocean (34–42° S). Mar Ecol 30:33–46
Genzano GN, Rodriguez C, Pastorino G et al (2009b) The hydroid and medusa of Corymorpha
januarii in temperate waters of the Southwestern Atlantic Ocean. Bull Mar Sci 84:229–235
Glibert PM, Wilkerson FP, Dugdale RC et  al (2014) Phytoplankton communities from San
Francisco Bay Delta respond differently to oxidized and reduced nitrogen substrates-even
under conditions that would otherwise suggest nitrogen sufficiency. Front Mar Sci 1:1–16
Gsell AS, de Senerpont Domis LN, Verhoeven KJF (2013) Chytrid epidemics may increase genetic
diversity of a diatom spring-bloom. ISME J 7:2057–2059
Guerrero MA, Izquierdo M, Canelo S (1976) Observaciones mensuales sobre algunos aspecto
biológicos de la ría de Bahía Blanca entre pto Ing. White y Pto. Cuatreros. Contr Cient
IADO 36:1–7
Guinder VA, Popovich CA, Perillo GME (2009) Particulate suspended matter concentrations in the
Bahía Blanca Estuary, Argentina: implication for the development of phytoplankton blooms.
Estuar Coast Shelf S 85:157–165
Guinder VA, Popovich CA, Molinero JC et al (2010) Long-term changes in phytoplankton phenol-
ogy and community structure in the Bahía Blanca Estuary, Argentina. Mar Biol 157:2703–2716
Guinder VA, Molinero JC, Popovich CA et  al (2012) Dominance of the planktonic diatom
Thalassiosira minima in recent summers in the Bahia Blanca Estuary, Argentina. J Plankton
Res 34:995–1000
Guinder VA, Popovich CA, Molinero JC et al (2013) Phytoplankton summer bloom dynamics in
the Bahía Blanca Estuary in relation to changing environmental conditions. Cont Shelf Res
52:150–158
Guinder VA, López-Abbate MC, Berasategui AA et al (2015) Influence of the winter phytoplank-
ton bloom on the settled material in a temperate shallow estuary. Oceanologia 57:50–60
Guinder VA, Molinero JC, López Abbate MC (2016) Phenological changes of blooming diatoms
promoted by compound bottom-up and top-down controls. Estuar Coast 40:95–104. https://
doi.org/10.1007/s12237-­016-­0134-­9
Guinder VA, Carcedo MC, Buzzi N et al (2017) Ephemeral parasitism on blooming diatoms in a
temperate estuary. Mar Fresh Res 69(1):128–133
Guinder VA, Tillman U, Krock B et  al (2018) Plankton multiproxy analyses in the Northern
Patagonian Shelf, Argentina: community structure, phycotoxins and characterization of
Alexandrium strains. Front Mar Sci 5:394
Hack LA, Tremblay LA, Wratten SD et  al (2008) Toxicity of estuarine sediments using a full
life-cycle bioassay with the marine copepod Robertsonia propinqua. Ecotox Environ Safe
70:469–474
Haddad V Jr, Zara F, Marangoni S et  al (2014) Identification of two novel cytolysins from the
hydrozoan Olindias sambaquiensis (Cnidaria). J Venom Anim Toxins Incl Trop Dis 20:10.
https://doi.org/10.1186/1678-­9199-­20-­10
Haddock SH (2004) A golden age of gelata: past and future research on planktonic ctenophores
and cnidarians. Hydrobiologia 530(13):549–556
Hoffmeyer MS (1990) Algunas observaciones sobre la alimentación de Mnemiopsis mccradyi
Mayer (Ctenophora-Lobata). Iheringia Ser Zool 70:55–65
108 A. A. Berasategui et al.

Hoffmeyer MS (1994) Seasonal succession of Copepoda in the Bahía Blanca Estuary. In: Ferrari
and Bradley (eds) Ecology and morphology of copepods, developments in hydrobiology,
hydrobiology 292(293): 303–308
Hoffmeyer M (2004) Decadal change in zooplankton seasonal succession in the Bahía Blanca
estuary, Argentina, following the introduction of two zooplankton species. J Plankton Res
26(2):181–189
Hoffmeyer MS (2007) Mesozooplancton. In Piccolo MC. Hoffmeyer MS (eds). Ecosistema del
Estuario de Bahía Blanca. FUNS press, pp 133–141
Hoffmeyer MS, Cervellini PM (2007) Zooplancton temporario. In: Piccolo MC, Hoffmeyer MS
(eds) Ecosistema del Estuario de Bahía Blanca, 3rd edn. EdiUNS, Bahía Blanca, p 153
Hoffmeyer MS, Mianzan HW (2007) Macrozooplankton del estuario y aguas costeras adyacen-
tes. In: Piccolo MC, Hoffmeyer MS (eds) Ecosistema del Estuario de Bahía Blanca, 3rd edn.
EdiUNS, Bahía Blanca, p 143
Hoffmeyer MS, Fernandez-Severini MD, Menendez MC et al (2008) Composition and dynamics
of mesozooplankton assemblages in the Bahía Blanca Estuary. In: Neves R, Baretta J, Mateus
M (eds) Perspectives on integrated coastal zonemanagement in South America. ISTPress:
Scientific Publishers, Lisboa, p 303
Hoffmeyer MS, Menendez MC, Biancalana F et al (2009a) Ichthyoplankton spatial pattern in the
inner shelf off Bahía Blanca Estuary, SW Atlantic Ocean. Estuar Coast Shelf S 84:383–392
Hoffmeyer MS, Berasategui AA, Beigt D et al (2009b) Environmental regulation of the estuarine
copepods Acartia tonsa and Eurytemora americana during coexistence period. J Mar Biol
Assoc U K 89:355–361
Kokelj F, Mianzan H, Avian M et  al (1993) Dermatitis due to Olindias sambaquiensis: a case
report. Cutis 51:339–342
Lancelot C, Muylaert K (2011) Trends in estuarine phytoplankton. Ecology:5–15
Lanfredi NW, D'Onofrio EE, Mazio CA (1988) Variations of the mean sea level in the Southwest
Atlantic Ocean. Cont Shelf Res 8:1211–1220
Litchman E, Klausmeier CA (2008) Trait-based community ecology of phytoplankton. 599. Annu
Rev Ecol Evol S. 39:615–639
Lopez Cazorla A (2007) Peces. In: Piccolo MC, Hoffmeyer MS (eds) Ecosistema del Estuario de
Bahía Blanca, 3rd edn. EdiUNS, Bahía Blanca, p 191
Lopez Cazorla A, Pettigrosso RE, Tejera L et al (2011) Diet and food selection by Ramnogaster
arcuata (Osteichthyes, Clupeidae). J Fish Biol 78(7):2052–2066
López-Abbate MC, Molinero JC, Guinder VA et al (2015) Microplankton dynamics under heavy
anthropogenic pressure. The case of the Bahia Blanca Estuary, Southwestern Atlantic Ocean.
Mar Pollut Bull 9(1):305–314
López-Abbate MC, Barría de Cao MS, Pettigrosso RE (2016) Seasonal changes in microzoo-
plankton feeding behavior under varying eutrophication level in the Bahía Blanca estuary (SW
Atlantic Ocean). J Exp Mar Biol Ecol 481:25–33
López-Abbate MC, Molinero JC, Guinder VA et al (2017) Time-varying environmental control of
phytoplankton in a changing estuarine system. Sci Total Environ 609:1390–1400
López-Abbate MC, Molinero JC, Barría de Cao MS et al (2019a) Eutrophication disrupts sum-
mer trophic links in an estuarine microbial food web. Food Webs 20:00121. https://doi.
org/10.1016/j.fooweb.2019.e00121
López-Abbate MC, Molinero JC, Perillo GM et al (2019b) Long-term changes on estuarine ciliates
linked with modifications on wind patterns and water turbidity. Mar Environ Res 144:46–55
Marcovecchio JE, Botte SE, Delucchi F et al (2008) Pollution processes in Bahía Blanca Estuary
environment. In: Neves R, Baretta J, Mateus M (eds) Perspectives on integrated coastal zone
management in South America. IST Press, Lisboa, p 301
Marcus N (2004) An overview of the impacts of eutrophication and chemical pollutants on cope-
pods of the coastal zone. Zool Stud 43(2):211–217
5  Plankton Ecology and Biodiversity in the Bahía Blanca Estuary 109

Marrari M, Signorini S, McClain CR et  al (2013) Reproductive success of the Argentine


anchovy, Engraulis anchoita, in relation to environmental variability at a mid-shelf front 875
(Southwestern Atlantic Ocean). Fish Oceanogr 22:247–261
Marrari M, Piola AR, Valla D (2017) Variability and 20-year trends in satellite-derived surface
chlorophyll concentrations in large marine ecosystems around South and Western Central
America. Front Mar Sci 4:372
Menéndez MC, Piccolo MC, Hoffmeyer MS (2012) Short-term variability on mesozooplankton
community in a shallow mixed estuary (Bahía Blanca, Argentina): influence of tidal cycles and
local winds. Estuar Coast Shelf Sci 112:11–22
Mianzan HW (1986a) Estudio sistemático y bioecológico de algunas medusas Scyphozoa de la
región subantártica. PHD thesis, Universidad Nacional de La Plata, La Plata
Mianzan HW (1986b) Beroe ovata en aguas de la Bahía Blanca, Argentina (Ctenophora).
Spheniscus 2:29–32
Mianzan HW (1989a) Las medusas Scyphozoa de la Bahía Blanca. Boletim do Instituto
Oceanografico São Paulo 37:29–32
Mianzan HW (1989b) Sistemática y zoogeografía de Scyphomedusae en aguas neríticas argen-
tinas. Inv Mar CICIMAR 4:15–34
Mianzan HW (1989c) Distribución de Olindias sambaquiesis Müller, 1861 (Hydrozoa,
Limnomedusae) en el Atlántico Sudoccidental. Iheringia Sér Zoo 69:155–157
Mianzan HW, Ramírez F (1996) Olindias sambaquiensis stings in the South West Atlantic. In:
Williamson JAH, Fenner PJ et al (eds) Venomous and poisonous marine animals: a medical and
biological handbook. Univ. New South Wales Press, Brisbane, pp 206–208 and 301
Mianzan HW, Sabatini M (1985) Estudio preliminar sobre distribución y abundancia de Mnemiopsis
maccradyi en el estuario de Bahía Blanca (Ctenophora). Spheniscus 1:53–68
Mianzan HW, Zamponi MO (1988) Estudio bioecológico de Olindias sambaquiensis Muller, 1861
(Limnomedusae, Olindiidae) en el área de Monte Hermoso, II. Factores meteorológicos que
influyen en su aparición. Iheringia 2:63–68
Mianzan HW, Fenner PJ, Cornelius PFS et al (2001) Vinegar as a disarming agent to prevent fur-
ther discharge of the nematosyst of the stinging hydromedusa Olindias sambaquiensis. Cutis
68:45–48
Mosovich JH, Young P (2012) Picadura de medusa Olindias sambaquiensis. Análisis de 49 casos.
Medicina 72:380–388
Negrin VL, Spetter CV, Asteasuain RO et al (2011) Influence of flooding and vegetation on carbon,
nitrogen, and phosphorus dynamics in the pore water of a Spartina alterniflora salt marsh. J
Environ Sci 23(2):212–221
Negrin VL, Spetter CV, Guinder VA et al (2013) The role of Sarcocornia perennis and tidal flood-
ing on sediment biogeochemistry in a South American wetland. Mar Biol Res 9(7):703–715
Parmar TK, Rawtani D, Agrawal YK (2016) Bioindicators: the natural indicator of environmental
pollution. Front Life Sci 9:110–118
Patry W, Knowles T, Christianson I, Howard M (2014) The hydroid and early medusa stage of
Olindias formosus (Cnidaria, Hydrozoa, Limnomedusae). J Mar Biol Assoc UK 94:1409–1415
Perillo GME (2009) Tidal courses: classification, origin and functionality. In: Perillo GME,
Wolanski E, Cahoon DR et  al (eds) Coastal wetlands: an integrated ecosystem approach.
Elsevier, Amsterdam, p 185
Perillo GME, Sequeira ME (1989) Geomorphologic and sediment transport characteristics of the
middle reach of the Bahia Blanca estuary (Argentina). J Geophys Res Oceans 94:14351–14362
Pettigrosso RE (2003) Planktonic ciliates Choreotrichida and Strombidiida from the inner zone of
the Bahía Blanca estuary, Argentina. Iheringia 93:117–126
Pettigrosso RE, Popovich CA (2009) Phytoplankton-aloricate ciliate community in the Bahía Blanca
Estuary (Argentina): seasonal patterns and trophic groups. Braz J Oceanogr 57(3):215–227
Pettigrosso RE, Garcia MD, Uibrig RA et al (2016) Mixotrophic ciliate dynamics in two zones of
a temperate and highly turbid estuary in South America, Argentina. Ecol Austral 26:107–119
110 A. A. Berasategui et al.

Piccolo MC, Hoffmeyer MS (2007) Ecosistema del Estuario de Bahía Blanca, 3rd edn. EdiUNS,
Bahía Blanca, p 233
Popovich CA, Gayoso AM (1999) Effect of irradiance and temperature on the growth rate of
Thalassiosira curviseriata Takano (Bacillariophyceae), a bloom diatom in Bahía Blanca estu-
ary (Argentina). J Plankton Res 21:1101–1110
Popovich CA, Spetter CV, Marcovecchio JE et  al (2008) Dissolved nutrient availability during
winter diatom bloom in a turbid and shallow estuary (Bahía Blanca, Argentina). J Coast Res
24:95–102
Pratolongo P, Mazzon C, Zapperi G et al (2013) Land cover changes in tidal salt marshes of the
Bahía Blanca estuary (Argentina) during the past 40 years. Estuar Coast Shelf S 133:23–31
Pratolongo P, Piovan MJ, Cuadrado DG et al (2017) Coastal landscape evolution on the western
margin of the Bahía Blanca Estuary (Argentina) mirrors a non-uniform sea-level fall after the
mid-Holocene highstand. Geo-Mar Lett 37:373–384
Regoli F, Gorbi S, Frenzilli G et al (2002) Oxidative stress in ecotoxicology: from the analysis of
individual antioxidants to a more integrated approach. Mar Environ Res 54:419–423
Resgalla C Jr, Goncalves VC, Klein AHF (2005) The occurrence of jellyfish stings on the Santa
Catarina coast, southern Brazil. Brazil. Braz J Oceanogr 53:183–186
Resgalla C Jr, Rosseto AL, Haddad JV (2011) Report of an outbreak of stings caused by Olindias
sambaquiensis Muller, 1861 (Cnidaria: Hydrozoa) in southern Brazil. Braz J Oceanogr
59:391–396
Resgalla C Jr, Petri L, Teodoro da Silva BG et al (2019) Outbreaks, coexistence, and life cycle
of jellyfish species in relation to abiotic and biological factors along a South American coast.
Hydrobiologia 839:87–102
Rodriguez C, Genzano G, Mianzan H (2007) First record of Eutonina scintillans Bigelow, 1909
(Hydrozoa: Leptomedusae: Eirenidae) in temperate waters of the southwestern Atlantic Ocean.
Investig Mar 35:135–138
Rodriguez C, Miranda TP, Marques AC et al (2012) The genus Hybocodon (Cnidaria, Hydrozoa)
in the southwestern Atlantic Ocean, with a revision of Hybocodon species recorded in the area.
Zootaxa 3523:39–48
Rodriguez CS, Pujol MG, Mianzan HW et al (2014) First record of the invasive stinging medusa
Gonionemus vertens in the southern hemisphere (Mar del Plata, Argentina). Lat Am J Aquat
Res 42:653–657
Rodriguez CS, Marques AC, Mianzan HW et  al (2017) Environment and life cycles influence
distribution patterns of hydromedusae in austral South America. Mar Biol Res 13:659–676
Sabatini ME (1989) Ciclo anual del copépodo Acartia tonsa Dana 1849 en la zona interna de Bahía
Blanca (Provincia de Buenos Aires). Sci Mar 53:847–856
Salomon PS, Granéli E, Neves MHCB, Rodriguez EG (2009) Infection by Amoebophrya spp.
parasitoids of dinoflagellates in a tropical marine coastal area. Aquat Microb Ecol 55:143–153
Schiariti A, Kawahara M, Uye SI et  al (2008) Life cycle of the jellyfish Lychnorhiza lucerna
(Scyphozoa: Rhizostomeae). Mar Biol 156:1–12
Schiariti A, Sal Moyano MP, Giberto DA et  al (2012) First record of the association between
Lychnorhiza lucerna (Scyphozoa, Rhizostomeae) and Cyrtograpsus affinis (Decapoda,
Varunidae). Lat Am J Aquat Res 40:1090–1093
Schiariti A, Morandini AC, Jarms C et al (2014) Asexual reproduction strategies and blooming
potential in Scyphozoa. Mar Ecol Prog Ser 510:241–253
Schmoker C, Russo F, Drillet G et al (2016) Effects of eutrophication on the planktonic food web
dynamics of marine coastal ecosystems: the case study of two tropical inlets. Mar Environ Res
119:176–188
Sieburth JMCN, Smetacek V, Lenz J (1978) Pelagic ecosystem structure: heterotrophic com-
partments of the plankton and their relationship to plankton size fractions. Limnol Oceanogr
23:1256–1263
Silva AP, Neumann-Leitão S, Schwamborn R et al (2004) Mesozooplankton of an impacted bay in
North Eastern Brazil. Braz Arch Biol Tech 47(3):485–493
5  Plankton Ecology and Biodiversity in the Bahía Blanca Estuary 111

Smyth K, Elliott M (2016) Effects of changing salinity on the ecology of the marine environment.
In: Solan M, Whiteley NM (eds) Stressors in the marine environment. Oxford University Press
Oxford University Press, London
Stoecker DK, Capuzzo JM (1990) Predation on Protozoa: its importance to zooplankton. J
Plankton Res 12:891–908
Suderman BL, Marcus NH (2002) The effects of Orimulsion and Fuel Oil #6 on the hatching suc-
cess of copepod resting eggs in the seabed of Tampa Bay, Florida. Environ Pollut 120:787–795
Toshino S, Tanimoto M, Minemizu R (2019) Olindias deigo sp, a new species (Hydrozoa,
Trachylinae, Limnomedusae) from the Ryukyu Archipelago, southern Japan. Zoo Keys
900:1–21
Truchet MD, Buzzi NS, Carcedo C et al (2019) First record of the fiddler crab Leptuca (Uca) uru-
guayensis in the Bahía Blanca estuary (Buenos Aires, Argentina) with comments on its biology
in South America. Regional Studies in Marine. https://doi.org/10.1016/j.rsma.2019.100539
Tsangaris C, Kormasb K, Strogyloudia E et al (2010) Multiple biomarkers of pollution effects in
caged mussels on the Greek coastline. Comp Biochem Phys 151(3):369–378
Turner JT (2004) The importance of small planktonic copepods and their roles in pelagic marine
food webs. Zool Stud 43:255–266
Uriarte I, Villate F (2004) Effects of pollution on zooplankton abundance and distribution in two
estuaries of the Basque coast (Bay of Biscay). Mar Pollut Bull 49:220–228
Wang WX (2002) Interactions of trace metals and different marine food chains. Mar Ecol Prog
Ser 243:295–309
Weston AJ, Chung R, Dunla WC, Morandini AC et al (2013) Proteomic characterisation of tox-
ins isolated from nematocysts of the South Atlantic jellyfish Olindias sambaquiensis. Toxicon
71:11–17
Whitfield AK, Elliott M, Basset A et al (2012) Paradigms in estuarine ecology –a review of the
Remane diagram with a suggested revised model for estuaries. Estuar Coast Shelf S 97:78–90
Winder M, Cloern JE (2010) The annual cycles of phytoplankton biomass. Philos Trans R Soc
B. https://doi.org/10.1098/rstb.2010.0125
Winder M, Carstensen J, Galloway AWE et al (2017) The land-sea interface: a source of high-­
quality phytoplankton to support secondary production. Limnol Oceanogr 62:S258–S271.
https://doi.org/10.1002/lno.10650
Worden AZ, Follows MJ, Giovannoni SJ et al (2015) Rethinking the marine carbon cycle: factoring
in the multifarious lifestyles of microbes. Science 347(6223):1257594. https://doi.org/10.1126/
science.1257594
Zamponi M, Facal O (1987) Estudio Bioecológico de Olindias sambaquiensis Muller, 1861 en el
área de Monte Hermoso. l. Ciclo de vida. (Limnomedusae, Olindidae). Neotropica 33:119–126
Zamponi MO, Mianzan HW (1985) La mecánica de captura y alimentación de Olindias sambaqui-
ensis Müller, 1861 (Limnomedusae) en el medio natural y en condiciones experimentales. Hist
Nat 5:269–278
Zapperi G, Piovan MJ, Pratolongo P (2017) Community structure and spatial zonation of benthic
macrofauna in mudflats of the Bahía Blanca Estuary, Argentina. J Coast Res 34(2):318–327
Chapter 6
Biology and Ecology of the Benthic Algae

M. Emilia Croce, M. Cecilia Gauna, Carolina Fernández, Ailen M. Poza,


and Elisa R. Parodi

6.1  What Are Algae?

Algae form a heterogeneous group of organisms that range from single microscopic
cells to giant seaweeds belonging to diverse evolutionary lineages. The term algae
has no formal taxonomic standing; it is routinely used to designate a polyphyletic,
non-cohesive, and artificial group of eukaryotes and prokaryotic photosynthetic
organisms. Algae include two main groups, according to their body size, macro-
scopic algae (macroalgae) and a highly diversified group of microorganisms known
as microalgae (Barsanti and Gualtieri 2006). The profound diversity of forms, sizes,
ecological niches, levels of organization, photosynthetic pigments, storage prod-
ucts, structural polysaccharides, and life histories reflects the separate evolutionary
origins of this diverse group.
By definition, algae are considered photoautotrophs; they use sunlight and CO2
to produce carbohydrates and ATP, depending entirely upon their photosynthetic

M. E. Croce () · M. C. Gauna


Instituto Argentino de Oceanografía, IADO (Universidad Nacional del Sur-CONICET),
Bahía Blanca, Argentina
Departamento de Biología, Bioquímica y Farmacia, Universidad Nacional del Sur,
Bahía Blanca, Argentina
e-mail: ecroce@criba.edu.ar
C. Fernández
Instituto Argentino de Oceanografía, IADO (Universidad Nacional del Sur-CONICET),
Bahía Blanca, Argentina
Centro de Emprendedorismo y Desarrollo Territorial Sostenible (CEDETS), Universidad
Provincial del Sudoeste (UPSO) – Comisión de Investigaciones Científicas de la Provincia de
Buenos Aires (CIC), Bahía Blanca, Argentina
A. M. Poza · E. R. Parodi
Instituto Argentino de Oceanografía, IADO (Universidad Nacional del Sur-CONICET),
Bahía Blanca, Argentina

© Springer Nature Switzerland AG 2021 113


S. M. Fiori, P. D. Pratolongo (eds.), The Bahía Blanca Estuary,
https://doi.org/10.1007/978-3-030-66486-2_6
114 M. E. Croce et al.

apparatus for metabolic processes. However, most algal lineages include colorless
heterotrophic species that obtain organic carbon from the environment, either by
taking up dissolved substances or by engulfing bacteria and other cells. Some spe-
cies of algae are called mixotrophs, as they combine both photoautotrophy and het-
erotrophy as nutritional strategies. There is a gradient among these three strategies;
thus, the algae can be classified as obligate heterotrophic, facultative mixotrophic
(e.g., some dinoflagellates), obligate phototrophic (e.g., Dinobryon Ehrenberg spe-
cies), and obligate mixotrophic (e.g., Euglena Ehrenberg species) (Graham
et al. 2016).

6.1.1  Habitats Colonized by Algae

Algae are mainly aquatic organisms, except for subaerial species that live exposed
to the atmosphere and rely on liquid or vapor to carry out their metabolic functions.
Aquatic algae display a broad range of tolerance to pH, temperature, turbidity, and
concentrations of O2 and CO2, being able to colonize almost every aquatic environ-
ment; however, obligate photoautotrophic species are limited to shallow areas
because of the rapid attenuation of light with depth. Aquatic algae can be found
suspended in water bodies (planktonic algae), even under ice in polar areas, or
attached to substrates or within sediments (benthic algae).
Benthic algae, either macroscopic or microscopic, can grow on stones (epilithic),
on mud or sand (epipelic), on the thallus of other algae or plants (epiphytic), or on
animals (epizoic). The term microphytobenthos is used to designate the community
composed of eukaryotic microalgae, mainly diatoms, and cyanobacteria, that live
on illuminated surfaces of a wide variety of aquatic habitats, ranging from tidal flats
and marshes to submerged aquatic vegetation beds and subtidal sediments
(Macintyre et al. 1996; Miller et al. 1996). The term periphyton is also used to des-
ignate this group of algae; however, it is more frequently used to refer to freshwater
benthic microalgae. Even though microphytobenthos is less conspicuous than mac-
roalgae or vascular plants, it contributes significantly to total primary productivity
in coastal areas, and in many shallow aquatic systems, the biomass of the microphy-
tobenthic community far exceeds that of phytoplankton (Pinckney and Zingmark
1993; Daehnick et al. 1992). The distribution of benthic microscopic algae varies
extensively from place to place depending on the presence of suitable substrata,
water depth, light availability, and physical perturbation (Sand-Jensen and Borum
1991). If strongly attenuated irradiances reach the sea bottom because of deep and/
or turbid water, then benthic primary producers will be absent. In contrast, if high
irradiances reach the bottom, as in shallow transparent waters, then the primary
production of benthic microalgae can be dominant (Stevenson 1996).
Microphytobenthic algae occur in marine and freshwater habitats, while benthic
macroalgae mainly habit marine environments. Large marine macroalgae are usu-
ally called seaweeds. Macroalgae that live in estuaries and marine coasts are usually
classified according to the levels of the coast at which they occur. Supralittoral
6  Biology and Ecology of the Benthic Algae 115

macroalgae are those that grow above the high-tide level, within the reach of waves
and spray; intertidal macroalgae are those that grow on shores exposed to tidal
cycles; and sublittoral macroalgae are those that grow in the benthic environment
from the extreme low-water level to around 200 m deep (Graham et al. 2016). Their
distribution along the coastal gradient is related with their photosynthetic capacities
(red and brown macroalgae have accessory pigments that allow them to photosyn-
thesize in regions where the light is attenuated) and their tolerance to salinity and
desiccation. The influence of these factors on macroalgal assemblages will be
described in Sect. 6.3.1.

6.1.2  Forms and Body Types of Algae

Many microalgal species occur solitary as unicellular individuals with or without


flagella, hence motile or non-motile, in a variety of shapes. Other algae exist as
aggregates of several single cells held together loosely or in a highly organized
fashion, the colony. In these types of aggregates, the cell number is indefinite,
growth occurs by cell division of its components, and each cell can survive on its
own. When the number and arrangement of cells are determined at the time of origin
and remain constant during the lifespan of the individual colony, this is termed coe-
nobium (e.g., Volvox Linnaeus). Solitary and colonial species propelled by flagella
are referred to as flagellates; however, not all flagellates include algal species.
Another common type of algal body is the filament, formed by a series of cells
arranged end to end, with adjacent cells sharing a common cross wall, and forming
chains where the daughter cells are connected to each other by their end wall.
Filaments may be formed by one, two, or several rows of cells, and they may be
unbranched as in Oscillatoria Vaucher ex Gomont (Cyanophyceae) or Ulothrix
Kützing (Ulvophyceae), have false branching as in Tolypothrix Kützing ex É. Bornet
and C. Flahault (Cyanophyceae), or have true branching as in Cladophora Kützing
(Ulvophyceae) (Barsanti and Gualtieri 2006).
Macroalgal body is called thallus; it may be constructed by branched or
unbranched filaments, coenocytic siphons (siphonous thallus), parenchyma, or
pseudoparenchyma. Coenocytic thalli are composed of one giant multinucleate cell,
lacking internal cell walls, except for septa that occur in reproductive states.
Examples of this morphology can be found in the family Bryopsidales such as
Bryopsis J.V. Lamouroux and Codium Stackhouse. Parenchymatous thallus is com-
posed of tissues with a three-dimensional array of cells often connected by intercel-
lular connections; this tissue organization is found in Ulva Linnaeus (Ulvophyceae)
and many of the brown algae such as kelps. Pseudoparenchymatous algae are made
up of a loose or close aggregation of numerous, intertwined, branched filaments that
collectively form the thallus, held together by mucilages, especially in red algae
such as Gelidium J.V.  Lamouroux (Florideophyceae). Thallus construction is
entirely based on a filamentous construction with little or no internal cell differen-
tiation (Graham et al. 2016).
116 M. E. Croce et al.

6.2  Classification of Algae

6.2.1  Taxonomic Classification

Among algal systematists, a wide range of classification schemes have been pro-
posed (Bold and Wynne 1978; South and Whittick 1987; Margulis et al. 1990; Van
den Hoek et  al. 1995; Graham and Wilcox 2000). Considering the polyphyletic
nature of algae as a group of organisms, algal classification is somewhat difficult to
cope with the traditional taxonomic systems. However, it is still useful to represent
the general characters and levels of organization, despite the fact that taxonomic
opinion may change as new information accumulates.
The algae are traditionally classified into lineages or phyla, according to numer-
ous characteristics: the photosynthetic pigments; the chemical nature of the energy
storage product; the organization of the thylakoid membranes and other features of
the chloroplasts; the composition and structure of the cell wall; the number, arrange-
ment, and ultrastructure of the flagella; and the life cycle. Cyanobacteria make up a
particular phylum of the domain Eubacteria, whereas eukaryotic algae are classified
into more than 10 phyla of the domain Eukarya; these are Chlorophyta (green
algae), Rhodophyta (red algae), Glaucophyta (glaucophytes), Cryptophyta (crypto-
monads), Dinophyta (dinoflagellates), Bacillariophyta (diatoms), Phaeophyta
(brown algae), Haptophyta (haptophytes), Chlorarachniophyta (chlorarachnio-
phytes), and Euglenophyta (euglenoids). Molecular studies aiming to assess the
internal genetic coherence of nuclear genes and ribosomal RNA are increasingly
unraveling the relationships among these major lineages (Graham et al. 2016).
Cyanophyta (also known as blue-green algae, cyanobacteria, or cyanophytes) are
unique algae, as they are prokaryotes. This is the most ancient algal lineage, with
fossils dating back almost 3000 million years. The blue-green algae exhibit a variety
of forms and are the most widely distributed of algal groups. Their cellular organi-
zation can be unicellular, branched or unbranched filamentous, or unspecialized
colonial. The cyanophytes are distributed in marine and freshwater environments,
occasionally forming blooms in eutrophic waters. They are an important component
of benthic systems forming mats on soil, mudflats, and hot springs but are less con-
spicuous in soils along reef margins. Cyanobacteria can be found as symbiotic
organisms in diatoms, ferns, lichens, cycads, sponges, algae, and other systems
(Stewart et al. 1983; Costa and Lindblad 2002; Charpy et al. 2012; Rikkinen 2013).
Glaucophyta algae can be unicellular flagellates or form colonies in freshwater
habitats, although some species can be found in soil samples. They are dorsoven-
trally constructed and have blue-green photosynthetic plastids. Glaucophytes are of
particular importance in evolutionary studies because their plastids differ from
those of other eukaryotic algae and resemble cyanobacteria in some way (Barsanti
and Gualtieri 2006).
Red algae, or formally, Rhodophyta, occur as single cells, individual filaments,
aggregations of filaments, or sheets of cells. Photosynthetic pigments are present in
the red plastids of most species, except in certain parasitic genera. They are unusual
6  Biology and Ecology of the Benthic Algae 117

among eukaryotes because of the lack of flagella in any stages of their life cycle and
the presence of accessory phycobiliproteins organized in phycobilisomes. Although
a few studies have noticed subtle motility of red algal propagules (Rosenvinge 1927;
Geitler 1944; Lin et al. 1975; Hill et al. 1980; Pickett-Heaps et al. 2001), it is gener-
ally accepted that they are non-motile, especially compared with brown and green
algae. The dispersal and settlement of red algal propagules are strongly dependent
on the abiotic factors of the benthic environment (Ogata 1953; Luther 1976; Harlin
and Lindbergh 1977). In most species, cytokinesis is incomplete; thus, the daughter
cells are separated by a proteinaceous plug that fills the junction between cells (pit
connection), which successively becomes a plug. Most red algae have sexual life
cycles which usually involve alternation of three generations. Red algae are espe-
cially diverse and abundant in tropical and subtropical marine waters, but they are
also present in freshwater and terrestrial environments. Red algae can be classified
under two main classes, the Bangiophyceae that retain morphological characters
that are found in the ancestral pool of red algae, ranging from single cells to multi-
cellular filaments or sheetlike thalli, and the Florideophyceae that include morpho-
logically complex red algae and are widely considered to be a derived, monophyletic
group. One of the most striking features of the Florideophyceae is that they are the
most diverse algal group regarding the thallus construction.
The phylum Phaeophyta is defined by one particular feature, that is, the presence
of two different flagella in the cells. Flagellate cells are thus termed heterokont, pos-
sessing a long mastigonemate flagellum, and a short smooth one that points back-
ward along the cell. This division includes several classes, Phaeophyceae,
Bacillariophyceae, Cryptophyceae, Dictyochophyceae, Haptophyceae, and
Xanthophyceae, of which the first two are  the most relevant in coastal environ-
ments. The class Phaeophyceae (brown algae) is almost exclusively marine and is
dominant in temperate waters. They include more than 250 genera, ranging from
microscopic filaments to giant kelps that can reach 80  m long. Some species of
Phaeophyceae display the greatest complex organization of tissues and cells among
algae. Many have photosynthetic blades, as well as specialized blade-bearing stipes,
holdfasts, and specialized conductive cells. The class Bacillariophyceae, integrated
by organisms informally called diatoms, represents the most common algal group.
They are very abundant and thus significant primary producers responsible for an
estimated 20% of global carbon fixation. Diatoms dominate the phytoplankton of
cold, nutrient-rich waters, such as upwelling areas of the oceans and recently circu-
lated lake waters. Diatoms are the most significant producers of biogenic silica,
dominating the marine silicon cycle. It is estimated that over 30  million km2 of
ocean floor is covered with sedimentary deposits of diatom frustules (Harris et al.
2006). These organisms are also important components of the benthic estuarine
environment forming the microphytobenthic community.
The members of the phylum Dinophyta (dinoflagellates) are typically unicellular
flagellates. Dinoflagellates have two flagella with independent beating patterns,
conferring a characteristic rotatory swimming motion. Flagella are apically inserted
in a region close to the midpoint of the ventral side of the cell. They are common
components of the freshwater and marine habitats (Barsanti and Gualtieri 2006).
118 M. E. Croce et al.

The phylum Euglenophyta includes mostly unicellular widely distributed flagel-


lates; predominantly occupants of interfaces, such as the air-water and sediment-­
water boundaries. They are especially abundant in highly eutrophic environments
and are known to be tolerant to extreme conditions of desiccation, low pH, and heat
(Walne and Kivic 1990).
Chlorarachniophyta are  amoeboid, coccoid or  flagellate cells with secondary
green plastids (Hibberd and Norris 1984). They are phototrophic and phagotrophic,
engulfing bacteria, flagellates, and eukaryotic algae. The name chlorarachniophytes
refers to the green color of their plastids and the spider shape of the cells.
Chlorarachniophytes occur in temperate and tropical marine waters, growing among
sand grains, on mud, in tide pools, on seaweeds, or in plankton.
The phylum Chlorophyta (green algae) displays a large range of somatic differ-
entiation varying from flagellates to complex multicellular thalli differentiated into
macroscopic structures that resemble organs. The different levels of thallus organi-
zation (unicellular, colonial, filamentous, siphonous, and parenchymatous) have
served as the basis of their classification. Chlorophytes include at least nine lineages
of early-diverging, unicellular prasinophytes (Lemieux et  al. 2014) and a more-­
derived, monophyletic assemblage known as the “core Chlorophyta” (Fučikova
et  al. 2014). Traditionally, the chlorophytes were divided into the classes
Ulvophyceae (ulvophyceans), Trebouxiophyceae (trebouxiophyceans), and
Chlorophyceae (chlorophyceans).
The taxonomic treatment that will be used throughout this chapter for the algae
occurring in the Bahía Blanca Estuary originates from morphological identifica-
tions of recently collected specimens or from local coastal reports. For this reason,
some algae will be mentioned to the genus level, while others to the species level.

6.2.2  Morpho-functional Classification

In every biological community, each species occupies a unique ecological niche and
commonly there are groups of species that utilize the resources in similar ways.
That is, there are species that may be geographically distant and evolutionarily dis-
tinct, but have similar ecological functions in the ecosystem and thus occupy similar
adaptive zones. As mentioned before, benthic algae can be single cells or large sea-
weeds with internal structural complexity. This variability and diversity can be sim-
plified by classifying algae into functional-morphological categories. In oceanic
systems, this classification is used to describe macroalgal assemblages (Vanderklift
and Lavery 2000; Konar and Iken 2009), to address physiological questions (Littler
and Arnold 1982; Johansson and Snoeijs 2002), and to examine the impact of sev-
eral types of disturbance on benthic communities (Dethier 1981). The recognition
of such aggregates by ecologists enables them to understand and predict the out-
come of interspecific interactions and to interpret patterns in community structure
6  Biology and Ecology of the Benthic Algae 119

without studying individual species (Steneck and Watling 1982). Littler and Littler
(1980) and Steneck and Watling (1982) proposed models for algae, where the over-
all form of the thallus was hypothesized to predict aspects of its physiology and
ecology, and resistance to consumers. The general models hypothesize that the type
of growth and mineralization of algae dictates relative rates of primary productivity,
growth rate, competitive ability, resistance to consumption by grazers, resistance to
physical disturbance, tolerance to physiological stress, and successional stage, and
that all of these functions are correlated with each other.
The model proposed by Littler and Littler (1984) considers algal form groups:
sheet, filamentous, coarsely branched, thick leathery, jointed calcareous, and crus-
tose. Steneck and Watling (1982) later created a slightly different ranking that
included filamentous, foliose, corticated, leathery, articulated calcareous, and
crustose forms. A more recent model of classification, proposed by Steneck and
Dethier (1994), is based on the productivity and susceptibility to grazing, incorpo-
rating algal groups such as microalgae, filamentous, crustose, foliose, corticated
foliose, corticated macrophyte, leathery macrophyte, and articulated calcareous
(Table 6.1). Balata et al. (2011) proposed a new expanded classification of mor-
phological functional groups based on several characteristics of the thallus like
structure, growth form, branching pattern, and also on the taxonomic affinities of
the alga (Table 6.2).

Table 6.1  Classification of algae into morpho-functional groups


Functional External
groups morphology Internal anatomy Texture Representatives
Microalgae Single cell Diatoms – Cyanobacteria
Filamentous Delicately Uniseriate, Soft, Gelidium, Giffordia,
branched multiseriate, or flexible Cladophora, Ectocarpus,
lightly corticated Acrochaetium,
Chaetomorpha
Sheetlike Thin tubular or Uncorticated, Soft, Colpomenia, Dictyota,
flattened one-several cells flexible Ulva, Porphyra
thick
Coarsely Coarsely Corticated Fleshy-­ Codium, Colpomenia
branched branched, wiry
upright, terete
Thick Thick blades Differentiated, Leathery, kelps
leathery and branches heavily corticated, rubbery
thick walled
Jointed Articulated, Calcified genicula, Stony Corallina
calcareous calcareous, flexible
upright intergenicula
Crustose Epilithic, Calcified or Stony or Ralfsioid crust
prostrate, uncalcified parallel tough
encrusting cell rows
Adapted from Littler and Littler (1984) and Steneck and Dethier (1994)
120 M. E. Croce et al.

Table 6.2 Classification of macroalgae from the Bahía Blanca Estuary into morpho-­
functional groups
Functional group Description Taxa
Chlorophyta
1. Filamentous Uniseriate filaments, either Cladophora surera,
uniseriate branched or not Chaetomorpha linum
2. Bladelike Thalli forming blades one to few Ulva lactuca, Ulva prolifera,
layers of cells thick Ulva intestinalis, Ulva flexuosa
3. Tubular Macroscopic thalli consisting of Blidingia marginata, Blidingia
one-layered hollow tubes minima
4. Siphonous with thin Algae consisting of branched, Bryopsis plumosa
separate filaments single-celled filaments
Phaeophyta
1. Filamentous Uniseriate branched filaments Ectocarpus siliculosus,
uniseriate Hincksia hincksiae
2. Compressed with Corticated algae producing Punctaria latifolia,
bladelike habit expanded blades with erect habit Planosiphon nakamurae
Rhodophyta
1. Filamentous Thin filaments (up to 2 mm wide), Polysiphonia abscissa,
uniseriate and forming bushy branched thalli Polysiphonia morrowii,
pluriseriate with erect devoid of (or with limited) Ceramium diaphanum,
thallus prostrate parts Acrochaetium sp.
2. Bladelike Thalli forming blades one to few Pyropia sp.
layers of cells thick
3. Larger-sized Branched corticated algae with Gelidium pusillum, Gelidium
corticated erect habit and comparatively thick crinale, Gracilaria verrucosa
axes (more than 2 mm wide)
4. Smaller-sized Branched corticated algae with Neosiphonia harveyi
corticated erect habit and relatively thin axes
(up to 2 mm wide)
According to Balata et al. (2011)

6.3  F
 actors Affecting Algal Assemblages
in Estuarine Environments

6.3.1  Abiotic Factors

Estuarial environments are fluctuating habitats, characterized by a mixture of fresh-


water and seawater which originates unstable conditions; hence, the communities
that develop in estuaries are well adapted to this habitat. Physical and chemical fac-
tors of the estuarine environment clearly influence the functional morphology and
behavior of the benthic communities. For example, in intertidal flats community
structure is regulated by sediment particle size, oxygen concentration, salinity, and
temperature, to which the species may respond differently (Coull 1999; Schweiger
et al. 2008). Periodic inundation and exposure to air is harsh in these environments.
Light intensity, salinity, pH, nutrient levels, grazing, and sediment stability limit the
6  Biology and Ecology of the Benthic Algae 121

productivity of algae (Kennish 2017). In estuaries, high biogeochemical rates


account for a relatively low number of species (Costanza et al. 1993); this is because
only a limited number of species are adapted to salinity variation (Webb et al. 1997).
Microphytobenthic and macroalgal communities are regulated by light, tempera-
ture, and nutrient availability, which are the major parameters controlling their
reproduction (Denis et al. 2012). Seaweeds disperse either as free-floating macro-
scopic forms or as microscopic propagules (e.g., spores, gametes, and zygotes). The
spatial patterns of seaweed in estuaries are contingent upon the dispersal capabili-
ties of the populations; hence, tidal currents play a crucial role in these habitats as
they influence the dispersal, settlement, and attachment of macroalgae
(Kennish 2017).

6.3.2  Biotic Interactions

Community structure of estuarine environments is influenced by various biotic


interactions, involving grazing and competition (Buffan-Dubau and Carman 2000).
The trophic dynamics of estuaries are complex because this environment is occu-
pied by different types of primary producers such as phytoplankton, salt marsh
plants, submersed seagrass, and benthic algae. Unlike the open sea, where practi-
cally all phytoplankton is consumed alive, in estuaries, primary producers are not
heavily grazed, but die and begin to decompose before being consumed, contribut-
ing to the detritus food web (Omstedt et al. 2014). Algae generate abundant organic
detritus constituting an important food source available all year round; they also can
provide new habitats where some species can find a refuge from predators (Valiela
et al. 1997; Raffaelli et al. 1998). Thus, organic matter from benthic macroalgae
together with phytoplankton may support both benthic and pelagic food webs in the
intertidal and adjacent subtidal areas of shallow bay systems (Kang et  al. 2003).
Macroalgal thallus tends to have higher nitrogen and phosphorus contents than sea-
grass tissues (Atkinson and Smith 1983; Duarte 1995). When macroalgal decompo-
sition occurs, the derived nutrients accumulate and increase the nutritional value of
the sediment, stimulating the growth of other primary producers, microalgae and
bacteria (Rossi 2006), therefore promoting the increase in abundance and biomass
of grazers and deposit feeders (Hull 1987; Ford et al. 1999; Rossi and Underwood
2002). In the case of organic matter transported through the water column, this is
used by filter feeders such as oysters, clams, and mussels; while the organic mat-
ter that accumulates in the sediment is later used by deposit feeders such as worms,
amphipods, and a many other small organisms (Day et al. 2013).
In estuarine soft sediments, macroalgae are often spatially distributed forming a
mosaic of patches of different species interspersed with bare substratum (Berezina
et  al. 2007). Several studies have detected that benthic macroalgal communities
support higher abundances of both epibenthos and infauna than do other compara-
ble unvegetated bottoms (Summerson and Peterson 1984; Irlandi and Peterson
1991). The effects of macroalgae on benthic organisms are probably density
122 M. E. Croce et al.

dependent, because at moderate densities, macroalgae create a more heterogeneous


environment (Thiel and Watling 1998). This favors the increase of subsurface
deposit feeders and the abundance of epifaunal species, which find abundant shelter
and food in the new algal substratum (Raffaelli et al. 1998). At high densities, mac-
roalgal blooms alter the natural balance between production and decomposition of
organic matter and can have dramatic effects on the local fauna (Valiela et al. 1997).
The extent of these effects depends on the composition of the dominant faunal
assemblages at the bloom onset, the magnitude and nature (drifting or steady) of the
bloom, the season, and, to a lesser extent, the type of bloom-forming algae (Raffaelli
et al. 1998).
Isaksson and Pihl (1992) observed that when algae cover approximately 30–50%
of the surface in soft-bottom marine ecosystems, the abundance of epibenthic fauna
associated with vegetation increases, whereas with algal coverages of 90%, the
abundance of mobile epibenthic fauna declines. The decomposition of large depos-
its of macroalgae can affect the availability of oxygen in the sediment and cause
episodes of anoxia and sulfide production, accentuated by reduced water flow under
the algae and increased sedimentation rate (Nedergaard et al. 2002). In these situa-
tions, the abundance and diversity of the assemblages can decrease because animals
can be displaced or killed or because of an alteration between predator and prey
relationships (Renaud et al. 1999; Kelaher and Levinton 2003). In areas with fluctu-
ating oxygen levels, mobile predators can temporarily leave hypoxic areas and then
return when oxygen rises to capture the infaunal invertebrates that emerged during
low oxygen (Nestlerode and Diaz 1998).
Infauna from shallow soft bottoms, for example burrowing adult bivalves,
respond to high biomass of bloom-forming algae by migrating on the surface.
Juvenile bivalves, on the other hand, can reach high abundance among algal strands
and filaments, while highly mobile species, such as the gastropod Hydrobia
W. Hartmann, 1821, often reach high densities in algal mats (Raffaelli et al. 1998).
In contrast, Franz and Friedman (2002) observed that the blooms of Ulva lactuca
Linnaeus drastically reduced the abundance of epibenthic copepods, attributing that
reduction to the anoxic conditions registered within the algal mats. In drifting algae,
Norkko et al. (2000) found high epifaunal abundance on the macroalgae Ectocarpus
siliculosus (Dillwyn) Lyngbye and Pilayella littoralis (Linnaeus) Kjellman, with
intermediate oxygen levels within the drifting seaweed, and only found hypoxic
conditions at the algal-sediment interface. In the case of the mobile infauna, fila-
mentous algae provide a refuge from physical stresses. However, for sedentary
infauna, the cover by laminar algae forms a barrier between the animals and the
oxygenated water column generating an adverse effect on survivorship (Perkins and
Abbott 1972; Bach and Josselyn 1979; Dauer and Conner 1980). The physical
structure provided by algae serves as a refuge and reduces the predation rate of
mobile epibenthic species in soft-sediment communities (Lenanton et  al. 1982;
Robertson and Lenanton 1984; Marx and Herrnkind 1985; Wilson et al. 1990; Smith
and Herrnkind 1992). For instance, the increased cover of Ulva and Cladophora
reduces the predation rates of the crustaceans Crangon crangon (Linnaeus, 1758)
and Carcinus maenas (Linnaeus, 1758) by the cod Gadus morhua (Linnaeus, 1758)
6  Biology and Ecology of the Benthic Algae 123

(Isaksson et al. 1994; Pihl et al. 1995). The juveniles of many fish species also find
shelter and abundant food between macroalgal communities, and, after reaching a
certain size, they swim to deeper waters (Coles et al. 1993; Ross and Moser 1995;
Rooker et al. 1998). Seagrass meadows thus play an important role as “nurseries”
for numerous fish species (Heck et al. 2003).
In summary, anoxic conditions have an overall negative effect on the abundance
of the faunal community in spite of all the other possible concomitant positive
effects of the bloom (i.e., enhanced refuge, food, and structure for recruitment)
(Sagasti et al. 2001). The relative tolerance of organisms to hypoxic conditions gen-
erated by algae may determine the increase or decrease of  predation rates. For
example, predation on larval fish by jellyfish can increase during hypoxic episodes,
because jellyfish tolerate hypoxia, but larval fish are unable to evade predators in
low oxygen. Concurrently, predation on larval fish by adult fish decreases, because
the adult fish are intolerant to hypoxia (Breitburg et al. 1994).
On the other hand, it is important to mention the impact of herbivores on the
biomass of benthic algae (i.e., top-down control). Microphytobenthic algae are
affected by grazing (Hillebrand et al. 2000). Several studies report that the extent of
herbivory on benthic algae is highly variable (Cebrian et al. 1998; Cebrian 1999,
2002). In some cases, herbivores may remove a large percentage of the production
of microphytobenthic communities (Nicotri 1977; Baird and Ulanowicz 1993), but
in other instances, herbivory only represents a minor loss for the community
(Admiraal et al. 1983; Montagna 1984). In spite of the large variability found within
each community type, microphytobenthos and macroalgae tend to lose higher per-
centages of primary production due to grazing  than do seagrass communities
(Bennett et al. 1999; Blanchard et al. 2001), influencing the dynamics of producers
biomass and nutrient recycling.

6.4  Adaptations Mechanisms of the Benthic Algae

Estuaries are habitats with high variability; hence, the communities that live in them
are frequently disturbed. Some species of algae rapidly recolonize the substrates
after a disturbance; a new generation can appear within a few weeks, and several
generations may develop in one year. Such a fast succession of generations allows
the adaptation of algal species that are able to tolerate the abrupt changes that occur
in estuarine conditions (Larsen and Sand-Jensen 2006). These changes favor the
development of fast-growing, short-lived, and morphologically simple algae, like
phytoplankton, epiphytic algae, and ephemeral macroalgae, whereas benthic mac-
rophytes such as kelp do not occur (Pedersen and Borum 1996).
Macroalgal species tend to exhibit vertical patterns of distribution, from the
upper to the lower tidal levels, originating different zonation patterns. This is
because different species have different adaptive responses to the physical, chemi-
cal, and biotic conditions. The lack of hard substrates, the low light penetration, the
variations of salinity, pollution, competition, and grazing are factors that explain the
124 M. E. Croce et al.

reduced seaweed richness of estuaries and the occurrence of some annual species
(Druehl 1967; Larsen and Sand-Jensen 2006). Although estuaries are stressful envi-
ronments to most marine algae, a few adapted species can flourish. For example,
eutrophic estuaries promote the proliferation of opportunistic and tolerant macroal-
gae like Ulva, Chaetomorpha Kützing, Cladophora, Monostroma Thuret, Ceramium
Roth, Gracilaria Greville, Porphyra C.  Agardh, Pyropia J.  Agardh, Ectocarpus
Lyngbye, and Pilayella Bory, which can bloom into large proportions (Morand and
Merceron 2005; Scanlan et al. 2007).
Wilkinson (1981) exposed three fundamental points to explain macroalgal distri-
bution patterns in estuaries: (1) The colonization occurs almost wholly by marine
species, with freshwater ones abundant only in the uppermost reaches of the estuary.
(2) In the uppermost regions of estuaries, there is a reduction in species number and
diversity due to the decrease of red algae (Rhodophyta) and then of brown algae
(Phaeophyta). Green algae (Chlorophyta) and blue-green algae (Cyanophyta) do
not necessarily become more numerous, in terms of species, but they constitute a
much greater proportion of the algal community because they extend further inland.
(3) Colonization of the mid-reaches is dominated by brackish water algae.
The algae that live in estuaries generally proliferate under a broad range of tem-
peratures, irradiances, salinities, and nutrient conditions (Martins et al. 1999; Taylor
et al. 2001; Sousa et al. 2007; Choi et al. 2010; Kim et al. 2011). For example, the
growth rate of Ulva increases at low salinity and high nutrient levels in laboratory
culture conditions (Taylor et al. 2001). Ulva has physiological adaptations to grow
as free-floating mats. Some species of Ulva have partially hollow branched tubular
thalli filled with air generated through photosynthesis, which under unfavorable
conditions produce thalli that can float for 2 or 3 months favoring survival (Kim
et al. 2011).
The filamentous algae are well adapted to overcome the low light intensities of
estuarine waters. They have thin photosynthetic tissues, with high contents of pig-
ments per cell volume that reduce the trajectory of light through the tissue. As they
have high area/volume ratio, they use efficiently the light and nutrients per biomass
unit (Kirk 1994; Niklas 1992; Nielsen and Sand-Jensen 1990; Duarte 1995; Duarte
et al. 1995).
Estuarine filamentous green algae with heterotrichous growth such as Cladophora
present densely pigmented assimilatory cells that penetrate upward through the mud
and the covering mats that contrast greatly with the weakly pigmented cells of the
prostrate system. This morphological characteristic constitutes an ecological adap-
tation to burial by soft sediment, as it helps retain moisture during low tides
(Boedeker and Hansen 2010).
Cladophorales algae have other strategies for protection against desiccation and
fluctuating salinities, for example, the formation of a thick gelatinous cover pro-
duced by swelling of the outer cell wall layers (Wille 1909), endophytic habit
(Polderman 1976), and the presence of hematochrome/oil droplets (Wille 1909).
Chaetomorpha has been mainly studied for its ecological role as a possible regu-
lator of nutrient availability in estuarine habitats (Krause-Jensen et al. 1996, 1999;
Menendez 2005) and for its ability to tolerate a wide range of salinities. The
6  Biology and Ecology of the Benthic Algae 125

identification of ascorbate oxidase activity in Chaetomorpha linum (O.F. Müller)


Kützing suggests a novel mechanism of adaptation to increased salinity, because
this enzyme could have a role in salt stress by catalyzing intracellular production of
water, which could mitigate the stress (Caputo et al. 2010).
Algae that are unicellular, colonial, or have a thallus with 1–4 cell layers can
change their pigment concentration, affecting light absorbance as a mechanism of
adaptation to irradiance variations (Agusti et al. 1994).
In most estuaries, microphytobenthic communities are composed of a mixture of
different taxa that can form conspicuous biogenic structures on intertidal and supra-
tidal sediments commonly called biofilms and microbial mats, the latter being
among the oldest ecological structures on Earth (van Gemerden 1993). The main
difference between biofilms and mats is that the first one is formed by a single layer
of microorganisms embedded in an organic matrix, while microbial mats are verti-
cally stratified microbial communities dominated by cyanobacteria. Microbial mats
are regarded as advanced stages of biofilms (Noffke 2010). Biofilms and microbial
mats result from the growth and activity of microphytobenthic organisms that trap
sediment particles and bind them in extracellular polymeric substances (EPS) pro-
duced by the organisms (Margulis et  al. 1980; Krumbein 1994). Although many
microorganisms are capable of secreting EPS, phototrophs are especially important
since they produce them de novo through CO2 fixation, while chemotrophic organ-
isms form EPS by converting other organic compounds, which may be limited
(Stal 2006).
Extracellular polymeric substances are a complex mixture of high-molecular-­
weight polymers consisting of 90% or more of polysaccharides (Hoagland et  al.
1993). They allow the locomotion of microorganisms and also provide protection
against changes in salinity, temperature, UV radiation, and desiccation (Decho
2000). At the same time, they generate high cohesion in the sediment since they
form an adherent cover on the particles (de Winder et  al. 1999; Decho 2000).
Microphytobenthos develops mainly in marine sediments of the intertidal and
supratidal region because the extreme and strongly fluctuating environmental con-
ditions that prevail in this region exclude, or at least reduce the abundance of grazers
(Stal 2006). In open coastal environments, coarse sand is deposited over high-­
energy areas, becoming unsuitable for the development of microbial mats or bio-
films because of the damage caused by the abrasion of sand particles induced by the
action of wind and waves (Eckman et al. 2008). On the other hand, silt and clay
sediments are deposited in low-energy areas such as estuaries and deltas; these sedi-
ments are characterized by low light penetration and high sedimentation rate, which
difficult the phototactic migration response of cyanobacteria. In addition, this type
of sediment usually contains large amounts of nutrients, conditions under which the
cyanobacteria are outcompeted by opportunistic organisms, mainly diatoms, which
have high growth rates under high nutrient conditions. This is why cyanobacteria
prefer sediments of fine to medium sand as a substrate for the formation of micro-
bial mats, while in silt-clay sediments, biofilms of diatoms predominate (Watermann
et al. 1999).
126 M. E. Croce et al.

In extreme conditions, microphytobenthic organisms are able to regulate their


photosynthesis to avoid photoinhibition (Cartaxana et al. 2011), thereby maintain-
ing relatively high abundance in the sediment. For example, epipelic life forms are
able to migrate vertically in the sediment to position themselves in favorable light
conditions, and diatoms can adapt their photosynthetic apparatus efficiently to the
light conditions in a few minutes (Glud et al. 2002). In these ways, physiological
photoinhibition is avoided at high light levels of solar radiation. However, the
microgradients of sediment grain sizes and organic particle sizes, percentages of
organics, porosity, light attenuation, and oxygen govern vertical microalgal distri-
bution patterns (Kennish 2017).

6.5  The Benthic Algae of the Bahía Blanca Estuary

Among the benthic algae that occur in the Bahía Blanca Estuary, microscopic
epipelic algae are the most studied. The first studies of the microphytobenthos
appeared in the 1980s (i.e. Cicerone 1987; Farías 1988) which initially depicted the
diversity of the microscopic algae of this region. These preliminary studies were
subsequently followed by qualitative studies conducted on specific coasts of the
estuary (Parodi and Barria 2003). The most recent studies focus on ecological
aspects (Da Rodda 2004) or were motivated by the anthropic impact caused by the
expansion of the industrial area (Pizani 2009).
Thanks to these studies, there is a fair knowledge of the diversity of microalgal
taxa that can be found in many of the coastal regions of the estuary. The relevance
that diatoms and Cyanophyceae have in the process of sediment stabilization and
their contribution to the formation of microbial mats has led the phycological stud-
ies in this region. These topics will be discussed later in the chapter and in Box 6.1.
A different scenario stands for the macroscopic algae (including their micro-
scopic life stages), to which less attention has been paid. There are fair explanations
for this situation. The Bahía Blanca Estuary is located in a coastal region considered
to be poor in macroalgal flora (Liuzzi et al. 2011). From a biogeographic point of
view, the Bahía Blanca Estuary is located in a transition zone between the
Argentinean and the Magellanic biogeographic provinces that is characterized by a
reduction of macroalgal biodiversity (Balech and Ehrlich 2008). This scarcity of
macroscopic algae is a consequence of two factors that are related with the charac-
teristics of the sediment and its dynamics: (1) in the majority of the coasts, the
dominant fraction of the sediment is composed of small-sized particles, which may
prevent the settlement of the macroalgal reproductive cells and/or compromise the
successful attachment of the microscopic life stages of macroalgae (Parodi 2004);
and (2) the turbid waters resulting from the suspended fine sediment reduce the
penetration of light into the lower layers of the water column, limiting the depth at
which the macroalgae are able to photosynthesize. These combined conditions
restrict the occurrence of the microscopic and macroscopic thalli to the hard sub-
strata submerged in shallow waters or exposed to air during low tide, where the
6  Biology and Ecology of the Benthic Algae 127

reproductive cells, either motile or non-motile, can fix and access to light intensities
that allow them to photosynthesize. Examples of appropriate substrates are the con-
solidated sediments and any artificial or natural solid object set on the coast. Despite
that the turbid waters significantly reduce the penetration of light in the water, the
shallow tidal flats that span in some sectors of the Bahía Blanca Estuary allow the
development of small intertidal communities of macroalgae, where several species
can thrive.
As a consequence of these unfavorable conditions, the macroalgal assemblages
of the Bahía Blanca Estuary are less conspicuous and go unnoticed, unlike the
assemblages of other coasts of Argentina. The estuarial coasts have high sedimenta-
tion rates; therefore, some macroalgae are usually buried under the sediment. This
adds to the fact that the actual diversity of macroalgal species is usually obscured by
the existence of cryptic species. As the worldwide trend suggests, it is expected that
the number of macroalgal taxa in the Bahía Blanca Estuary increases as the coastal
industrialization advances, providing available artificial substrate, increasing the
nutrient loads in the water, and promoting the ingression of alien marine species
through ballast water.

6.5.1  D
 iversity and Composition of Soft-Bottom
Microalgal Assemblages

Many studies about different microphytobenthic communities have been carried out
in the Bahía Blanca Estuary to examine the biodiversity, structure, and dynamics of
this estuarine benthic ecosystem (Cicerone 1987; Parodi and Barría de Cao 2003;
Da Rodda and Parodi 2005; Pizani 2009). A particular feature of the tidal flats of the
Bahía Blanca Estuary is that they are colonized by extensive microbial mats
(Cuadrado and Pizani 2007; Cuadrado et al. 2011, 2012). They have been widely
studied to explore the relationship between microphytobenthos, sediment, and
physical factors, such as irradiance, temperature, sedimentation rate, and wave
height (Cuadrado et al. 2012, 2013; Pan et al. 2013a). Most of these studies have
been conducted in Puerto Rosales and, to a lesser extent, in Puerto Cuatreros,
Villarino Viejo, and Almirante Brown locations (Figs. 2.2, 2.3, and 2.4; Chap. 2).
Even though the surface of the mudflats of the Bahía Blanca Estuary is often
apparently devoid of vegetation, the richness of microphytobenthic algae is high
since a total of 144 taxa have been recorded by different authors (Parodi and Barría
de Cao 2003; Da Rodda 2004; Da Rodda and Parodi 2005; Pizani 2009; Fernández
et  al. 2018). Diatoms are the dominant group with 109 taxa, whereas 34 taxa of
Cyanobacteria and only 1 taxon of Euglenophyta have been registered. Regarding
diatoms, Nitzschia Hassall and Navicula Bory are the best represented genera, with
27 and 13 taxa, respectively. Figure 6.1 shows some common benthic microalgae
found in the Bahía Blanca Estuary.
128 M. E. Croce et al.

Fig. 6.1  Common benthic microalgae of the Bahía Blanca Estuary, (a) Euglena sp. (scale bar:
6 μm), (b) trichomes of Coleofasciculus chthonoplastes (scale bar: 20 μm), (c) biofilm of diatoms
and filamentous cyanobacteria developing on hard substrate (scale bar: 50 μm), (d) centric diatom
Melosira and filamentous cyanobacteria (scale bar: 6  μm), (e) centric diatoms on Blidingia sp.
(scale bar: 60 μm), (f) Nitzschia clausii (scale bar: 50 μm), (g) Gyrosigma spencerii (scale bar:
50 μm), (h) Cylindrotheca closterium (scale bar: 6 μm), (i) Navicula sp. (scale bar: 6 μm). (Photos
by (a) Natalia Pizzani, (b, d, f, g, h, and i) Carolina Fernandez, (c and e) M. Emilia Croce)
6  Biology and Ecology of the Benthic Algae 129

The unique characteristics of estuarine environments allow the coexistence of


freshwater, estuarine, and marine species; according to this, the diatom species
found in the microphytobenthos of the Bahía Blanca Estuary are mostly holo-­
euryhaline forms, which can tolerate large changes in the salinity of water from
hypotonic to hypertonic. On the other hand, freshwater species are also present,
which arrive from the discharges of freshwater tributaries (Pizani 2009). In addition,
some of the species found in surface sediments are typically planktonic forms,
namely, species of Triceratium Ehrenberg, Podosira stelligera (Bailey) A. Mann,
Planothidium delicatulum (Kützing) Round and Bukhtiyarova (reported as
Achnanthes delicatula (Kützing) Grunow), Luticola mutica (Kützing) D.G. Mann,
Craticula halophila (Grunow) D.G. Mann (cited as Navicula halophila (Grunow)
Cleve), Navicula salinicola Hustedt (reported as Navicula incertata Lange-­
Bertalot), Gyrosigma attenuatum (Kützing) Rabenhorst, Paralia sulcata (Ehrenberg)
Cleve, and Petrodictyon gemma (Ehrenberg) D.G. Mann (cited as Surirella gemma
(Ehrenberg) Kützing), or are species that are usually found in both planktonic and
benthic habitats, such as Entomoneis alata (Ehrenberg) Ehrenberg, Cylindrotheca
closterium (Ehrenberg) Reimann and J.C. Lewin, Gyrosigma fasciola (Ehrenberg)
J.W. Griffith and Henfrey, Nitzschia sigma (Kützing) W. Smith, and Petrodictyon
gemma (Ehrenberg) D.G.  Mann (Parodi and Barría de Cao 2003; Da Rodda and
Parodi 2005; Pizani 2009). This is because the surface of the sediment in the tidal
mudflat is constantly subjected to the action of waves and tidal currents, which
results in the resuspension of benthic individuals, as well as in the incorporation of
planktonic species by sedimentation.
The microphytobenthic community of Puerto Rosales consists mainly of fila-
mentous cyanobacteria and small diatoms. Among filamentous cyanobacteria,
Coleofasciculus chthonoplastes (Thuret ex Gomont) M. Siegesmund, J.R. Johansen,
and T. Friedl (reported as Microcoleus chthonoplastes Thuret ex Gomont) is domi-
nant, whereas species of Oscillatoria and Arthrospira Sitzenberger ex Gomont are
less abundant. C. chthonoplastes typically has many trichomes within a common
sheath threaded into a spiral arrangement; the resulting mesh of interweaving cya-
nobacterial filaments together with the microbially secreted EPS traps the sand
grains and significantly increases the cohesiveness of sediments (Stal et al. 1985).
In that sense, the dominance of C. chthonoplastes is indicative of well-developed
microbial mat, presenting an elevated resistance to erosion, and a protective cover
to the underlying sediments. The microbial mats in which this cyanobacterium is
dominant were termed “epibenthic mats” by Noffke (2010) and are typically found
in the supratidal zone.
The diatoms recorded in the Bahía Blanca Estuary are mainly small pennate
diatoms; the genera Nitzschia, Navicula, Diploneis Ehrenberg ex Cleve, and
Amphora Ehrenberg ex Kützing are quantitatively dominant. Other larger diatoms
of the genera Pleurosigma W.  Smith, Gyrosigma Hassall, and Cylindrotheca
Rabenhorst are also present, but they are less frequent. Among central diatoms, the
genera Thalassiosira Cleve and Coscinodiscus Ehrenberg and the species Cyclotella
meneghiniana Kützing and Paralia sulcata have been mentioned by different
authors (Pizani 2009; Pan et  al. 2013a, b). The dominance of small diatoms is
130 M. E. Croce et al.

attributed to the fact that small cells have higher growth and nutrient uptake rates
than bigger cells since they have higher surface/volume ratio, which allows small
cells to outcompete bigger cells when they are subjected to frequent physical distur-
bances (Williams 1964; Snoeijs et al. 2002).

Box 6.1: Ecosystem Engineers in the Bahía Blanca Estuary: The Crab
Neohelice granulata
Tidal currents are responsible for sediment transport, and waves produce
either sediment deposition or erosion. These factors interact with the biologi-
cal components of the coast determining whether deposition or erosion is the
dominant process in a specific site. These complex processes ultimately deter-
mine the type and abundance of organisms in the sediment (Blanchard et al.
2000; Dyer et al. 2000).
Bioturbation is defined as the biological reworking of sediments and soils
through animal activities like feeding and burrowing that generates changes in
chemical gradients and relocates resources and microorganisms. Such sedi-
ment restructuring also promotes physical alterations, by changing the bal-
ance of material transported, and affects the structure and functioning of
assemblages (Meysman et al. 2006; Kristensen et al. 2012). Organisms that
directly or indirectly modify the physical environment and regulate the avail-
ability of resources for other species are called ecosystem engineers (Statzner
et al. 2000; Gutiérrez et al. 2003; Berkenbusch and Rowden 2003; Cardinale
et al. 2004).
Neohelice granulata (Dana, 1851) (= Chasmagnathus granulata) is an
estuarine crab that excavates semipermanent burrows generating extensive
burrowing beds which cover up to 80% of the intertidal areas of SW Atlantic
estuaries and bays (Botto et al. 2006; Iribarne et al. 2005). N. granulata dis-
tributes from the northeastern coast of Patagonia, Argentina (42°25′S,
64°36′W), to Río de Janeiro, Brazil (22°57′S, 42°50′W) (Spivak 2010). In the
muddy salt marshes of the Bahía Blanca Estuary, an association between the
crab and the halophyte plant Sarcocornia perennis (Miller) A. J. Scott was
described by Perillo and Iribarne (2003). This association has a particular
configuration, where the plants form a ring surrounding a non-vegetated salt
pan densely excavated by the crab. These ring-shaped configurations are
1.5–8 m in diameter and have high water retention at the inner part. Such ring-­
shaped configuration of the halophyte vegetation is the macroscopic evidence
of the changes in salinity, humidity, and hardness that occur in the sediment
and which are a consequence of the plant-crab interaction (Escapa et al. 2007).
The composition and structure of the microphytobenthic assemblage differ
considerably among the different environments composing the rings. A thin
diatom biofilm, characterized by high abundance of diatoms and reduced
abundance of cyanobacteria, is observed in the inner part of the rings,
6  Biology and Ecology of the Benthic Algae 131

associated with small grain sediment, while well-developed microbial mats,


characterized by the presence of Coleofasciculus chthonoplastes, are present
in the outer region of the rings. Both structures show differences in the EPS
matrix, since the biofilms dominated by diatoms are less developed and more
irregular than those formed in the presence of cyanobacteria (Fig. 6.5a). In
cyanobacterial mats, EPS are found as an embedded continuous matrix,
whereas in diatom biofilms, they develop as a web of spongelike fibrils with
void spaces (Fernández et al. 2018) (Fig. 6.5b). Such differences in the micro-
phytobenthic assemblages are the result of changes in the physical properties
of sediments caused by the bioturbation of the burrowing crab.
A common consequence of the activity of burrowing deposit feeders is the
increase in softness and water content of the sediment. Also, the burrows con-
tribute to the accumulation of fine particles since the sediment with a high
percentage of clay accumulates at the burrow tunnel during high tide (Davis
1993; Botto and Iribarne 2000). This physical mixture would restrain the for-
mation of large and well-developed microbial mats on the superficial sedi-
ments of the salt marsh with abundant crabs. Then, the death of the plants in
the inner part of the rings can also be associated with permanent modifications
in the development of the microphytobenthos succession resulting from the
biodisturbing action of crabs (Fernández et al. 2018).
Given the important role of microphytobenthos in recycling of nutrients,
biofiltration, and sediment stabilization in coastal ecosystems, the study of the
distribution patterns of micro- and macroorganisms provides valuable infor-
mation for the formulation of integrated management plans, aiming to reduce
the ecosystem erosion and contamination.

6.5.2  Hard Substrate Available for Algal Settlement

In the Bahía Blanca Estuary, the hard substrate available for the settlement of algae
comprises natural and artificial structures (Fig. 6.2). The hard substrate of natural
origin consists of rocks made of consolidated fine sedimentary particles (henceforth
named outcrops) and the mollusk shells (oysters, mussels, snails, and barnacles).
The outcrops are made up of compacted sand and clay material giving rise to a rela-
tively hard substrate (Spalleti 1980; Aliotta and Lizasoain 2004). These outcrops are
common in the southeastern coasts of the estuary like Villa del Mar (Fig. 2.4;
Chap. 2). These substrates are called outcrops because they appear interrupting the
large tidal flats or sand beaches. In the Bahía Blanca Estuary, macroalgal assem-
blages inhabit the depressions (tidal pools) that form on these outcrops, which
remain filled with water during low tide (Fig. 6.2b).
The main artificial hard substrate consists of wooden docks, concrete barriers,
metal ladders, floating platforms and bridges, mooring ropes, sewer pipes, buoys,
132 M. E. Croce et al.

Fig. 6.2  Artificial and natural substrata of the Bahía Blanca Estuary colonized by benthic algae,
(a) general view of a bloom of green macroalgae on a salt marsh in Villa del Mar, (b) macroalgal
assemblage in a tidal pool located in the outcrops of Villa del Mar, (c) macroalgae attached to mol-
lusk shells, (d) macroalgae attached to a floating platform in CNBB, (e) green and red macroalgae
growing on concrete rocks in the upper intertidal zone of Puerto Rosales, (f) detail of Blidingia
sp. on a concrete rock in Puerto Rosales, (g) green macroalgae on a wheel in Villa del Mar, (h)
green macroalgae covering the surface of a sewer pipe in Club Náutico Bahía Blanca (CNBB).
(Photos by M. Emilia Croce)
6  Biology and Ecology of the Benthic Algae 133

and rubber wheels. All these surfaces appear in the different coasts of the Bahía
Blanca Estuary and are colonized by micro- and macroalgae. Although the hard
substrate is scarce in the Bahía Blanca Estuary, the industrial growth in the region
has modified the environment in such a way that the surfaces available for algal
attachment have increased in number by the artificial structures constructed
by humans.
Although microbial mats develop conspicuous biogenic structures on soft sedi-
ments, the biofilms composed of diatoms and filamentous cyanobacteria can also be
found covering hard natural or artificial substrates in the intertidal zone (Fig. 6.1c).
In contrast to microalgae, macroscopic algae are always found attached to hard
substrates (except for the drifting species). This is because the main condition for a
macroalga to colonize a substrate is that the substrate is relatively stable, for the
algae to remain attached and to avoid being flushed away by waves or currents.
Consequently, macroscopic algae are virtually able to occupy any type of substrate
as long as the surface is suitable to attach for the microscopic propagules (spores or
gametes), or a fragment of the thallus, in the case of vegetative propagation (Amsler
et al. 1992; Fletcher and Callow 1992). Because of the different tolerances of each
species of macroalgae to desiccation stress (Dromgoole 1980; Davison and Pearson
1996), the success to colonize a stable substrate greatly depends on the location of
the substrate with respect to the coastline (Lobban and Harrison 1994). The estab-
lishment and persistence of benthic algal populations depend on the reproductive
performance of the species but also on the successful survival of their propagules.
Any factor that influences the recruitment, the settlement, and/or the post-settlement
of algal propagules becomes an important factor determining the establishment,
dynamics, and structure of algal communities (Wahl and Hoppe 2002). For exam-
ple, substratum microtopography has been recognized as one of the major factors
structuring marine benthic communities (Emson and Faller-Fristch 1976; Woodin
1978; Menge et al. 1983; Brault and Bourget 1985; Bergeron and Bourget 1986). In
the case of artificial substrate, the type of material that constitutes the substrate also
influences the colonization and survival of the macroalgal species. For example,
substrates that retain water like ropes are favorable for species that are less tolerant
to desiccation (Nienhuis 1969).
In general, marine macroalgae have a wide capacity of dispersal through a vari-
ety of forms, from unicellular to multicellular propagules, either sexual or asexual
(Santelices 1990; Norton 1992). Due to their potential for colonization of new habi-
tats, any modification of the habitat can quickly lead to changes in the macroalgal
diversity of the coast. The factors associated with the expansion of the industrial
area, such as the increased availability of substrate, the introduction of exotic spe-
cies through ballast water, and the increase of nutrient loads into the water, are
promoting changes in the diversity and distribution of macroalgae in the Bahía
Blanca Estuary.
134 M. E. Croce et al.

6.5.3  Diversity and Composition of Macroalgal Assemblages

The richness of macroalgae in the Bahía Blanca Estuary is low compared with other
coasts of Argentina (Miloslavich et al. 2011). According to the literature, a total of
19 macroalgal taxa have been recorded on the coasts of this estuary (Perillo et al.
2001; Parodi 2004; Bremec et  al. 2004; Croce et  al. 2015; Hoshino et  al. 2020;
Koller 2021). Eighteen of those taxa were found in the most recent surveys from
2015 to 2019, together with three new taxa, Blidingia marginata (J.  Agardh)
P.J.L. Dangeard ex Bliding, Blidingia minima (Nägeli ex Kützing) Kylin, and a spe-
cies of Pyropia J. Agardh. The complete list of 22 taxa is shown in Table 6.2. Red
and green macroalgae are the dominant groups with nine taxa each. According to
the classification of functional groups proposed by Balata et al. (2011), six morpho-­
functional groups are recognized; the majority of the taxa belong to the categories
bladelike and filamentous uniseriate and pluriseriate with erect thallus.
Although the turbid waters in this region limit the growth of macroalgae because
of the low light penetration through the water column, some tolerant species such as
opportunistic species with rapid growth rates and small turf-like forms flourish in
these habitats. In terms of biomass, the most abundant red macroalgae are
Polysiphonia abscissa J.D. Hooker and Harvey, Polysiphonia morrowii Harvey, and
Ceramium diaphanum (Lightfoot) Roth. These three filamentous species grow in
the tidal pools of the outcrops in Villa del Mar, where they are present almost all
year round, although their bushy thalli are larger in winter (Fig. 6.3a, c). Species of
Polysiphonia Greville also live attached to floating platforms near the surface in the
recreational harbor of the Bahía Blanca Nautic Club (CNBB), where they are con-
tinuously submerged. The presence of Gracilaria verrucosa (Hudson) Papenfuss in
the tidal flats of Villa del Mar is reported in the literature (Parodi 2004); however,
this species has not been found in recent surveys. The macroscopic thalli of Pyropia
are abundant during the winter on the hard substrates of the upper intertidal of
Puerto Rosales. Pyropia grows in the mooring ropes that are frequently exposed to
air during low tide, on the wooden pillars that support the docks, and on concrete
rocks (Fig. 6.3d). In general, the distribution of benthic red macroalgae that inhabit
the Bahía Blanca Estuary is limited to the intertidal flats and salt marshes of Villa
del Mar; except for Pyropia, they have not been recorded in other coasts with hard
substrate available. The lack of mobility of the propagules may restrict the spread of
red macroalgae in this estuary. Two species of Gelidium J.V.  Lamouroux grow
exclusively in the tidal pools of the outcrops located in the tidal flats of Villa del
Mar. They form dense mats on the consolidated sediment or grow as epiphytes on
mollusk shells (Fig. 6.3b). Gelidium species are perennial in these coasts and repro-
duce sexually all year round. They also reproduce and spread vegetatively, due to
their ability for regeneration and reattachment from fragments by the formation of
rhizoids. This behavior has been observed during culture experiments (unpublished
data) and is reported for other species of Gelidium as well (Santelices and Varela
1994; Titlyanov and Titlyanova 2006; Otaiza et al. 2018).
6  Biology and Ecology of the Benthic Algae 135

Fig. 6.3  Common benthic red macroalgae of the Bahía Blanca Estuary, (a) Polysiphonia mor-
rowii, detail of the thallus and axes (scale bar: 500 μm), (b) Gelidium pusillum, general aspect of
the thallus (scale bar: 1 cm), (c) Ceramium diaphanum, general aspect of the thallus (scale bar:
1 cm), (d) Pyropia sp.: thalli in nature (scale bar: 1 cm). (Photos by (a, b, and d) M. Emilia Croce,
(c) Ailen Poza)

The green algae are the second dominant group, represented by nine taxa
(Table 6.2). Ulva Linnaeus is the most conspicuous genus, represented by four spe-
cies that reach high biomass in winter (Fig. 6.4c). They cover large surfaces of the
intertidal flats and tidal pools in Villa del Mar. They also colonize submerged objects
and cover the concrete walls and rocks located in the upper intertidal zone of the
harbors. Green algae of the family Ulvophyceae are the most common group of
macroalgae in estuarine habitats. This is because they withstand desiccation occa-
sioned by the strong irradiance and wind during low tide and tolerate a wide range
of salinities. The presence of the genus Enteromorpha Link was reported by Perillo
et al. (2001), Parodi (2004), and Bremec et al. (2004), although this genus has been
now merged into Ulva (Hayden et al. 2012). Isolated thalli of Cladophora surera
136 M. E. Croce et al.

Fig. 6.4  Common benthic green and brown macroalgae of the Bahía Blanca Estuary, (a) Blidingia
marginata, detail of the thallus (scale bar: 60 μm), (b) Bryopsis plumosa, general aspect of the
thallus (scale bar: 2  cm), (c) Ulva lactuca, general aspect of the thallus (scale bar: 1  cm), (d)
Punctaria latifolia (scale bar: 2 cm). (Photos by (a, c, and d) Ailen Poza, (b) M. Emilia Croce)

E.R.  Parodi and E.J.  Cáceres are usually found in tidal pools or on the concrete
structures, but are most frequently found as an epiphyte on other macroalgae.
C. surera and Ulva flexuosa Wulfen (= Enteromorpha flexuosa (Wulfen) J. Agardh)
are tolerant to salinity changes and more related to freshwater environments (Parodi
2004). The coenocytic algae Bryopsis plumosa (Hudson) C. Agardh forms dense
bushes during winter in the upper tidal pools of the outcrops of Villa del Mar
(Fig. 6.4b). Chaetomorpha linum grows as an epiphyte on other macroalgae, but it
also occurs on the shaded regions of the concrete walls of the harbor of the
CNBB. Green macroalgae such as Blidingia minima and B. marginata grow usually
on concrete rock in the upper intertidal zone where they can reach high densities
(Fig. 6.4a).
6  Biology and Ecology of the Benthic Algae 137

The brown algae are less frequent in this habitat; they are represented only by
four taxa, the Ectocarpaceae Ectocarpus siliculosus and Hincksia hincksiae (Harvey)
P.C. Silva, the Chordariaceae Punctaria latifolia Greville, and the Scytosiphonaceae
Planosiphon nakamurae M.  Hoshino, M.E.  Croce, Hanyuda, and Kogame
(Table 6.2). The Ectocarpaceae have been almost exclusively found as epiphytes on
other macroalgae. Macroscopic thalli of P. nakamurae can be found in winter, grow-
ing on the walls and concrete blocks in the harbor of CNBB that are completely
exposed during low tide. P. latifolia occurs in the salt marshes dominated by
Spartina Schreb., located in Villa del Mar. The stems of this vascular plant offer a
temporary substrate for the attachment during the winter where the macroscopic
thalli of P. latifolia (Fig. 6.4d) attach sparsely (Parodi 2004, as Punctaria latifolia
var. crouanii).
Among the macroalgae that live in the Bahía Blanca Estuary, three of them are
considered exotic (or alien): Polysiphonia morrowii; Neosiphonia harveyi (Bailey)
M.-S. Kim, H.-G. Choi, Guiry, and G.W. Saunders; and Planosiphon nakamurae.
These species have been registered in other coasts of the Patagonian region as well.
P. morrowii is present in Puerto Madryn (Raffo et al. 2014), Bahía Anegada (Croce
and Parodi 2014), and Las Grutas (personal observation). There are different
hypotheses of the introduction of P. morrowii. The vectors may have been ballast
water (Hewitt et al. 2007) or other exotic marine organisms such as the Pacific oys-
ter in Bahía Anegada or the macroalgae Undaria pinnatifida (Harvey) Suringar, that
have been introduced in the 1980s and the 1990s (Verlaque 2001; Kim et al. 2004;
Geoffroy et al. 2012). There is no evidence about the initial site of their introduc-
tion, but it may have dispersed along the South Atlantic coast by shipping among
different harbors. N. harveyi is also present in the southern coasts of Argentina
(Raffo et al. 2014), and it is catalogued as introduced, given that it is worldwide
known as an invasive species. Planosiphon nakamurae was recently found in the
coasts of the South Atlantic Ocean. Its identity was confirmed by molecular tools,
which also evidenced its relationship with a Japanese haplotype (Hoshino
et al. 2020).
The exotic macroalgae are characterized by fast vegetative growth and dispersal,
usually by fragmentation, by a rapid completion of the life cycle, and wide toler-
ances to environmental variables (Nyberg and Wallentinus 2005). For several rea-
sons, exotic macroalgae are common in estuarine environments. First, because
estuaries are areas of entrance to the new environment, usually through ballast
water that is released by the international vessels. Second, the strong environmental
variations, mainly salinity changes and desiccation, are more easily tolerated by
exotic species. And third, because estuaries are regions occupied by human settle-
ments, and consequently, they constitute eutrophic environments where the exotic
species with fast-growing capabilities can exploit outcompeting with the native
species.
138 M. E. Croce et al.

6.6  Epiphytic Algae

Epibiosis is defined as a nonsymbiotic facultative association between an epibiont


(an organism that lives attached to a living surface) and a basibiont (a substrate
organism or host) (Wahl 1989). This phenomenon is common in aquatic environ-
ments and among macroalgae (epiphytes), which may live attached to animals, vas-
cular plants, or other macroalgae.
In the Bahía Blanca Estuary, there are natural substrates such as the leaves and
stems of halophyte plants, and the thalli of the macroalgae themselves, where epi-
phytic macroalgae and microalgae can attach. For example, P. latifolia grows
attached to the stems of Spartina. Although the population is relatively ephemeral,
this epiphytic association is persistent throughout the years, suggesting that it is
well established. The most frequent taxa that live attached to other macroalgae are
the filamentous red and green macroalgae Ceramium Roth, Acrochaetium Nägeli,
Cladophora, and Ulva sp. (Koller 2021); however, the diatoms Melosira C. Agardh,
Achnanthes Bory, and Cocconeis Ehrenberg frequently appear attached to the sur-
faces of the macroalgae thallus, solitary or in chains. Blidingia species are some-
times covered by epiphytic diatoms; the most frequent is Melosira that forms dense
brown tufts of long chains (Fig. 6.1e).
Epiphytic associations among macroalgae and benthic fauna are also common in
the Bahía Blanca Estuary. Polysiphonia and Gelidium species are epiphytes on two
dominant bivalves, the native mussel Brachidontes rodriguezii (d’Orbigny, 1842)
and the exotic Pacific oyster Magallana gigas (Thunberg, 1793) (= Crassostrea
gigas Thunberg, 1793). Succession and ecological studies in other coasts of the
northern Patagonian region have shown that M. gigas offers a suitable substrate for
the settlement of native and exotic macroalgae (Borges 2006; Croce and Parodi
2012) in areas where no suitable hard substrate is  available. Gelidium lives also
attached to the gastropod Crepidula aculeata (Gmelin, 1791). This association is
very frequent suggesting that the symbiosis may be beneficial to both species. The
macroalgae may benefit from nutrient loadings produced by the gastropod beneath,
and the gastropod may enhance the dispersal of the macroalgae by carrying them
from one place to another. From the perspective of the mollusk, the benefit may be
related to the avoidance of predators, since the macroalgae cover the shell completely.

6.7  Mat-Forming Macroalgae

The term “turf” is widely used in marine ecology to identify a layer of short and
densely branched algae that is several millimeters to a few centimeters tall (Connell
et al. 2014). Several macroalgal species fall under this definition, although the more
common are small-sized species of red algae representatives of the orders
Ceramiales, Gelidiales, and Corallinales. A  more specific term, the word “mat”,
identifies a small group of algae defined as short and densely branched and formed
6  Biology and Ecology of the Benthic Algae 139

by prostrate and erect axes and which grow entangled into a thick mass (Hay 1981).
These groups of algae, either turf-forming or mat-forming species, are relevant to
the dynamics of benthic ecosystems (Airoldi et  al. 1995; Bulleri and Benedetti-­
Cecchi 2006; Gorman and Connell 2009). One reason for that is that these cushions
trap large amounts of sediment particles in relation with their small size, influencing
the transportation of energy and organic matter in intertidal marine environments
(Airoldi 2003). It is known from several studies that the macroalgae that form this
type of mats have an important influence on the nutrient dynamics. Besides their
role in the dynamics of the mentioned abiotic factors, they are also key components
of biotic assemblages as they provide an excellent niche for little crustaceans, poly-
chaetes, annelids, mollusks, and other small invertebrates (Prathep et  al. 2003).
Through many years of surveys in the intertidal regions of the Bahía Blanca Estuary,
it has been noticed that the populations of the mat-forming Gelidiales may be key
components of the benthic communities in this coast. The vegetative growth of
these algae has been studied in culture conditions, and preliminary results showed
that these algae have a high capacity of regeneration by producing numerous
branches, rhizoids, and rhizoidal filaments, and they can also withstand a high load
of epiphytes for long periods of time (unpublished data).

Box 6.2: Potential Use of Native Estuarine Macroalgae for Biomitigation


Marine macroalgae are used in a variety of domestic and industrial processes.
The more ancestral use is as fresh food (Lee et al. 2017), but due to their vary-
ing intrinsic characteristics and chemical composition, they can also be used
for the production of a variety of algal products, for example, fertilizers
(Selvam and Sivakumar 2014), biogas by anaerobic digestion (Hinks et  al.
2013), phycocolloids (Chan and Matanjun 2017), and polymers that can be
incorporated into conventional plastic formulations to develop biodegradable
plastics (Freile-Pelegrín et al. 2007). Macroalgae are also a source of bioac-
tive compounds (polysaccharides, proteins, lipids, polyphenols, carotenoids,
and vitamins). These phytochemicals have different functional groups includ-
ing carboxyl, hydroxyl, phosphate, and amine that can bind pollutants (Areco
and dos Santos 2010, Sanjeewa et al. 2016). The presence of sulfated polysac-
charides in the cell wall of macroalgae, mainly in their fibrous matrix and
intercellular spaces, is the main reason for their high capacity to bind con-
taminants. In fact, hydroxyl, sulfate, and carboxyl groups of the polysaccha-
ride chains are strong ion exchangers; therefore, they are the important sites
of complexation of metal cations (Vasconcelos and Leal 2001). Biosorption is
one of the most promising remediation technologies for aquatic areas that are
polluted with heavy metal ions (Gupta et al. 2015). The major advantages of
biosorption using the macroalgal biomass for wastewater treatment are the
low cost and investment needed, the simple design and easy operation, and the
use of nontoxic substances. Hence, recently, the interest in using seaweed as
140 M. E. Croce et al.

biomitigators or for bioremediation of marine environments is increasing


(Kim et al. 2017). For this reason, macroalgae can be exploited as a resource
at the same time as it is used for ecological services.
Along the eastern coast of the Bahía Blanca Estuary, there are several
human settlements. From the inner part of the estuary to the mouth, we find
the towns of Villarino Viejo, General Daniel Cerri, Bahía Blanca, and Punta
Alta. Of all these emplacements, the city of Bahía Blanca has the largest
demographic growth due to the settlement of a fertilizer industry, a large pet-
rochemical pole, and thermoelectric plants, as well as the expansion of the
harbor. As a consequence of this economic development, the estuary is the
receptor of waste discharges from industrial origin (oil derivatives, pesticides,
heavy metals, etc.) as well as untreated domestic sewage, which have gener-
ated problems of contamination (Marcovecchio et al. 2010).
The macroalgal assemblages that inhabit the Bahía Blanca Estuary are
potentially useful for implementing methodologies of pollutants remotion and
eutrophication management. The most promising candidate for biomitigation
is the green alga Ulva, which is particularly useful in the biosorption of heavy
metals and other compounds due to its high surface area, relatively simple
structure, and uniform distribution of binding sites (Sari and Tuzen 2008;
Turner et al. 2007).
The multiple functions and uses of seaweeds discussed above would pro-
mote the cultivation of seaweeds to obtain high-quality raw materials for dif-
ferent applications. Biosorption by seaweeds is a promising method that
utilizes efficiently the naturally existing raw material. It is noticeable that very
few studies have used real wastewater for the treatment and most of the exper-
iments have used simulated wastewater. Therefore, it is recommended that
future studies consider the use of real wastewater, especially in impacted envi-
ronments such as the Bahía Blanca Estuary.

Fig. 6.5  Structure of microphytobenthic assemblages, (a) scanning electron photomicrograph of


biofilms dominated by pennate diatoms embedded in a compacted matrix of EPS (scale bar:
20 μm), (b) microbial mat dominated by filamentous cyanobacteria (arrows) (scale bar: 6 μm).
(Photos by Constanza Da-Rodda)
6  Biology and Ecology of the Benthic Algae 141

Glossary

Amoeboid Type of cell organization that lacks a cell wall and the pro-
toplasm undergoes frequent changes in shape.
Blade (= lamina) Flattened structure that is somewhat leaflike.
Bloom Massive or conspicuous growth of algae, usually a large
percentage of the total cells belong to one or a few species.
Coccoid Simple cell type that is spherical, subspherical, or
rod-shaped.
Cytokinesis The process by which one cell physically divides into
two cells.
Epilithic Organisms living attached to the surfaces of rocks
and stones.
Epipelic Organisms living attached to the surfaces of mud or sand.
Epiphytic Organisms living on the surfaces of plants or algae, using it
for support but not for nutrition.
Epizoic Organisms living on the surfaces of animals.
Eukaryotic Organisms composed of cells with membrane-bounded
nucleus. Most contain cells with a complex organization,
with microtubules, membrane-­ bounded organelles, and
chromatin organized into more than a single chromosome.
Eutrophic Waters that contain relatively high levels of dissolved nutri-
ents (e.g., nitrate, phosphate); typically exhibit high levels
of primary productivity.
Flagellum Long, threadlike organelle that projects out of the cell and
functions in motility. In eukaryotic cells, they consist of a
9 doublet + 2 central singlet array of microtubules.
Frustule In diatoms, the silica cell wall or test, composed of
two valves.
Holdfast A cell or multicellular structure that functions in attachment
to a substrate.
Kelp Large phaeophytes that are members of the Laminariales.
Mastigonemate Stiff, lateral hairs borne by a flagellum, consisting of a base,
a tubular shaft, and several terminal hairs.
Monophyletic Evolutionary term referring to a trait or group of organisms
that evolved directly from a common ancestor.
Phagotrophic Mode of nutrition referring to heterotrophic protoctists or
tissue cells that ingest solid food particles by phagocytosis.
Phycobiliproteins Complex of phycobilins with protein found in cyanobacte-
ria, rhodophytes, glaucocystophytes, and some
cryptophytes.
Phycobilisome Cellular structure containing phycobilin pigments and
arranged as protrusions on the surface of the thylakoids of
cyanobacteria, rhodophytes, and glaucocystophytes, but
within the thylakoids (between membranous stacks) in the
plastids of cryptophytes.
142 M. E. Croce et al.

Photosynthetic pigments Pigments are chemical compounds present in algae by


which the energy of sunlight is captured for
photosynthesis.
Plastids Generic term for photosynthetic organelles present in
algae. Plastids contain the enzymes and pigments for
photosynthesis, ribosomes, and other structures.
Polyphyletic Evolutionary term referring to a trait or group of organ-
isms that  derived from more than one common evolu-
tionary ancestor or ancestral group.
Prokaryotic Typically unicellular organism lacking a distinct nucleus
and membrane-bound organelles.
Propagules Generative structure, any unicellular or multicellular,
capable of survival, dissemination, and further growth.
Somatic General term referring to the body (soma) of an organ-
ism. The part not involved in reproduction or
germination.
Spore Reproductive cell (motile or not) that originates by mito-
sis (mitospore) or meiosis (meiospore).
Stipe Stalk, stemlike portion of the algal thallus.
Symbiotic Physical association between organisms of different
species.
Thallus The body of an alga that is not differentiated into roots,
stalks, and leaves as in other plants.
Thylakoid Flat saclike structure formed by membranes that is pres-
ent in the cytoplasm of cyanobacterial cells and in plas-
tids of eukaryotic algae and plants.

References

Admiraal W, Bouwman LA, Hoekstra L et  al (1983) Qualitative and quantitative interactions
between microphytobenthos and herbivorous meiofauna on a brackish intertidal mudflat. Int
Rev Ges Hydrobiol 68:175–191
Agusti S, Enriques S, Frost-Christensen S et  al (1994) Light harvesting among photosynthetic
organisms. Funct Ecol 8:273–279
Airoldi L (2003) The effects of sedimentation on rocky coast assemblages. Oceanogr Mar Biol
41:161–236
Airoldi L, Rindi F, Cinelli F (1995) Structure, seasonal dynamics and reproductive phenology
of a filamentous turf assemblage on a sediment influenced, rocky subtidal shore. Bot Mar
38:227–237
Aliotta S, Lizasoain GO (2004) Tipos de fondos y su caracterización geológica por métodos sis-
moacústicos. In: Piccolo MC, Hoffmeyer M (eds) Ecosistema del estuario de Bahía Blanca.
Instituto Argentino de Oceanografía, Bahía Blanca, pp 51–59
Amsler CD, Reed DC, Neushul M (1992) The microclimate inhabited by macroalgal propagules.
J Br Phycol 27:253–270
6  Biology and Ecology of the Benthic Algae 143

Areco MM, dos Santos A (2010) Copper, zinc, cadmium and lead biosorption by Gymnogongrus
torulosus. Thermodynamics and kinetics studies. Colloid Surface B 81:620–628
Atkinson MJ, Smith SV (1983) C:N:P ratios of benthic marine plants. Limnol Oceanogr
28:568–574
Bach SD, Josselyn MN (1979) Production and biomass of Cladophora prolifera (Chlorophyta,
Cladophorales) in Bermuda. Bot Mar 22:163–168
Baird D, Ulanowicz RE (1993) Comparative study on the trophic structure, cycling and ecosystem
properties of four tidal estuaries. Mar Ecol Prog Ser 99:221–237
Balata D, Piazzi L, Rindi F (2011) Testing a new classification of morphological functional groups
of marine macroalgae for the detection of responses to stress. Mar Biol 158:2459–2469
Balech E, Ehrlich MD (2008) Esquema biogeográfico del Mar Argentino. Rev Invest Des Pesq
19:45–75
Barsanti L, Gualtieri P (2006) Algae: anatomy, biochemistry, and biotechnology. Taylor & Francis,
Boca Raton
Bennett A, Bianchi TS, Means JC et  al (1999) The effects of polycyclic aromatic hydrocarbon
contamination and grazing on the abundance and composition of microphytobenthos in salt
marsh sediments. J Exp Mar Biol Ecol 242:1–20
Berezina N, Tsiplenkina IG, Pankova ES et  al (2007) Dynamics of invertebrate communities
in stony littoral of the Neva Estuary (Baltic Sea) under macroalgal blooms. Transit Waters
Bull 1:49–60
Bergeron P, Bourget E (1986) Shore topography and spatial partitioning of crevice refuges by ses-
sile epibenthos in an ice disturbed environment. Mar Ecol Prog Ser 28:129–145
Berkenbusch K, Rowden AA (2003) Ecosystem engineering-moving away from “just-so” stories.
N Z J Ecol 27:67–73
Blanchard GF, Paterson DM, Stal LJ et  al (2000) The effect of geomorphological structures
on potential biostabilisation by microphytobenthos on intertidal mudflats. Cont Shelf Res
20:1243–1256
Blanchard GF, Guarini J-M, Orvain F et al (2001) Dynamic behaviour of benthic microalgal bio-
mass in intertidal mudflats. J Exp Mar Biol Ecol 264:85–60
Boedeker C, Hansen GI (2010) Nuclear rDNA sequences of Wittrockiella amphibia (Collins)
comb. nov. (Cladophorales, Chlorophyta) and morphological characterization of the mat-like
growth form. Botanica Marina 53(4):351–356
Bold HC, Wynne MJ (1978) Introduction to the algae: structure and reproduction. Prentice-Hall,
Inc., Englewood Cliffs
Borges ME (2006) Ecología de las ostras en ambientes del sur bonaerense: cultivo y manejo de sus
poblaciones. PhD thesis, Universidad Nacional del Sur, Bahía Blanca, p 247
Botto F, Iribarne O (2000) Contrasting effects of two burrowing crabs (Chasmagnathus granu-
lata and Uca uruguayensis) on sediment composition and transport in estuarine environments.
Estuar Coast Shelf Sci 51:141–151
Botto F, Iribarne O, Gutiérrez J et al (2006) Ecological importance of passive deposition of organic
matter into burrows of the SW Atlantic crab Chasmagnathus granulatus. Mar Ecol Prog Ser
312:201–226
Brault S, Bourget E (1985) Structural changes in an estuarine subtidal epibenthic community:
biotic and physical causes. Mar Ecol Prog Ser 21:63–73
Breitburg DL, Steinberg N, DuBeau S et al (1994) Effects of low dissolved oxygen on predation
on estuarine fish larvae. Mar Ecol Prog Ser 64:235–246
Bremec CS, Martinez DE, Elias R (2004) Asociaciones bentónicas de fondos duros y comunidades
incrustantes. In: Piccolo MC, Hoffmeyer M (eds) Ecosistema del estuario de Bahía Blanca.
Instituto Argentino de Oceanografía, Bahia Blanca, pp 171–178
Buffan-Dubau E, Carman KR (2000) Diel feeding behaviour of meiofauna and their relationships
with microalgal resources. Limnol Oceanogr 2:381–395
Bulleri F, Benedetti-Cecchi L (2006) Mechanisms of recovery and resilience of different compo-
nents of mosaics of habitats on shallow rocky reefs. Oecologia 149:482–492
144 M. E. Croce et al.

Caputo E, Ceglie V, Lippolis M et al (2010) Identification of a NaCl-induced ascorbate oxidase


activity in Chaetomorpha linum suggests a novel mechanism of adaptation to increased salin-
ity. Environ Exp Bot 69:63–67
Cardinale BJ, Gelmann ER, Palmer MA (2004) Net spinning caddisflies as stream ecosystem
engineers: the influence of Hydropsyche on benthic substrate stability. Funct Ecol 18:381–387
Cartaxana P, Ruivo M, Hubas C et  al (2011) Physiological versus behavioral photoprotection
in intertidal epipelic and epipsammic benthic diatom communities. J Exp Mar Biol Ecol
405:120–127
Cebrian J (1999) Patterns in the fate of production in plant communities. Am Nat 154:449–468
Cebrian J (2002) Variability and control of carbon consumption, export and accumulation in
marine communities. Limnol Oceanogr 47:11–22
Cebrian J, Williams M, McClelland J et al (1998) The dependence of heterotrophic consumption
and C accumulation on autotrophic nutrient content in ecosystems. Ecol Lett 1:165–170
Chan PT, Matanjun P (2017) Chemical composition and physicochemical properties of tropical red
seaweed, Gracilaria changii. Food Chem 221:302–336
Charpy L, Casareto BE, Langlade MJ et  al (2012) Cyanobacteria in coral reef ecosystems: a
review. J Mar Biol 2012:1–9
Choi TS, Kang EJ, Kim JH, Kim KY (2010) Effect of salinity on growth and nutrient uptake of
Ulva pertusa (Chlorophyta) from an eelgrass bed. Algae 25:17–26
Cicerone D (1987) Estimación de la biomasa de diatomeas bentónicas en áreas mesolitorales de la
porción interna del Estuario de Bahía Blanca, a partir del análisis de pigmentos fotosintéticos.
Lic thesis. Universidad Nacional del Sur, Bahía Blanca
Coles RG, Lee Long WJ, Watson RA (1993) Distribution of seagrasses, and their fish and penaeid
prawn communities, in Cairns Harbour, a tropical estuary, Northern Queensland, Australia.
Aust J Mar Fresh Res 44:193–126
Connell SD, Foster MS, Airoldi L (2014) What are algal turfs? Towards a better description of
turfs. Mar Ecol Prog Ser 495:299–307
Costa JL, Lindblad P (2002) Cyanobacteria in symbiosis with cycads. In: Rai AN, Bergman B,
Rasmussen U (eds) Cyanobacteria in symbiosis. Springer, Dordrecht
Costanza R, Kemp WM, Boynton WR (1993) Predictability, scale, and biodiversity in coastal and
estuarine ecosystems-implications for management. Ambio 22:88–96
Coull BC (1999) Role of meiofauna in estuarine soft-bottom habitats. Aust J Ecol 24:327–343
Croce ME, Parodi ER (2012) Seasonal dynamic of macroalgae in intertidal pools formed by beds
of Crassostrea gigas (Mollusca, Bivalvia) on the north Patagonian Atlantic coast. Bot Mar
55:49–58
Croce ME, Parodi ER (2014) The Japanese alga Polysiphonia morrowii (Rhodomelaceae,
Rhodophyta) on the South Atlantic Ocean: first report of an invasive macroalga inhabiting
oyster reefs. Helgol Mar Res 68(2):241–252
Croce ME, Gauna MC, Fernández C et  al (2015) Intertidal seaweeds from North Atlantic
Patagonian coasts, Argentina. Check List 11(5):1739
Cuadrado DG, Pizani NV (2007) Identification of microbially induced sedimentary structures over
a tidal flat. Latin Am J Sedimentol Basin Anal 14:65–116
Cuadrado DG, Carmona NB, Bournod CN (2011) Biostabilization of sediments by microbial
mats in a temperate siliciclastic tidal flat, Bahía Blanca estuary (Argentina). Sediment Geol
237:95–61
Cuadrado DG, Carmona NB, Bournod CN (2012) Mineral precipitation on modern siliciclastic
tidal flats colonized by microbial mats. Sediment Geol 271–272:58–66
Cuadrado DG, Bournod CN, Pan J et al (2013) Microbially-induced sedimentary structures (MISS)
as record of storm action in supratidal modern estuarine setting. Sediment Geol 296:1–8
Da Rodda C (2004) Biodiversidad y distribución espacial de las comunidades microalgales bentóni-
cas de la marisma “El Cangrejal” en el Estuario de Bahía Blanca. Lic Thesis, Universidad
Nacional del Sur, Bahía Blanca, p 51
6  Biology and Ecology of the Benthic Algae 145

Da Rodda C, Parodi ER (2005) Cyanophyceae Epipélicas de la Marisma El Cangrejal en el estu-


ario de Bahía Blanca (Buenos Aires, Argentina). Bol Soc Argent Bot 40:157–168
Daehnick AE, Sullivan MJ, Moncreiff CA (1992) Primary production of the sand microflora in
seagrass beds of Mississippi sound. Bot Mar 35:131–139
Dauer DM, Conner WG (1980) Effects of moderate sewage input on benthic polychaete popula-
tions. Estuar Coast Mar Sci 6:335–346
Davis WR (1993) The role of bioturbation in sediment resuspension and its interaction with physi-
cal shearing. J Exp Mar Biol Ecol 171:187–200
Davison IR, Pearson GA (1996) Stress tolerance in intertidal seaweeds. J Phycol 32:197–211
Day JW, Yáñez-Arancibia A, Kemp WM et al (2013) Introduction to estuarine ecology. In: Day
JW, Crump BC, Kemp WM et al (eds) Estuarine ecology, 2nd edn. Wiley, Hoboken, pp 1–18
De Winder B, Staats N, Stal LJ et al (1999) Carbohydrate secretion by phototrophic communities
in tidal sediments. J Sea Res 42:131–146
Decho AW (2000) Exopolymer microdomains as a structuring agent for heterogeneity within
microbial biofilms. In: Riding RE, Awramik SM (eds) Microbial sediments. Springer,
Berlin, pp 1–9
Denis L, Gevaert F, Spilmont N (2012) In situ intertidal variability of microphytobenthic produc-
tion in a macrotidal temperate estuarine system. J Soils Sediments:1517–1529
Dethier M (1981) Heteromorphic algal life histories: the seasonal pattern and response to her-
bivory of the brown crust Ralfsia californica. Oecologia 49:333–339
Dromgoole FI (1980) Desiccation resistance of intertidal and subtidal algae. Bot Mar 23:149–159
Druehl LD (1967) Vertical distribution of some benthic marine algae in a British Columbia inlet,
as related to some environmental factors. J Fish Res Bd Can 24:33–46
Duarte CM (1995) Submerged aquatic vegetation in relation to different nutrient regimes. Ophelia
41:87–11
Duarte CM, Sand-Jensen K, Nielsen SL et al (1995) Comparative functional ecology. Trends Ecol
Evol 6:418–421
Dyer KR, Christie MC, Wright EW (2000) The classification of intertidal mudflats. Cont Shelf
Res 20:639–660
Eckman JE, Andres MS, Marinelli RL et al (2008) Wave and sediment dynamics along a shallow
subtidal sandy beach inhabited by modern stromatolites. Geobiology 6:21–32
Emson RH, Faller-Fristch RJ (1976) An experimental investigation into the effect crevice availabil-
ity on abundance and size-structure in a population of Littorina rudis (Manton) (Gastropoda;
Prosobranchia). J Exp Mar Biol Ecol 23:285–297
Escapa M, Minkoff DR, Perillo GM et  al (2007) Direct and indirect effects of burrowing crab
Chasmagnathus granulatus activities on erosion of Southwest Atlantic Sarcocornia dominated
marshes. Limnol Oceanogr 52:2340–2349
Farías L (1988) Estudio del microfitobentos en áreas mesolitorales de la porción interna del
Estuario de Bahía Blanca (Puerto Cuatreros). Seminario Lic. Oceanografía. Universidad
Nacional del Sur, Bahia Blanca
Fernández C, Da-Rodda C, Gauna MC et al (2018) The role of plant-crab interaction in structuring
microphytobenthic communities in a shallow temperate estuary. Braz J Oceanogr 66:307–314
Fletcher RL, Callow ME (1992) The settlement, attachment and establishment of marine algal
spores. J Br Phycol 27:303–329
Ford RB, Thrush SF, Probert PK (1999) Macrobenthic colonisation of disturbances on an intertidal
sandflat, the influence of season and buried algae. Mar Ecol Prog Ser 191:163–174
Franz DR, Friedman I (2002) Effects of macroalgal mat (Ulva lactuca) on estuarine sand flat cope-
pods: an experimental study. J Exp Mar Biol Ecol 271:209–226
Freile-Pelegrín Y, Madera-Santana T, Robledo D et al (2007) Degradation of agar films in a humid
tropical climate: thermal, mechanical, morphological and structural changes. Polym Degrad
Stab 92:244–252
146 M. E. Croce et al.

Fučikova K, Lewis PO, Lewis LA (2014) Widespread desert affiliation of trebouxiophycean algae
(Trebouxiophyceae, Chlorophyta) including discovery of three new desert genera. Phycol Res
62:294–305
Geitler L (1944) Furchungsteilung, simultane Mehrfachteilung, Lokomotion, Plasmoptse und
Ókologie der Bangiacee Porphyridium cruentum. Flora 37:300–333
Geoffroy A, Le Gall L, Destombe C (2012) Cryptic introduction of the red alga Polysiphonia
morrowii Harvey (Rhodomelaceae, Rhodophyta) in the North Atlantic Ocean highlighted by a
DNA barcoding approach. Aquat Bot 60:67–71
Glud RN, Kühl M, Wenzhöfer F et al (2002) Benthic diatoms of a high Arctic fjord (Young Sound,
NE Greenland): importance for ecosystem primary production. Mar Ecol Prog Ser 238:15–29
Gorman D, Connell SD (2009) Recovering subtidal forests on human-dominated landscapes. J
Appl Ecol 46:1258–1265
Graham LE, Wilcox LW (2000) Algae. Prentice Hall, Upper Saddle River
Graham LE, Graham JM, Wilcox LW et al (2016) Algae, 3rd edn. LJLM Press, LLC, Madison
Gupta VK, Nayak A, Agarwal S (2015) Bioadsorbents for remediation of heavy metals: current
status and their future prospects. Environ Eng Res 20:1–18
Gutiérrez JL, Jones CG, Strayer DL et al (2003) Mollusks as ecosystem engineers: the role of shell
production in aquatic habitats. Oikos 61:79–90
Harlin MM, Lindbergh JM (1977) Selection of substrata by seaweed: optimal surface relief. Mar
Biol 40:33–40
Harris MA, Cumming BF, Smol JP (2006) Assessment of recent environmental changes in New
Brunswick (Canada) lakes based on paleolimnological shifts in diatoms species assemblages.
Can J Botany 84:151–163
Hay ME (1981) The functional morphology of turf-forming seaweeds: persistence in stressful
marine habitats. Ecology 62:739–750
Hayden H, Blomster J, Maggs CA et al (2012) Linnaeus was right all along: Ulva and Enteromorpha
are not distinct genera. Eur J Phycol 38:277–294
Heck KL Jr, Hays G, Orth RJ (2003) Critical evaluation of the nursery role hypothesis for seagrass
meadows. Mar Ecol Prog Ser 253:123136
Hewitt CL, Campbell ML, Schaffelke B (2007) Introductions of seaweeds: accidental transfer
pathways and mechanisms. Botanica Mar 50:326–337
Hibberd DJ, Norris RE (1984) Cytology and ultrastructure of Chlorarachnion reptans
(Chlorarachniophyta) division nova, Chlorarachniophyceae, (Clasis nova). J Phycol 20:36–330
Hill SA, Towill LR, Sommerfeld MR (1980) Photomovement responses of Porphyridium purpu-
reum. J Phycol 16:444–448
Hillebrand HB, Worm B, Lotze HK (2000) Marine microbenthic community structure regulated by
nitrogen loading and grazing pressure. Mar Ecol Prog Ser 204:27–38
Hinks J, Edwards S, Sallis PJ et al (2013) The steady state anaerobic digestion of Laminaria hyper-
borea. Effect of hydraulic residence on biogas production and bacterial community composi-
tion. Bioresour Technol 143:221–230
Hoagland KD, Rosowski JR, Gretz MR et al (1993) Diatom extracellular polymeric substances:
function, fine structure, chemistry, and physiology. J Phycol 29:537–566
Hoshino M, Croce ME, Hanyuda T, Kogame K (2020) Species delimitation of Planosiphon graci-
lis morphospecies (Scytosiphonaceae, Phaeophyceae) from Japan and the description of Pl.
nakamurae sp. nov. Phycologia 59:116–126
Hull SC (1987) Macroalgal mats and species abundance, a field experiment. Estuar Coast Shelf
Sci 25:519–532
Iribarne O, Bruschetti M, Escapa M et al (2005) Small and large-scale effect of the SW Atlantic
burrowing crab on habitat use by migratory shorebirds. J Exp Mar Biol Ecol 315:87–61
Irlandi EA, Peterson CH (1991) Modification of animal habitat by large plants mechanisms by
which seagrasses influence clam growth. Oecologia 87:307–318
Isaksson I, Pihl L (1992) Structural changes in benthic macrovegetation and associated epibenthic
faunal communities. Neth J Sea Res 30:131–140
6  Biology and Ecology of the Benthic Algae 147

Isaksson I, Pihl L, van Montfrans J (1994) Eutrophication-related changes in macro vegetation


and foraging of young cod (Gadus morhua L.): a mesocosm experiment. J Exp Mar Biol Ecol
177:203–217
Johansson G, Snoeijs P (2002) Macroalgal photosynthetic responses to light in relation to thallus
morphology and depth zonation. Mar Ecol Prog Ser 244:63–72
Kang CK, Kim JB, Lee KS et al (2003) Trophic importance of benthic microalgae to macrozoob-
enthos in coastal bay systems in Korea: dual stable C and N isotope analyses. Mar Ecol Prog
Ser 259:79–92
Kelaher BP, Levinton JS (2003) Variation in detrital enrichment causes spatio-temporal variation
in soft-sediment assemblages. Mar Ecol Prog Ser 261:8597
Kennish M (2017) Ecology of estuaries: volume 2: biological aspects, 1st edn. CRC Press,
Boca Raton
Kim MS, Yang CE, Mansilla A et  al (2004) Recent introduction of Polysiphonia morrowii
(Ceramiales, Rhodophyta) to Punta Arenas, Chile. Bot Mar 47:389–394
Kim JH, Kang EJ, Park MG et al (2011) Effects of temperature and irradiance on photosynthesis
and growth of a green-tide-forming species (Ulva linza) in the Yellow Sea. J Appl Phycol
23:421–432
Kim JK, Yarish C, Hwang EK et al (2017) Seaweed aquaculture: cultivation technologies, chal-
lenges and its ecosystem services. Algae 32:1–13
Kirk JTO (1994) Light composition and photosynthesis in aquatic communities. Cambridge
University Press, Cambridge
Koller L (2021) Cultivo de una especie de Gelidium (Gelidiales, Rhodophyta) del Estuario de
Bahía Blanca, Thesis, Universidad Nacional del Sur, Bahía Blanca
Konar B, Iken K (2009) Influence of taxonomic resolution and morphological functional groups in
multivariate analyses of macroalgal assemblages. Phycologia 48:24–31
Krause-Jensen D, McGlathery K, Rysgaard S et al (1996) Production within dense mats of the
filamentous macroalga Chaetomorpha linum in relation to light and nutrient availability. Mar
Ecol Prog Ser 134:207–216
Krause-Jensen D, Christensen PB, Rysgaard S (1999) Oxygen and nutrient dynamics within mats
of the filamentous macroalga Chaetomorpha linum. Estuaries 22:31–38
Kristensen E, Penha-Lopes G, Delefosse M et al (2012) What is bioturbation? The need for a pre-
cise definition for fauna in aquatic sciences. Mar Ecol Prog Ser 446:285–302
Krumbein WE (1994) The year of the slime. In: Krumbein WE, Paterson D, Stal LJ (eds)
Biostabilization of sediments. Bibliotheks und Informationssystem der Carl von Ossietzky
Universitat, Oldenburg, pp 1–7
Larsen A, Sand-Jensen K (2006) Salt tolerance and distribution of estuarine benthic macroalgae in
the Kattegat–Baltic Sea area. Phycologia 45:13–23
Lee WK, Lim YY, Leow ATC et al (2017) Factors affecting yield and gelling properties of agar. J
Appl Phycol 29:1527–1540
Lemieux C, Otis C, Turmel M (2014) Six newly sequenced chloroplast genomes from prasinophyte
green algae provide insights into the relationship among prasinophyte lineages and the diver-
sity of streamlined genome architecture in picoplanktonic species. BMC Genomics 15:857
Lenanton RCJ, Robertson AI, Hansen JA (1982) Nearshore accumulations of detached macro-
phytes as nursery areas for fish. Mar Ecol Prog Ser 9:51–57
Lin HP, Sommerfeld MR, Swafford JR (1975) Light and electron microscope observations on
motile cells of Porphyridium purpureum (Rhodophyta). J Phycol 11:452–457
Littler MM, Arnold KE (1982) Primary productivity of marine macroalgal functional-form groups
from southwestern North America. J Phycol 18:307–311
Littler MM, Littler DS (1980) The evolution of thallus form and survival strategies in benthic
marine macroalgae: field and laboratory tests of a functional forma model. Am Nat 116:25–44
Littler MM, Littler DS (1984) Relationships between macroalgal funtional form groups and sub-
strata stability in a subtropical rocky-intertidal system. J Exp Mar Biol Ecol 74(1):13–34
148 M. E. Croce et al.

Liuzzi MG, Lopez Gappa J, Piriz ML (2011) Latitudinal gradients in macroalgal biodiversity in
the Southwest Atlantic between 36 and 55°S. Hidrobiologia 673(1):205–214
Lobban C, Harrison PJ (1994) Seaweed ecology and physiology. Cambridge University Press,
New York
Luther G (1976) Bewuchsuntersuchungen auf Natursteinsubstraten im Gezeitenbereich des
Nordsylter Wattenmeeres: Algen. Helgolander wiss Meeresunters 28:318–351
Macintyre HL, Geider RJ, Miller DC (1996) Microphytobenthos: the ecological role of the “secret
garden” of unvegetated, shallow-water marine habitats. I. Distribution, abundance and primary
production. Estuaries 19:186–201
Marcovecchio JE, Botté SE, Fernandez Severini MD et al (2010) Geochemical control of heavy
metal concentrations and distribution within Bahía Blanca Estuary (Argentina). Aquat
Geochem 16(2–3):251–266. https://doi.org/10.1007/s10498-­009-­9076-­1
Margulis L, Barghoorn ES, Ashendorf D et  al (1980) The microbial community in the layered
sediments at Laguna Figueroa, Baja California, Mexico: does it have Precambrian analogues?
Precambrian Res 11:93–123
Margulis L, Corliss JO, Melkonian M et  al (eds) (1990) Handbook of Protoctista. Jones and
Bartlett Publishers, Boston
Martins I, Oliveira JM, Flindt MR et  al (1999) The effect of salinity on the growth rate of the
macroalgae Enteromorpha intestinalis (Chlorophyta) in the Mondego estuary (West Portugal).
Acta Oecol 20:259–265
Marx JM, Herrnkind WF (1985) Macroalgae (Rhodophyta: Laurencia spp.) as habitat for young
juvenile spiny lobsters, Panulirus argus. Bull Mar Sri 36:423–431
Menendez M (2005) Effect of nutrient pulses on photosynthesis of Chaetomorpha linum from a
shallow Mediterranean coastal lagoon. Aquat Bot 82:181–192
Menge BA, Ashkenas LR, Matson A (1983) Use of artificial holes in studying community devel-
opment in cryptic marine habitats in a tropical rocky intertidal region. Mar Biol 77:128–141
Meysman FJR, Middelburg JJ, Heip CHR (2006) Bioturbation: a fresh look at Darwin’s last idea.
Trends Ecol Evol 21:688–695
Miller DC, Geider RJ, Macintyre HL (1996) Microphytobenthos: the ecological role of the “secret
garden” of unvegetated, shallow-water marine habitats. II Role in sediment stability and
shallow-­water food webs. Estuaries 19:202–212
Miloslavich P, Klein E, Diaz JM et al (2011) Marine biodiversity in the Atlantic and Pacific coasts
of South America: knowledge and gaps. PLoS One 6(1):e14631
Montagna PA (1984) In situ measurement of meiobenthic grazing rates on sediment bacteria and
edaphic diatoms. Mar Ecol Prog Ser 18:119–130
Morand P, Merceron M (2005) Coastal eutrophication and excessive growth of macroalgae. In:
Pandalai SG (ed) Recent research developments in environmental biology. Research Signpost,
Trivandrum, pp 395–449
Nedergaard RI, Risgaard-Peterson N, Finster K (2002) The importance of sulfate reduction associ-
ated with Ulva lactuca thalli during decomposition: a mesocosm experiment. J Exp Mar Biol
Ecol 275:15–29
Nestlerode JA, Diaz RJ (1998) Effects of periodic environmental hypoxia on predation of a teth-
ered polychaete, Glycera americana: implications for trophic dynamics. Mar Ecol Prog Ser
172:185–195
Nicotri ME (1977) Grazing effects of four marine intertidal herbivores on the microflora. Ecology
58:620–632
Nielsen SL, Sand-Jensen K (1990) Allometric scaling of maximal photosynthetic growth rate to
surface/volume ratio. Limnol Oceanogr 251:77–180
Nienhuis PH (1969) The significance of the substratum for intertidal algal growth on the artificial
rocky shore of the Netherlands. Int Revue Ges Hydrobiol 54(2):207–215
Niklas KJ (1992) Plant Allometry: the scaling of form and process. University of Chicago
Press, Chicago
Noffke N (2010) Microbial mats in sandy deposits from the Archean era to today. Springer, Berlin
6  Biology and Ecology of the Benthic Algae 149

Norkko JE, Bonsdorff E, Norkko A (2000) Drifting algal mats as an alternative habitat for benthic
invertebrates. Species specific responses to a transient resource. J Exp Mar Biol Ecol 248:79–64
Norton TA (1992) Dispersal by macroalgae. Br Phycol J 27:293–301
Nyberg CD, Wallentinus I (2005) Can species traits be used to predict marine macroalgal introduc-
tions? Biol Invasions 7(2):265–279
Ogata E (1953) Some experiments on the settlement of spores of red algae. Jpn J Ecol 3:128
Omstedt A, Humborg C, Pempkowiak J et al (2014) Biogeochemical control of the coupled CO2–
O2 system of the Baltic Sea: a review of the results of Baltic-C. Ambio 43:49–59
Otaiza RD, Rodriguez CY, Caceres JH et al (2018) Fragmentation of thalli and secondary attach-
ment of fragments of the agarophyte Gelidium lingulatum (Rhodophyta, Gelidiales). J Appl
Phycol 30:1921–1931
Pan J, Bournod CN, Cuadrado DG et al (2013a) Interaction between estuarine microphytobenthos
and physical forcings: the role of atmospheric and sedimentary factors. Int J Geosci 4:352–361
Pan J, Bournod CN, Pizani NV et al (2013b) Characterization of microbial mats from a siliciclastic
tidal flat (Bahía Blanca estuary, Argentina). Geomicrobiol J 30:665–674
Parodi ER (2004) Marismas y algas bentónicas. In: Piccolo MC, Hoffmeyer M (eds) Ecosistema
del estuario de Bahía Blanca. Instituto Argentino de Oceanografía, Bahia Blanca, pp 61–67
Parodi ER, Barría de Cao S (2003) Benthic microalgal communities in the inner part of the Bahía
Blanca estuary (Argentina): a preliminary qualitative study. Oceanol Acta 25:279–284
Pedersen M, Borum J (1996) Nutrient control of algal growth in estuarine waters. Nutrient limita-
tion and the importance of nitrogen requirements and nitrogen storage among phytoplankton
and species of macroalgae. Mar Ecol Prog Ser 142:261–272
Perillo GME, Iribarne O (2003) Processes of tidal channels develop in salt and freshwater marshes.
Earth Surf Proc Land 28:1473–1482
Perillo GME, Piccolo MC, Parodi ER et al (2001) Bahia Blanca Estuary ecosystem: a review. In:
Seeliger U, Kjerfve B (eds) Coastal marine ecosystems of Latin America. Springler Verlag,
Berlín, pp 205–217
Perkins EJ, Abbott OJ (1972) Nutrient enrichment and sand flat fauna. Mar Pollut Bull 3:70–72
Pickett-Heaps J, West J, Wilson S et al (2001) Time-lapse videomicroscopy of cell (spore) move-
ment in red algae. Eur J Phycol 36(1):9–22
Pihl L, Isaksoon I, Wennhage H et  al (1995) Recent increases of filamentous algae in shallow
Swedish Bays: effects on the community structure of epibenthic fauna and fish. Neth J Aquat
Ecol 29:349–358
Pinckney JL, Zingmark RG (1993) Modelling the annual production of intertidal benthic microal-
gae in estuarine ecosystems. J Phycol 29:396–407
Pizani N (2009) Valorización de las interacciones microfitobentos-sedimentos en planicies de
marea impactadas por el dragado hidráulico. Ph.D Thesis, Universidad Nacional del Sur,
Bahía Blanca
Polderman PJG (1976) Wittrockiella paradoxa Wille (Cladophoraceae) in n.w. European salt-
marshes. Hydrobiol Bull 6:98–63
Prathep A, Marrs R, Norton T (2003) Spatial and temporal variations in sediment accumulation in
an algal turf and their impact on associated fauna. Mar Biol 142(2):381–390
Raffaelli DG, Raven JA, Poole LJ (1998) Ecological impact of green macroalgal blooms. Oceanogr
Mar Biol Annu Rev 36:97–125
Raffo MP, Geoffroy A, Destombe C et al (2014) First record of the invasive red alga Polysiphonia
morrowii Harvey (Rhodomelaceae, Rhodophyta) on the Patagonian shores of the Southwestern
Atlantic. Botanica Marina 57(1):21–26
Renaud PE, Syster DA, Ambrose WGJ (1999) Recruitment patterns of continental shelf benthos
of North Carolina, USA, effects of sediment enrichment and impact on community structure.
J Exp Mar Biol Ecol 237:89–66
Rikkinen J (2013) Molecular studies on cyanobacterial in lichen symbioses. Mycokeys 6:3–32
150 M. E. Croce et al.

Robertson AI, Lenanton RCJ (1984) Fish community structure and food chain dynamics in the
surf-zone of sandy beaches: The role of detached macrophyte detritus. J Exp Mar Biol Ecol
84:265–283
Rooker JR, Holt SA, Soto MA et al (1998) Post settlement patterns of habitat use by Sciaenid
fishes in subtropical seagrass meadows. Estuaries 21:318–327
Rosenvinge LK (1927) On mobility in the reproductive cells of the Rhodophyceae. Bot Tidsskrift
40:72–80
Ross SW, Moser ML (1995) Life history of juvenile gag, Mycteroperca microlepis, in North
Carolina estuaries. Bull Mar Sci 56:222–237
Rossi F (2006) Small-scale burial of macroalgal detritus in marine sediments: effects of Ulva spp.
on the spatial distribution of macrofauna assemblages. J Exp Mar Biol Ecol 332:84–89
Rossi F, Underwood AJ (2002) Small-scale disturbance and increased nutrients as influences on
intertidal macrobenthic assemblages experimental burial of wrack in different intertidal envi-
ronments. Mar Ecol Prog Ser 241:29–39
Sagasti A, Schaffner LC, Duffy JE (2001) Effects of periodic hypoxia on mortality, feeding and
predation in an estuarine epifaunal community. J Exp Mar Biol Ecol 258:257–283
Sand-Jensen K, Borum J (1991) Interactions among phytoplankton, periphyton, and macrophytes
in temperate freshwaters and estuaries. Aquat Bot 41:137–175
Sanjeewa KKA, Kim EA, Son KT et al (2016) Bioactive properties and potentials cosmeceutical
applications of phlorotannins isolated from brown seaweeds: a review. J Photochem Photobiol
B 162:60–65
Santelices B (1990) Patterns of organization of intertidal and shallow subtidal vegetation in wave
exposed habitats of Central Chile. Hidrobiologia 192(1):35–57
Santelices B, Varela D (1994) Abiotic control of reattachment in Gelidium chilense (Montagne)
Santelices & Montalva (Gelidiales; Rhodophyta). J Exp Mar Biol Ecol 177:145–155
Sarı A, Tuzen M (2008) Biosorption of Pb (II) and Cd (II) from aqueous solution using green alga
(Ulva lactuca) biomass. J Hazard Mater 152(1):302–308
Scanlan CM, Foden J, Wells E et al (2007) The monitoring of opportunistic macroalgal blooms for
the water framework directive. Mar Poll Bull 55:162–171
Schweiger O, Settele J, Kudrna O et al (2008) Climate change can cause spatial mismatch of tro-
phically interacting species. Ecol 89:3472–3479
Selvam GG, Sivakumar K (2014) Influence of seaweed extract as an organic fertilizer on the
growth and yield of Arachis hypogea L. and their elemental composition using SEM-Energy
Dispersive Spectroscopic analysis. Asian Pac J Reprod 3:18–22
Smith KN, Herrnkind WF (1992) Predation on early juvenile spiny lobsters Panulirus argus
(Latreille): influence of size and shelter. J Exp Mar Biol Ecol 157:3–18
Snoeijs P, Busse S, Potapova M (2002) The importance of diatom cell size in community analysis.
J Phycol 38:265–272
Sousa AI, Martins I, Lilleb AI et al (2007) Influence of salinity, nutrients and light on the germina-
tion and growth of Enteromorpha sp. spores. J Exp Mar Biol Ecol 341:142–150
South GR, Whittick A (1987) Introduction to phycology. Blackwell, Oxford
Spalleti L (1980) Paleoambientes sedimentarios en secuencias silicoclásticas. Asoc Geológ
Argent 8:1–175
Spivak ED (2010) The crab Neohelice (=Chasmagnathus granulata): an emergent animal model
from emergent countries. Helgoland Mar Res 64:149–154
Stal LJ (2006) Microphytobenthos as a biogeomorphological force in intertidal sediment stabiliza-
tion. Ecol Eng 36:236–245
Stal LJ, van Gemerden H, Krumbein WE (1985) Structure and development of a benthic marine
microbial mat. FEMS Microbiol Ecol 31:111–125
Statzner B, Fièvet E, Champagne JY et al (2000) Crayfish as geomorphic agents and ecosystem
engineers: biological behavior affects sand and gravel erosion in experimental streams. Limnol
Oceanogr 45:630–640
6  Biology and Ecology of the Benthic Algae 151

Steneck RL, Dethier MN (1994) A functional group approach to the structure of algal-dominated
communities. Oikos 69:476–498
Steneck RL, Watling L (1982) Feeding capabilities and limitation of herbivorous mollusks: a func-
tional group approach. Mar Biol 68:299–319
Stevenson RJ (1996) An introduction to algal ecology in freshwater benthic habitats. In: Stevenson
RJ, Bothwell ML, Lowe RL (eds) Algal ecology: freshwater benthic ecosystems. Academic,
San Diego, pp 3–30
Stewart WDP, Rowell P, Rai AN (1983) Cyanobacteria-eukaryotic plant symbioses. Ann Inst
Pasteur Microbiol 134(1):205–228
Summerson HC, Peterson CH (1984) The role of predation in organizing benthic communities of
a temperate-zone seagrass bed. Mar Ecol Prog Ser 15:63–77
Taylor R, Fletcher RL, Raven JA (2001) Preliminary studies on the growth of selected ‘green tide’
algae in laboratory culture: effects of irradiance, temperature, salinity and nutrients on growth
rate. Bot Mar 44:327–336
Thiel M, Watling L (1998) Effects of green algal mats on infaunal colonization of a New England
mud flat–long-lasting but highly localized effects. Hydrobiologia 375:177–189
Titlyanov EA, Titlyanova TV (2006) Production of plantlets of the red alga of Gelidium genus
(Rhodophyta) from thallus fragments. Russ J Mar Biol 32(5):284–288
Turner A, Lewis MS, Shams L et  al (2007) Uptake of platinum group elements by the marine
macroalga, Ulva lactuca. Mar Chem 65(3–4):271–280
Valiela I, McClelland J, Hauxwell J et al (1997) Macroalgal blooms in shallow estuaries, controls
and ecophysiological and ecosystem consequences. Limnol Oceanogr 42:165–1118
Van den Hoek C, Mann DG, Jahns HM (1995) Algae – an introduction to phycology. Cambridge
University Press, Cambridge
van Gemerden H (1993) Microbial mats: a joint venture. Mar Geol 113:3–25
Vanderklift MA, Lavery PS (2000) Patchiness in assemblages of epiphytic macroalgae on
Posidonia coriacea at a hierarchy of spatial scales. Mar Ecol Prog Ser 192:127–135
Vasconcelos MTSD, Leal MFC (2001) Seasonal variability in the kinetics of Cu, Pb, Cd and Hg
accumulation by macroalgae. Mar Chem 74:65–85
Verlaque M (2001) Checklist of the macroalgae of Thau Lagoon (Herault, France), a hot spot of
marine species introduction in Europe. Oceanol Acta 24:29–49
Wahl M (1989) Marine epibiosis. I. Fouling and antifouling: some basic aspects. Mar Ecol Prog
Ser 58:175–189
Wahl M, Hoppe K (2002) Interactions between substratum rugosity, colonization density and peri-
winkle grazing efficiency. Mar Ecol Prog Ser 225:239–249
Walne PL, Kivic PA (1990) Euglenida. In: Margulis L, Corlissis JO, Melkonian M et  al (eds)
Handbook of Protoctista. Jones & Bartlett Publishers, Boston, pp 270–287
Watermann F, Hillebrand H, Gerdes G et al (1999) Competition between benthic cyanobacteria and
diatoms as influenced by different grain sizes and temperatures. Mar Ecol Progr Ser 187:77–87
Webb P, Wooldridge T, Schlacher T (1997) Osmoregulation and spatial distribution in four species
of mysid shrimps. Comp Biochem Physio A 117:427–431
Wilkinson M (1981) Survival strategies of attached algae in estuaries. In: Jones NV, Wolff WJ
(eds) Feeding and survival strategies of estuarine organisms. Springer, Boston, pp 29–38
Wille N (1909) Algologische Notizen XV. Über Wittrockiella nov. gen. Nytt Mag Naturvidensk
47:209–225
Williams RB (1964) Division rates of salt marsh diatoms in relation to salinity and cell size.
Ecology 45:877–880
Wilson KA, Able KW, Heck JKL (1990) Predation rates on juvenile blue crabs in estuarine nurs-
ery habitats: evidence for the importance of macroalgae (Ulva lactuca). Mar Ecol Prog Ser
58:243–251
Woodin SA (1978) Refuges, disturbance, and community structure: a marine soft-bottom example.
Ecology 59:274–284
Chapter 7
The Intertidal Meiobenthos of the Bahía
Blanca Estuary

Verónica N. Bulnes, Agustín G. Menechella, Kevin A. Rucci,


and Michel Sciberras

7.1  Introduction

The meiofauna is present in every aquatic environment, from freshwater to marine


habitats, in all kinds of sediments from muddy soft to coarse gravel, even associated
to surface structures like vegetation, sea ice and biogenetic structures, like corals
and worm tubes. Most meiofaunal organisms live in the interstitial spaces between
the sediment particles, especially near the surface, but they may be also found living
commensally in animal tubes, hydrozoan colonies and mollusc pallial cavities. In
marine habitats, their representatives inhabit from high beach sediments to the
deepest depths of the sea.
The organisms gathered in the meiofaunal realm have been known since the
eighteenth century, but it was not until 1942 that Mare defined the term

V. N. Bulnes ()
Instituto de Ciencias Biológicas y Biomédicas del Sur, INBIOSUR (Universidad Nacional del
Sur, CONICET), Bahía Blanca, Argentina
Zoología de Invertebrados I. Departamento de Biología, Bioquímica y Farmacia. Universidad
Nacional del Sur, Bahía Blanca, Argentina
e-mail: vebulnes@criba.edu.ar
A. G. Menechella
Zoología de Invertebrados I. Departamento de Biología, Bioquímica y Farmacia. Universidad
Nacional del Sur, Bahía Blanca, Argentina
Instituto Argentino de Oceanografía, IADO (Universidad Nacional del Sur, CONICET),
Bahía Blanca, Argentina
K. A. Rucci
Instituto de Investigaciones Biológicas y Tecnológicas, Universidad Nacional de Córdoba,
CONICET, IIBYT, Córdoba, Argentina
M. Sciberras
Zoología de Invertebrados I. Departamento de Biología, Bioquímica y Farmacia. Universidad
Nacional del Sur, Bahía Blanca, Argentina

© Springer Nature Switzerland AG 2021 153


S. M. Fiori, P. D. Pratolongo (eds.), The Bahía Blanca Estuary,
https://doi.org/10.1007/978-3-030-66486-2_7
154 V. N. Bulnes et al.

meiobenthos, as the animals smaller than macrofauna but larger than microfauna.
Almost all along the twentieth century, the study of these organisms was centred
on describing and classifying them, and in the 1970s, specialists showed a growing
interest in the ecology of these animals. From that moment on, defining meiofauna
for quantitative studies required the establishment of practical limits relevant to
the size and not always agreeing with taxonomic classifications. The formal size
boundaries of meiofauna were operationally defined, based on the standardized
mesh width of sieves with 500 μm (1000 μm) as upper and 44 μm (63 μm) as lower
limits (Giere 2009). All benthic organisms that pass through the coarse sieve but are
retained by the finer sieve during processing are considered meiobenthos. Since this
chapter will deal only with benthic marine metazoan phyla, the term meiobenthos
will be used as a synonym of meiofauna.
After 20  years of quantitative research, the specialist hypothesized that some
marine taxa may have developed singular meiofaunal traits over evolutionary time
and what we define today as meiofauna is not only a definite size category but also
a separate biological and ecological unit. The most important characteristics are
interpreted as consequences of the miniaturization of the body regarding the inter-
stitial habitus, and they are mostly related to reproductive aspects, like the com-
pletely benthic direct life cycle, the short generation times and the semelparity.
They are usually motile forms and feed by seeking food particles in a highly dis-
criminate manner, and their dispersion phase take place only as adults. Opposite to
the meiofaunal traits, the macrofauna is characterized by a planktonic larval devel-
opment and dispersal, with long generation times, iteroparity, and either sedentary
or motile, they seem to feed more unselectively on particles of food (Warwick
et al. 2006).
From the 36 recognized animal phyla, 22 are considered meiofaunal taxa. Most
of them constitute the ‘permanent meiofauna’, since their representatives remain
within the meiofaunal size limits throughout their life. The rest of them are part of
the ‘temporary meiofauna’ and include the taxa with immature stages that fall
within the meiofaunal size range, but when they achieve sexual maturity, they reach
macrofaunal dimensions.
The meiofauna consume a wide spectrum of food sources, including microalgae,
eukaryotic microbes, small metazoans, bacteria and organic detritus, although at
species level they often show a specialized diet. The meiofaunal activity modifies a
series of physical, chemical and biological sediment properties, positively and nega-
tively affecting various ecosystem services including sediment stabilization, bio-
chemical cycling, waste removal and food web dynamics, at various spatial and
temporal scales. This evidence is still scattered, but the interest is growing and it is
expected that unravelling and quantifying the role of the meiofauna will support
future environmental management policies (Schratzberger and Ingels 2017).
Although ubiquitous in marine habitats, the meiofaunal taxa are not evenly dis-
tributed. On a large scale, the highest abundances are recorded in estuarial environ-
ments, while the lowest are usually found at the deep sea. On a smaller scale, for
example, on a beach slope, the meiofauna distributes patchily. The causes of such
distributions are difficult to define, but benthologists agree that primarily the
7  The Intertidal Meiobenthos of the Bahía Blanca Estuary 155

sediment particle size, the temperature and the salinity are the physical factors limit-
ing the abundance and species composition (Coull 1999). The grain size is the key
factor since it directly determines spatial and structural conditions and indirectly
determines the physical and chemical conditions of the sediment. In muddy estua-
rine sediments and intertidal flats, the meiofauna is restricted to the narrow 2–3 cm
from the surface, where it faces highly variable and severe conditions. The freshwa-
ter supply varies seasonally, while the seawater supply does it twice a day.
Additionally, the temperature, the intensity of light and the exposition vary in accor-
dance with the season of the year, the tides, the time of the day and the air tempera-
ture. On top of that, the existence of anthropogenic pressures like a high local
population density, the presence of harbours and dredging activities limits the eco-
system availability for the meiobenthic organisms.
The specialist estimates there is a world average of 106 meiofaunal organisms per
square metre in every uncontaminated estuarine sediment, with a biomass of
0.75–2 mg per square metre (Coull 1999), which is low compared to other benthic
components. On the other hand, the throughput of carbon can be orders of magni-
tude higher than its standing stock, which makes the meiofauna the main responsi-
ble for the production and flow of energy (rather than its storage) in the ecosystem
(Schratzberger and Ingels 2017).
In sediments, the nematodes are usually the most abundant taxon comprising
60–90% of the total fauna and copepods are typically second at 10–40%.
Occasionally, a group other than nematodes may predominate (e.g. turbellarians) or
copepods are not second in abundance (Coull 1999). Meiofauna can be locally and
temporally very abundant, particularly after a large spawning or a settling event.
The diversity and composition of the nondominant taxa vary depending on sediment
particle size, exposure, etc.
In South America, there is a serious gap in the knowledge of the marine micro-
scopic animals. The biodiversity of meiofaunal taxa is heavily underestimated and
even less is known about the interstitial community. There are a few isolated exam-
ples in central Chile and Brazil (Rodriguez et al. 2001; Albuquerque et al. 2007;
Dupuy et al. 2015) where most meiofaunal organisms are gathered in higher taxa
like phylum, class and order, without further discrimination.
In Argentina, the meiofaunal biodiversity is scarcely known, and even in some
comprehensive contributions of the Argentinean marine fauna (Boschi and Cosseau
2004; Calcagno 2014), they remain neglected. Except for the free-living nematodes,
there is a lack of information about the interstitial marine biodiversity, both spatial
and temporal distribution, as well as its role in the ecosystems, especially in estua-
rine environments, like the Bahía Blanca Estuary.
With extensive tidal plains that are discovered at low tide, a salinity continuous
gradient and a predominance of muddy to sandy sediment with silt and clay, the
Bahía Blanca Estuary houses a benthic community with high taxonomic diversity.
When it comes to the knowledge of the estuarine communities of the southeast
Buenos Aires province, there are some long-term studies centred in the macrofaunal
component of the benthos (Bremec et al. 2007; Elías et al. 2007), overlooking the
meiofaunal component, probably due to the small size of the meiobenthic
156 V. N. Bulnes et al.

organisms, the time-consuming sample sorting, as well as the extreme complexity


of their taxonomy. It was not until 2007 when the study of the meiofaunal taxa
begun with the taxonomic contributions on microturbellarians (Bulnes 2007), nem-
atodes (Martelli 2010; Villares et al. 2013) and harpacticoid copepods (Sciberras
2018; Sciberras et al. 2014, 2018, 2021). These contributions revealed many species
new to science suggesting that the contribution of meiofauna to the biodiversity of
the Bahía Blanca Estuary remains underestimated.

7.2  Nematoda

The Nematoda constitute a highly diverse group of animals that have been exten-
sively studied, especially due to the economical and medical importance associated
to the parasitic representatives. The phylum comprises about 25,000 described spe-
cies, including the parasitic taxa, although this represents a small percentage of the
estimated biodiversity. The marine roundworms include 4000–5000 species, and
they are considered the most diverse and widespread group of nematodes, occurring
from shallow shores to the abyss. They are permanent meiobenthic inhabitants, and
at a local scale, they are highly abundant. A typical surface area meiocore
(10  cm2)  may contain about 60–90% of the total faunal diversity, followed by
harpacticoid copepods (10–40%) and other less represented taxa like Platyhelminthes,
Gastrotricha, etc. (Schmidt-Rhaesa 2014).
The nematodes’ life cycle is strictly bound to the substrate. They are easily rec-
ognized in the sediment samples, since they move undulating in the dorsoventral
plane, pushing against a substratum, such as a sand grain or any submerged surface.
They are non-segmented, cylindrical, threadlike worms, with a simple body plan
and morphologically highly homogeneous among species (Higgins and Thiel 1988).
For years, the taxonomy of the nematodes has been established on the external
characteristics, shape and cuticularization of the buccal cavity. Nevertheless, the
recent discovery that many morphospecies are in fact complexes of several geneti-
cally distinct species that are hard or impossible to discriminate based on morphol-
ogy has forced specialists to integrate both morphological and molecular information
to obtain a more accurate biodiversity status in the ecosystems (Schmidt-­
Rhaesa 2014).
Meiobenthic marine nematodes are small-sized animals ranging between 0.5 and
3 mm long. Nematodes inhabiting sandy habitats tend to be more slender, whereas
nematodes from muddy habitats are generally more robust, probably because the
former move through the interstitial apertures, while the latter borrow through the
sediments.
Despite their high diversity and abundance, their total biomass is much lower
than that of bacteria and macrobenthos, and there is a still ongoing discussion
whether nematodes significantly contribute to ecosystem processes. Nevertheless,
there is no doubt that nematodes contribute to particular ecosystem functions, such
as the decomposition of organic matter and the stabilization of intertidal muds.
7  The Intertidal Meiobenthos of the Bahía Blanca Estuary 157

In Argentina, the marine Nematoda of the Magellanic zoogeographical region


(42° to 55° South) have been extensively studied over 40 years by Catalina Pastor
de Ward and collaborators (Pastor de Ward 2001, 2003, 2004; Pastor de Ward and
Lo Russo 2007; Harguinteguy et  al. 2012; Pastor de Ward et  al. 2014, 2015a, b;
Villares et al. 2015, 2016; Martelli et al. 2017). On the other hand, the biodiversity
of nematodes from the Argentinean zoogeographical region (20° to 42° South),
including the Bahía Blanca Estuary, is still poorly known. During December 2012,
Martelli sampled in three sites along the northern coast of the Bahía Blanca Estuary
and recorded over 60 different genera (Table 7.3). Until today, only 13 taxa have
been identified to species level (see species list), two of them exclusive from the
estuary, Campylaimus bonariensis Villares, Martelli, Lo Russo and Pastor, 2013,
and Campylaimus arcuatus Villares, Martelli, Lo Russo and Pastor, 2013 (Table 7.2).
The fact that 50% of the species registered by Pastor and collaborators in the adjoin-
ing zoogeographical region were new to science led us think the biodiversity of
Nematoda in the Bahía Blanca Estuary is heavily underestimated.
The community structure, biodiversity and distribution of meiobenthic nema-
todes along the intertidal mudflats in the Bahía Blanca Estuary are a result of the
combination of environmental variables, such as the mean grain size, the proportion
of silt and clay, the food availability, the eutrophication degree and the presence of
deposited heavy metals. Their abundance is related to the organic matter concentra-
tion, and their role in the regulation of biogeochemical cycling of nutrients is not
completely understood, but their ubiquity in the sediments of the Bahía Blanca
Estuary suggest they may function as mediators associated with the strength of
energy flux through trophic webs.
The meiofaunal taxa are distributed in small patches (Giere 2009). These pat-
terns are also observed in nematodes, and for that reason, parameters like the
monthly abundance in one site, even with a high number of samples, are not good
enough to describe the community. The results from systematic sampling during
2013 and 2014 show sharp differences in the standard deviation of each sample
(Table 7.1), hiding the seasonal differences, and revealing the study of this commu-
nity must be complemented with other parameters.
Aside from the abundances and according to the environmental variables and the
distance to the estuary’s mouth, the nematode communities can be clustered in three
different groups: the inner estuary group, the mid-portion estuary group and the
outer estuary group. These differences are partially explained by their feeding hab-
its. Following Wieser’s (1953) classification, nematodes from the Bahía Blanca
Estuary can be divided into four different feeding types. The first two types are
deposit feeders – the selective deposit feeders, devoid of a developed buccal cavity;
and the nonselective deposit feeders, with a conical buccal cavity but without any
armature. Both types are exploiting the particulate organic matter. A third type are
the epistratum feeders, possessing a buccal cavity armed with denticles and other
small sclerotized structures, predominantly grazing over surface diatoms and other
microalgae. The fourth type gather the omnivores and predators, and they possess
spacious mouth openings and buccal cavities with well-developed teeth and other
cuticularized structures.
158 V. N. Bulnes et al.

Table 7.1  Monthly abundance of Nematoda and Harpacticoida in Arroyo Pareja beach (Bahía
Blanca Estuary), based on 10 samples each month (n = 10)
Nematoda Harpacticoida
Mean ± SD [min – max] Mean ± SD [min – max]
April 2013 734 ± 269.79 [448–1347] 232.2 ± 50.87 [172–300]
May 2013 535.9 ± 283 [12–933] 41.2 ± 30 [5–88]
August 2013 556.8 ± 295.77 [963–92] 28.8 ± 21.46 [8–78]
October 2013 612.9 ± 160.85 [359–903] 15.6 ± 9.7 [3–21]
December 2013 899.66 ± 443.82 [7–1370] 14.88 ± 11.37 [1–38]
January 2014 192 ± 230.91 [1–762] 48.4 ± 54.48 [4–184]
February 2014 502.4 ± 165.84 [212–660] 169.2 ± 71.69 [77–286]
The data are presented as mean ± standard deviation and [minimum–maximum] range (Sciberras,
unpublished data)

Table 7.2  List of recorded species of Nematoda in the Bahía Blanca Estuary
List of recorded species
Campylaimus arcuatus Villares, Martelli, Lo Russo and Pastor, 2013
Campylaimus bonariensis Villares, Martelli, Lo Russo and Pastor, 2013
Daptonema laxum (Wieser, 1956)
Enoploides vectis Gerlach, 1957
Molgolaimus typicus Furstenberg and Vincx, 1992
Nudora besnardi (Gerlach, 1956)
Paraethmolaimus dahli (Gerlach, 1953)
Paramonohystera parabutschlii (Timm, 1961)
Prochromadorella ditlevseni (de Man, 1922)
Sabatieria wieseri Platt, 1985
Setosabatieria hilarula (de Man, 1922)
Thalassoalaimus macrosmaticus Wieser, 1953
Thalassomonhystera parva (Bastian, 1865)
Adapted from Martelli (2013)

The communities inhabiting the inner and mid portion of the estuary are strongly
dominated by genera belonging to the nonselective deposit feeders (Fig. 7.1a, b). In
these microambients, a low oxygen level, small grain size and high organic matter
concentrations coexist. This combination of factors seems to reduce the survival
possibilities of many nematodes, resulting in communities characterized by a low
diversity and richness, and only two genera, Paraethmolaimus Jensen, 1994, and
Terschellingia de Man, 1888, dominated the biodiversity of the inner estuary
portion.
7  The Intertidal Meiobenthos of the Bahía Blanca Estuary 159

Table 7.3  List of recorded genera of Nematoda in the Bahía Blanca Estuary
List of recorded Nematoda genera
Adoncholaimus Filipjev, 1918 Monoposthia de Man, 1889
Aegialoalaimus de Man, 1907 Neochromadora Micoletzky, 1924
Anticyathus Cobb, 1920 Odontophora Bütschli, 1874
Araeolaimus de Man, 1888 Oncholaimellus de Man, 1890
Calyptronema Marion, 1870 Oncholaimus Dujardin, 1845
Campylaimus Cobb, 1920 Oxystomina Filipjev, 1918
Ceramonema Cobb, 1920 Paracyatholaimoides Gerlach, 1953
Chromadora Bastian, 1865 Paracyatholaimus Micoletzky, 1922
Chromadorina Filipjev, 1918 Paraethmolaimus Jensen, 1994
Cobbia de Man, 1907 Linhomoeus Bastian, 1865
Comesoma Bastian, 1865 Paramonhystera Steiner, 1916
Cyartonema Cobb, 1920 Polysigma Cobb, 1920
Daptonema Cobb, 1920 Prochromadorella Micoletzky, 1924
Diplolaimella Allgén, 1929 Parachromadorita Blome, 1974
Diplopeltula Gerlach, 1950 Promonhystera Wieser, 1956
Enoplolaimus de Man, 1893 Pselionema Cobb, 1933
Halalaimus de Man, 1888 Pseudocella Filipjev, 1927
Halichoanolaimus de Man, 1886 Pseudosteineria Wieser, 1956
Hopperia Vitiello, 1969 Richtersia Steiner, 1916
Laimella Cobb, 1920 Sabatieria Rouville, 1903
Leptolaimus de Man, 1876 Setosabatieria Platt, 1985
Linhystera Juario, 1974 Siphonolaimus de Man, 1893
Marylynnia (Hopper, 1972) Spilophorella Filipjev, 1917
Mesacanthion Filipjev, 1927 Spirinia Gerlach, 1963
Metachromadora Filipjev, 1918 Steineria Micoletzky, 1922
Metalinhomoeus de Man, 1907 Synonchiella Cobb, 1933
Metoncholaimus Filipjev, 1918 Terschellingia de Man, 1888
Microlaimus de Man, 1880 Thalassoalaimus de Man, 1893
Molgolaimus Ditlevsen, 1921 Thoracostomopsis Ditlevsen, 1918
Monhystera Bastian, 1865 Viscosia de Man, 1890
Adapted from Martelli (2013)

On the other hand, towards the estuary’s mouth, the sediments grain size and
therefore the permeability rises and the animals do not have to face the low tensions
of oxygen present in the inner portion of the estuary. Although this portion is still
dominated by the nonselective deposit feeders (Fig  7.1c), in particular the genus
Setosabatieria Platt, 1985, there is a growing presence of epistratum feeders, fol-
lowed by the remaining feeding types. The increase in the abundance and diversity
of nematodes suggests the presence of a wider range of food resources in this area.
To see a complete list of recorded species and genera, please check Tables 7.2
and 7.3 (Martelli 2013).
160 V. N. Bulnes et al.

Fig. 7.1  Nematoda community based on their feeding habits in the different regions of the Bahía
Blanca Estuary: (a) inner zone, (b) middle zone, (c) external zone. Adapted from Martelli (2013)

7.3  Interstitial Copepods

The subclass Copepoda possesses more than 12,500 described species. It is a rela-
tively small group of arthropods, but it is incredibly abundant and probably consti-
tutes the main group of animals in terms of the number of individuals. In marine
sediments, the copepods are usually the second most abundant taxon in the meio-
fauna after the nematodes, and they can turn out the dominant group as the particle
size of the sediment becomes coarser. Between the copepods, the order Harpacticoida
assembles most of their benthic representatives, although some members of the
order Cyclopoida have also colonized this environment. Among these benthic
organisms, those who have successfully exploited the interstitial space between the
sediment particles are usually characterized by their small size and elongate worm-
like bodies (Huys et al. 1996; Boxshall and Halsey 2004).
Typically, marine interstitial sediment communities consist of around 30 species
of copepods. Usually most individuals belong from three to five species, revealing a
high degree of dominance. Most species are present throughout the year, and their
abundances often show seasonal cyclical changes. Often, copepod abundance is
high from the end of the spring to the beginning of the autumn (warm period) and
remains low along the winter (Table  7.1). Nonetheless, some contributions have
reported abundance peaks during the autumn and over the winter (Feller 1980; Coull
and Dudley 1985; Davidovich and Chepurnov 1991). Along 1  year, some of the
7  The Intertidal Meiobenthos of the Bahía Blanca Estuary 161

species display simple life cycles, with a single abundance peak, while others
undergo more complex life cycles, with multiple abundance peaks. In general,
before an abundance peak, a reproduction peak occurs, where there is an increasing
number of females carrying eggs. When these ovigerous females are present
3–6 months of the year, the species is considered to exhibit a restricted reproductive
period. On the other side of the spectrum, this period is considered prolonged, if the
egg-carrying females remain present over 3 months, but disappear after 12 months.
Eventually, some species exhibit a continuous reproductive period, and the oviger-
ous females are present all year long. Even when there are a few examples of species
displaying continuous reproductive behaviour, most meiobenthic copepod species
have either restricted or prolonged reproductive periods (Coull and Dudley 1985).
The study of the copepods has been attached to the study of the meiofauna since
the beginning. The best-known fauna is that of the northern hemisphere. Most con-
tributions regarding the ecology of meiobenthic harpacticoids describe the results
obtained in Europe and North America. This knowledge continues to grow since
they are now considered trustworthy bioindicators of pollution (along with the nem-
atodes) (see Box 7.1).

Box 7.1: Invisible Yet Powerful: Meiofauna as a Bioindicator


Since its birth, meiobenthology has been a neglected discipline, since it deals
with tiny and almost invisible animals. Unfortunately, a defined trend took
over: the human mind seemed to be focused only on the study of ‘big ani-
mals’, such as large invertebrates or the well-known vertebrates. For a long
time, meiofaunal animals were not very interesting, not glamorous, and not
dangerous, and they look all the same – simply unimpressive tiny animals. In
recent years, molecular studies brought new interest to these old known phyla.
Meiofaunal taxa are now considered cornerstones in the explanation of animal
phylogeny and world awarded champions when it comes to explaining the
molecular machinery of the metazoan development (model organisms of three
Nobel-Prize-rewarded investigations).
Meiobenthic animals are interstitial and live between the sediment grains.
The absence of dispersal phases, like planktonic larvae, keep them their entire
life, and even for generations, in the near vicinity. These features are shared
with all meiofaunal taxa, and they have been a key factor why these organisms
gained ecological importance in the last decade.
As a consequence of the increase in human activities, marine environments
have been strongly affected. Pipeline inputs, domestic wastes, regular port
and harbour dredging, oil spills and shipping and pollutants derived from
these sources, many of which are lethal or have long-term deleterious effects,
are trapped in marine sediments where they accumulate over long periods.
This hidden danger affects all organisms in the ecosystem, and in front
the line, the invisible and neglected meiofauna.
Since the 1970s, the benthos, which are the animals inhabiting the sedi-
ments, have been identified as a suitable ecological group for biomonitoring
162 V. N. Bulnes et al.

the effects of pollution, especially the macrobenthos. Since the specialist real-
ized the meiofauna represent a separate biologically and ecologically defined
group of animals, rather than an arbitrarily defined size range of benthic inver-
tebrates, the scientist has been increasingly using the meiofauna as a bioindi-
cator of the environmental quality on coastal marine systems.
What are the features that make these animals promising candidates for
biomonitoring?
• They are ubiquitous.
• They are found in a high number of species and individuals (rich and abun-
dant), even in small volume samples.
• Some taxa display a high sensitivity to several pollution agents.
• They are characterized by rapid life cycles and thus rapid generation turn-
over and lack larval dispersion.
• They demand lower costs related to field sampling.
This combination of traits provides the scientist with rapid and reproduc-
ible information regarding marine pollutants and their effects on ecosystem
quality. Their application spectrum is high including all known methods
applied in pollution studies:
• Field studies and computer-based models
• Toxicity essays based on living organisms, obtaining uptake rates or deter-
mining sublethal reactions to deterioration or recovering over generations
• Micro- and mesocosm experiments, which allows the obtention of ‘realis-
tic’ results
• Analytical, histological and genetical approaches applied to single animals
The meiofauna is diverse, and finding a suitable model organism for these
kinds of studies is difficult. For that matter, environmental surveys have been
using the most abundant and richest taxa. The nematodes and harpacticoid
copepods come in first line, followed by other phyla like foraminifers, etc.
For example, some researchers, based on their abundance, have proposed
the nematode to copepod ratio (N:C) as a fast, easy and reliable tool for moni-
toring the effect of organic matter enrichment, a common marine contami-
nant. Since the nematodes seem more resistant to the environmental stress
induced by the increasing dissolved organic matter content, a higher ratio in
polluted areas is expected.
Another parameter used to evaluate the effects of pollutants is the index of
trophic diversity (ITD) of nematodes. This index is based on the proportion of
the different feeding groups (see Sect. 7.2). In a given area, any environmental
disturbance will affect the food supply. When the ITD grows, it means a single
trophic group dominates the community, which is the expected outcome, asso-
ciated with increased stress related to the changing nature or proportion of the
food items. However, the biomonitoring with meiofaunal taxa seems a promis-
ing tool and at the same time is at serious risk, particularly since the number of
scientists dealing with their complex taxonomy is sinking every year.
7  The Intertidal Meiobenthos of the Bahía Blanca Estuary 163

Despite the available amount of literature worldwide that discloses the relevance
of copepods as components of the meiobenthos, in Argentina, only a few ecological
studies have includes them. During a period of 15  years in the late 1970s, Rosa
Pallares made the first contributions to the taxonomic knowledge of this group in the
south of Argentina, surveying the Ría Deseado (47°45′S), the Staten Island (54°46′S)
and Tierra del Fuego (˃54°48′S), including scattered data on its ecology (e.g. Pallares
1968a, b, 1969, 1975, 1982). The state knowledge of meiobenthic copepods from
the Bahía Blanca Estuary is even scarcer. Sciberras and collaborators described
three new species since 2014 (Sciberras et al. 2014, 2017, 2021). A more compre-
hensive study of their ecology is available in her PhD thesis (Sciberras 2018), which
provides the first descriptive analysis of the seasonal variation of the abundance,
diversity and reproductive period of the meiobenthic copepod community in Arroyo
Pareja beach, a locality placed on the northern coast of the Bahía Blanca Estuary
(Fig. 2.4; Chap. 2).
The study conducted by Sciberras (2018) revealed the presence of 13 harpacti-
coid species, five of them displaying temporal and numerical dominance. They were
present ten out of 12 months and displayed the highest abundances. Halectinosoma
parejae Sciberras, Huys, Bulnes and Cazzaniga, 2017 (Ectinosomatidae), was the
dominant species during most part of the year and constitutes about the 68% of the
total of harpacticoid community in Arroyo Pareja. Regarding the abundances of the
species, Halectinosoma parejae was followed by one species of the genus Delavalia
(Miraciidae), after that Nannopus sp. (Nannopodidae) (Sciberras et  al. 2021, in
press), followed by a still undetermined species of the family Miraciidae and finally
by Quinquelaophonte aestuarii Sciberras, Bulnes and Cazzaniga, 2014
(Laophontidae). All the abovementioned species together attained the 94,06% of the
harpacticoid community in this locality. The remaining 6% was represented by
other eight species, always in a very low proportion (Fig. 7.2).
Only one species of the genus Delavalia may be found present throughout the
whole year; nevertheless, its density remains always low (Fig. 7.3c). Halectinosoma
parejae (Fig. 7.3b) was the dominant species throughout the summer (December–
February) and autumn (March–May). In July, at the beginning of the winter, the

Fig. 7.2  Total composition of benthic copepods of the Bahía Blanca Estuary
164 V. N. Bulnes et al.

Fig. 7.3  Benthic copepods of the Bahía Blanca Estuary. (a) Abundance of the harpacticoid com-
munity from Arroyo Pareja (13 putative species), (b) abundance of Halectinosoma parejae
Sciberras, Huys, Bulnes and Cazzaniga, 2017, (c) abundance of Delavalia sp., (d) abundance of
species of Miraciidae, (e) abundance of Nannopus sp., (f) abundance of Quinquelaophonte aestua-
rii Sciberras, Bulnes and Cazzaniga, 2014

adult harpacticoids disappeared almost completely. Towards the end of the coldest
season and the beginning of the spring (August–October), the species less repre-
sented won importance, until the end of the spring, when the abundance of
Halectinosoma parejae regained its dominancy. There was a strong seasonality in
terms of density in the harpacticoid community from Arroyo Pareja. The highest
abundances were registered in summer and autumn, and reached their deepest point
during August, in the middle of the winter (Fig. 7.3a). The seasonal changes of the
whole community were determined by the changes in the population of
Halectinosoma parejae. This species was more abundant in summer, although there
was a small peak of abundance in winter. The ovigerous females (females carrying
eggs) of Halectinosoma parejae were present nine out of 12 months, recording their
highest abundances during the spring, just before the population abundance peak of
February–April (Fig. 7.3b). In terms of abundance, Delavalia sp. followed second
(Fig. 7.2), showing a similar seasonal pattern to Halectinosoma parejae, with the
7  The Intertidal Meiobenthos of the Bahía Blanca Estuary 165

most important peak in March and a small peak in July (Fig.  7.3c). Similar to
Halectinosoma parejae, the ovigerous females are present 9  months of the year,
being more abundant at the end of the summer (February–March).
The species belonging to the genus Nannopus, which represents around 8% of
the harpacticoid community (Fig.  7.2), has low densities during most part of the
year, with an abrupt peak towards the end of the summer (February–March), a
period when most part of ovigerous females are registered (Fig. 7.3e).
An undetermined species belonging to the family Miraciidae (Fig. 7.2), repre-
senting around 6% of the harpacticoid community, shows a similar pattern to
Nannopus sp. since it is found in low densities during most part of the year but it
shows an abrupt peak in autumn (April), although most part of ovigerous females
are registered during the small abundance peak during the winter (August)
(Fig. 7.3d).
Regarding Quinquelaophonte aestuarii (Fig. 7.2), just a few data can be pointed
out since it represents just 3% of the harpacticoid community and it is present in
very low densities during the whole year. The samples never contained more than 17
specimens per 30 cubic centimetres, even in the months of highest abundance; thus,
only the presence was registered during the end of the winter to the beginning of the
spring (August–September) (Fig. 7.3f).
According to these data, the number of harpacticoid species in Arroyo Pareja
agrees with the expected number for marine interstitial environments. Besides, the
harpacticoid community follows the high dominance model since Halectinosoma
parejae is the dominant species and, together with just a few more species, they
represent more than 90% of the total harpacticoid community. The temporal pattern
of higher abundances in the warm period is verified, although a lower abundance
peak occurs in winter (August), followed by the disappearance of practically all
harpacticoid adults in June and July.
The absence of Halectinosoma parejae in the winter samples suggests that this
species has an annual life cycle. The peaks of abundance of three of the less abun-
dant species in August (Delavalia sp., Nannopus sp. and the undetermined species
belonging to the family Miracidae) could be related to the absence of the dominant
species (Halectinosoma parejae) or with the development of an algal bloom, char-
acteristic of this environment (Freije and Gayoso 1988; Gayoso 1998; Popovich
et al. 2008). Although these blooms have been extensively studied associated to the
phytoplankton of the estuary, the sudden availability of nutrients that causes it prob-
ably has a similar effect in microfitobenthos.
The four most abundant species appear to have different reproductive periods.
Halectinosoma parejae ovigerous females can be found in all the months that adults
are present. Delavalia sp. and Nannopus sp. also have prolonged reproductive peri-
ods, but with a much higher proportion of ovigerous females in summer, while the
unidentified species of Miraciidae seems to concentrate its reproduction in
the winter.
The meiobenthic copepod community of the estuary is diverse, and its study adds
another dimension to the community dynamics in a complex and changing environ-
ment that remains still little explored. Although the data exposed in this section arise
166 V. N. Bulnes et al.

from short-term standardized surveys on a single locality of the Bahía Blanca


Estuary (Sciberras 2018), it constitutes the basis for planning the next steps regard-
ing this well-represented but little-known zoological group.

7.4  The Turbellarians

For over 140  years the turbellarians have been important research objects. They
display an extraordinary regeneration power and  they are considered one of the
early-most spiralian Eubilateria lineages. For a long time, the turbellarians were
gathered in a class within the Platyhelminthes. In the last two decades, with the
development and use of advanced molecular studies to unravel the metazoan evolu-
tionary relationships, strong evidence is  supporting that the turbellarians are not
only a paraphyletic group of Platyhelminthes, but they also gather more than one
phylum of metazoan animals.
In this chapter, the phylum Acoelomorpha Ehlers, 1985, comprising the clade
Acoela Uljanin, 1870, and Nemertodermatida Karling, 1970, as well as some tradi-
tional Platyhelminthes families, will be considered turbellarians, as it is usually
considered in the scientific literature.
According to their size, turbellarians can be divided into two major groups, with
no taxonomic association: the macroturbellarians or large worms, including speci-
mens with a length between 1  cm and 30  cm, and the microturbellarians, which
usually range in length from 10 μm to 300 μm. Marine triclads and polyclads con-
stitute the macroturbellarians, which generally occur on hard seafloor, under stones,
or associated with macrophytes, algae or gravel. The microturbellarians usually
occur on sandy or muddy bottoms, in sheltered tide flats, in the lower beach slope
and in the swash zone (Giere 2009).
There are more than 3000 described species of turbellarians, most of them inter-
stitial microturbellarians inhabiting the marine littoral. They are small-sized bilate-
rian acoelomate worms, oval to round in cross section, extremely flexible that
combine muscular and ciliary action to achieve movement, with a ventrally posi-
tioned mouth, without an anus and usually with an epidermal dual-gland adhesive
system, which make them well adapted to the interstitial life. Additionally, they
display direct benthic development, dispersal only as adults, short generation times
and semelparity.
Their diversity is high, but their small size, as well as their taxonomic complex-
ity, usually excludes them from ecological surveys. The determination always
requires an assortment of separation methods, the study of live material under light
microscope in delicate squash preparations, time-consuming sectioning and inter-
pretation of histological slides and, finally, comparison with, in some cases, rather
incomplete or imprecise descriptions often hidden in obscure journals rarely acces-
sible in our country.
7  The Intertidal Meiobenthos of the Bahía Blanca Estuary 167

The density and biomass of turbellarians are usually higher in sandy habitats
than in muddy bottoms. In sand, their diversity can be of the same magnitude as that
of nematodes, and turbellarian biomass can be even higher.
In the meiofaunal community, the majority of the turbellarian species are consid-
ered to be predators of small zoobenthos, representing the top consumers, where
they usually play the role, that in muddy sediments are carried out by the macrofau-
nal predators (Martens and Schockaert 1986). Based on gut analysis and observa-
tions on feeding, there are also grazers feeding on bacteria and microalgae.
Occasionally, diatoms are found in the gut of predators, but it remains unclear
whether they entered the gut via prey animals or they show some grade of mixotro-
phic behaviour (Reise 1984).
The turbellarians are not beneficiated by the sediment destabilization, and they
usually avoid living in funnels and mounds, but they inhabit the vicinity of burrows,
especially near the feeding pocket of other borrowing taxa, like annelids (Reise 1985).
The first monographic treatises on turbellarians are almost 130 years old, and
since then, a lot has been achieved in numerous aspects including taxonomy, sys-
tematic, biogeography, evolutionary development, physiology and ecology.
Nevertheless, most of the studies have been carried out in the northern hemisphere
(Rieger 1998). In South America, most studies were done on Brazilian turbellarians
between 1940 and 1960, and in the Galapagos Islands, in the 1970s. In Argentina,
most of the microturbellarians surveys have been carried out by Dr. Damborenea
and Dr. Brusa in La Plata Museum, focusing their interest mainly on freshwater spe-
cies (Brusa 2006a, b; Damborenea et al. 2007; Brusa et al. 2008).
Since 2004 Bulnes has taken over the study of the meiobenthic marine turbellar-
ian community. The first studies were carried out in two localities of the northern
coast of the Bahía Blanca Estuary: Arroyo Pareja beach and Baterías beach. The
first locality represents a highly anthropized area, where a number of industrial
activities as well as seasonal touristic and sport activities are performed. On the
other hand, Baterías beach shows a low anthropogenic influence, since it is located
in a restricted area of the Naval Base Puerto Belgrano (Fig. 2.4; Chap. 2). Between
November 2004 and February 2006, a systematic survey on these localities revealed
some ecological information of the Bahía Blanca Estuary’s community of meioben-
thic turbellarians. This community displayed a great degree of spatial, temporal and
abundance diversity. The sampling was aimed to examine the first five superficial
centimetres of the surf zone, and in the intertidal and supralittoral zone, the sedi-
ments from the surface, down to the water table depth (McLachlan and Defeo 2018).
During this survey, 3278 specimens were extracted and taxonomically deter-
mined. The results revealed the presence of over 28 species of interstitial free-living
Platyhelminthes and Acoelomorpha (Bulnes 2007).
Although the turbellarians are present all year long in both localities, two maxi-
mum abundance peaks were registered during the autumn and the spring in April
and May and October. The lowest abundances occur in summer (January and
February) and at the end of the winter (August and September) (Fig. 7.4a).
To see a complete list of recorded species and genera, please check Tables 7.4
and 7.5.
168 V. N. Bulnes et al.

Fig. 7.4  Turbellarians of the Bahía Blanca Estuary. (a) Total abundance in Baterías and Arroyo
Pareja beaches, (b) Shannon (H′) and Pielou (J′) index in Baterías beach, (c) Shannon (H′) and
Pielou (J′) index in Arroyo Pareja beach, (d) dominance index variation in Baterías and Arroyo
Pareja beaches

Most of the species were distributed independently from each other, and they
were present for at least 3 months of the year displaying gradual taxon substitution.
In Baterías beach, the highest diversity index (H′) was found in January (ten spe-
cies), during one of the lowest abundance peaks, while in September the lowest
diversity index (three species) agreed with a low abundance peak. The highest
Pielou equity indexes (J′) were detected in February, April, October and November,
whereas the lowest in May, June, July and September (Fig. 7.4b). In Arroyo Pareja
beach, the highest diversity indexes occurred in November, January and June (eight
species) and reached the lowest diversity index (two species) in July. The highest
values of the Pielou equity index were detected in September and November,
7  The Intertidal Meiobenthos of the Bahía Blanca Estuary 169

Table 7.4  List of turbellarian species registered in the Bahía Blanca Estuary
Acoelomorpha
Philocelis robrochai Hooge and Rocha, 2006
Platyhelminthes Proseriata
Vannuccia talea Marcus, 1954
Macrostomorpha
Karlingia lutheri (Marcus, 1948)
Myozona evelinae Marcus, 1949
Archimacrostomum brasiliensis (Marcus, 1952)
Dalytyphloplanida
Kalyla gabriellae Marcus, 1951
Kalyptorhynchia
Cheliplana uruguayensis Van Steenkiste, Volonterio,
Schockaert and Artois, 2008

Table 7.5  List of turbellarian Acoelomorpha


genera registered in the Bahía
Deuterogonaria Dörjes, 1968
Blanca Estuary
Kuma Marcus, 1950
Symsagittifera Kostenko and
Mamkaev, 1990
Platyhelminthes
Macrostomum Schmidt, 1848
Microstomum Schmidt, 1848
Promonotus Beklemischev, 1927
Meidiama Marcus, 1946
Carcharodorhynchus Meixner, 1938
Cheliplana de Beauchamp, 1927
Gnathorhynchus Meixner, 1929

whereas the lowest in April and July (Fig. 7.4c). Although Baterías beach showed
higher dominance indexes than Arroyo Pareja beach, both showed the highest diver-
sity in autumn and spring, and the lowest were in winter and summer.
The calculated dominance indexes (Fig. 7.4d) suggest that the species founded
favourable conditions, especially regarding the food availability, leading to a fast
growth in the abundance of fewer species that exploit rapidly the resources, proba-
bly multiplying asexual reproduction cycles, producing a high number of gametes
or even achieving sexual maturity in a short period (Heitkamp 1988). On the other
hand, the high diversity values calculated in the summer and the winter may be a
consequence of the fast development of highly specialized species, present only in
low numbers, but adapted to explore this temporarily impoverished environment.
In Baterías beach, 50% of the specimens belong to the clade Proseriata, followed
by the Acoela, the Macrostomorpha, the Kalyptorhynchia and the Neodalyellida
(Fig. 7.5a), while in Arroyo Pareja beach, 75% of the animals were representatives
170 V. N. Bulnes et al.

Fig. 7.5  Turbellarians of the Bahía Blanca Estuary. (a) Relative abundances of turbellarian groups
in Baterías beach, (b) in Arroyo Pareja beach

of the clade Acoela, followed by the Proseriata, the Macrostomorpha and the
Kalyptorhynchia (Fig. 7.5b).
The dominancy values remain high, since Deuterogonaria sp. represented the
50.24% of the collected specimens, being completely dominant in Arroyo Pareja
beach during March and April, although absent in July and August (Fig.  7.6a).
Regarding the abundance, a new species of Symsagittifera and a species belonging
to the genus Promonotus followed the abovementioned acoel. Both Symsagittifera
sp. and Promonotus sp. showed high abundance peaks in different months. The
14.3% of all the turbellarians were Symsagittifera sp., while 13.44% were
Promonotus sp. The Proseriata Promonotus displayed an abundance peak in autumn,
like Deuterogonaria sp., while Symsagittifera sp. was more abundant in spring
(October), exhibiting one of the examples of the already mentioned gradual taxon
substitution (Fig. 7.6b). In Baterías beach, the dominancy values of Deuterogonaria
sp. (23.02%) and Meidiama sp. (26.5%) were alike, although lower than in Arroyo
Pareja beach. Moreover, Meidiama sp. was absent during the months when
Deuterogonaria sp. was dominant, and Deuterogonaria was absent during the
period when Meidiama sp. was dominant, suggesting there is some degree of com-
petition between these two species (Fig. 7.6c).
In general, the Acoela were dominant in Arroyo Pareja beach. This may be
directly connected to the fact that Arroyo Pareja beach sediments are a mixture of
fine-grain sand with a high content of clay, since mudflats and shallow tidal chan-
nels surround this locality, where the benthic microalgae proliferate. This combina-
tion of small-sized interstitial spaces and the abundance of diatoms as a food source
may condition the development of the small-sized Acoela over the other larger tur-
bellarian taxa. In Baterías, the beach profile is slightly more exposed to the wave
action, and thus, the clay content is lower and the interstitial water currents are
stronger. Here, the development of the turbellarians with adhesive dual glands for
fixation to the substrate seems favoured. The high energy of this environment condi-
tions the proliferation of diatoms, and the interstitial microhabitats are dominated
by the predacious Proseriata and Kalyptorhynchia (Wellner and Reise 1989).
7  The Intertidal Meiobenthos of the Bahía Blanca Estuary 171

Fig. 7.6  Turbellarians of the Bahía Blanca Estuary. (a) Relative abundance of Deuterogonaria sp.
in Arroyo Pareja beach, (b) Promonotus sp. and Symsagittifera sp. abundances in Arroyo Pareja
beach, (c) relative abundance of Deuterogonaria sp. and Meidiama sp. in Baterías beach

The direct observation of diatoms in the gut content of Macrostomorpha and


Proseriata suggests that although these Rhabditophora usually predate on copepods,
nematodes and other turbellarians, they are also able to exploit different food
sources, especially when the preys are scarce. Although the reproductive behaviour
of the turbellarian agrees with the generalization made for meiobenthic organisms,
the seasonal low selectivity seems to collide with the generalization made by
Warwick and collaborators, when they characterized the meiofaunal taxa as selec-
tive feeders, being temporarily opportunistic, under given conditions (Warwick
et al. 2006).
172 V. N. Bulnes et al.

7.5  Other Taxa

7.5.1  Tardigrades

The water bears belong to the phylum Tardigrada. They live associated with a coat-
ing of water, which allows the animals to move and achieve all their metabolic
requirements. The marine tardigrade may be found from the tide line up to the abys-
sal depths. They are cylindrical-bodied animals with four pair of limbs and possess
numerous sensory structures, usually concentrated on the head, a complete diges-
tive system and they are gonochoristic. Since 2015, Menechella and collabora-
tors described three new species of marine tardigrades, all from a city in the direct
vicinity of the Bahía Blanca Estuary (Menechella et al. 2015, 2017).
In Arroyo Pareja beach, Bulnes (unpublished data) found tardigrades in the sam-
ples with a higher sand content, from March to June and always in low numbers, not
higher than five individuals per 25 cubic centimetres of sediment. The most frequent
species is Batillipes amblypyge Menechella, Bulnes and Cazzaniga, 2017 (Fig. 7.6a);
and rarely Batillipes lingularum Menechella, Bulnes and Cazzaniga, 2017, was
present. The latter is smaller and it is always covered with some debris on its dorsal
surface, which make the exemplars difficult to sort out from the samples.

7.5.2  Kinorhynchs

The Kinorhyncha or mud dragons are marine microscopic animals, usually more
diverse in muddy sediments, rather than in sandy habitats. They have a short neck
and a trunk divided into 11 segments, usually with several sensory structures like
spines, sensory and glandular spots. The head displays numerous sensory append-
ages, the scalids, usually protected by a set of flattened appendages, the placids,
which combine to form a movable protective structure of the head.
In the Bahía Blanca Estuary, only two species have been registered. In Arroyo
Pareja beach, in sediments with higher sand content, Rucci and colaborators found
Franciscideres cf. kalenesos Dal Zotto, Di Domenico, Garrafoni and Sørensen,
2013 (Rucci et al. 2020), a species with some unique adaptations to life in sandy
habitats. In areas dominated by finer sediments, another species belonging to the
clade Allomalorhagida has been collected, but they have been accidental observa-
tions, since the sampling and separation of kinorhynchs for taxonomic studies
require specific techniques.

7.5.3  Annelids

The interstitial life seems to have evolved many times in different taxonomic groups,
especially between the polychaetes. In many microscopic annelids, the segmental
parapodia, the cirri as well as head appendages are reduced or even lacking. Many
7  The Intertidal Meiobenthos of the Bahía Blanca Estuary 173

species seem to have retained some paedomorphic features associated with their
miniaturization (Higgins and Thiel 1988).
In the Arroyo Pareja beach, a species of Polygordius Schneider, 1868, has been
observed, especially in the spring months (Fig.  7.7d). Several other archiannelid
species have been observed, but there is still a lack of taxonomic information regard-
ing this group.

Fig. 7.7  Meiofaunal taxa of the Bahía Blanca Estuary. (a) Batillipes amblypyge Menechella,
Bulnes and Cazzaniga, 2017, (b) gastrotrich Chaetonotoidea, (c) Pseudostomella sp., (d)
Polygordius sp., (e) Ototyphlonemertes sp. (Photos by Verónica N. Bulnes)
174 V. N. Bulnes et al.

7.5.4  Gastrotrichs

They are small, aquatic and interstitial worms. They are characterized by the dense
ciliature on their ventral surface, as well as an anterior terminal mouth and a distal
pair of adhesive tubes. Their identification depends on the characterization of the
ornamentation of the body, like scales, plates and spines covering the trunk. The
gastrotrichs are gathered in two orders: the Macrodasyda and the Chaetonotida
(Fig. 7.7b). In Arroyo Pareja, representatives of both orders have been recognized,
and the most frequent gastrotrich belongs to the genus Pseudostomella Swedmark,
1956 (Fig. 7.7). The knowledge of the South American marine gastrotrichs is still
poor, and the lack of specialists in this group in Argentina will not allow us to char-
acterize the community of this taxon in the near future.

7.5.5  Nemerteans

The nemerteans are non-segmented worms, a protrusible proboscis, with complete


digestive system, a well-developed nervous system and a closed blood vascular sys-
tem. They are found in all marine habitats and they are considered voracious preda-
tors, making them top consumers in the meiofaunal food chain.
Like with many other animals, the collection, separation and preparation of
nemerteans require specific methods, still not implemented. Nevertheless, the meth-
ods used until now allowed us to obtain some preliminary data.
In the Bahía Blanca Estuary, we have registered a species belonging to the genus
Ototyphlonemertes Diesing, 1863 (Fig. 7.7e). This genus is considered cosmopoli-
tan, and a detailed taxonomic characterization of this species may provide valuable
data, since there are major gaps in the geographic records of this genus.
Microscopic free-living animals have been considered somehow cosmopolitan
for a long time. In contrast to macrofauna, samples from new studied areas have
often revealed microscopic animals that could be ascribed to familiar species. It is
clear we possess very few data to make inferences about the distribution and many
cosmopolitan species could represent complexes of sibling species, each with prob-
ably a much-restricted area of distribution. Only intensive sampling and morpho-
logical studies combined with molecular techniques may throw some light on the
biogeography of the meiofauna (Artois et al. 2011). In the Bahía Blanca Estuary, the
study of the meiofauna started a few years ago, but although the description of the
biodiversity will take some more time, many meiobenthic taxa have been recog-
nized and await for a proper characterization.
7  The Intertidal Meiobenthos of the Bahía Blanca Estuary 175

References

Albuquerque EF, Brandão Pintol AP, d’Alcântara de Queiroz Perez A et  al (2007) Spatial and
temporal changes in intertidal meiofauna on a sandy beach of South America. Braz J Oceanogr
55(2):121–131. https://doi.org/10.1590/S1679-­87592007000200005
Artois T, Fontaneto D, Hummon WD et  al (2011) Ubiquity of microscopic animals? Evidence
from the morphological approach in species identification. In: Fontaneto D (ed) Biogeography
of Microscopic Organisms. Is everything small everywhere? The systematic association special
79:244–283
Boschi EE, Cousseau MB (eds) (2004). La vida entre mareas: vegetales y animales de las costas de
Mar del Plata, Argentina. Publicaciones Especiales INIDEP, Argentina
Boxshall GA, Halsey SH (2004) An introduction to copepod diversity, vol I. Ray Society, London
Bremec CS, Martínez DE, Elías R (2007) Asociaciones bentónicas de fondos duros y comunidades
incrustantes. In: Piccolo MC, Hoffmeyer M (eds) Ecosistema del estuario de Bahía Blanca.
EDIUNS, Bahía Blanca
Brusa F (2006a) First record of Prolecithophora (Platyhelminthes) from Argentina: Plagiostomum
cilioejaculator sp. n. Zootaxa 1223:47–53. https://doi.org/10.11646/zootaxa.1223.1.4
Brusa F (2006b) Macrostomida (Platyhelminthes: Rhabditophora) from Argentina, with descrip-
tions of two new species of Macrostomum and of stylet ultrastructure. Zool Sci 23:853–862.
https://doi.org/10.2108/zsj.23.853
Brusa F, Damborenea MC, Noreña C (2008) Dalyellioida (Platyhelminthes, Rhabdocoela) from
the Rio de la Plata estuary in Argentina, with the description of two new species of Gieysztoria.
Zootaxa 1861:1–16. https://doi.org/10.11646/zootaxa.1861.1.1
Bulnes VN (2007) Microturbelarios (Acoelomorpha y Platyhelminthes) litorales del estuario de
Bahía Blanca. PhD Dissertation. Universidad Nacional del Sur, Argentina
Calcagno JA (ed) (2014) Los invertebrados marinos. Fundación de Historia Natural Félix de
Azara, Argentina
Coull BC (1999) Role of meiofauna in estuarine soft-bottom habitats. Austral Ecol 24(4):327–343.
https://doi.org/10.1046/j.1442-­9993.1999.00979.x
Coull BC, Dudley BW (1985) Dynamics of meiobenthic copepod populations: a long-term study
(1973–1983). Mar Ecol Prog Ser 24:219–229
Damborenea MC, Brusa F, Noreña C (2007) New Dalyelliidae (Platyhelminthes, Rhabditophora)
from Buenos Aires province, Argentina, and their stylet ultrastructure. Zool Sci 24:803–810.
https://doi.org/10.2108/zsj.24.803
Davidovich NA, Chepurnov V (1991) Periodical components of harpacticoids abundance dynam-
ics in upper Black-sea sublittoral near Karadag. Okeanologiya 31(1):140–145
Dupuy C, Nguyen TH, Mizrahi D et al (2015) Structure and functional characteristics of the meio-
fauna community in highly unstable intertidal mud banks in Suriname and French Guiana
(North Atlantic coast of South America). Cont Shelf Res 110:39–47. https://doi.org/10.1016/j.
csr.2015.09.019
Elías R, Iribarne O, Bremec CS et  al (2007) Comunidades bentónicas de fondos blandos. In:
Piccolo MC, Hoffmeyer M (eds) Ecosistema del estuario de Bahía Blanca. EDIUNS, Bahía
Blanca, pp 179–190
Feller RJ (1980) Quantitative cohort analysis of a sand-dwelling meiobenthic harpacticoid cope-
pod. Est Coast Mar Sci 11(4):459–476. https://doi.org/10.1016/S0302-­3524(80)80068-­5
Freije RH, Gayoso AM (1988) Producción primaria del estuario de Bahía Blanca. Informes
UNESCO, Ciencias del Mar, Uruguay 7:112–114
176 V. N. Bulnes et al.

Gayoso AM (1998) Long-term phytoplankton studies in the Bahía Blanca Estuary, Argentina.
ICES J Mar Sci 55:655–660
Giere O (2009) Meiobenthology. The microscopic motile fauna of aquatic sediments, 2nd edn.
Springer, Berlin
Harguinteguy CA, Cofré MN, Pastor de Ward CT (2012) Change in the meiofaunal commu-
nity structure of sandy beaches of the Nuevo Gulf (Chubut, Argentina). Pap Avulsos Zool
52(34):411–422. https://doi.org/10.1590/S0031-­10492012021400001
Heitkamp U (1988) Life-cycles of microturbellarians of pools and their strategies of adaptation to
their habitats. Prog Zool 36:449–455
Higgins RP, Thiel H (1988) Introduction to the Study of Meiofauna. Smithsonian Institution Press,
Washington, DC
Huys R, Gee JM, Moore CG, Hamond R (1996) Marine and brackish water harpacticoid copepods.
Part 1. Keys and notes for identification of the species, Synopses of the British Fauna (new
series). Academic Press, Linnean Society of London, London
Martelli A (2010) Estudio taxonómico preliminar de nematodos de vida libre del estuario de Bahía
Blanca. Degree thesis. Universidad Nacional del Sur, Argentina
Martelli A (2013) Technical report. Comisión de Investigaciones Científicas de la Provincia de
Buenos Aires. La Plata, Argentina
Martelli A, Lo Russo V, Villares G et  al (2017) Two new species of free-living marine nema-
todes of the family Oxystominidae Chitwood, 1935 (Enoplida) with a review of the genus
Thalassoalaimus de Man, 1893 from the Argentine coast. Zootaxa 4250(4):347–357. https://
doi.org/10.11646/zootaxa.4250.4.5
Martens P, Schockaert ER (1986) The importance of turbellarians in the marine meiobenthos: a
review. Hydrobiologia 132:295–303. https://doi.org/10.1007/BF00046263
McLachlan A, Defeo O (2018) The ecology of sandy shores, 3rd edn. Academic Press, London
Menechella AG, Bulnes VN, Cazzaniga NJ (2015) A new Batillipedidae (Tardigrada,
Arthrotardigrada) from Argentina. Zootaxa 4032(3):339–344. https://doi.org/10.11646/
zootaxa.4032.3.11
Menechella AG, Bulnes VN, Cazzaniga NJ (2017) Two new species of Batillipes (Tardigrada,
Arthrotardigrada, Batillipedidae) from the Argentinean Atlantic coast, and a key to all known
species. Mar Biodivers 48:239–247. https://doi.org/10.1007/s12526-­017-­0640-­4
Pallares RE (1968a) Copépodos marinos de la Ría Deseado (Santa Cruz, Argentina). Contribución
sistemática-ecológica I. Servicio de Hidrografía Naval H-1024:11–25
Pallares RE (1968b) Copépodos marinos de la Ría Deseado (Santa Cruz, Argentina). Contribución
sistemática-ecológica II. Physis 27(75):245–262
Pallares RE (1969) El género Scutellidium en la Ría Deseado (Crustacea, Copepoda). Physis
29(78):51–72
Pallares RE (1975) Copépodos marinos de la Ría Deseado (Santa Cruz, Argentina). Contribución
sistemática-ecológica IV. Physis 34(89):213–227
Pallares RE (1982) Copépodos harpacticoides marinos de Tierra del Fuego (Argentina). IV. Bahía
Thetis. Centro de Investigación de Biología Marina (CIBIMA). Contribución Científica
186:1–39
Pastor de Ward CT (2001) A new nematode from West Patagonian coast, Biarmifer madrynensis
sp. n. with redefinition of the genus Biarmifer Wieser 1954 (Nematoda, Cyatholaimidae). Bull
Inst R Sci Nat Belg Biol 71:139–149
Pastor de Ward CT (2003) Two new species of Sabatieria (Nematoda, Comesomatidae) from Golfo
Nuevo, Chubut (Argentina). Zootaxa 172:1–12. https://doi.org/10.11646/zootaxa.172.1.1
Pastor de Ward CT (2004) New species of Hopperia (Nematoda, Comesomatidae) and
Metachromadora (Nematoda, Desmodoridae) from Patagonia, Chubut, Argentina. Zootaxa
542:1–7. https://doi.org/10.11646/zootaxa.542.1.1
Pastor de Ward CT, Lo Russo V (2007) A review of the genus Richtersia (Nematoda:
Selachinematidae): new species from Golfo San José and Golfo San Matías, Chubut (Argentina).
J Mar Biol Assoc UK 87(5):1153–1160. https://doi.org/10.1017/S0025315407056755
7  The Intertidal Meiobenthos of the Bahía Blanca Estuary 177

Pastor de Ward CT, Lo Russo V, Villares G (2014) Two new species of Prasaveljevia Wieser,
1953 (Thoracostomopsidae, Nematoda) from Argentinean coasts (Chubut, Argentina). Zootaxa
4027(4):551–564. https://doi.org/10.11646/zootaxa.3999.4.2
Pastor de Ward CT, Lo Russo V, Villares G (2015a) Patagonema iubatum gen. Nov. and sp. nov.
(Nematoda, Oncholaimidae, Octonchinae) from Patagonia sandy beach (Argentina). J Mar
Biol Assoc UK 95(1):101–109. https://doi.org/10.1017/S0025315414000708
Pastor de Ward CT, Lo Russo V, Villares G et  al (2015b) Free-living marine nematodes from
San Julián Bay (Santa Cruz, Argentina). ZooKeys 489:1–133. https://doi.org/10.3897/
zookeys.489.7311
Popovich CA, Guinder VA, Pettigrosso RE (2008) Composition and dynamics of phytoplank-
ton and aloricate ciliate communities in the Bahía Blanca estuary. In: Neves R, Baretta JW,
Mateus M (eds) Perspectives on integrated coastal zone management in South America. IST
Press, Lisboa
Reise K (1984) Free-living Platyhelminthes (Turbellaria) of a marine sand flat: an ecological study.
Microfauna Mar 1:1–62
Reise K (1985) Tidal flat ecology. An experimental approach to species interactions, Ecological
studies 54. Springer, Berlin/New York
Rieger RM (1998) 100 years of research on ‘Turbellaria’. Hydrobiol 383:1–27. https://doi.org/1
0.1023/A:1003423025252
Rodriguez JG, Lopez J, Jaramillo E (2001) Community structure of the intertidal meiofauna
along a gradient of morphodynamic sandy beach types in southern Chile. Rev Chil Hist Nat
74:885–897. https://doi.org/10.4067/S0716-­078X2001000400015
Rucci KA, Neuhaus B, Bulnes VN et al (2020) New record of the soft-bodied genus Franciscideres
(Kinorhyncha) from Argentina, with notes on its movement and morphological variation.
Zootaxa 4780(1):107–131. https://doi.org/10.11646/zootaxa.4780.1.5
Schmidt-Rhaesa A (ed) (2014) Handbook of zoology. Gastrotricha, Cycloneuralia and Gnathifera.
Volume 2 Nematoda. De Gruyter, Berlin/Boston
Schratzberger M, Ingels J (2017) Meiofauna matters: the roles of meiofauna in benthic ecosys-
tems. J Exp Mar Biol Ecol 502:12–25. https://doi.org/10.1016/j.jembe.2017.01.007
Sciberras M (2018) Biodiversidad de copépodos harpacticoideos bentónicos en un área del estu-
ario de Bahía Blanca. PhD dissertation. Universidad Nacional del Sur, Argentina. http://reposi-
toriodigital.uns.edu.ar/handle/123456789/4460
Sciberras M, Bulnes VN, Cazzaniga NJ (2014) A new species of Quinquelaophonte (Copepoda:
Harpacticoida) from Argentina. Zoologia (Curitiba) 31:496–502. https://doi.org/10.1590/
S1984-­46702014000500010
Sciberras M, Huys R, Bulnes VN et al (2017) A new species of Halectinosoma Vervoort, 1962
(Copepoda: Harpacticoida) from Argentina, including a key to species with unusual leg arma-
ture patterns, notes on wrongly assigned taxa and an updated key to ectinosomatid genera. Mar
Biodivers 48:407–422. https://doi.org/10.1007/s1252601708060
Sciberras M, Cazzaniga NJ, Huys R (2021) Description of a new species of Nannopus Brady,
1880 (Copepoda: Harpacticoida) from Argentinean waters, including an updated key to spe-
cies. Zootaxa. (in press)
Villares G, Martelli A, Lo Russo V et  al (2013) Three new species and one new record of
Campylaimus (Diplopeltidae, Nematoda) from argentine coasts (Buenos Aires and Santa Cruz,
Argentina). Zootaxa 3613(1):083–096. https://doi.org/10.11646/zootaxa.3613.1.4
Villares G, Lo Russo V, Pastor CT (2015) A new species of Pontonema (Oncholaimidae, Nematoda)
and a redescription of Pontonema incisum Wieser, 1953 from Santa Cruz and Chubut Provinces,
Argentina. Zootaxa 4058(3):417–428. https://doi.org/10.11646/zootaxa.4058.3.8
Villares G, Lo Russo V, Pastor de Ward CT et  al (2016) Free-living marine nematodes from
San Antonio Bay (Río Negro, Argentina). ZooKeys 574:43–55. https://doi.org/10.3897/
zookeys.574.7222
Warwick RM, Dashfield SL, Somerfield PJ (2006) The integral structure of a benthic infaunal
assemblage. J Exp Mar Biol Ecol 330:12–18. https://doi.org/10.1016/j.jembe.2005.12.013
178 V. N. Bulnes et al.

Wellner G, Reise K (1989) Plathelminth assemblages from an exposed and a sheltered Sandy
Beach of the Island of Sylt in the North Sea. Microfauna Mar 5:277–294
Wieser W (1953) Die Beziehung zwischen Mundhöhlengestalt, Ernährungsweise und Vorkommen
bei freilebenden marinen Nematoden. Ark Zool 2:439–484
Chapter 8
The Intertidal Soft-Bottom Macrobenthic
Invertebrates

M. Cecilia Carcedo, Sabrina Angeletti, Georgina Zapperi, Eder P. Dos Santos,


and Sandra M. Fiori

8.1  T
 he Macrobenthic Invertebrates and Their Distribution
in Intertidal Areas of the Estuary

Sheltered coasts of bays and estuaries are depositional environments with fine-­
grained sediments and gently sloping intertidal areas, commonly known as tidal
flats. These environments form in those places where tides and tidal currents domi-
nate over other hydrodynamic forces (Klein 1985). Tidal flats are inhabited by a
diverse assemblage of invertebrates, ranging from microscopic organisms within
interstitial spaces of sediment particles to large forms such as crabs and shrimps.
Specifically, those invertebrates larger than 500 μm and living associated with the
bottom are called macrobenthic invertebrates. They constitute a diverse group of
animals in coastal areas, in terms of its number of species, habitat preferences and
feeding habits. Based on their position relative to the water/sediment interface, the
macrobenthic invertebrates can be classified into three categories: (1) the endoben-
thos, that is, animals living buried within the sediment; (2) the epibenthos, that is,
animals living on top of the sediment surface; and (3) the hyperbenthos, that is,
animals living in the water layer close to the sea bed, but also found within the

M. C. Carcedo () · S. M. Fiori


Instituto Argentino de Oceanografía IADO, (Universidad Nacional del Sur-CONICET),
Bahía Blanca, Argentina
Departamento de Biología, Bioquímica y Farmacia, Universidad Nacional del Sur,
Bahía Blanca, Argentina
e-mail: ccarcedo@iado-conicet.gob.ar
S. Angeletti
Instituto de Ciencias Biológicas y Biomédicas del Sur, Universidad Nacional del Sur,
CONICET, INBIOSUR, Bahía Blanca, Argentina
G. Zapperi · E. P. Dos Santos
Instituto Argentino de Oceanografía IADO, (Universidad Nacional del Sur-CONICET),
Bahía Blanca, Argentina

© Springer Nature Switzerland AG 2021 179


S. M. Fiori, P. D. Pratolongo (eds.), The Bahía Blanca Estuary,
https://doi.org/10.1007/978-3-030-66486-2_8
180 M. C. Carcedo et al.

sediments due to the close proximity to the bottom (Pearson and Rosenberg 1987;
Mees and Jones 1997). All these organisms, including slow-moving, relatively
mobile and burrowing forms, have a range of effects on the physical structure and
dynamics of the system, contributing as architects of their own habitat (Hansell 2005).
In relation to the feeding habits, protected areas such as estuarine tidal flats are
typically inhabited by deposit feeders, while suspension feeders tend to be the most
abundant group in high-energy environments like exposed sandy beaches (Snelgrove
1999). Dense deposit-feeding macrobenthic populations found on tidal flats sub-
stantially contribute to nutrient cycling by affecting sediment transport processes
through bioturbation activities (Meysman et  al. 2006; Van Colen et  al. 2010).
Moreover, these populations represent important trophic linkages through forage
provision for fishes, birds and other epibenthic invertebrates, on one hand, and by
feeding on benthic algae and bacteria, on the other hand (Van Oevelen et al. 2006).
Reise (1985) showed that large carnivorous fishes and birds had little effect on the
benthic community in sand flats (protected beaches, nearly estuarine environments).
However, small epibenthic predators such as crabs, shrimps and juvenile fishes had
significant effects on the small macrofauna and on juveniles of the larger macro-
fauna. Through this study, Reise (1985) suggested a conceptual hierarchy of bio-
logical processes influencing community structure, pointing out the importance of
biological interactions on tidal flats (Defeo and McLachlan 2005; McLachlan and
Defeo 2018). It is considered that in these environments, the lower limit of the inter-
tidal area is mainly controlled by biotic factors such as the presence of competitors
or predators, whereas the upper limit is mainly governed by abiotic processes such
as immersion period, salinity and desiccation (Paterson et al. 2019).
The sheltered coast of the Bahía Blanca Estuary gives rise to an extensive, gently
sloping and low-energy intertidal area, characterized by salt marshes and unvege-
tated tidal plains. Both environments are inhabited by a diverse community of mac-
robenthic invertebrates, with a great representation of deposit feeders, being
polychaetes, crustaceans and molluscs the most represented groups. Early studies
conducted in the Bahía Blanca Estuary by Elías (1985) and Elías and Bremec (1986)
described for the first time the macroinfaunal associations of intertidal areas, spe-
cifically in four points covering the inner and mid sections of the Principal Channel
(Fig. 2.1; Chap. 2). In the inner area, Puerto Cuatreros (Figs. 2.1 and 2.2; Chap. 2)
was characterized by an intertidal area with salt marshes and mudflats where the
benthic macrofaunal community was dominated by the polychaete Laeonereis
acuta (Fig.  8.1a) in the upper intertidal zone and Leodamas verax (Fig.  8.1b),
Scoletoma (= Lumbrineris) tetraura and Kinbergonuphis sp. in the lower intertidal
zone. Another site studied in the inner section of the estuary was Puerto Galván
(Figs. 2.1 and 2.3; Chap. 2). This site was characterized by the polychaetes
Laeonereis acuta and Eteone sp. occupying the lower and middle zones of the inter-
tidal area, and sometimes replaced by the snail Heleobia australis, mainly in middle
and higher elevations within the intertidal area. The crab Cyrtograpsus altimanus
was distributed along the entire mesolittoral area.
In the mid section of the estuary, in Puerto Rosales (Figs. 2.1 and 2.4; Chap. 2), the
community was dominated by the snail Heleobia australis along the entire intertidal
8  The Intertidal Soft-Bottom Macrobenthic Invertebrates 181

Fig. 8.1  Some of the species found in intertidal soft-bottom areas in the Bahía Blanca Estuary.
The polychaetes (a) Laeonereis acuta and (b) Leodamas verax; the bivalves (c) Tagelus plebeius
and (d) Nucula semiornata; the amphipods (e) Monocorophium insidiosum and (f) Heterophoxus
videns. (Photos by M. Cecilia Carcedo (a, b, d, e, f) and © Guido & Philippe Poppe – www.
conchology.be (c))

area, and sometimes it is replaced by Laeonereis acuta at middle and lower eleva-
tions. Other species found were the priapulid Priapulus tuberculatospinosus, the
snail Buccinanops deformis, the polychaetes Scoletoma tetraura and Ninoe sp., the
clam Tagelus plebeius (Fig. 8.1c) and the crab Cyrtograpsus altimanus. The lower
intertidal level was characterized by several deposit feeder bivalves such as Nucula
182 M. C. Carcedo et al.

semiornata (Fig. 8.1d), Malletia sp. and Pitar rostratus and other organisms such as
the cnidarian Stylatula darwinii. Finally, the fourth site studied was Los Pichones
(38°56′46″S, 62°17′56″W), located far from industries and urban settlements and
therefore considered less influenced by human activities. This site was characterized
by a low intertidal area dominated by polychaetes such as Kinbergonuphis sp.,
Leodamas verax, Ninoe sp. and Axiothella sp. The middle and higher levels of the
intertidal area were exclusively inhabited by the crab Neohelice granulata.
A recent study, carried out in Villarino Viejo, in the inner section of the Principal
Channel (Figs. 2.1 and 2.2; Chap. 2), characterized the community structure and the
spatial zonation of benthic macrofauna in intertidal mudflats. This study mentions
the crab Neohelice granulata as the typical species representative of the upper lev-
els, along with the polychaete Polydora sp. (Zapperi et  al. 2017). Both species
showed a clear seasonal pattern of higher crab densities during spring to early
autumn and higher polychaete density during late autumn and winter. The same
study also remarked the presence of the dominant polychaete Laeonereis acuta in
both sections of the mudflat, with higher abundances in the lower zone. The poly-
chaete Eteone sp. and the priapulid Priapulus tuberculatospinosus, both species
mentioned for the intertidal areas 30 years earlier (Elías 1985), were also found in
the lower mudflat.
The burrowing crab N. granulata is largely considered the most conspicuous
benthic species in tidal flats from the southwestern Atlantic coast (Spivak 2010). It
has a broad distribution in the Bahía Blanca Estuary, occurring on both salt marshes
and mudflats. The term ‘cangrejales’, from the Spanish word ‘cangrejo’ (crab), is
commonly used locally to collectively denote the entire mudflat, due to the high
density of burrows made by these crabs, with records of up to 172 burrows/m2 in
unvegetated areas, and up to 88 burrows/m2 in salt marshes (Angeletti 2017). Crab
burrows have a diameter of up to 12 cm, and they can reach up to 150 cm deep into
the sediment (Escapa 2007). In these environments, N. granulata represents an
important species within the benthic macrofaunal community in relation to its eco-
logical role and its high population density (see Box 8.1).

Box 8.1: The Burrowing Crab Neohelice granulata: A Key Species in


South American Estuaries
Neohelice granulata (Dana, 1851) is a crab from the northeastern coast of
Patagonia, Argentina (42°25′S, 64°36′W), to Rio de Janeiro, Brazil (22°57′S,
42°50′W) (Spivak 2010), with a characteristic burrowing behaviour (Fig. 8.3).
It digs and keeps open semipermanent burrows, allowing it to adopt a semiter-
restrial life. Its presence grants a special physiognomy to the regions where it
is found, called ‘cangrejal’ in Argentina and Uruguay, and ‘caranguejal’ in
Brazil (Spivak 1997). Due to their semiterrestrial habits, the species are active
8  The Intertidal Soft-Bottom Macrobenthic Invertebrates 183

in spring and summer but remain inactive or hidden inside their burrows
throughout most autumn and winter. Consequently, the time available for
feeding and growth should be limited by cold periods (Bas et  al. 2005;
Angeletti and Cervellini 2015).
Regarding the controversy of the systematics of this species, it can be men-
tioned briefly that Rathbun (1918) redescribes the species under the name
Chasmagnathus granulata, which was used during most of the twentieth cen-
tury, despite the disagreement of the scientific community in the masculine
gender of specific names. However, the name Chasmagnathus granulatus has
also been used by some authors (Lozada et  al. 1988). Sakai et  al. (2006)
reviewed and reclassified all species previously attributed to the genera Helice
and Chasmagnathus, redefined both genera and restricted them for the species
of the seas of East Asia. They also introduced a new genus for the species of
South America (Neohelice), New Zealand (Austrohelice) and the tropical and
subtropical Indo-Pacific (Pseudohelice). Therefore, the species is currently
called Neohelice granulata, belonging to the Grapsoidea superfamily and
Varunidae family (Schubart et al. 2002).
This species has a broad feeding spectrum. It is considered omnivorous as
it feeds on both algae and plants as well as polychaetes and small molluscs. It
is also a deposit feeder, since crabs ingest sediments with a mixture of organic
and inorganic matter. Stomach content studies showed traces of plants in
crabs inhabiting marshes, but sediments (along with polychaetes, diatoms and
nematodes) in crabs inhabiting mudflats, thus indicating that crabs act as
deposit feeders in mudflats and herbivores in salt marshes (Iribarne et  al.
1997; Botto et al. 2005; Méndez Casariego et al. 2011b). They can ingest on
the surface and eliminate faeces in depth or vice versa, thus producing an
important vertical mixture of sediments and generating important conse-
quences in the processing of organic matter and nutrient flows (Reise 2002;
Alberti et al. 2011).
Neohelice granulata lives in burrows to protect itself from wave action,
extreme temperatures and desiccation. The burrows also provide shelter from
aerial and terrestrial predators during periods of low tide, and from aquatic
predators at high tide, thus avoiding the stress of finding refuge or escaping
predators. Moreover, burrows are the places where two important events in a
crab’s life occur, moulting and reproduction, and also where young recruits
are safeguarded until they reach larger sizes (Milner et al. 2010; Sal-Moyano
et al. 2012). When building and maintaining their burrows, crabs bring sedi-
ments to the surface, forming mounds near burrow entrances (Murray et al.
2002). This bioturbation affects sediment structure because the cohesive
nature of the organic matrix is disrupted by this process. Apart from directly
affecting sediment porosity and permeability, this process has ecological sig-
nificance in the aeration of soils containing anoxic sediments and in the
184 M. C. Carcedo et al.

distribution of halophytes. Consequently, high densities of active burrowers


can increase erosion rates and sediment mobility (Botto and Iribarne 2000).
As a result of this disturbance in the sediment, Neohelice granulata may
directly or indirectly affect other species that inhabit the same environment.
This interaction can be advantageous for some organisms by promoting their
growth or being harmful. Organisms that are sensitive to disturbance or exca-
vation can suffer high mortality due to abrasion or dispersion (Kneib 1991;
Billick and Case 1994), and predators of the coastal environment can attack
areas where the infauna is exposed (Auster and Crockett 1984). The affected
species include halophytic plants such as Spartina densiflora, Spartina alter-
niflora and Sarcocornia perennis (Bortolus and Iribarne 1999; Perillo and
Iribarne 2003; Minkoff et  al. 2006; Escapa et  al. 2007; Daleo and Iribarne
2009; Alberti et  al. 2008, 2011, 2014); other decapod crabs such as
Cyrtograpsus angulatus and Leptuca uruguayensis (Spivak et al. 1994, 1996;
Botto and Iribarne 2000; César et al. 2005; Méndez Casariego et al. 2009);
polychaetes such as Laeonereis acuta and Heteromastus similis (Palomo et al.
2003; Escapa et al. 2004b; Rosa and Bemvenuti 2005); clams such as Tagelus
plebeius (Gutiérrez and Iribarne 1998; Lomovasky et al. 2006); birds such as
Calidris fuscicollis, Charadrius falklandicus, Pluvialis squatarola,
Haematopus palliatus, Larus atlanticus and Tringa melanoleuca (Botto et al.
1998; Palomo et al. 2003; Iribarne et al. 2005; Berón et al. 2011); and rodents
such as Akodon azarae and Oligoryzomys flavescens (Canepuccia et al. 2008),
among others.
Therefore, the presence of the burrowing crab Neohelice granulata has
important implications for the functioning of the estuarine ecosystem, being
of great importance in the study of biological and ecological problems. In
addition, the species is being used simultaneously for toxicology, biochemis-
try, neurophysiology and neurobiology and behaviour studies, becoming one
of the most studied crab species, and considered as an emergent animal model
for biological research at a global scale (Spivak 2010). At a local scale,
N. granulata was declared as an ‘emblematic’ species by the Honorable
Deliberative Council of Bahía Blanca city, according to ordinance N° 12.671
(2004). Other species reached by this ordinance were the olrog’s gull (Larus
atlanticus), the silver or Franciscan dolphin (Pontoporia blainvillei), the
sharks (Carcharias taurus, Carcharhinus brachyurus) and migratory birds.
The main objective of the ordinance was to recognize the species and through
them to be able to educate the community and foster local identity.
8  The Intertidal Soft-Bottom Macrobenthic Invertebrates 185

Fig. 8.2  Association between salt marshes and benthic macrofauna. (a) The burrowing crab
Neohelice granulata in salt marshes of Spartina alterniflora, (b) burrow of Neohelice granulata in
salt marshes of Sarcocornia perennis, (c and d) the hydrobiid snail Heleobia australis. (Photos by
Sandra M. Fiori (a), Sabrina Angeletti (b) and Walter Nievas (c, d))

Fig. 8.3  Individuals of the


burrowing crab Neohelice
granulate around a burrow
entrance in mudflats of the
Bahía Blanca Estuary.
(Photo by Sabrina
Angeletti)
186 M. C. Carcedo et al.

8.2  A
 ssociations Between Salt Marshes
and Benthic Macrofauna

Salt marshes are highly productive systems, and therefore the biomass exported from
these environments is often considered the basis of estuarine food webs (Chalmers
et al. 1985; Day et al. 1989). They are recognized worldwide for the various ecosys-
tem services they provide, including atmospheric carbon sequestration (Chmura
et al. 2003), reduction of eutrophication through nutrient uptake (Sousa et al. 2008)
and metal sequestration (Weis and Weis 2004). In addition, marsh plants interact
with benthic nutrients, modifying the nitrogen and phosphorus cycles (Pedersen
et al. 2004). Their own presence also reduces the water energy with the consequent
resistance to surface erosion. These environmental changes can have, in many cases,
strong effects on the macrobenthic community structure. Jones et al. (1994) define
the ecosystem engineers as the organisms that directly or indirectly modulate the
availability of resources (other than themselves) to other species, by causing physical
state changes in biotic or abiotic materials. By doing so, they can modify, maintain
and/or create habitats. A way the salt marshes can act over community is ameliorat-
ing harsh physical conditions and therefore providing a less stressful microenviron-
ment, e.g. the release of oxygen with the consequent changes in the redox potential
and the pH in the rhizosphere (Pedersen et al. 2004), the reduction of thermal stress,
favouring the establishment of other organisms by shading (Callaway 1995; Bortolus
et al. 2002) and the availability of refuge (Hovel et al. 2001; Lewis and Eby 2002)
and food availability (Kneib 1984). Therefore, salt marsh plants can have large
effects on species distributions in intertidal communities. In the Bahía Blanca
Estuary, salt marshes are distributed along all the margins of channels and on all the
islands. The dominant salt marsh plant is the smooth cordgrass Spartina alterniflora,
a species recently reported as an invasive species for South America (Bortolus et al.
2015). Other important salt marsh plants in the area are Spartina densiflora and
Sarcocornia perennis, both species typical of substrates with longer air exposure,
which are therefore more restricted to the supralittoral areas.

8.2.1  T
 he Interaction Between Salt Marsh Plants
and the Burrowing Crab Neohelice granulata

The amelioration of harsh physical conditions by plants plays an important role in


dictating the abundance and distribution patterns of the crab Neohelice granulata.
In Mar Chiquita Lagoon, located in the north coast of Buenos Aires Province (Fig.
1.2; Chap. 1), a strong relationship between plant cover and the spatial distribution
of N. granulata was registered in the high marsh, where the spatial distribution of
these crabs strongly overlapped the distribution of both Spartina densiflora and
Sarcocornia perennis (Fig. 8.2a, b). Plant cover drives the spatial distribution of the
dominant macroinvertebrate through facilitation, by buffering crabs from harsh,
8  The Intertidal Soft-Bottom Macrobenthic Invertebrates 187

stressful environmental conditions. However, this relationship was not observed in


the low marsh, where crabs established in both vegetated and unvegetated areas.
This suggests that environmental conditions are more stressful on bare than on veg-
etated substrates but only in the high marsh, where stressful temperature and dehy-
dration are important factors controlling the establishment and permanence of crabs
(Bortolus et al. 2002). N. granulata density decreases towards the upper intertidal
zone, as noted by Bortolus and Iribarne (1999). However, the colonization of the
high marsh by N. granulata is facilitated by the presence of plants. Sarcocornia
perennis is a good colonizer of high and saline intertidal areas. It generates a shad-
ing area allowing the sediment to stay wetter, therefore softer and more likely to be
excavated. It also dampens various environmental factors, such as desiccation and
high surface temperatures (Bortolus et al. 2002; Angeletti 2012).
The herbivorous habit of Neohelice granulata in marshes can determine the
lower limit of the distribution of Spartina densiflora (Alberti et al. 2010). In addi-
tion, burrows of N. granulata can increase nutrient circulation and promote the
production of S. densiflora through increased oxygenation that facilitates the growth
of a fungus that colonizes its roots (Daleo et al. 2007; Daleo and Iribarne 2009).
Regarding the herbivorous habit of the burrowing crab, it has been shown that it
preferentially consumes S. densiflora more than Sarcocornia perennis (Costa et al.
2003). A complex process generated by the biological interaction between N. gran-
ulata and S. perennis has been described by Perillo and Iribarne (2003) in the inner
section of the Bahía Blanca Estuary. Due to the crab’s burrowing activity, plants
grow by generating ring-formed patches from 1.5 to 8 m in diameter. Plant clones
concentrate at the edges, while the inner sector is dominated by crab burrows. This
interaction facilitates the erosion of the surface, arising from a decreased soil resis-
tance due to the presence of burrows, and induces the formation of salt ponds in salt
marshes. As a result of continuous sediment removal by crabs, the ponds remain
sunken and accumulate water (even at low tide). These circular shallow ponds are
continuously expanding in size and can join each other, forming different configura-
tions. Water trapped in ponds can drain to nearby areas or pre-existing tidal chan-
nels (Escapa 2007).

 he Joint Distribution of Spartina alterniflora


8.2.2  T
and the Snail Heleobia australis

Another species frequently found in salt marshes where cordgrasses such as Spartina
spp. are the dominant plant is the hydrobioid snail Heleobia australis (d’Orbigny,
1835) (Fig. 8.2c, d). This snail is a generally dominant species in abundance and
biomass constituting a food item of relevance for fishes and crustaceans. In turn,
since it incorporates the organic matter present in the sediments through its diet, it
is a fundamental link in the nutrient cycle through coastal food webs (Figueiredo-­
Barros et al. 2006). Heleobia australis has a broad geographical distribution from
188 M. C. Carcedo et al.

Rio de Janeiro, Brazil (22°54′S), to San Antonio Oeste, northern Argentinean


Patagonia (40°84′S) (Aguirre and Farinati 2000), and it is mostly associated to mud,
high salinity and low-energy coastal areas (Canepuccia et al. 2007). In the Bahía
Blanca Estuary, it represents the only hydrobioid snail, being one of the most abun-
dant species of intertidal mudflats and salt marshes in this ecosystem (Elías et al.
2004; Carcedo and Fiori 2011).
At the microscale, its spatial distribution is determined by a set of complex
behaviours that allow the active selection of areas with certain physical and chemi-
cal qualities. The adults of H. australis are capable of floating away from the sedi-
ment itself by creating a gas bubble inside its shell. This behaviour was observed by
Little and Nix (1976) for Hydrobia ulvae and then by Echeverría et al. (2010) for
H. australis as a dispersion strategy. Through this mechanism, adults can temporar-
ily enter the water column, and be carried along by the tide- and wind-driven cur-
rents, being able to escape from excessively stressful situations in search of better
environmental conditions. Indeed, within intertidal areas of the Bahía Blanca
Estuary, individuals of H. australis can actively select vegetated areas of Spartina
alterniflora, resulting in a joint distribution between both species (Canepuccia et al.
2007; Carcedo and Fiori 2011). The explanation of this distribution pattern could be
related to different processes, and Canepuccia et  al. (2007) evaluate different
hypotheses. One of them was related to the capability of S. alterniflora stems to
generate microhabitats that favour the establishment of species by reducing preda-
tion rates (Lewis and Eby 2002), but authors conclude that predation seems to be a
weak force to explain the association between species. Another hypothesis was
related to the feeding preferences of the snail, assuming that S. alterniflora indi-
rectly provides food resources (as epiphytes) to the snails. Stable isotopic analysis
shows that neither S. alterniflora nor its epiphytes directly contribute to the diet of
H. australis, being the sediment its main food source. Therefore, the association
was neither explained by this hypothesis. Canepuccia et al. (2007) finally suggest
that the distribution and survival of H. australis in marshes could be regulated by
the temperature and the dehydration stress (Hovel et al. 2001). This conclusion was
based on the observations that H. australis established more in experimental shaded
areas and also on the higher mortality registered under more intense sun irradiance
in bare areas. Therefore, the reduction of temperature provided by the shading of
S. alterniflora stems leads to the selection of these microhabitats by the snails.
Although H. australis typically inhabits soft bottoms, it is capable to colonize
hard substrates and reefs. The hard substrates of the Bahía Blanca Estuary show, in
most cases, intermediate densities of H. australis between salt marshes and tidal
flats (Carcedo and Fiori 2011). Given the structural complexity of the rocks, these
areas could give certain advantages for snails similar to what occurs in reefs of the
Pacific oyster, Magallana gigas (ex Crassostrea gigas) in Anegada Bay, Argentina
(Fig. 1.2; Chap. 1) (Escapa et al. 2004a), and in tubes of the polychaete Ficopomatus
enigmaticus in Mar Chiquita (De Francesco and Isla 2003; Luppi and Bas 2002).
The unvegetated tidal flats were the areas less selected by H. australis. Their high
conductance, reflectivity and caloric capacity could make these areas reaching
8  The Intertidal Soft-Bottom Macrobenthic Invertebrates 189

higher temperatures than those with vegetation cover or with shaded spaces within
hard substrates.
The active selection of microenvironments, described for adults of H. australis,
was also suggested for juveniles and recruits. Snails of different sizes are differen-
tially distributed across microenvironments in intertidal areas of the Bahía Blanca
Estuary (Carcedo and Fiori 2011). Salt marshes of S. alterniflora were the only
microenvironments where individuals smaller than 2.5 mm were found. In addition,
juveniles and sub-adult snails are also able to colonize the hard substrates, while
adults are distributed in all microenvironments. A similar behaviour was observed
in Mar Chiquita Lagoon by the crab Cyrtograpsus angulatus, whose adults and
juveniles reach high densities in muddy substrates, while its megalops, recruits and
small juveniles are almost exclusively found in reefs of the polychaete F. enigmati-
cus, due to their importance as refugees against predation and cannibalism (Spivak
et al. 1994; Luppi and Bas 2002). Although the hypothesis of dehydration stress was
not tested between snails of different sizes, the effect of higher temperatures on the
snails could be exacerbated in juveniles and recruits, leading to the observed size
distribution between microenvironments.

8.3  Processes Modulated by Local Macroinfauna

Benthic organisms have a direct relationship with the substrate, and their interac-
tions with the bottom environments generate both stabilization and destabilization
of sediments. Some benthic microalgae excrete extracellular polymeric substances
and build an organic matrix that binds sediment grains forming the so-called bio-
films, recognized for their stabilizing function (Decho 2000; Cuadrado et al. 2011).
Macroinfauna, on the other hand, can promote the destabilization of cohesive sedi-
ments through the production of pellets and the formation and maintenance of dif-
ferent types of constructions, such as burrows and caves (Herman et  al. 1999;
Braeckman et al. 2011). In addition to causing a physical disturbance, local macro-
infauna can change the characteristics of the sediment affecting the porosity, perme-
ability, resistance to erosion and its mobility (Botto and Iribarne 2000; Escapa et al.
2008). Thus, both bio-stabilizing and bio-destabilizing organisms influence the
properties of intertidal sediments, significantly affecting geomorphology (Widdows
et al. 2000; Murray et al. 2002).
Macroinfaunal activity can have considerable effects on the structure of both ter-
restrial and marine sediments. Its importance was first demonstrated by Charles
Darwin in 1881, who dedicated his latest scientific book to this subject: The
Formation of Vegetable Mould, Through the Action of Worms, with Observations on
Their Habits. The content of this book focused on sediment alterations caused by
the activity of burrowing organisms, especially by worms, a process that was later
called ‘bioturbation’ (Richter 1952). Darwin was the first person to realize that bio-
logical activity at the local level, carried out by small invertebrates, could have
consequences on a landscape scale, due to its influence on processes that model
190 M. C. Carcedo et al.

geomorphology. Numerous studies of a variety of disciplines, including ecology,


edaphology, hydrology, geomorphology and even archaeology, cite Darwin’s book
as the original reference.
Macroinfauna in marine ecosystems modify the environment at different spatial
and temporal scales. Many of these modifications are initially at a microscale, but
they are likely to have large-scale effects on benthic and pelagic seascapes (Meadows
et al. 2012). Tidal flats and salt marshes are habitats of a great variety of macroin-
fauna, subject to abiotic and biotic processes that actively shape them and define
their functions. These organisms, through their different ways of life and feeding
strategies, modify the size of the particles and the chemistry of the soil (Cadée 2001;
Méndez Casariego et al. 2011a).

8.3.1  The Bioturbation

In modern ecological theory, bioturbation is recognized as an archetypal example of


ecosystem engineering, capable of modifying geochemical gradients, redistributing
food resources, viruses, bacteria, a variety of larval stages and eggs. From an evolu-
tionary perspective, it has been shown that bioturbation played a key role in the
evolution of metazoan life at the end of the Precambrian era (Meysman et al. 2006).
In recent decades, bioturbation has been rediscovered as an important factor in the
evolution of the earth’s surface, through its influence on soil formation, erosion and
horizon stability (Dietrich and Perron 2006; Kristensen et al. 2012).
In sediments of estuarine environments, oxygen penetration and aerobic decom-
position are limited to the first surface millimetres, and the deeper layers are char-
acterized by anaerobic decomposition processes (Revsbech et al. 1980; Kostka et al.
2002). However, the bioturbation activity of the fauna and roots of plants can
increase the depth of the sediment’s oxic layer, favouring aerobic processes (Fanjul
et al. 2015). In environments of low or intermediate energy, such as the intertidal
areas of the Bahía Blanca Estuary, bioturbating macrobenthic fauna is the main
agent for mixing sedimentary particles (Rhoads and Young 1970; Jones and Jago
1993; Botto and Iribarne 2000). It has been proven that the activity of macrobenthic
invertebrates that build and maintain their burrows affects the oxygenation, geo-
chemistry and transport of particulate materials in the sediment column, mixing the
upper layers with deeper ones, in a few days (Botto et al. 2006; Fanjul et al. 2007).
The burrowing and semiterrestrial crab Neohelice granulata actively and con-
stantly build its burrows in the intertidal zone during low tide. Crabs dig and keep
open semipermanent burrows, allowing this invertebrate to adopt a semiterrestrial
life. In the Bahía Blanca Estuary, burrows are tunnel-shaped, with ramifications and
multiple entrances, which is directly associated with the type of substrate on which
the burrows are built. However, there are differences in the structural morphology of
burrows and the burrowing activities in contrasting microhabitats (e.g. salt marshes
and mudflats), indicating that this species has an adaptive burrowing behaviour
(Escapa et al. 2007; Angeletti et al. 2018a). Burrows from salt marshes have narrow
8  The Intertidal Soft-Bottom Macrobenthic Invertebrates 191

entrances and deep tunnels, and burrows from mudflats present very narrow
entrances and surface tunnels. Crabs would change the depth of their burrows in
relation to water table depth, to reach groundwater levels. In this way, the burrows
would contain water throughout the tide cycle, an adaptation to maintain the humid-
ity in high marsh areas (Iribarne et al. 1997; Escapa et al. 2007). These multiple
entrances may provide easy access to shelter, especially in unvegetated mudflats. In
addition, there is evidence suggesting that a larger number of burrow entrances in
intertidal mudflats increases the ability to entrap organic particles that may serve as
food (Iribarne et al. 1997; Botto et al. 2006). In salt marshes, chambers were discov-
ered inside tunnels, possibly used for reproductive purposes (Angeletti et al. 2018a).
Neohelice granulata is commonly considered as an ecosystem engineer due to its
great burrowing activity and its potential effect on intertidal sediment characteris-
tics. When building and maintaining their burrows, crabs bring sediments to the
surface forming biogenic mounds near burrow entrances (Murray et al. 2002). This
bioturbation affects sediment structure because the cohesive nature of the organic
matrix is disrupted by this process. Besides affecting sediment porosity and perme-
ability, this process has ecological significance in the aeration of soils containing
anoxic sediments and in the distribution of halophytes. Consequently, high densities
of active burrowers can increase erosion rates and sediment mobility (Botto and
Iribarne 2000).
Studies developed in the Bahía Blanca Estuary showed that biogenic mounds
removed by crabs did not have a specific orientation. No general pattern was
observed in their position in relation to the mouth of the burrow or to the location
relative to the subtidal zone (Angeletti 2017). Regarding the penetrability of sedi-
ments, i.e. the force necessary to penetrate the substrate, biogenic mounds had
higher water content and penetrability than sediments not removed by crab’s activ-
ity (Escapa et al. 2004b; Angeletti et al. 2018a). This trend has been demonstrated
by controlled laboratory experiments, carried out by Escapa et al. (2007), where the
inclusion of crabs modified the sediment penetrability. After the bioturbation activ-
ity, the average force required to penetrate the sediments was reduced to less than
20% of that of the control sediments. However, no differences were found in the
organic matter content between biogenic mounds and control sediments, because
remixing keeps organic matter homogeneously dispersed in the sedimentary col-
umn (Bortolus and Iribarne 1999; Gutiérrez et al. 2006). The activity of Neohelice
granulata transports sediments with a large amount of silt and clay from deep layers
to the surface, recovering these inorganic materials along with the associated organic
matter, and making them available at the surface and the water column (Escapa
et al. 2007; Angeletti et al. 2018a).
Bioturbation by Neohelice granulata interacts with local physical processes and
hydrodynamic conditions and can affect a wide variety of sedimentological param-
eters. Biologically induced roughness modifies hydrodynamics above the sediment
layer, which in turn affects erosion and resuspension (Boudreau and Jørgensen
2001). Even in coastal systems, which are traditionally seen as modelled only by the
physical forces of currents and waves, hydraulic engineers have recognized that
bioturbation is a crucial component in sedimentary dynamics models (Murray et al.
192 M. C. Carcedo et al.

2002; Paarlberg et al. 2005). Several studies have linked biological data on biotur-
bation to adjust parameters in bioturbation models (Gilbert et al. 2003; Solan et al.
2004; Jarvis et al. 2010; Schiffers et al. 2011).
The suspended sediment contribution arising from the bioturbation activity of
N. granulata in intertidal sites of the Bahía Blanca Estuary was analysed by
Angeletti et al. (2018b) using several approaches, ranging from field experiments to
numerical modelling. It was found that crabs from mudflats remove, trap and erode
more sediment per unit area than crabs from salt marshes, due to their higher popu-
lation density and the mobility of cohesive sediments. Results obtained through
MOHID (water modelling system) simulations showed that the sediments in the
inner section of the estuary were maintained in the water column much longer than
sediments from the middle zone. This longer residence time in the inner area could
be attributed to the geomorphological and hydrodynamic characteristics of this sec-
tion of the estuary, where numerous tidal channels coexist and phenomena of ‘reten-
tion’ occur before entrance into the main channel. By contrast, in the middle section
of the estuary, the sediments are affected by greater water depth and higher tidal
current speeds. In addition, wind-generated waves can be a determining factor in the
spatio-temporal evolution of the bioavailable sediment in the water column.

8.3.2  The Benthic-Pelagic Coupling

Benthic-pelagic coupling is the process which links bottom sediments and the water
column in aquatic systems and reveals as the exchange of energy, mass or nutrients
(Griffiths et al. 2017). This phenomenon plays a major role (along with terrigenous
input and upwelling) in determining the production and biological structure of these
aquatic habitats (e.g. Sommer 1989; Valiela 1995). Estuaries are commonly shallow
ecosystems, where the occurrence of strong coupling of matter and energy between
the water column and the sediment is enhanced. Their strong horizontal gradients of
salinity, water current velocity, turbulence and turbidity affect the coupling between
ecological and biogeochemical processes in the water column and the sediment
(Burdige 2006; Griffiths et  al. 2017). This interaction highly influences nutrient
recycling and overall productivity of the ecosystem.
Processes in the benthic boundary layer are key components in the dynamics of
coastal ecosystems (Chauvaud et al. 2000; Dale and Prego 2002; Marcus and Boero
1998). This is the layer in which the exchange of materials between sediments and
the water column takes place and where biological communities may influence
these fluxes through bioturbation, excretion and respiration, as well as organic
deposition (Norkko et al. 2001). Benthic and pelagic communities, therefore, can
significantly modify physical-chemical properties of sediments, changing redox
conditions and enriching the sediments with organic matter.
The fate of the organic matter produced in the water column depends on many
physical-chemical and biological factors, such as currents, salinity, temperature,
8  The Intertidal Soft-Bottom Macrobenthic Invertebrates 193

algal density and aggregate formation (Alldredge and Jackson 1995; Kiørboe and
Jackson 2001). However, the shallowness of the environment strengthens benthic-­
pelagic interactions. The sink of phytodetritus and deposition of pelagic material
may increase the organic matter input into the benthic habitat, inducing microbial
remineralization. In response, the bottom sediments will contribute dissolved nutri-
ents to the water column, and it is estimated that 30–80% of the nitrogen required in
shallow coastal waters is supplied by benthic release (Blackburn and Henriksen
1983; Boynton and Kemp 1985; Nixon 1981). The most important productivity
event in the pelagic food webs is phytoplankton blooms (Legendre 1990) which
also have a large influence on benthic nutrient release or regeneration (Hansen and
Blackburn 1992; Jensen et al. 1990; Overnell et al. 1995), as well as in their fast
consumption up to depletion (Boyer et al. 2009; Bricker et al. 2008).
Ecosystem dynamics in the Bahía Blanca Estuary is dominated by a winter-early
spring phytoplankton bloom that commonly occurs in the shallow waters of the
inner zone. Studies carried out along 30  years have demonstrated that the lower
turbidity observed in late autumn-early winter is a repetitive pattern which has been
previously associated with the start of the phytoplankton bloom (Guinder 2011;
Guinder et al. 2009). In this shallow environment, the sink of phytoplankton cells
may represent an important contribution of organic matter to the benthic communi-
ties, and nutrient regeneration in the surface sediments may contribute significantly
to the pelagic environment. Except for isolated observations of material settlement
during the blooming season, few studies were performed regarding benthic-pelagic
coupling in the area. Guinder et  al. (2015) have indirectly addressed the subject
when they evaluated the exportation of particulate suspended matter to the benthos
through the analysis of settled material inside sediment collectors. Their results
showed that the sedimentation of viable and large diatoms may represent an impor-
tant source of allochthonous carbon for the benthic community.
The evaluation of nutrient fluxes using microcosm experiments revealed impor-
tant variations in flux magnitude and direction that were related to the phytoplank-
ton dynamics (Zapperi et  al. 2016). Moderate release of nitrite and nitrate
(NO3– + NO2−) from the sediment to the water column was observed in the incuba-
tion experiments performed for the upper and lower zones of the intertidal mudflats.
This flux towards the water column represents a vital input after depletion of these
nitrogen (N) species during the post-phytoplankton bloom period. However, the
most dramatic response registered in the incubation experiments was ammonium
(NH4+) recovery. After the chlorophyll-a peak during the post-bloom stage, the
lower mudflats showed extremely high rates of NH4+ regeneration, whereas the
upper mudflats continued the uptake trend they showed during pre-bloom and
bloom stages. Low macrofaunal densities were recorded in the upper mudflats dur-
ing that same period, while lower mudflats showed high densities of the polychaete
Laeonereis acuta. Breakdown of organic matter to produce NH4+ is a major mecha-
nism of N remineralization (McGlathery et al. 2004), and the results suggest a key
role of benthic fauna in N recycling, as well as in the rapid recovery of NH4+ levels
in the water column.
194 M. C. Carcedo et al.

In the Bahía Blanca Estuary, macrobenthic communities in the upper and lower
mudflats show differences in terms of their species composition as well as their pat-
terns of activity throughout the year. Some common species in the upper mudflats,
like Neohelice granulata and Polydora sp., are virtually absent in lower mudflats,
whereas frequent species in lower mudflats, like Laeonereis acuta and Eteone sp.,
appear at much lower abundances in upper mudflats (Zapperi et al. 2016). In the
same study, it was observed that lower mudflats showed maximum faunal densities
during the chlorophyll-a peak, right after the bloom, whereas the upper mudflats
recovered macrobenthic activity several months after the phytoplankton bloom.
Biological activity of benthic macrofauna is expected to have a large effect on ben-
thic fluxes (Hughes et al. 2000; Michaud et al. 2006; Webb and Eyre 2004), and it is
reported that bioturbation would enhance benthic fluxes up to one order of magni-
tude (Burdige 2006), which can be related to the NH4+ recovery patterns
described above.
Another clear representation of the coupling, mediated by biota, is the variable
abundance of benthic macrofauna through the year. Neohelice granulata showed a
minimum activity in June, at the same time that Laeonereis acuta attained maxi-
mum density, in agreement with the lowest turbidity registered. Erosion rates in the
study area are closely related to bioturbation activity by the burrowing crab N. gran-
ulata (Escapa et al. 2007). On the contrary, polychaetes might contribute to sedi-
ment stability, reducing susceptibility to physical resuspension (Mortimer et  al.
1999). Studies have suggested that feeding activity of L. acuta promotes sediment
stability by deposition of cohesive, organically enriched sediments on the surface
(Palomo and Iribarne 2000). On the other hand, light limitation has been proposed
as a major control of winter bloom development in the Bahía Blanca Estuary
(Guinder et al. 2009) because of its elevated turbidity. Hence, sufficient light pene-
tration in the water column, promoted by the turbidity decrease associated to
reduced N. granulata activity, is a potential trigger for phytoplankton cell prolifera-
tion (Cloern 1987; Irigoien and Castel 1997; May et al. 2003).
Summarizing, benthic-pelagic coupling plays a key role in determining produc-
tion and biological structure in the Bahía Blanca Estuary. Its importance is reflected
in how benthic-pelagic coupling mediated by biological activity plays a significant
role in creating the window of lower turbidity that allows a phytoplankton bloom,
which sustains the high productivity of this ecosystem. As the bloom develops, dis-
solved nutrients in the water column are consumed, and organic matter is produced.
As a counterpart, the local coupling between remineralization in the surface sedi-
ments and benthic flux to the water column as dissolved nutrients allows the recov-
ery of nutrient levels and supports primary production in the forthcoming cycle.

8.4  N
 ew Registers of Benthic Macrofaunal Species During
the Last Decade and Their Implications

Because of the vast surface of the intertidal area in the Bahía Blanca Estuary and the
instability of its soft sediments, it is a largely unexplored environment. In addition,
the great majority of the animals living in intertidal flats have a cryptic lifestyle,
8  The Intertidal Soft-Bottom Macrobenthic Invertebrates 195

remaining hidden in the muddy bottoms inside their burrows in deeper layers of
sediments and emerging only when necessary to feed or to perform other vital func-
tions (e.g. crabs, polychaetes). Some species even spend their whole lifespan inside
their burrows (e.g. pholadid clams), which are recognizable through small openings
visible on the surface. Relatively little is known about the biology and ecology of
benthic macrofaunal invertebrates in intertidal areas of the Bahía Blanca Estuary,
and the recent discovery of newly established populations supports this assumption.
The most important findings are developed in the next paragraphs together with a
brief discussion of the feasible causes and consequences of their presence and the
processes that could be implicated in the extension of their distribution ranges.

8.4.1  H
 idden in the Mud: The Exotic Piddock
Barnea truncata

The family Pholadidae, commonly known as piddocks and angel wings, are repre-
sented by bivalve species that build tubular burrows by using their shell to mechani-
cally erode the substratum. The Atlantic Mud-piddock Barnea truncata that belongs
to this family is restricted to Atlantic Ocean margins. In the western Atlantic, B. trun-
cata is distributed from 24°S to 45.4°N, from southeastern Brazil (west of São
Paulo) to northern USA (Maine). In the eastern Atlantic, it has been recorded from
15°N to 34°S latitude, from Senegal to South Africa (COSEWIC  2009;  Hebda
2011). During 2010, individuals of B. truncata were found in the intertidal zone of
the Bahía Blanca Estuary (Fiori et  al. 2012), and this finding represents the first
record of the species at more than 2000 km south from the edge of its native range
in South America. The native species Barnea lamellosa, with a distribution limited
to the coasts of Uruguay and Argentina (Turner 1954), was also registered in the
same intertidal area where B. truncata was found, but in a lower number than its
exotic congener.
As for many of the non-indigenous invasive species registered in the Bahía Blanca
Estuary, the accidental introduction through ballast water from ships was proposed
as the most probable entry route to the ecosystem. Fiori et al. (2012) indicate several
assumptions that support this hypothesis; one was related to the location where the
population of B. truncata was found, in the intertidal zone near Ingeniero White
(Figs. 2.1 and 2.2; Chap. 2), a port with heavy maritime traffic. Another assumption
was related to the long permanence of their larvae in the water column as indicated
by Turner (1954) and Chanley (1965), and finally due to its absence in intermediate
locations between its southern native range and the Bahía Blanca Estuary. However,
Fiori et al. (2012) also suggest that the natural expansion of this species is a possibil-
ity that should be also considered. The lack of records in intermediate locations
could be associated with the cryptic lifestyle of this species, which further restricts
the chances of being discovered. This was the case at the northern limits of its distri-
bution range, where a single population was detected in Nova Scotia, approximately
350 km north and east of the nearest record in Maine (Hebda 2011).
196 M. C. Carcedo et al.

Invasive bivalve species usually become dominant in their communities, and


therefore, bivalve activities such as shell production, filter feeding and bioturbation
can markedly affect ecosystem structure and functioning. In the case of burrowing
bivalves, their ecosystem effects include possible changes in grain size, organic
matter and nutrient concentrations via bioturbation, as well as an increase in oxygen
penetration to deep sediment layers because of deep burrowing (Sousa et al. 2009).
In addition, the surfacing shells can cause changes in hydrodynamic conditions and
an increase in substrate complexity by creating a diversity of new microhabitats
(Gutiérrez et al. 2003), in the form of available substrates for epibionts as well as
shelter to other organisms in the interstices between their shells. However, the spe-
cific effects of a non-indigenous species on the environment and in the local biota
should be evaluated. Further studies are needed to understand the consequences of
the presence of B. truncata in the intertidal area of the Bahía Blanca Estuary, espe-
cially in relation to its engineering activities.

8.4.2  A
 nother Crab in the Estuary: The Fiddler Crab
Leptuca uruguayensis

Fiddler crabs are widely distributed throughout the tropics and subtropics of the
world, especially in South America, where they form dense populations in silty or
silt-sandy ecosystems, like sheltered bays and estuaries (Spivak et al. 1991), where
they construct burrows in the high levels of the intertidal area (Crane 1975). Adults
of fiddler crabs are generally small, with only a few species exceeding 25 mm in
carapace width. The most notable feature of these crabs is the sexual dimorphism of
the chelae: adult males developed an enlarged cheliped, not present in females or in
the early stages of males. This cheliped is developed during the period of sexual
maturation, being useful for combat and territorial defence (Yamaguchi and Henmi
2008), and its red vibrant colouration, acquired during the reproductive season, is
useful to define territory and to attract females (Crane 1975). It is the waving of this
large chela that gives the popular name of ‘fiddlers’ to these crabs.
The southwestern Atlantic fiddler crab Leptuca uruguayensis (Nobili, 1901) is
distributed from southern Brazil (33°S) to the northern coast of Argentina (38°S). It
is restricted to the intertidal zones of sheltered bays and estuaries, which results in
a puzzling distribution. One of the southernmost permanent population inhabits
Samborombón Bay (Fig. 1.2; Chap. 1), in the southern limits of the Río de la Plata
Estuary (35°30′–36°22′S), where it is the dominant intertidal species, reaching den-
sities of up to 140 crabs per m−2 (Boschi 1964; Iribarne and Martínez 1999). Other
populations are found in the northern coast of Buenos Aires Province at the Mar
Chiquita Lagoon, at the Quequén Grande Estuary and at the Quequén Salado
Estuary (Spivak et al. 1991; Boschi et al. 1992) (Fig. 1.2; Chap. 1), but with lower
densities than those found in Samborombón Bay. In fact, a pattern of decreasing
8  The Intertidal Soft-Bottom Macrobenthic Invertebrates 197

density with increasing latitude was proposed for the species, over a land distance
of less than 300 km.
During February 2019, Truchet et al. (2019) reported for the first time the pres-
ence of Leptuca uruguayensis in the Bahía Blanca Estuary, in the upper intertidal
zone of an unvegetated tidal flat in Puerto Rosales (Fig. 2.4; Chap. 2), at approxi-
mately 200 km of its southern record. The species was registered in a density of
144  ±  48 individuals/m2, higher than those found in the Quequén Grande and
Quequén Salado Estuaries and close to those reported in northern locations like the
Samborombón Bay, where they reach 142 individuals/m2 (César et al. 2005) and
San Antonio Cape (the southern extreme of Samborombón Bay) where densities
reach 133 to 207 individuals/m2 (Torres Jorda and Roccatagliata 2002). However,
these studies report populations that persist over extensive areas, unlike the popula-
tion inhabiting the Bahía Blanca Estuary, which encompasses only in a small patchy
distribution (Truchet et al. 2019).
Bogazzi et al. (2001) suggest that the southern distribution limit of L. uruguay-
ensis is mainly due to wind patterns. They indicate that during the period of larval
release and recruitment, the contribution of wind drift transport is less frequent
coastwards from northern to southern locations. Given that the wind is the main
force moving the upper layer of the water column and consequently crab larvae,
authors expect a decrease in the chances of carrying megalopae to southern loca-
tions and that develop into adults. The presence of adults of L. uruguayensis in the
Bahía Blanca Estuary indicates that crab larvae effectively arrived in the coast. The
high density found in a small patchy distribution together with the scarce variation
in carapace width suggests that these individuals may come from a batch of larvae
that reached the coast at once (Truchet et al. 2019).
A possible explanation of the presence of L. uruguayensis in the Bahía Blanca
Estuary could be the introduction of larvae into the estuary via ship ballast water
(Truchet et al. 2019). The movement of commercial and transport ships and boats,
exacerbated by the increased globalization of commerce, has facilitated the spread
of species. The most studied cases are those related to invasive species; neverthe-
less, some cases correspond to native species that found new places to settle accord-
ing to their biological requirements. Another possible explanation could be related
to the natural shift in species distributions to higher latitudes. Predicted biological
and ecological responses to rising temperatures in the oceans are expected to cause
shifts in species distribution ranges (Hoegh-Guldberg and Bruno 2010; Burrows
et  al. 2014; García Molinos et  al. 2015; Sunday et  al. 2015). Particularly, in the
southwestern Atlantic Ocean, warming-favoured species could benefit from future
climate change scenarios and expand southwards their geographical distribution.
Future studies will be necessary to analyse the permanence of this population in the
Bahía Blanca Estuary, especially in relation to its viability and the possible expan-
sion to other sites within the estuary.
198 M. C. Carcedo et al.

8.4.3  The Ephemeral Polyp Corymorpha januarii

Dense populations of polyps have been observed by local fishermen for more than
20 years in the margin of muddy-bottom channels of the Bahía Blanca Estuary. This
information remained unnoticed by local scientists until November 2013, when pol-
yps collected in a small channel of the estuary were analysed for identification pur-
poses (Dutto et al. 2019). The species was Corymorpha januarii Steenstrup, 1855
(Cnidaria: Hydrozoa), a corymorphid hydroid native from the southwestern Atlantic
coasts. It lives in tropical and temperate shallow waters from Rio de Janeiro (Brazil)
to Puerto Madryn (northern Argentinean Patagonia, 42°48′S), being endemic to this
region (da Silveira and Migotto 1992; Genzano et al. 2009). The channel where the
polyps were collected is called El Saco (Fig. 2.2; Chap. 2), a small and shallow
NW-SE oriented channel (3 km long, 110 m wide and a maximum of 5 m depth in
high tide) usually used for artisanal fishing, located in the inner section of the Bahía
Blanca Estuary, south of the Principal Channel (Fig. 2.1; Chap. 2).
The Corymorpha januarii population observed along an area of 250 m long and
3 m wide was composed of an estimated number of 31,500 polyps (42 ± 5 polyps/
m2). Almost all visible polyps were lying on the mud, completely exposed to the sun
and air during the ebb tide (3 h). The accompanying fauna consisted of an epiben-
thic species, the octocoral Stylatula darwinii as well as by endobenthic species such
as the bivalves Pitar rostrata, Malletia sp. and Nucula sp.; the gastropod Buccinanops
deformis; amphipods like Monocorophium insidiosum (Fig. 8.1e) and Heterophoxus
videns (Fig. 8.1f); and several species of polychaetes, being Leodamas verax and
Axiothella sp., the dominant ones (Dutto et al. 2019).
Polyps play an important role in energy transfer from the water column to the
benthos (Gili and Coma 1998) because of their great abundances and high predation
impact on a wide food spectrum, from detritus, diatoms and benthic microplankton
to egg and fish larvae (e.g. Gili and Hughes 1995; Gili et  al. 2008). Dutto et  al.
(2019) indicate that the most frequently trophic items consumed by Corymorpha
januarii were particulate organic matter (POM) and certain harpacticoid copepod
species, with Microarthridion aff. littorale being the most highly consumed species,
followed by the calanoid copepod Acartia tonsa and the mysid Neomysis ameri-
cana. Considering the total density of polyps, its consumption represented a daily
carbon consumption of 75 mg C/m2/day, and the daily mass-specific ingestion rate
of polyps of C. januarii was 20.5% of the polyp biomass.
The study of intertidal hydroids has been largely neglected, particularly of those
in southwestern Atlantic coasts (Genzano et  al. 2017). Therefore, the fortuitous
finding of this species substantially contributes to understanding the role of these
populations as important biomass contributors in benthic communities, being an
essential part of the ‘animal forest’ (Rossi et al. 2017, 2019) on coastal ecosystems.
8  The Intertidal Soft-Bottom Macrobenthic Invertebrates 199

8.5  Threats to Macrobenthic Species

The intertidal areas of estuaries represent important areas for recruitment, breeding,
feeding and refugee of several species of fishes and birds, mainly due to their low
wave energy. The Bahía Blanca Estuary is not an exception, as several species entry
into the estuary for reproductive or feeding purposes and the great majority feed on
macrobenthic invertebrates found in intertidal areas (Elías et al. 2004). These organ-
isms tend to be highly vulnerable because they are permanently exposed to several
toxic organic and inorganic pollutants found in seawater and sediments (Kennish 2002).
The economic and demographic development of coastal regions, with the conse-
quent increase of urban, industrial and port settlements, generates great loads of
waste, which are discharged into the water. This discharge of anthropogenic con-
taminants to the environment can have profound effects on natural ecosystems
(Vitousek et  al. 1997). In the Bahía Blanca Estuary, several studies reflect the
anthropogenic impact to which macrofaunal communities are subject. Most studies
evaluated the impacts of cloacal sewage discharge (Fiori et  al. 2019), polycyclic
aromatic hydrocarbons (Arias et al. 2009), heavy metals (Buzzi and Marcovecchio
2016; Simonetti et al. 2012, 2013, 2018) and microplastics (Villagrán et al. 2019).
Some of them were conducted at the community level, but the vast majority of stud-
ies evaluate impacts on targeted populations as the razor clam Tagelus plebeius and
the burrowing crab Neohelice granulata.
Several macrobenthic species found in intertidal areas of the estuary, such as the
burrowing crab Neohelice granulata, are known to have important ecosystem
effects, in relation to its capability to modify energy fluxes in the salt marshes. Crab
bioturbation and herbivory, for example, may account for variability in the uptake
of atmospheric CO2 by salt marshes and can also be relevant to ‘Blue Carbon’ stud-
ies (i.e. carbon that is sequestered and stored in coastal marine environments;
Nellemann et al. 2009), having an impact on mitigation of climate change. Therefore,
the preservation and restoration of salt marshes, along with the other vegetated
coastal ecosystems (i.e. mangroves and seagrass meadows), represent an essential
piece of the solution to global climate change. Through herbivory, crabs could
potentially modulate the amount of carbon stored in salt marshes by reducing above-­
ground biomass. The effects mediated by bioturbation comprise a complex
interplay between direct and indirect effects with both positive and negative out-
comes in terms of carbon sequestration and storage (Martinetto et al. 2016).
Therefore, the preservation and restoration of these coastal ecosystems represent
a win-win scenario for multiple objectives as biodiversity conservation and climate
change mitigation. Some efforts have been made to conserve biodiversity in the
Bahía Blanca Estuary. The Nature Reserve Bahía Blanca, Bahía Falsa and Bahía
Verde (Fig. 2.1; Chap. 2) was created in 1998 with the primary purpose of safe-
guarding the marine ecosystem as a whole as well as to protect and preserve an
important refuge and reproduction site for several species. In 2016, the Bahía Blanca
Estuary was declared as ‘site of regional importance’ of the Western Hemisphere
200 M. C. Carcedo et al.

Shorebird Reserve Network (WHSRN). This conservation strategy aims to protect


the nesting, breeding and staging habitats of migratory shorebirds.
Another important threat to macrobenthic species in the Bahía Blanca Estuary is
the introduction of exotic species which can lead to important changes in soft-­
bottom communities (see Box 8.2). While the introduction of invasive species to
estuarine systems worldwide continues unabated, more data are needed to

Box 8.2: Islands of Heterogeneity and Complexity in the Intertidal


Area: The Hard Substrates
Opposite to the soft-bottom environments, natural hard substrates occupy
relatively small areas in the Argentine coast. Particularly, in the Bahía Blanca
Estuary, the most important hard substrate areas are located in the northern
and external coast, near Punta Alta city (Fig. 2.4; Chap. 2). Unfortunately,
little attention has been paid to study the benthic macrofaunal community that
lives in these habitats. These large rocky outcrops, called ‘beachrocks’, have
relevant geological features, due to the fact that its genesis is linked to the
coastal morphosedimentological evolution (Aliotta et  al. 2009). One of the
most important outcrops is in the coastal zone of Villa del Mar (Fig. 2.4;
Chap. 2), in the upper intertidal zone. These rocks are dominated by the exotic
barnacles Balanus glandula and Amphibalanus amphitrite (Fiori and
Bieckzinsky 2008) and the Pacific oyster Magallana gigas, with higher densi-
ties in lower intertidal levels (Dos Santos and Fiori 2010).
Another important rocky outcrop found in intertidal areas of the Bahía
Blanca Estuary is located near Puerto Rosales, distributed along the interme-
diate and lower levels of the intertidal zone. Here, individuals of Magallana
gigas were registered ranging from 7.9 ± 4.7 individuals/m2 (attached in scat-
tered gravels and shells) to 82.4 ± 23.8 individuals/m2 (forming oyster reefs)
(Dos Santos, pers. comm.). Other species found inhabiting these hard sub-
strates were the mussel Brachidontes rodriguezii, the snail Heleobia australis
(Carcedo and Fiori 2011) and several invertebrate species reported for regional
communities of intertidal mussel beds such as the polychaete Alitta succinea
and the amphipod Monocorophium insidiosum (Dos Santos et al. 2018).
Organisms inhabiting natural hard substrates can colonize artificial struc-
tures like breakwaters, boats, rafts, buoys and seawater intakes. These artifi-
cial structures are widely distributed along the Bahía Blanca Estuary and
represent the most extensive hard-bottom surfaces for the settlement of marine
organisms. Therefore, the vast majority of studies of hard-bottom macrofauna
in the Bahía Blanca Estuary have focused on these types of structures. The
first studies were conducted during the 1970s, in the two main port areas:
Ingeniero White (Fig. 2.3; Chap. 2), located in the middle zone of the estuary,
and Naval Base Puerto Belgrano, in the external zone of the estuary (Fig. 2.4;
Chap. 2). These studies covered the species identification, intertidal zonifica-
tion, annual succession of subtidal communities and the main epibiotic phe-
nomena (Bastida and Torti 1971; Bastida et  al. 1974a, 1974b; Bastida and
Lichtschein de Bastida 1978; Valentinuzzi de Santos 1971).
8  The Intertidal Soft-Bottom Macrobenthic Invertebrates 201

Valentinuzzi de Santos (1971) indicated that the intertidal fouling commu-


nity was dominated by the balanid Amphibalanus amphitrite, the mussel
Brachidontes rodriguezii and several species of polychaetes. During the
2000s, the dominance of the community was due to two balanid species,
A. amphitrite and Balanus glandula, together with the bryozoan Amathia (=
Bowerbankia) sp. (Bremec et al. 2004). More than 10 years later, the domi-
nance of B. glandula increased over A. amphitrite, and a high abundance of
the anemone Diadumene lineata (Roldán 2014) was also detected. Floating
substrates such as rafts in the Port of Ingeniero White showed a significant
increase in the biomass of the hydrozoan Tubularia crocea during the warm
season, within a community dominated by passive and active filters of cosmo-
politan distribution (Dos Santos et al. 2006). The presence of exotic and cos-
mopolitan species is the main feature of the subtidal hard-bottom community
of the Bahía Blanca Estuary, like the polychaete Hydroides dianthus and the
bryozoans Bugulina flabellata, Bugula neritina, Bugulina simplex, Bugulina
stolonifera and Cryptosula pallasiana (Orensanz et  al. 2002), included,
among others, in a more extensive list of species (Bremec et al. 2004).
During the last few decades, a rapid proliferation of artificial hard structures
(e.g. energy infrastructure, aquaculture, coastal defences) was observed in
coastal settings, with the consequent replacement of natural, often sedimen-
tary, environments with hard substrata (Firth et al. 2016). Artificial hard sub-
strates play a particularly important role in the dynamics of marine invasions,
being possibly the first substrates colonized by non-native species arriving in a
bay or estuary. For example, in coastal North America, approximately 90% of
the alien species inhabiting hard substrata have been reported from docks and
marinas (Mineur et al. 2012). It may result partly from their predominance on
developed shorelines, where vectors and supply of propagules are presumably
concentrated (Ruiz et  al. 2009). In addition, hard-­bottom communities may
facilitate initial invasions through greater availability of resources over those
on natural substrata. These areas can, therefore, provide seed populations that
persist and increase in abundance and, as a result, spread regionally to other
environments, including natural substrata (Glasby et al. 2007; Ruiz et al. 2009).
A dispersion of the non-native balanid Balanus glandula from the
Argentine Port of Mar del Plata to neighbouring rocky coasts has been
observed since the 1970s (Spivak 2005), reaching southern Patagonian ports
years later (Rico et al. 2001). B. glandula entered the Bahía Blanca Estuary in
the 1980s, detected through the observation of its larvae in the local zooplank-
ton community (Hoffmeyer 1983). Regarding the other non-native balanid,
Amphibalanus amphitrite, it entered local port before 1970, as indicated by
early records in intertidal areas of Port of Ingeniero White (Valentinuzzi de
Santos 1971) and intertidal-subtidal areas of Naval Base Puerto Belgrano
(Bastida and Torti 1971). In 2008, the anemone Diadumene lineata was reg-
istered in the Bahía Blanca Estuary, attached to stems and roots of Spartina
alterniflora, in an association never registered before (Molina et al. 2008).
202 M. C. Carcedo et al.

The more recent invasion registered in the Bahía Blanca Estuary was the
oyster Magallana (= Crassostrea) gigas (Dos Santos and Fiori 2010) (Fig. 8.4),
a species native in the northwestern Pacific Ocean and the Sea of Japan, and
introduced illegally in Argentina in 1982 for aquaculture purposes (Escapa
et al. 2004a; Borges 2005, 2006). Since then, it has spread both northwards
until reaching the Bahía Blanca Estuary and southwards to the northern
Patagonia Region, indicating its sustained expansion, mainly due to its capa-
bility to tolerate a wide range of environmental conditions (Shatkin et al. 1997).
Even though the natural and artificial hard substrates, i.e. the primary habi-
tat for the settlement of the Pacific oyster, are confined to a few sectors in the
Bahía Blanca Estuary, the species is able to settle on rocks, debris and shells
from the lower intertidal zone to depths of 40 m. Once it settles on a shell or
a small stone, clumps of oysters may merge to form dense aggregations and
potentially a reef (Herbert et al. 2016), facilitating further settlement on soft-
bottom substrates (Fiori et al. 2016). Thus, the soft-bottom environments of
the estuary with presence of gravels and/or shells could be modified towards
bivalve reef as a result of the settlement of the Pacific oyster on these sub-
strates (Bravo 2013).
The establishment of oyster reefs on soft bottoms can promote structural
changes with effects on infaunal invertebrates, such as an increase in the
abundance of epifaunal organisms that use the reef as a refuge and a decrease
in those that do not need shelters. A higher abundance of preys can also affect
habitat selection by coastal birds (Escapa et al. 2004a). Due to the high adap-
tation and expansion capability of the species, a negative impact of M. gigas
on environmental values and human health in the estuary should be expected.
The impacts can be estimated through a recently developed risk index, which
evaluates the effectiveness to manage the environmental matrix as the best
option for preventing or reducing undesired effects of the oyster (Fiori et al.
2016). For example, the risk index could evaluate projections of direct control
operations as eradication of adult oysters at specific locations or the impacts
of specific management practices.
Important changes in the ecosystem are expected due to the establishment
of Magallana gigas, both in the pelagic and benthic environments. The spe-
cies has a high filtration rate, consuming suspended organisms, as well as
detrital organic matter, being able to modify the plankton community. In the
case of the benthic environment, oyster populations in high densities as those
found on the Bahía Blanca Estuary have the potential to (1) reduce food avail-
ability for other native filter feeding, (2) modify populations due to changes in
abundance of their planktonic life stages and (3) change richness and diversity
of community due to the increased bottom complexity (Crooks 2002; Ruesink
et al. 2005; Troost et al. 2009; Wilkie et al. 2013). These changes not only
impact at a local level, but its successful establishment increases the chances
of a possible expansion to other places vectorized by national and interna-
tional ships that constantly circulate in the estuary (Dos Santos and Fiori
2010). In fact, studies have predicted the dispersion of the Pacific oyster along
extensive areas of the Argentine coast (Carrasco and Baron 2010).
8  The Intertidal Soft-Bottom Macrobenthic Invertebrates 203

Fig. 8.4  The Pacific oyster Magallana gigas in artificial hard substrates of the Bahía Blanca
Estuary. (Photo by M. Emilia Croce)

effectively assess their long-term environmental impacts (Kennish 2002).


Furthermore, in the current scenario of global climate change, future studies should
tend to analyse specific responses of organisms into global and integrative
approaches. Whole-­system perspectives are needed for understanding how this
complex and dynamic ecosystem will respond to multiple anthropogenic drivers.

References

Aguirre ML, Farinati EA (2000) Aspectos sistemáticos, de distribución y paleoambientales de


Littoridina australis (D’Orbigny 1835) (Mesogastropoda) en el Cuaternario marino de
Argentina (Sudamérica). Geobios 33:569–597
Alberti J, Escapa M, Iribarne O, Silliman B, Bertness M (2008) Crab herbivory regulates plant
facilitative and competitive processes in Argentinean marshes. Ecology 89(1):155–164
Alberti J, Casariego AM, Daleo P, Fanjul E, Silliman B, Bertness M, Iribarne O (2010) Abiotic
stress mediates top-down and bottom-up control in a Southwestern Atlantic salt marsh.
Oecologia 163(1):181–191
Alberti J, Cebrian J, Casariego AM, Canepuccia A, Escapa M, Iribarne O (2011) Effects of nutrient
enrichment and crab herbivory on a SW Atlantic salt marsh productivity. J Exp Mar Biol Ecol
405(1):99–104
Alberti J, Daleo P, Fanjul E, Escapa M, Botto F, Iribarne O (2014) Can a single species challenge
paradigms of salt marsh functioning? Estuar Coast 38(4):1178–1188
Aliotta S, Spagnuolo J, Farinati E (2009) Origen de una roca de playa en la región costera de Bahía
Blanca, Argentina. Pesquisas em Geociências 36(1):107–116
Alldredge AL, Jackson GA (1995) Preface: aggregation in marine system. Deep Sea Res II Top
Stud Ocean 42(1):1–7
Angeletti S (2012) Composición y estructura poblacional de Neohelice granulata (Varunidae) en
una planicie de marea del estuario de Bahía Blanca (Villa del Mar). Tesis de Licenciatura en
Ciencias Biológicas. Univ Nac del Sur, Argentina, p 50
Angeletti S (2017) Efecto bioturbador del cangrejo Neohelice granulata sobre la distribución y
transporte de sedimento en ambientes intermareales próximos al límite sur de su distribución
geográfica: Un estudio poblacional comparado. PhD thesis. Univ Nac del Sur, Argentina, p 180
204 M. C. Carcedo et al.

Angeletti S, Cervellini PM (2015) Population structure of the burrowing crab Neohelice gran-
ulata (Brachyura, Varunidae) in a Southwestern Atlantic salt marsh. Lat Am J Aquat Res
43(3):539–547
Angeletti S, Cervellini PM, Lescano L (2018a) Burrowing activity of Neohelice granulata crab
(Brachyura, Varunidae) in Southwest Atlantic intertidal areas. Cienc Mar 44(3):155–167
Angeletti S, Pierini JO, Cervellini PM (2018b) Suspension sediment contribution resulting from
bioturbation in intertidal sites of SW Atlantic mesotidal estuary: data analysis and numerical
modelling. Sci Mar 82(4):245–256
Arias A, Spetter C, Freije H, Marcovecchio J (2009) Polycyclic aromatic hydrocarbons (PAHs)
distribution in water column, native mussels (Brachydontes sp and Tagelus sp) and fish
(Odontesthes sp) from an industrialized south American estuary. Estuar Coast Shelf Sci
85:67–81
Auster PJ, Crockett LR (1984) Foraging tactics of several crustacean species for infaunal prey. J
Shellfish Res 4:139–143
Bas C, Luppi T, Spivak E (2005) Population structure of the South American estuarine crab,
Chasmagnathus granulatus (Brachyura: Varunidae) near the southern limit of its geographical
distribution: comparison with northern populations. Hydrobiologia 537:217–228
Bastida R, Lichtschein de Bastida V (1978) Las incrustaciones biológicas de Puerto Belgrano.
III Estudio de los procesos de epibiosis registrados sobre paneles acumulativos. CIDEPINT
Anales 3:55–97
Bastida RO, Torti MR (1971) Estudio preliminar de las incrustaciones biológicas de Puerto
Belgrano (Argentina). LEMIT Anales, Ser II 3:45–75
Bastida R, Spivak E, L’Hoste SG, Adabbo HE (1974a) Las incrustaciones biológicas de Puerto
Belgrano. I Estudio de la fijación sobre paneles mensuales, período 1971/72. LEMIT Anales
Ser II(274):97–165
Bastida R, Spivak E, L’Hoste SG, Adabbo HE (1974b) Las incrustaciones biológicas de Puerto
Belgrano. II Estudio de los procesos de epibiosis registrados sobre paneles mensuales. LEMIT
Anales Ser II(274):25–33
Berón MP, García GO, Luppi T, Favero M (2011) Age-related prey selectivity and foraging effi-
ciency of Olrog’s Gulls (Larus atlanticus) feeding on crabs in their nonbreeding grounds. Emu
111(2):172–178
Billick I, Case TJ (1994) Higher order interactions in ecological communities: what are they and
how can they be detected? Ecology 75(6):1529–1543
Blackburn TH, Henriksen K (1983) Nitrogen cycling in different types of sediments from Danish
waters. Limnol Oceanogr 28(3):477–493
Bogazzi E, Iribarne O, Guerrero R, Spivak E (2001) Wind pattern may explain the southern limit
of distribution of a southwestern Atlantic fiddler crab. J Shellfish Res 59:483–500
Borges ME (2005) La ostra del Pacífico Crassostrea gigas (Thumberg, 1793) en la Bahía Anegada
(Provincia de Buenos Aires). In: Penchaszadeh PE et al (eds) Invasores: Invertebrados exóticos
en el Río de la Plata y región marina aledaña. EUDEBA, Buenos Aires, pp 311–367
Borges ME (2006) Ecología de las ostras en ambientes del sur bonaerense: cultivo y manejo de sus
poblaciones. PhD thesis. Univ Nac del Sur, Argentina, p 247
Bortolus A, Iribarne OO (1999) Effects of the SW Atlantic burrowing crab Chasmagnathus granu-
lata on a Spartina salt marsh. Mar Ecol Prog Ser 178:79–88
Bortolus A, Schwindt E, Iribarne OO (2002) Positive plant-animal interactions in the high marsh
of an Argentinean coastal lagoon. Ecology 83:733–742
Bortolus A, Carlton JT, Schwindt E (2015) Reimagining South American coasts: unveiling the
hidden invasion history of an iconic ecological engineer. Divers Distrib 21(11):1267–1283
Boschi EE (1964) Los crustáceos decápodos Brachyura del litoral bonaerense (R Argentina) In:
Boletín del Instituto de Biología Marina, V6, pp 3–99
Boschi EE, Fischbach CE, Iorio MI (1992) Catálogo ilustrado de los crustáceos estomatópodos y
decápodos marinos de Argentina. Frente Marítimo 10, 7(Sec. A):–94
Botto F, Iribarne OO (2000) Contrasting effects of two burrowing crabs (Chasmagnathus granu-
lata and Uca uruguayensis) on sediment composition and transport in estuarine environments.
Estuar CoastShelf Sci 51(2):141–151
8  The Intertidal Soft-Bottom Macrobenthic Invertebrates 205

Botto F, Iribarne OO, Martínez MM, Delhey K, Carrete M (1998) The effect of migratory shore-
birds on the benthic species of three southwestern Atlantic Argentinean estuaries. Estuaries
21(4):700–709
Botto F, Valiela I, Iribarne O, Martinetto P, Alberti J (2005) Impact of burrowing crabs on C and
N sources, control, and transformations in sediments and food webs of SW Atlantic estuaries.
Mar Ecol Progr Ser 293:155–164
Botto F, Iribarne O, Gutierrez J, Bava J, Gagliardini A, Valiela I (2006) Ecological importance
of passive deposition of organic matter into burrows of the SW Atlantic crab Chasmagnathus
granulatus. Mar Ecol Progr Ser 312:201–210
Boudreau BP, Jørgensen BB (2001) The benthic boundary layer: transport processes and biogeo-
chemistry. Oxford University Press, London, p 404
Boyer JN, Kelble CR, Ortner PB, Rudnick DT (2009) Phytoplankton bloom status: chlorophyll
a biomass as an indicator of water quality condition in the southern estuaries of Florida,
USA. Ecol Indic 9(6):S56–S67
Boynton WR, Kemp WM (1985) Nutrient regeneration and oxygen consumption by sediments
along an estuarine salinity gradient. Mar Ecol Progr Ser 23(1):45–55
Braeckman U, Van Colen C, Soetaert K, Vincx M, Vanaverbeke J (2011) Contrasting macrobenthic
activities differentially affect nematode density and diversity in a shallow subtidal marine sedi-
ment. Mar Ecol Progr Ser 422:179–191
Bravo ME (2013) Estado de la invasión de la ostra del Pacífico, Crassostrea gigas, en el estu-
ario de Bahía Blanca. Tesis de Licenciatura en Ciencias Biológicas. Univ Nac del Sur, Bahía
Blanca, p 40
Bremec CS, Martínez DE, Elías R (2004) Asociaciones bentónicas de fondos duros y comunidades
incrustantes. In: Piccolo MC, Hoffmeyer MS (eds) Ecosistema del estuario de Bahía Blanca.
Instituto Argentino de Oceanografía, Bahía Blanca, pp 171–178
Bricker SB, Longstaff B, Dennison W, Jones A, Boicourt K, Wicks C, Woerner J (2008) Effects of
nutrient enrichment in the nation’s estuaries: a decade of change. Harmful Algae 8(1):21–32
Burdige DJ (2006) Geochemistry of marine sediments. Princeton University Press, Princeton, p 630
Burrows MT, Schoeman DS, Richardson AJ, Molinos JG, Hoffmann A, Buckley LB et  al
(2014) Geographical limits to species-range shifts are suggested by climate velocity. Nature
507:492–495
Buzzi NS, Marcovecchio JE (2016) A baseline study of the metallothioneins induction and its
reversibility in Neohelice granulata from the Bahía Blanca estuary (Argentina). Mar Poll Bull
112(1–2):452–458
Cadée GC (2001) Sediment dynamics by bioturbating organisms. In: Ecological comparisons of
sedimentary shores. Springer, Berlin/Heidelberg, pp 127–148
Callaway RM (1995) Positive interactions among plants. Bot Rev 61:306–349
Canepuccia AD, Escapa M, Daleo P, Alberti J, Botto F, Iribarne OO (2007) Positive interactions of
the smooth cordgrass Spartina alterniflora on the mud snail Heleobia australis, in southwest-
ern Atlantic salt marshes. J Exp Mar Biol Ecol 353:180–190
Canepuccia AD, Fanjul MS, Fanjul E, Botto F, Iribarne OO (2008) The intertidal burrowing crab
Neohelice (= Chasmagnathus) granulata positively affects foraging of rodents in south western
Atlantic salt marshes. Estuar Coast 31(5):920–930
Carcedo MC, Fiori SM (2011) Patrones de distribución y abundancia de Heleobia australis
(Caenogastropoda: Cochliopidae) en el estuario de Bahía Blanca, Argentina. In: Cazzaniga
NJ (ed) El Género Heleobia (Caenogastropoda: Cochliopidae) en América del Sur. Amici
Molluscarum (Núm Esp), pp 33–35
Carrasco MF, Baron PJ (2010) Analysis of the potential geographic range of the Pacific oyster
Crassostrea gigas (Thunberg, 1793) based on surface seawater temperature satellite data and
climate charts: the coast of South America as a study case. Biol Invasions 12:2597–2607
César II, Armendáriz LC, Becerra RV (2005) Bioecology of the fiddler crab Uca uruguayensis and
the burrowing crab Chasmagnathus granulatus (Decapoda, Brachyura) in the Refugio de Vida
Silvestre Bahía Samborombón, Argentina. Hydrobiologia 545(1):237–248
206 M. C. Carcedo et al.

Chalmers GC, Wiegert RG, Wolff PL (1985) Carbon balance in a salt marsh: interactions of dif-
fusive export, tidal deposition and rainfall-caused erosion. Estuar Coast Shelf Sci 21:757–771
Chanley PE (1965) Larval development of a boring clam, Barnea truncata. Chesap Sci 6:162–166
Chauvaud L, Jean F, Ragueneau O, Thouzeau G (2000) Long-term variation of the bay of Brest
ecosystem: benthic-pelagic coupling revisited. Mar Ecol Progr Ser 200:35–48
Chmura GL, Anisfeld S, Cahoon D, Lynch J (2003) Global carbon sequestration in tidal, saline
wetland soils. Global Biogeochem Cy 17:1–12
Cloern JE (1987) Turbidity as a control on phytoplankton biomass and productivity in estuaries.
Cont Shelf Res 7(11):1367–1381
COSEWIC (2009) COSEWIC assessment and status report on the Atlantic Mud-piddoc Barnea
truncate in Canada. Committee on the Status of Endangered Wildlife in Canada. Ottawa:
vii + 42 p
Costa CS, Marangoni JC, Azevedo AM (2003) Plant zonation in irregularly flooded salt marshes:
relative importance of stress tolerance and biological interactions. J Ecol 91(6):951–965
Crane J (1975) Fiddler crabs of the world, Ocypodidae: genus Uca. Princeton University Press,
Princeton, p 736
Crooks JA (2002) Characterizing ecosystem-level consequences of biological invasions: the role
of ecosystem engineers. Oikos 97:153–166
Cuadrado DG, Carmona NB, Bournod C (2011) Biostabilization of sediments by microbial
mats in a temperate siliciclastic tidal flat, Bahia Blanca estuary (Argentina). Sediment Geol
237(1):95–101
da Silveira FL, Migotto AE (1992) Rediscovery of Corymorpha januarii Steenstrup, 1854
(Hydrozoa, Corymorphidae) on the southeastern and southern coasts of Brazil. Steenstrupia
18:81–89
Dale AW, Prego R (2002) Physico-biogeochemical controls on benthic-pelagic coupling of nutri-
ent fluxes and recycling in a coastal upwelling system. Mar Ecol Progr Ser 235:15–28
Daleo P, Iribarne OO (2009) The burrowing crab Neohelice granulata affects the root strate-
gies of the cordgrass Spartina densiflora in SW Atlantic salt marshes. J Exp Mar Biol Ecol
373(1):66–71
Daleo P, Fanjul E, Casariego AM, Silliman BR, Bertness MD, Iribarne OO (2007) Ecosystem
engineers activate mycorrhizal mutualism in salt marshes. Ecol Lett 10(10):902–908
Day JW, Hall CAS Jr, Kemp WM, Yañez-Arancibia A (1989) Estuarine ecology. Wiley-Interscience
Publication/Wiley, New York
De Francesco CG, Isla FI (2003) Distribution and abundance of hydrobiid snails in a mixed estuary
and a coastal lagoon, Argentina. Estuaries 26:790–797
Decho AW (2000) Microbial biofilms in intertidal systems: an overview. Cont Shelf Res
20(10):1257–1273
Defeo O, McLachlan A (2005) Patterns, processes and regulatory mechanisms in sandy beach
macrofauna: a multi-scale analysis. Mar Ecol Progr Ser 295:1–20
Dietrich WE, Perron JT (2006) The search for a topographic signature of life. Nature
439(7075):411–418
Dos Santos EP, Fiori SM (2010) Primer registro sobre la presencia de Crassostrea gigas
(Thumberg, 1793) en el Estuario de Bahía Blanca (Argentina). Comunicaciones de la Sociedad
Malacológica del Uruguay 93:245–252
Dos Santos EP, Piccolo CM, Parodi ER (2006) Desarrollo de la biomasa de una comunidad
incrustante (fouling) relacionado con parámetros ambientales. VI Jornadas Nacionales de
Ciencias del Mar, Puerto Madryn, Argentina, p 175
Dos Santos EP, Carcedo MC, Zotelo C, Fiori SM (2018) Effects of erosion-accretion processes on
a biogenic reef formed by Brachidontes rodriguezii (Mollusca: Mytilidae) on a sandy beach
along the South Atlantic coast. J Marine Syst 187:146–155
Dutto MS, Carcedo MC, Nahuelhual EG, Conte AF, Berasategui AA, Garcia MD et  al (2019)
Trophic ecology of a corymorphid hydroid population in the Bahía Blanca estuary, Southwestern
Atlantic. Reg Stud Mar Sci 31:100746
8  The Intertidal Soft-Bottom Macrobenthic Invertebrates 207

Echeverría CA, Neves RAF, Pessoa LA, Paiva PC (2010) Spatial and temporal distribution of
the gastropod Heleobia australis in an eutrophic estuarine system suggests a metapopulation
dynamics. Nat Sci 2(8):860–867
Elías R (1985) Macrobentos del estuario de Bahía Blanca. I Mesolitoral. Spheniscus 1:1–33
Elías R, Bremec CS (1986) Macrobentos del área de la bahía Blanca (Provincia de Buenos Aires).
II Relaciones entre las asociaciones de sustrato móvil. Spheniscus 3:51–52
Elías R, Iribarne OO, Bremec CS, Martínez DE (2004) Comunidades bentónicas de fondos blan-
dos. In: Piccolo MC, Hoffmeyer MS (eds) Ecosistema del Estuario de Bahía Blanca. Instituto
Argentino de Oceanografía, Bahía Blanca, pp 179–190
Escapa M (2007) Efecto de las interacciones biológicas en la erosión de marismas. PhD thesis,
Univ Nac de Mar del Plata, Argentina, p 162
Escapa M, Iribarne OO, Navarro D (2004b) Effects of the intertidal burrowing crab Chasmagnathus
granulatus on infaunal zonation patterns, tidal behavior and risk of mortality. Estuaries
27(1):120–131
Escapa M, Isacch JP, Daleo P, Alberti J, Iribarne OO, Borges M, Dos Santos EP, Galiardini D, Lasta
M (2004a) The distribution and ecological effects of the invasive Pacific Oyster Crassostrea
gigas (Thunberg, 1793) in Northern Patagonia. J Shellfish Res 23:765–772
Escapa M, Minkoff DR, Perillo GM, Iribarne OO (2007) Direct and indirect effects of burrow-
ing crab Chasmagnathus granulatus activities on erosion of Southwest Atlantic Sarcocornia-­
dominated marshes. Limnol Oceanogr 52(6):2340–2349
Escapa M, Perillo GM, Iribarne OO (2008) Sediment dynamics modulated by burrowing crab
activities in contrasting SW Atlantic intertidal habitats. Estuar Coast Shelf Sci 80(3):365–373
Fanjul E, Grela MA, Iribarne OO (2007) Effects of the dominant SW Atlantic intertidal burrowing
crab Chasmagnathus granulatus on sediment chemistry and nutrient distribution. Mar Ecol
Progr Ser 341:177–190
Fanjul E, Escapa M, Montemayor D, Addino M, Alvarez MF, Grela MA, Iribarne OO (2015)
Effect of crab bioturbation on organic matter processing in south West Atlantic intertidal sedi-
ments. Jour Sea Res 95:206–216
Figueiredo-Barros MP, Leal JJF, Esteves F, De A, Rocha A, De M, Bozelli RL (2006) Life cycle,
secondary production and nutrient stock in Heleobia australis (d’Orbigny 1835) (Gastropoda:
Hydrobiidae) in a tropical coastal lagoon. Estuar Coast Shelf Sci 69:87–95
Fiori SM, Bieckzinsky F (2008) Cirripedios exóticos del estuario de Bahía Blanca. In: Cazzaniga
NJ, Arelovich HM (eds) Ambientes y Recursos Naturales del Sudoeste Bonaerense:
Producción, contaminación y conservación (Actas de las V jornadas Interdisciplinarias del
Sudoeste Bonaerense). EdiUNS, Bahía Blanca, pp 421–431
Fiori SM, Simonetti P, Dos Santos EP (2012) First record of Atlantic mud piddock, Barnea
(Anchomasa) truncata (Bivalvia, Pholadidae) in Argentina. Aquat Invasions 7(2):283–286
Fiori SM, Pratolongo PD, Zalba SM, Carbone ME, Bravo ME (2016) Spatially explicit risk assess-
ment for coastal invaders under different management scenarios. Mar Biol 163(12):245
Fiori SM, Bravo ME, Elías R, Serra AV, Carcedo MC, Dos Santos EP, Botté SE (2019) Effects of
sewage effluent on the subtidal macrobenthic assemblage in a urban estuary (Argentina). Ecol
Austral (in press)
Firth LB, Knights AM, Bridger D, Evans AJ, Mieszkowska N, Moore PJ, O’Connor NE, Sheehan
EV, Thompson RC, Hawkins SJ (2016) Ocean sprawl: challenges and opportunities for biodi-
versity management in a changing world. In: Oceanography and marine biology. CRC Press,
New York, pp 201–278
García Molinos J, Halpern BS, Schoeman DS, Brown CJ, Kiessling W, Moore PJ et  al (2015)
Climate velocity and the future of global redistribution of marine biodiversity. Nat Clim Chang
61:83–88
Genzano GN, Mianzan H, Rodríguez C, Diaz Briz LM (2009) The hydroid and medusa of
Corymorpha januarii in temperate waters of the Southwestern Atlantic Ocean. Bull Mar Sci
84:229–235
208 M. C. Carcedo et al.

Genzano GN, Bremec CS, Díaz-Briz LM, Costella JH, Morandini AC, Miranda T, Marques AC
(2017) Faunal assemblages of intertidal hydroids (Hydrozoa, Cnidaria) inhabiting salt marshes
and intertidal outcrops from Argentinean Patagonia (SW Atlantic Ocean). Lat Am J Aquat Res
45:177–187
Gilbert F, Aller RC, Hulth S (2003) The influence of macrofaunal burrow spacing and diffusive
scaling on sedimentary nitrification and denitrification: an experimental simulation and model
approach. J Mar Res 61:101–125
Gili JM, Coma R (1998) Benthic suspension feeders: their paramount role in littoral marine food
webs. Trends in Ecol Evol 13:316–321
Gili JM, Hughes RG (1995) Ecology of benthic hydroids. Oceanogr Mar Biol Annu Rev
33:351–422
Gili JM, Duró A, García-Valero J, Gasol JM, Rossi S (2008) Herbivory in small carnivores: benthic
hydroids as an example. J Mar Biol Assoc UK 88:1541–1546
Glasby TM, Connell SD, Holloway MG, Hewitt CL (2007) Nonindigenous biota on artificial struc-
tures: could habitat creation facilitate biological invasions? Mar Biol 151:887–895
Griffiths JR, Kadin M, Nascimento FJ, Tamelander T, Törnroos A, Bonaglia S, Bonsdorff E, Brüchert
V, Gårdmark A, Järnström M, Kotta J (2017) The importance of benthic–pelagic coupling for
marine ecosystem functioning in a changing world. Glob Change Biol 23(6):2179–2196
Guinder VA (2011) Dinámica del fitoplancton en el estuario de Bahía Blanca y su relación con
las variables ambientales en el marco del cambio climático. PhD thesis, Univ Nac del Sur,
Argentina, p 153
Guinder VA, Popovich CA, Perillo GM (2009) Particulate suspended matter concentrations in the
Bahía Blanca estuary, Argentina: implication for the development of phytoplankton blooms.
Estuar Coast Shelf Sci 85(1):157–165
Guinder VA, López-Abbate MC, Berasategui AA, Negrin VL, Zapperi G, Pratolongo PD, Severini
MDF, Popovich CA (2015) Influence of the winter phytoplankton bloom on the settled material
in a temperate shallow estuary. Oceanologia 57(1):50–60
Gutiérrez JL, Iribarne OO (1998) The occurrence of juveniles of the grapsid crab Chasmagnathus
granulata in siphon holes of the stout razor clam Tagelus plebeius. J Shellfish Res 17(4):925–930
Gutiérrez JL, Jones CG, Strayer DL et al (2003) Mollusks as ecosystem engineers: the role of shell
production in aquatic habitats. Oikos 101:79–90
Gutiérrez JL, Jones CG, Groffman PM, Findlay SE, Iribarne OO, Ribeiro PD, Bruschetti CM
(2006) The contribution of crab burrow excavation to carbon availability in surficial salt-marsh
sediments. Ecosystems 9(4):647–658
Hansell MH (2005) Animal architecture. Oxford University Press, Oxford, p 343
Hansen LS, Blackburn TH (1992) Effect of algal bloom deposition on sediment respiration and
fluxes. Mar Biol 112:147–152
Hebda A (2011) Information in support of a recovery potential assessment of Atlantic mudpiddock
(Barnea truncata) in Canada. DFO Can Sci Advis Sec Res Doc 2010/117: vi + 30 p
Herbert RJ, Humphreys J, Davies CJ, Roberts C, Fletcher S, Crowe TP (2016) Ecological impacts
of non-native Pacific oysters (Crassostrea gigas) and management measures for protected
areas in Europe. Biodivers Conserv 25(14):2835–2865
Herman PMJ, Middelburg JJ, Van de Koppel J, Heip CHR (1999) Ecology of estuarine macroben-
thos. Adv Ecol Res 29:195–240
Hoegh-Guldberg O, Bruno JF (2010) The impact of climate change on the world’s marine ecosys-
tems. Science 328:1523–1528
Hoffmeyer MS (1983) Zooplancton del área interna de la Bahía Blanca (Buenos Aires, Argentina).
I-Composición faunística. Hist Nat, Corrientes, Argentina 3(8):73–94
Hovel KA, Bartholomew A, Lipcius RN (2001) Rapidly entrainable tidal vertical migrations in the
salt marsh snail Littoraria irrorata. Estuaries 24:808–816
Hughes DJ, Atkinson RJA, Ansell AD (2000) A field test of the effects of megafaunal burrows
on benthic chamber measurements of sediment-water solute fluxes. Mar Ecol Progr Ser
195:189–199
8  The Intertidal Soft-Bottom Macrobenthic Invertebrates 209

Iribarne OO, Martínez MM (1999) Predation on the southwestern Atlantic fiddler crab (Uca uru-
guayensis) by migratory shorebirds (Pluvialis dominica, P. squatarola, Arenaria interpres and
Numenius phaeopus). Estuarine Res Fed 22:47–54
Iribarne OO, Bortolus A, Botto F (1997) Between-habitats differences in burrow characteristics
and trophic modes in the southwestern Atlantic burrowing crab Chasmagnathus granulatus.
Mar Ecol Progr Ser 155:132–145
Iribarne OO, Bruschetti M, Escapa M, Bava J, Botto F, Gutierrez J et al (2005) Small-and large-­
scale effect of the SW Atlantic burrowing crab Chasmagnathus granulatus on habitat use by
migratory shorebirds. J Exp Mar Biol Ecol 315(1):87–101
Irigoien X, Castel J (1997) Light limitation and distribution of chlorophyll pigments in a highly
turbid estuary: the Gironde (SW France). Estuar Coast Shelf Sci 44(4):507–517
Jarvis NJ, Taylor A, Larsbo M et  al (2010) Modelling the effects of bioturbation on the re-­
distribution of 137Cs in an undisturbed grassland soil. Eur J Soil Sci 61:24–34
Jensen MH, Lomstein E, Serrensen J (1990) Benthic NH4+ and NO3− flux following sedimenta-
tion of a spring phytoplankton bloom in Aarhus bight, Denmark. Mar Ecol Prog Ser 61:87–96
Jones S, Jago CF (1993) In situ assessment of modification of sediment properties by burrowing
invertebrates. Mar Biol 115(1):133–142
Jones CG, Lawton JH, Shachak M (1994) Organisms as ecosystem engineers. In: Ecosystem man-
agement. Springer, New York, pp 130–147
Kennish MJ (2002) Environmental threats and environmental future of estuaries. Environ Conserv
29(1):78–107
Kiørboe T, Jackson GA (2001) Marine snow, organic solute plumes, and optimal chemosensory
behavior of bacteria. Limnol Oceanogr 46(6):1309–1318
Klein G d V (1985) Intertidal flats and intertidal sand bodies. In: Davis RA (ed) Coastal sedimen-
tary environments, 2nd edn. Springer, New York, pp 187–224
Kneib RT (1984) Patterns of invertebrate distribution and abundance in the intertidal salt marsh:
causes and questions. Estuaries 2:392–412
Kneib RT (1991) Indirect effects in experimental studies of marine soft-sediment communities.
Am Zool 31(6):874–885
Kostka JE, Roychoudhury A, Van Cappellen P (2002) Rates and controls of anaerobic micro-
bial respiration across spatial and temporal gradients in saltmarsh sediments. Biogeochemistry
60(1):49–76
Kristensen E, Penha-Lopes G, Delefosse M, Valdemarsen T, Quintana CO, Banta GT (2012) What
is bioturbation? The need for a precise definition for fauna in aquatic sciences. Mar Ecol Progr
Ser 446:285–302
Legendre L (1990) The significance of microalgal blooms for fisheries and for the export of par-
ticulate organic carbon in oceans. J Plankton Res 12(4):681–699
Lewis DB, Eby LA (2002) Spatially heterogeneous refugia and predation risk in intertidal salt
marshes. Oikos 96:119–129
Little C, Nix W (1976) The burrowing and floating behavior of the gastropod Hydrobia ulvae.
Estuar Coast Mar Sci 4(5):537–544
Lomovasky BJ, Casariego AM, Brey T, Iribarne OO (2006) The effect of the SW Atlantic burrow-
ing crab Chasmagnathus granulatus on the intertidal razor clam Tagelus plebeius. J Exp Mar
Biol Ecol 337(1):19–29
Lozada M, Romano A, Maldonado H (1988) Effect of morphine and naloxone on a defensive
response of the crab Chasmagnathus granulatus. Pharmacol Biochem Be 30(3):635–640
Luppi TA, Bas CC (2002) Rol de los arrecifes del poliqueto invasor Ficopomatus enigmaticus
Fauvel 1923 (Polychaeta: Serpulidae) en el reclutamiento de Cyrtograpsus angulatus Dana,
1951 (Brachyura: Grapsidae). Cienc Mar 28(4):319–330
Marcus NH, Boero F (1998) Minireview: the importance of benthic-pelagic coupling and the for-
gotten role of life cycles in coastal aquatic systems. Limnol Oceanogr 43(5):763–768
Martinetto P, Montemayor DI, Alberti J, Costa CS, Iribarne OO (2016) Crab bioturbation and
herbivory may account for variability in carbon sequestration and stocks in South West Atlantic
salt marshes. Front Mar Sci 3:122
210 M. C. Carcedo et al.

May CL, Koseff JR, Lucas LV, Cloern JE, Schoellhamer DH (2003) Effects of spatial and temporal
variability of turbidity on phytoplankton blooms. Mar Ecol Progr Ser 254:111–128
McGlathery KJ, Sundback K, Anderson IC (2004) The importance of primary producers for ben-
thic nitrogen and phosphorus cycling. In: Nielsen SL, Banta GT, Pedersen F (eds) Estuarine
nutrient cycling: the influence of primary producers. Kluwer Academic, Dordrecht, pp 231–261
McLachlan A, Defeo O (2018) The ecology of sandy shores. Academic Press
Meadows PS, Meadows A, Murray JM (2012) Biological modifiers of marine benthic seascapes:
their role as ecosystem engineers. Geomorphology 157:31–48
Mees J, Jones MB (1997) The hyperbenthos. Oceanogr Mar Biol 35:35–212
Méndez Casariego A, Alberti J, Luppi T, Iribarne OO (2009) Stage-dependent interactions between
intertidal crabs: from facilitation to predation. J Mar Biol Assoc UK 89(4):781–788
Méndez Casariego A, Luppi T, Iribarne O, Daleo P (2011a) Increase of organic matter transport
between marshes and tidal flats by the burrowing crab Neohelice (Chasmagnathus) granulata
Dana in SW Atlantic salt marshes. J Exp Mar Biol Ecol 401(1–2):110–117
Méndez Casariego A, Alberti J, Luppi T, Daleo P, Iribarne OO (2011b) Habitat shifts and spa-
tial distribution of the intertidal crab Neohelice (Chasmagnathus) granulata Dana. J Sea Res
66(2):87–94
Meysman FJR, Middelburg JJ, Heip CHR (2006) Bioturbation: a fresh look at Darwin’s last idea.
Trends Ecol Evol 21:688–695
Michaud E, Desrosiers G, Mermillod-Blondin F, Sundby B, Stora G (2006) The functional group
approach to bioturbation: II.  The effects of the Macoma balthica community on fluxes of
­nutrients and dissolved organic carbon across the sediment-water interface. J Exp Mar Biol
Ecol 337(2):178–189
Milner RNC, Booksmythe I, Jennions MD, Backwell PRY (2010) The battle of the sexes? Territory
acquisition and defence in male and female fiddler crabs. Anim Behav 79(3):735–738
Mineur F, Cook EJ, Minchin D, Bohn K, MacLeod A, Maggs CA (2012) Changing coasts: marine
aliens and artificial structures. In: Oceanography and marine biology. CRC Press, New York,
pp 198–243
Minkoff DR, Escapa M, Ferramola FE, Maraschín SD, Pierini JO, Perillo GM, Delrieux C (2006)
Effects of crab–halophytic plant interactions on creek growth in a SW Atlantic salt marsh: a
cellular automata model. Estuar Coast Shelf Sci 69(3):403–413
Molina LM, Valinas MS, Pratolongo PD, Elías R, Perillo GME (2008) First record of the sea
anemone Diadumene lineata (Verrill 1871) associated to Spartina alterniflora roots and stems,
in marshes at the Bahia Blanca estuary, Argentina. Biol Invasions 11:409–416
Mortimer RJG, Krom MD, Watson PG, Frickers PE, Davey JT, Clifton RJ (1999) Sediment–
water exchange of nutrients in the intertidal zone of the Humber estuary, UK. Mar Poll Bull
37(3):261–279
Murray JM, Meadows A, Meadows PS (2002) Biogeomorphological implications of microscale
interactions between sediment geotechnics and marine benthos: a review. Geomorphology
47(1):15–30
Nellemann C, Corcoran E, Duarte CM, Valdes L, DeYoung C, Fonseca L, Grimsditch G (2009)
Blue Carbon. A Rapid Response Assessment. United Nations Environmental Programme,
GRID-Arendal, Birkeland Trykkeri AS, Birkeland
Nixon SW (1981) Remineralization and nutrient cycling in coastal marine ecosystems. In: Neilson
BJ, Cronin LE (eds) Estuaries and Nutrients. Humana Press, New York, pp 111–138
Norkko A, Hewitt JE, Thrush SF, Funnell GA (2001) Benthic-pelagic coupling and suspension
feeding bivalves: linking site-specific sediment flux and biodeposition to benthic community
structure. Limnol Oceanogr 46(8):2067–2072
Orensanz JM, Schwindt E, Pastorino G, Bortolus A, Casas G, Darrigran G, Elías R, Lopez Gappa
JJ, Obenat S, Pascual MS, Penchaszadeh P, Piriz ML, Scarabino F, Spivak ED, Vallarino EA
(2002) No longer the pristine confines of the world ocean: a survey of exotic marine species in
the southwestern Atlantic. Biol Invasions 4:115–143
8  The Intertidal Soft-Bottom Macrobenthic Invertebrates 211

Overnell J, Edwards A, Grantham BE, Harvey SM, Jones JW, Leftley JW, Smallman DJ (1995)
Sediment-water coupling and the fate of the spring phytoplankton bloom in Loch Linnhe, a
Scottish fjordic sea-loch. Sediment processes and sediment-water coupling. Estuar Coast Shelf
Sci 4:1–19
Paarlberg AJ, Knaapen MA, de Vries MB, Hulscher SJMH, Wang ZB (2005) Biological influences
on morphology and bed composition of an intertidal flat. Estuar Coast Shelf Sci 64(4):577–590
Palomo G, Iribarne OO (2000) Sediment bioturbation by polychaete feeding may promote sedi-
ment stability. Bull Mar Sci 67(1):249–257
Palomo G, Botto F, Navarro D, Escapa M, Iribarne OO (2003) Does the presence of the SW
Atlantic burrowing crab Chasmagnathus granulatus Dana affect predator–prey interactions
between shorebirds and polychaetes? J Exp Mar Biol Ecol 290(2):211–228
Paterson DM, Fortune I, Aspden RJ, Black KS (2019) Intertidal Flats: Form and Function. In:
GME P et al (eds) Coastal Wetlands. Elsevier, pp 383–406
Pearson TH, Rosenberg R (1987) Feast and famine: structuring factors in marine benthic commu-
nities. In: Gee JHR, Giller PS (eds) Organization of communities, past and present. Blackwell
Scientific, Oxford, pp 373–395
Pedersen MF, Nielsen SL, Banta GT (2004) Interactions between vegetation and nutrient dynam-
ics in coastal marine ecosystems: an introduction. In: Nielsen S, Banta G, Pedersen M (eds)
Estuarine nutrient cycling: the influence of primary producers. Kluwer Academic Publishers,
Dordrecht, pp 1–15
Perillo GME, Iribarne OO (2003) New mechanisms studied for creek formation in tidal flats: from
crabs to tidal channels. EOS Trans Am Geophys Union 84(1):1–5
Rathbun MJ (1918) The grapsoid crabs of America. US Government Printing Office
Reise K (1985) Tidal flat ecology: an experimental approach to species interactions. Springer, Berlin
Reise K (2002) Sediment mediated species interactions in coastal waters. J Sea Res 48(2):127–141
Revsbech NP, Sorensen J, Blackburn TH, Lomholt JP (1980) Distribution of oxygen in marine
sediments measured with microelectrodes. Limnol Oceanogr 25(3):403–411
Rhoads DC, Young DK (1970) The influence of deposit-feeding organisms on sediment stability
and community trophic structure. J Mar Res 28:150–178
Richter R (1952) Fluidal-Textur in Sediment-Gesteinen und über Sedifluktion überhaupt. Notizbl
Hess Landesamtes Bodenforsch Wiesb 6:67–68
Rico A, Lanas P, López Gappa J (2001) Temporal and spatial patterns in the recruitment of Balanus
glandula and Balanus laevis (Crustacea, Cirripedia) in Comodoro Rivadavia harbor (Chubut,
Argentina). Rev Mus Argentino Cienc Nat 3(2):175–179
Roldán JP (2014) Comunidades macrobentónicas incrustantes del sector portuario de Bahía
Blanca. Tesis de Licenciatura en Ciencias Biológicas. Univ Nac del Sur, Bahía Blanca, p 33
Rosa LC, Bemvenuti CE (2005) Effects of the burrowing crab Chasmagnathus granulata (Dana)
on meiofauna of estuarine intertidal habitats of Patos lagoon, southern Brazil. Braz Arch Biol
Technol 48(2):267–274
Rossi S, Bramanti L, Gori A, Orejas C (2017) An overview of the animal forests of the world. In:
Rossi S, Bramanti L, Gori A, Orejas C (eds) Marine animal forests. Springer, Cham, pp 1–26
Rossi S, Rizzo L, Duchêne JC (2019) Polyp expansion of passive suspension feeders: a red coral
case study. Peer J Preprints 7:e27490v1
Ruesink JL, Lenihan HS, Trimble AC, Heiman KW, Micheli F, Byers JE, Kay MC (2005)
Introduction of non-native oysters: ecosystem effects and restoration implications. Annu Rev
Ecol Evol Syst 36:643–689
Ruiz GM, Freestone AL, Fofonoff PW, Simkanin C (2009) Habitat distribution and heterogeneity
in marine invasion dynamics: the importance of hard substrate and artificial structure. In: Wahl
M (ed) Marine hard bottom communities. Springer, Berlin/Heidelberg, pp 321–332
Sakai K, Türkay M, Yang SL (2006) Revision of the Helice/Chasmagnathus complex (Crustacea:
Decapoda: Brachyura). Abh Senckenberg Naturforsch Ges 565:1–76
212 M. C. Carcedo et al.

Sal-Moyano MP, Gavio MA, Luppi TA (2012) Mating system of the burrowing crab Neohelice
granulata (Brachyura: Varunidae) in two contrasting environments: effect of burrow architec-
ture. Mar Biol 159(7):1403–1416
Schiffers K, Teal LR, Travis JMJ et al (2011) An open source simulation model for soil and sedi-
ment bioturbation. PLoS One 6:e28028
Schubart CD, Cuesta JA, Felder DL (2002) Glyptograpsidae, a new brachyuran family from
Central America: larval and adult morphology, and a molecular phylogeny of the Grapsoidea.
J Crustacean Biol 22(1):28–44
Shatkin GS, Shumway SE, Hawes R (1997) Considerations regarding the possible introduction of
the Pacific oyster (Crassostrea gigas) to the Gulf of Maine: a review of a global experience. J
Shellfish Res 16:463–477
Simonetti P, Botté SE, Fiori SM, Marcovecchio JE (2012) Heavy-metal concentrations in soft
tissues of the burrowing crab Neohelice granulata in Bahía Blanca estuary, Argentina. Arch
Environ Contam Toxicol 62(2):243–253
Simonetti P, Botté SE, Fiori SM, Marcovecchio JE (2013) Burrowing crab (Neohelice granu-
lata) as a potential bioindicator of heavy metals in the Bahía Blanca estuary, Argentina. Arch
Environ Contam Toxicol 64(1):110–118
Simonetti P, Botté SE, Marcovecchio JE (2018) Heavy metal bioconcentration factors in the bur-
rowing crab Neohelice granulata of a temperate ecosystem in South America: Bahía Blanca
estuary, Argentina. Environ Sci Pollut R 25(34):34652–34660
Snelgrove PV (1999) Getting to the bottom of marine biodiversity: sedimentary habitats: ocean
bottoms are the most widespread habitat on earth and support high biodiversity and key eco-
system services. Bioscience 49(2):129–138
Solan M, Wigham BD, Hudson IR et al (2004) In situ quantification of bioturbation using time
lapse fluorescent sediment profile imaging (f SPI), luminophore tracers and model simulation.
Mar Ecol Prog Ser 271:1–12
Sommer U (ed) (1989) Plankton ecology: succession in plankton communities. Springer,
Berlin, p 369
Sousa AI, Lillebø AI, Caçador I, Pardal MÂ (2008) Contribution of Spartina maritima to the
reduction of eutrophication in estuarine systems. Environ Pollut 156:628–635
Sousa R, Gutierrez JL, Aldrige DC (2009) Non-indigenous invasive bivalves as ecosystem engi-
neers. Biol Invasions 11:2367–2385
Spivak ED (1997) Cangrejos estuariales del Atlántico sudoccidental (25°-41°S) (Crustacea:
Decapoda: Brachyura). Investig Mar 25:105–120
Spivak E (2005) Los cirripedios litorales (Cirripedia, Thoracica, Balanomorpha) de la región del
Río de la Plata y las costas marinas adyacentes. In: Penchaszadeh PE et  al (eds) Invasores:
Invertebrados exóticos en el Río de la Plata y región marina aledaña. EUDEBA, Buenos Aires,
pp 251–309
Spivak ED (2010) The crab Neohelice (=Chasmagnathus) granulata: an emergent animal model
from emergent countries. Helgol Mar Res 64(3):149–154
Spivak ED, Gavio MA, Navarro CE (1991) Life history and structure of the world’s southernmost
Uca population: Uca uruguayensis (Crustacea, Brachyura) in Mar Chiquita lagoon (Argentina).
Bull Mar Sci 48:679–688
Spivak ED, Anger K, Luppi TA, Bas C, Ismael D (1994) Distribution and habitat preferences of
two grapsid crab species in Mar Chiquita lagoon (province of Buenos Aires, Argentina). Helgol
Meeresunters 48:59–78
Spivak ED, Anger K, Bas CC, Luppi TA, Ismael D (1996) Size structure, sex ratio, and breed-
ing season in two intertidal grapsid crab species from Mar Chiquita lagoon, Argentina.
Neritica 10:7–26
Sunday JM, Pecl GT, Frusher S, Hobday AJ, Hill N, Holbrook NJ et  al (2015) Species traits
and climate velocity explain geographic range shifts in an ocean-warming hotspot. Ecol Lett
18:944–953
8  The Intertidal Soft-Bottom Macrobenthic Invertebrates 213

Torres Jorda M, Roccatagliata D (2002) Population dynamics of Leidya distorta (Isopoda:


Bopyridae) infesting the fiddler crab Uca uruguayensis at the Río de la Plata estuary, Argentina.
J Crustacean Biol 22:719–727
Troost K, Gelderman E, Kamermans P, Smaal AC, Wolff W (2009) Effects of an increasing filter
feeder stock on larval abundance in the Oosterschelde estuary (SW Netherlands). J Sea Res
61:153–164
Truchet DM, Buzzi NS, Carcedo MC, Marcovecchio JE (2019) First record of the fiddler crab
Leptuca (= Uca) uruguayensis in the Bahía Blanca estuary (Buenos Aires, Argentina) with
comments on its biology in South America. Reg Stud Mar Sci 27:100539
Turner RD (1954) The family Pholadidae in the Western Atlantic and the Eastern Pacific. Part I -
Pholadinae. Johnsonia 3(33):1–63
Valentinuzzi De Santos S (1971) Estudio preliminar sobre las comunidades intercotidales del Pto
de Ing White (Pcia de Buenos Aires). Physis 30(81):407–417
Valiela I (ed) (1995) Marine ecological processes, 2nd edn. Springer, New York, p 686
Van Colen C, De Backer A, Meulepas G, Van Der Wal D, Vincx M, Degraer S, Ysebaert T (2010)
Diversity, trait displacements and shifts in assemblage structure of tidal flat deposit feeders
along a gradient of hydrodynamic stress. Mar Ecol Progr Ser 406:79–89
Van Oevelen D, Soetaert K, Middelburg JJ, Herman PMJ, others (2006) Carbon flows through a
benthic food web: integrating biomass, isotope and tracer data. J Mar Res 64:453–482
Villagrán DM, Truchet DM, Buzzi NS, Lopez ADF, Severini MDF (2019) A baseline study of
microplastics in the burrowing crab (Neohelice granulata) from a temperate southwestern
Atlantic estuary. Mar Poll Bull 110686
Vitousek PM, Mooney HA, Lubchenco J, Melillo JM (1997) Human domination of earth’s ecosys-
tems. Science 277:494–503
Webb AP, Eyre BD (2004) The effect of natural populations of the burrowing and grazing soldier
crab (Mictyris longicarpus) on sediment irrigation, benthic metabolism and nitrogen fluxes. J
Exp Mar Biol Ecol 309(1):1–19
Weis JS, Weis P (2004) Metal uptake, transport and release by wetland plants: implications for
phytoremediation and restoration. Environ Int 30:685–700
Widdows J, Brown S, Brinsley MD, Salkeld PN, Elliott M (2000) Temporal changes in inter-
tidal sediment erodability: influence of biological and climatic factors. Cont Shelf Res
20(10):1275–1289
Wilkie EM, Bishop MJ, O’Connor WA (2013) The density and spatial arrangement of the invasive
oyster Crassostrea gigas determines its impact on settlement of native oyster larva. Ecol Evol
3(15):4851–4860
Yamaguchi T, Henmi Y (2008) Cheliped differentiation and sex ratio of the fiddler crab Uca arcu-
ata. Crustacean Res 37:74–79
Zapperi G, Pratolongo P, Piovan MJ, Marcovecchio JE (2016) Benthic-pelagic coupling in an
intertidal mudflat in the Bahía Blanca Estuary (SW Atlantic). J Coast Res 32(3):629–637
Zapperi G, Piovan MJ, Pratolongo P (2017) Community structure and spatial zonation of benthic
macrofauna in mudflats of the Bahía Blanca estuary, Argentina. J Coast Res 34(2):318–327
Chapter 9
Taxonomic and Functional Assessment
of Subtidal Macrobenthic Communities
in the Bahía Blanca Estuary (Argentina)

M. Emilia Bravo, M. Cecilia Carcedo, Eder P. Dos Santos,


and Sandra M. Fiori

9.1  Introduction

The structural and functional characteristics of the macrobenthic soft-bottom inver-


tebrate communities are mainly controlled by properties of the habitat. This control
results from interplay of species-specific relationships with the physical and chemi-
cal properties of the sediment and surrounding seawater (Nikora 2010; Breine et al.
2018). Thus, the biomechanical properties of the benthic fauna are related to the
physical characteristics of the sediments (e.g., stability, shear strength, and texture)
and the water (e.g., intensity and dynamics of flow) (Nikora 2010). These variables,
together with the tolerance levels of each species to chemical characteristics (e.g.,
salinity, pH, organic matter content, dissolved oxygen), limit their distribution,
affecting the ecological configuration of the benthic communities they integrate
(Gray 2002; Teske and Wooldridge 2003; Breine et al. 2018). Other factors affecting
the ecological characteristics of benthic communities are interspecific relationships.
The relative importance of biotic versus abiotic factors in the structuring of com-
munities is poorly understood and is still being discussed, especially for the subtidal

M. E. Bravo ()
Instituto de Geociencias Básicas, Aplicadas y Ambientales de Buenos Aires (IGEBA-UBA-­
CONICET), Departamento de Ciencias Geológicas, Facultad de Ciencias Exactas y
Naturales, Universidad de Buenos Aires, Buenos Aires, Argentina
e-mail: mebravo@gl.fcen.uba.ar
M. C. Carcedo · S. M. Fiori
Instituto Argentino de Oceanografía, Universidad Nacional del Sur, CONICET, IADO,
Bahía Blanca, Argentina
Departamento de Biología, Bioquímica y Farmacia, Universidad Nacional del Sur,
Bahía Blanca, Argentina
E. P. Dos Santos
Instituto Argentino de Oceanografía, Universidad Nacional del Sur, CONICET, IADO,
Bahía Blanca, Argentina

© Springer Nature Switzerland AG 2021 215


S. M. Fiori, P. D. Pratolongo (eds.), The Bahía Blanca Estuary,
https://doi.org/10.1007/978-3-030-66486-2_9
216 M. E. Bravo et al.

benthic communities (Wilson 1990). Altogether, benthic communities integrate the


individual responses of each species to environmental properties in a dynamic way
across space and time. Due to their relative immobility, they are chronically exposed
to any environmental stressor, which makes them excellent indicators of environ-
mental health (Dean 2008). Understanding their response to environmental distur-
bances provides an insight into the level of stress and resilience of the marine
ecosystem (Ellis et al. 2017).
The characteristics of the environment can, in turn, be largely modified by the
benthic subtidal organisms. In estuaries, the structure of benthic communities regu-
lates ecosystem functioning through some biophysical processes such as bioturba-
tion and bioirrigation and biogeochemical processes such as nutrient cycling (Van
der Wal et al. 2017a). The benthic subtidal organisms are important secondary pro-
ducers and prey items of various benthic-feeding fish species (Van der Wal et al.
2017a; Bremec and Giberto 2017). Therefore, knowledge of their composition and
biological characteristics is important to establish the quality and quantity of food
available for species of fishing or conservation interest (Bremec and Giberto 2017).
At the same time, analyzing the ecosystem services in which benthic organisms are
involved may be a useful proxy of the resilience capacity of the marine system
(Dissanayake et al. 2018).
Human activities in estuarine environments often lead to large-scale changes
ranging from habitat modification to functional changes. For example, high rates of
sedimentation associated with anthropogenic activities can lead to loss of filter-­
feeding functional groups and replacement by deposit-feeders (Ellis et  al. 2017).
High-contaminant loads can lead to replacement by opportunistic fauna of smaller
body size and therefore with smaller cave depths, as well as with less longevity
(Dauer et al. 1992; Dean 2008). These changes lead to a lower biomass, oxygen-
ation, and reworking of sediments, among other consequences. Taxonomic and
functional changes in benthic communities can, in turn, lead to shifts in the ecosys-
tem services. Understanding the ecological and functional role of benthic organisms
is necessary for taking environmental management measures (Dissanayake
et al. 2018).
In Argentina, the subtidal environments of coasts have mainly been studied in
relation to the population dynamics of macrobenthic species with current or poten-
tial fishing value (Narvarte et al. 2007; Zaidman et al. 2016). For example, several
studies have been carried out on currently exploited commercial species, such as the
Tehuelche scallop Aequipecten tehuelchus (Soria et al. 2016), the spiral snail Zidona
dufresnei (Giménez et  al. 2005), the purple clam Amiantis purpurata (Morsán
2007), mussels Mytilus edulis platensis and Aulacomya atra, and the Patagonian
octopus Octopus tehuelchus (Narvarte et al. 2007). Only a small number of popula-
tion ecology studies have previously addressed non-commercial species, and these
were selected because of their dominance in the communities (Schiariti et al. 2006;
Güller and Zelaya 2017). Ecological studies of subtidal benthic communities have
mainly been focused on the relationship between benthic community distribution
and spatial heterogeneity in relation to sediment and salinity (Gilberto et al. 2004;
Diez et al. 2009; Kaminsky et al. 2018).
9  Taxonomic and Functional Assessment of Subtidal Macrobenthic Communities… 217

The subtidal macrobenthic communities of the Bahía Blanca Estuary were first
studied in the late 1980s (Elías 1987, 1992, 1995; Elías and Ieno 1993). These stud-
ies described the communities of subtidal macrobenthic invertebrates and their dis-
tribution patterns related to variations in water salinity and sediment grain size over
the estuary. After a 20-year period, a renewed interest in the status of the subtidal
macrobenthic communities was associated with the environmental monitoring pro-
gram required by Bahía Blanca Municipal Government (see Chap. 19). These moni-
toring programs were intended to update the knowledge base of the subtidal
communities in order to assess the status of the Bahía Blanca Estuary ecosystem in
relation to anthropogenic pressures. In addition, recent studies had focused on the
effect of wastewater discharge on benthic communities (Fiori et al. 2020). The sam-
pling sites studied by Elías (1992) and Fiori (2016) are shown in Fig. 9.1. The sub-
tidal seafloor of the Bahía Blanca Estuary has predominantly unconsolidated
sediments. Although there are sectors of the seabed with other morphological fea-
tures (sand dune fields, rocky outcrops, deep-holes, terraces), the seabed on the
shores of the channels analyzed in this study are the most representative of gradual
large-scale environmental variation (Fig. 9.1).

Fig. 9.1  Location of samples along the Principal Channel of the Bahía Blanca Estuary taken by
Elías in 1987 (yellow dots) and Fiori in 2016 (red dots)
218 M. E. Bravo et al.

9.2  P
 atterns of Abundance and Biomass of the Main
Taxonomic Groups

A central issue in community ecology is based on understanding the environmental


drivers of changes in community structure across space and how these determine the
distribution patterns of each species locally (Alves et  al. 2020). In estuaries, the
spatial variation of subtidal benthic communities is often related to salinity and
sediment texture gradients (van der Linden et al. 2012; Lana et al. 2018; Alves et al.
2020). This reflects the response of each species and therefore the gradual change in
community composition over a continuum (van der Linden et al. 2012; Lana et al.
2018; Alves et al. 2020). However, the relative importance of salinity with respect
to grain size varies between temperate estuaries. For example, abundance and bio-
mass in some northern hemisphere estuaries are related to salinity at the regional
level, but to variation in grain size and/or hydrodynamic conditions at the local level
(van der Linden et al. 2012; van der Wal et al. 2017a). On the other hand, in South
African estuaries, abundance and biomass patterns have a weak relationship with
the salinity gradient but a strong dependence on the sediment characteristics (Teske
and Wooldridge 2003). In Argentina, the strong variation in salinity and turbidity
along the Rio de la Plata Estuary is reflected in the appearance of a salt wedge, as
well as a turbidity front leading to complex patterns of spatial variation of benthic
communities along the gradient (Giberto et al. 2004).
In estuaries, polychaetes are often cited as the main group in terms of the abun-
dance and diversity of both species and biological features, whereas mollusks and
crustaceans are usually second-order groups (Pearson and Rosemberg 1978;
Ysebaert et al. 2002; Dean 2008; Van der Linden et al. 2012; Ellis et al. 2017), with
few exceptions, and they are usually associated with anthropogenic disturbance
(Muniz and Venturini 2015). Similarly, polychaetes are the dominant taxa over the
entire length of the Principal Channel in the Bahía Blanca Estuary, representing
more than 70% of the total abundance. Polychaetes are followed by mollusks in the
sectors located near the mouth of the second-order channels or by crustaceans at the
stations located in the Principal Channel (Fig. 9.2a). Not surprisingly, polychaetes
are the dominant group despite environmental changes in the estuary, considering
that they are one of the most abundant and diverse groups in marine communities.
They are capable of living in extremely diverse environments including some that
are extreme and inhospitable, such as highly sulfidic anoxic sediments (Dean 2008).
Also, the biodiversity of polychaetes is linked with diverse lifestyles. Some of the
most extreme cases of specialization cases include the bone-eating polychaetes
(Rouse et al. 2004) or those deriving energy from methane in deep-sea seeps and
vents (Levin et al. 2017). Polychaete specialization includes different species with
different adaptive strategies to pollutants, leading to specific responses to the differ-
ent levels of pollution (Dean 2008; Ellis et  al. 2017). Some of these specific
responses to pollution are shared between polychaetes belonging to the same fam-
ily, which is why several biological indices for assessing environmental quality are
based on polychaetes (Dean 2008).
9  Taxonomic and Functional Assessment of Subtidal Macrobenthic Communities… 219

Fig. 9.2  Spatial variability of macrobenthic communities along the Principal Channel with detail
of the relative contribution of the main taxa to (a) the total abundance and to (b) the total biomass

The relative contribution of secondary groups, like mollusks and crustaceans,


differs between the internal and external areas of the Principal Channel. In general,
mollusks are better represented in the inner zone of the Bahía Blanca Estuary, while
crustaceans reach greater abundances in the external zone. However, this pattern
associated with a linear environmental gradient is interrupted at the mouths of sec-
ondary channels. These local changes in benthic communities at the channel inter-
sections may be related to local changes in environmental conditions (van der Wal
et al. 2017a). The tidal channels and the Principal Channel are hydrodynamically
articulated and their intersections may have particular configurations of the seabed,
such as rocky outcrops, sand banks, and sand dunes (Ginsberg et  al. 2009). The
hydrodynamic and geomorphological complexity of the seafloor at the channel
intersections can drastically change the benthic communities at a local scale, even
leading to hot spots of diversity and/or biomass (van der Wal et al. 2017a). These
220 M. E. Bravo et al.

local interruptions of the large-scale estuarine environmental gradient can also be


found where the intersections of tidal channels lead to changes in salinity. This can
be generated by the presence of rivers and streams with daily or seasonal changes
in the amount of freshwater input van der Linden et al. 2012; van der Wal et al.
2017a; Lana et al. 2018). The seasonal changes in rainfall levels and consequent
freshwater discharges from rivers and streams in estuaries can create sectors known
as “micro-­estuaries” or sub-estuaries within euhaline or myxohaline zones (Lana
et al. 2018). As with biological responses to geomorphological and hydrodynamic
changes, species respond rapidly to these changes in salinity, and, consequently,
local changes in the composition of benthic communities can be observed (van der
Linden et al. 2012).
In addition, it is likely that local changes in overall patterns are related to anthro-
pogenic impact (Elias 1992; Fiori 2018; Fiori et al. 2020). Many studies on anthrop-
ically disturbed estuaries have found that the most stressful environments were
dominated by annelids (Dauer et al. 1992; Dean 2008; Kristensen et al. 2014). This
is a very diverse group, and a species-specific response to stress factors is expected
(Ellis et al. 2017) and will be discussed further (see Sect. 9.5). Some polychaetes are
often classified as tolerant or opportunistic to contamination (Dauer et  al. 1992;
Ellis et al. 2017), whereas amphipods are known to be sensitive to contaminated
sediments. On this basis, and since both groups are considered as ecological indica-
tors, the relationship between the relative abundances of both polychaetes and
amphipods is often used to measure environmental quality (Dauvin and Ruellet
2007). Similarly, in the Bahía Blanca Estuary, amphipods are the main group among
the crustaceans along the Principal Channel, but lower abundances were found at
the two stations that were closer to the sewage and industrial effluents (Fig. 9.2a).
The distribution pattern of biomass over the estuary shows that mollusks have a
higher biomass in the internal zone of the Principal Channel, whereas the biomass
of polychaetes, cnidarians, and crustaceans increases from the middle to the exter-
nal zone (Fig. 9.2b). Although crustaceans are relatively abundant, they do not rep-
resent an important contribution to the biomass due to their small body size (Fiori
2016). This is the opposite to the case of the cnidarians, because they are repre-
sented by the sea pen Stylatula darwini which is in low abundance, but it has a large
body size and therefore contributes more to the total biomass. This also happens in
the middle and outer zones of the Principal Channel, where the echinoderms, repre-
sented by ophiurans, contribute secondarily to the biomass in these zones (Fiori
2016). In the internal zone, the high dominance of mollusks is due to the bivalves
Corbula patagonica and Pitar rostratus (Fiori 2016), which in previous studies
were found to be dominant in the external zone (euhaline) in previous studies but
were related to very fine sandy sediments in all cases (Elías 1992). Both species
have also been mentioned as being abundant in relation to very fine sandy sediment
on the adjacent continental shelf “El Rincón” indicating a wide tolerance to varia-
tion in salinity by these species (Bremec 1990). Grain size preference may also
indicate that these species prefer bottom currents of low intensity, as expected for
some suspension-feeders (e.g., van der Wal et al. 2017a).
9  Taxonomic and Functional Assessment of Subtidal Macrobenthic Communities… 221

9.3  S
 patial Variations in Density, Biomass, Richness,
and Diversity

The mean density of macrobenthic invertebrates was similar along the Principal
Channel (Fig. 9.3a). The density is expressed here as a function of the sampling unit
(Van Veen grab) with a surface area of 0.18 m 2. The mean densities were similar in
the myxohaline and euhaline sectors, ranging from 83 individuals.grab−1 at the
innermost station to 104 individuals.grab−1 at the outermost station. Interrupting
this pattern, local changes with higher mean densities were found at the two stations
that were closer to the sewage and industrial effluents (Fig. 9.3a). This increase in
mean density associated with the sectors affected by sewage discharge is often
attributed to strategies of rapid recruitment of opportunistic benthic macrofauna
(Dauer et al. 1992).
The average biomass was similar along the Principal Channel with a slight
decrease from the internal to the external zone, except for changes at stations located
near to the port and sewage discharges where the total biomass was markedly lower
(Fig. 9.3b). In both cases, these are possibly stations with marked environmental
stress. The highest values of biomass were found at sites with the highest proportion
of fine sediments (Fig. 9.3e). In the estuary, these characteristics are associated with
environments with high rates of local deposition of sediments, nutrients, organic
matter, and contaminants (Carbone et al. 2016). Moreover, the internal zone has a
greater production of phytoplankton (Guinder et  al. 2012) that could favor
suspension-­feeding organisms, such as the bivalve Corbula patagonica, the main
contributor to the high biomass in the sector. Also, the high primary and secondary
productivity of the inner zone gives it unique qualities for feeding and reproduction
of several pelagic species, including species of commercial interest (Lopez Cazorla
2004). The area is also environmentally sensitive due to the combination of low
water renewal rates and high anthropogenic pressure (Carbone et  al. 2016).
Considering the high benthic biomass, it is important to evaluate its ecological func-
tions in relation to the retention and resuspension of pollutants (see Sect. 9.4).
The total species richness recorded in the Bahía Blanca Estuary was 155 species.
This value is intermediate between that found in closed estuaries, even in tropical
areas such as Brazil (Lana et al. 2018; Alves et al. 2020), and that of nearby open
coastal areas. This coincides with the geographical characteristics of the estuary,
which is semi-enclosed with little freshwater input concentrated in the interior zone,
resulting in a salinity gradient that is much less marked than in other estuaries
(Guinder et al. 2012). There are marked differences between myxohaline and euha-
line zones (Fig. 9.3c). The myxohaline zone of the Bahía Blanca Estuary has the
lowest species richness values, in ranges comparable to those found in other temper-
ate area estuaries (van der Linden et al. 2012; Bremec and Giberto 2017). The great-
est species richness is found in the external zone, where greater stability of
physical-chemical parameters can lead to a more balanced state in the macrobenthic
communities (van der Linden et al. 2012; van der Wal et al. 2017a). As for abun-
dance and biomass, species richness presented minimum values at the stations
affected by wastewater and industrial discharges (Fig. 9.3c). These stations also had
222 M. E. Bravo et al.

Fig. 9.3  Spatial variability of (a) mean density (ind.grab−1), (b) mean biomass (g.grab−1), (c)
Margalef species richness, (d) Shannon-Wiener diversity index (H′) of benthic communities in
relation to the variation of (e) grain size of sediments (phi units) and organic matter content (per-
centage) over the Principal Channel in the Bahía Blanca Estuary
9  Taxonomic and Functional Assessment of Subtidal Macrobenthic Communities… 223

the lowest species richness values of the three stations in the myxohaline zone, as
well as those in the euhaline zone, so they may lack species with the physiological
and behavioral adaptations necessary to overcome stress conditions, leading to
lower species richness (Dauer et al. 1992; Ellis et al. 2017).
The soft-bottom subtidal areas of the Principal Channel in the Bahía Blanca
Estuary were found to be inhabited by about 69 species according to Elias (1987),
but the latest research listed about 155 different taxa (Fiori 2016, 2018; Bravo
2019). All the species listed to date for the area are found in Table 9.1, with informa-
tion on the reference sources. Diversity by major taxa is also greater for the annelids
represented by 58 taxa, followed by mollusks (40 taxa) and crustaceans (30 taxa)
(Elias 1987; Fiori 2014, 2016, 2018). Other groups that constitute the benthic com-
munities are the ophiuroids, pennatulaceans, priapulids, bryozoans, hydrozoans,
pycnogonids, ostracods, nemerteans, and sponges (Elias 1987; Fiori 2014, 2016,
2018) (Fig. 9.4). The differences between the number of species found in the studies
carried out by Elias (1987) and that of more recent studies (Fiori 2014, 2016, 2018)
are due to the use of different mesh sizes during the separation of the organisms
from the sediment. Elias (1987) used a 2  mm mesh, whereas Fiori (2014, 2016,
2018) used a 0.5 mm mesh.
In the inner zone (myxohaline), with muddy sediments and very fine sand, the
species that dominate were the polychaetes Nicolea sp. and Leodamas verax.
Leodamas verax is an ubiquitous and abundant species that can be found along the
entire course of the Principal Channel and minor channels (Fiori 2016, 2018). Other
polychaetes such as Polydora sp. and Aphelocheta sp. are important in the interior
zone, decreasing towards the mouth of the estuary (Fiori 2016, 2018), while in the
external zone (euhaline), Axiothella sp. and the bivalve Corbula patagonica domi-
nate in the sandy mud sediments whose sand fraction is classified as fine to very fine
(Elías 1987; Fiori 2014, 2016, 2018). Finally, the middle zone is characterized by
Terebellides totae, associated to the sandy mud sediments (Elias 1987). Polychaetes
also dominate in the stations affected by sewage and industrial discharges, and it is
interesting that different families dominate in the different successional stages as
will be discussed later (Sect. 9.6).
Although most of the changes are observed when evaluating polychaetes, when
analyzing the secondary groups in abundance and diversity such as mollusks, they
were represented by 40 species (20 Gastropoda, 18 Bivalvia, 1 Polyplacophora, 1
Scaphopoda, Table 9.1). The bivalves Corbula patagonica, Nucula semiornata, and
Pitar rostratus were considered ubiquitous, mainly between 4 and 5  m depth.
Slightly higher abundances of the gastropod Buccinanops deformis and the bivalves
Macoma uruguayensis and Tagelus plebeius were found in shallower channels
(0.50  cm depth, Fiori 2018). In the channels affected by wastewater discharges,
mollusks represented by the bivalves C. patagonica, N. semiornata, P. rostratus,
Malletia cumingii, and Tagelus plebeius and the gastropod B. deformis reached the
same biomass as polychaetes (Fiori 2018; Fiori et al. 2020).
Species of the genus Corbula such as Corbula gibba have been classified as tol-
erant to a wide range of environmental disturbances such as coastal eutrophication,
and successful recruitment has also been recorded in areas subject to mass
224 M. E. Bravo et al.

Table 9.1  List of the species described for soft-bottom subtidal areas of the Bahía Blanca Estuary
Species Source
Annelida
Aglaophamus sp. Bremec et al. (1990)
Amphicteis sp. Bremec et al. (1990)
Amphicteis gunneri Elias et al. (2004)
Aphelocheta sp. Bravo et al. (2018) and Fiori (2018)
Aricidea sp. Bravo et al. (2018) and Fiori (2018)
Axiotella cf. constricta Elias et al. (2001)
Axiotella sp. Bremec et al. (1990), Elias et al. (2004) and Fiori
(2018)
aff. Chone sp. Bravo et al. (2018) and Fiori (2018)
Chone cf. striata Elias et al. (2004)
Capitella capitata Elias et al. (2001)
Cirratulidae Elias et al. (2004) and Fiori (2018)
Dodecaceria sp. Fiori (2018)
Dorvilleidae Fiori (2018)
Drilonereis orensanzi Bremec et al. (1990)
Eteone sp. Elias et al. (2004), Bravo et al. (2018) and Fiori
(2018)
Eusyllinae indet Bravo et al. (2018) and Fiori (2018)
Flabelligeridae Elias et al. (2004) and Fiori (2018)
Glycera americana Elias et al. (2004) and Bravo et al. (2018)
Goniadidae Elias et al. (2004)
Halosydna patagonica Bravo et al. (2018)
Halosydnella australis Elias et al. (2004)
Harmothoe sp. Elias et al. (2004)
Harmothoinae indet Bravo et al. (2018)
Kimbergonuphis sp. Elias et al. (2004)
Kimbergonuphis tenuis Fiori (2018)
Laeonereis acuta Elias et al. (2004), Bravo et al. (2018) and Fiori
(2018)
Leodamas verax Elias et al. (2004), Bravo et al. (2018) and Fiori
(2018)
Lepidonotinae Fiori (2018)
Lumbrineriopsis mucronata Bravo et al. (2018)
Lumbrineris tetraura Elias et al. (2004), Bravo et al. (2018) and Fiori
(2018)
Malacoceros sp. Bravo et al. (2018) and Fiori (2018)
Melinna uruguayi Bravo et al. (2018)
Myxicola infundibulum Elias et al. (2004)
Neanthes sp. Elias et al. (2004)
Nephtyidae Elias et al. (2004)
Nereididae Elias et al. (2004), Bravo et al. (2018) and Fiori
(2018)
(continued)
9  Taxonomic and Functional Assessment of Subtidal Macrobenthic Communities… 225

Table 9.1 (continued)
Species Source
Nicolea chilensis Elias et al. (2004)
Ninoe brasiliensis Bremec et al. (1990)
Ninoe falklandica Elias et al. (2004)
Onuphis setosa Bravo et al. (2018) and Fiori (2018)
Opheliidae Elias et al. (2004) and Fiori (2018)
Paraprionospio sp. Bremec et al. (1990)
Paraonidae Elias et al. (2001), Bravo et al. (2018) and Fiori
(2018)
Pectinaridae Fiori (2018)
Phylo felix Elias et al. (2004)
Polydora sp. Bravo et al. (2018) and Fiori (2018)
Scolecolepides sp. Elias et al. (2004)
Sabellariidae Elias et al. (2004)
Sabellaria nanella Fiori (2018)
Serpulidae Fiori (2018)
Spionidae Elias et al. (2004) and Fiori (2018)
Syllidae Elias et al. (2004)
Syllis sp. Bravo et al. (2018) and Fiori (2018)
aff. Telephus sp. Fiori (2018)
Terebella plagiostoma (ex Telephus Elias et al. (2004) and Bravo et al. (2018)
plagiostoma)
Terebellides totae Elias et al. (2004), Bravo et al. (2018) and Fiori
(2018)
Travisia sp. Bremec et al. (1990)
Tricobranchidae Fiori (2018)
Arthropoda
Artemesia longinaris Fiori (2018)
Austinixa patagoniensis Elias and Ieno (1993), Elias et al. (2004) and Fiori
(2018)
Cyrtograpsus affinis Fiori (2018)
Cyrtograpsus altimanus Elias (1992), Elias and Ieno (1993), Elias et al.
(2004), Bravo et al. (2018) and Fiori (2018)
Cyrtograpsus angulatus Elias and Ieno (1993), Elias et al. (2004) and Fiori
(2018)
Loxopagurus loxochelis Bremec (1990)
Neohelice granulata Elias and Ieno (1993) and Elias et al. (2004)
Pagurus criniticornis Bremec (1990)), Bravo et al. (2018) and Fiori (2018)
Pagurus exilis Bremec (1990)
Pagurus sp. Elias (1992), Elias and Ieno (1993) and Elias et al.
(2004)
Peisos petrunkevitchi Bravo et al. (2018)
Pilumnus reticulatus Elias et al. (2004) and Fiori (2018)
Pinnotheres sp. Elias and Ieno (1993) and Elias et al. (2004)
(continued)
226 M. E. Bravo et al.

Table 9.1 (continued)
Species Source
Pleoticus muelleri Fiori (2018)
Arthromysis magellanica Bravo et al. (2018) and Fiori (2018)
Neomysis americana Bravo et al. (2018) and Fiori (2018)
Ampithoe sp. Fiori (2018)
Caprella equilibra Fiori (2018)
Heterophoxus sp. Fiori (2018)
Heterophoxus videns Fiori (2018)
Monocorophium insidiosum (= Elias and Ieno (1993), Elias et al. (2004), Bravo et al.
Corophium sp.) (2018) and Fiori (2018)
Anthuridae Bremec (1990)
Rectarcturidae Fiori (2018)
Cristaserolis marplatensis (= Serolis Bremec (1990)
marplatensis)
Idotea sp. Fiori (2018)
Macrochiridothea sp. Bremec (1990)
Sphaeroma serratum Fiori (2018)
Synidotea marplatensis Fiori (2018)
Thysanoserolis elliptica (= Serolis Fiori (2018)
elliptica)
Monokalliapseudes schubarti (= Elias et al. (2001) and Bremec et al. (2017)
Kalliapseudes schubarti)
Anoplodactylus sp. Bravo et al. (2018) and Fiori (2018)
Mollusca
Aclis sp. (ex Pherusa sp.) Bremec (1990)
Adrana electa Bremec (1990), Elias et al. (2004) and Fiori (2018)
Amiantis purpurata Bremec (1990)
Anachis aff. Isabellei Bravo et al. (2018)
Barnea sp. Fiori (2018)
Bostrycapulus odites Fiori (2018)
Brachidontes rodriguezii Fiori (2018)
Buccinanops cochlidium (ex Bremec (1990)
Buccinanops gradatus)
Buccinanops deformis Bremec (1990), Elias et al. (2004), Bravo et al.
(ex Buccinanops globulosus) (2018) and Fiori (2018)
Buccinanops monilifer (ex Dorsanum Bremec (1990)
monilifer)
Carditamera plata Elias et al. (2004) and Fiori (2018)
Chaetopleura angulata Fiori (2018)
Columbellidae Fiori (2018)
Corbula patagonica Bremec (1990), Elias et al. (2004) and Fiori (2018)
Crepidula protea Fiori (2018)
Crepidula unguiformis Bremec et al. (1990)
(continued)
9  Taxonomic and Functional Assessment of Subtidal Macrobenthic Communities… 227

Table 9.1 (continued)
Species Source
Duplicaria gemmulata (ex Terebra Bremec et al. (1990)
gemmulata)
Ennucula puelcha Fiori (2018)
Entodesma patagonicum Fiori (2018)
Epitonium georgettinum Fiori (2018)
Eurytellina gibber Bremec (1990) and Elias et al. (2004)
Glycera americana Fiori (2018)
Heleobia australis Elias et al. (2004) and Bravo et al. (2018)
Kellia suborbicularis Fiori (2018)
Lyonsia alvarezii Elias et al. (2004) and Fiori (2018)
Mactra marplatensis Bremec (1990)
Macoma uruguayensis Elias et al. (2004)
Malletia cumingii Elias et al. (2004), Bravo et al. (2018) and Fiori
(2018)
Mitrella moleculina (ex Anachis Bremec (1990)
moleculina)
Notocochlis isabelleana (ex Natica Bremec (1990)
isabelleana)
Nucula semiornata Elias et al. (2004) and Fiori (2018)
Olivancillaria carcellesi Bremec (1990)
Olivella puelcha (ex Olivella plata) Bremec (1990) and Fiori (2018)
Olivella tehuelcha Elias et al. (2004)
Ostrea stentina Bravo et al. (2018) and Fiori (2018)
Parvanachis paessleri Elias et al. (2004)
Periploma compressum Elias et al. (2004)
Pitar rostratus Bremec et al. (1990), Elias et al. (2004), Bravo et al.
(2018) and Fiori (2018)
Pleurobranchaea inconspicua Fiori (2018)
Polyschides tetraschistus (ex Cadulus Bremec (1990)
quadridentatus)
Marginella prunum Bremec (1990)
Solen tehuelchus Bremec (1990)
Sphenia fragilis Elias et al. (2004) and Fiori (2018)
Thracia sp. Elias et al. (2004)
Tagelus plebeius Fiori (2018)
Tellina petitiana Fiori (2018)
Transenpitar americana Fiori (2018)
Echinodermata
Amphioplus albidus Bremec et al. (1990)
Amphiura crassipes Bremec et al. (1990)
Amphiura eugeniae Fiori (2018)
Encope emarginata Bremec et al. (1990)
Ophioplocus januarii Elias et al. (2004)
(continued)
228 M. E. Bravo et al.

Table 9.1 (continued)
Species Source
Amphipholis squamata Brogger pers. com.
Cnidaria
Amphisbetia operculata Fiori (2018)
Clytia sp. Fiori (2018)
Laomedea (Obelia) sp. Fiori (2018)
Obelia spp. Bravo et al. (2018)
Obelia bicuspidata Bremec et al. (1990)
Plumularia setacea Bremec et al. (1990)
Stylatula darwini Elias et al. (2004) and Bravo et al. (2018)
Symplectoscyphus sp. Fiori (2018)
Nemertea
Nemertino indet Elias et al. (2004)
Priapulida
Priapulus tuberculatospinosus Elias et al. (2004)
Bryozoa
Biflustra puelcha (ex Membranipora Bremec (1990)
puelcha)
Amathia imbricata Bravo (2019)
Scruparia ambigua Bravo (2019)
Anguinella palmata Bravo et al. (2018) and Fiori (2018)
Bugula neritina Bravo et al. (2018)
Bugulina simplex Bravo et al. (2018)
Bugulina stolonifera Bravo et al. (2018) and Fiori (2018)
Celleporella hyalina Bremec et al. (1990)
Conopeum sp. Bravo et al. (2018) and Fiori (2018))
Crisia sp. Bravo et al. (2018) and Fiori (2018)
Electra monostachys Bremec et al. (1990)
Membranipora sp. Bremec (1990) and Fiori (2018)

mortalities of other species. The association of Corbula patagonica with other


bivalves such as Ennucula puelcha, Malletia cumingii, and Macoma uruguayensis
was found in the Bahía Blanca Estuary and in the coast of Uruguay (Scarabino et al.
2006), while C. patagonica was associated with the bivalves Mactra marplatensis,
Angulus gibber, Adrana electa, Solen tehuelchus, and Pitar rostratus on the adja-
cent continental shelf (Bremec 1990). Special attention has been paid to the taxo-
nomic identification of the Nudibranchia Pleurobranchaea inconspicua in the Bahía
Blanca Estuary, due to the registration of the neurotoxin TTX (tetrodotoxin) in
specimens of Pleurobranchaea sp. aff. Maculata in Puerto Quequén (Farias et al.
2015), a coastal site very close to the Bahía Blanca Estuary (Fig. 1.2; Chap. 1).
Intake of only 1–2  mg of TTX can cause death in adult humans (Noguchi and
Arakawa 2008). High levels of toxins have been found in specimens of all ages of
Pleurobranchaea maculata (McNabb et  al. 2010), but so far no TXT has been
reported in specimens of P. inconspicua.
9  Taxonomic and Functional Assessment of Subtidal Macrobenthic Communities… 229

Fig. 9.4  Polychaetes found in subtidal soft-bottoms of the Bahía Blanca Estuary (a) Terebellides
totae (b) Leodamas verax (c) Corbula patagonica (d) Anachis sp. (e) Amphipholis squamata (f)
Anoplodactylus sp. (Photos by ME Bravo)

Crustaceans represent the third group in abundance and species richness, after
polychaetes and mollusks, with 30 species distributed in 6 sites along the Principal
Channel (Fiori 2018). The only species of crustaceans with a notable abundance
was the amphipod Corophium sp. in shallow subtidal areas of the in the inner zone
of the estuary, where it reached densities of up to 650 individuals/m2 (Elias and Ieno
1993). This low representation of crustaceans in the Bahía Blanca Estuary contrasts
with the adjacent continental shelf, “El Rincón.” In this area, crustaceans are domi-
nant in abundance, with tanaid (Kalliapseudes schubarti, sensu Elias et al. 2001;
Bremec et al. 2017) being the dominant species (Bremec 1990), reaching a domi-
nance of ~83%.
230 M. E. Bravo et al.

To date, only limited information is available on the temporal changes in the


ecological patterns of the benthic subtidal communities in the Bahía Blanca Estuary,
which includes some analyses of seasonal variation (Elias et al. 2004; Bravo 2019).
The highest values of density, biomass, richness, and diversity of species are found
during the spring (Elías et al. 2004; Bravo 2019), after the winter phytoplankton
bloom (Guinder et al. 2015). As in other temperate estuaries (Chainho et al. 2006;
Van der Linden et al. 2012; Quiroga et al. 2016), the maximum values of density,
biomass, species richness, and diversity are usually found in spring and summer,
after the blooms of primary producers (Chainho et al. 2006; Quiroga et al. 2016).
We are currently at the beginning of baseline studies, conducted systematically
since 2013 (see Chap. 19). The continuation of these studies will facilitate the evalu-
ation of temporal variations in the medium term. Alternatively, sampling with meth-
ods comparable to those of the studies carried out by Elias (1987) would make it
possible to evaluate some of the main changes that have occurred in recent decades.

9.4  Functional Ecology of Macrobenthic Species

Several studies focused on understanding the ecological functions of species and


their distribution patterns along environmental gradients in estuaries (e.g., van der
Linden et al. 2012; van der Wal et al. 2017a), mainly with coastal management and
conservation aims. This concept was introduced in conservation biology by Ehrlich
and Walker (1998) who proposed the protection of redundant functional groups in
order to maximize ecosystem resilience. Quantitative analysis for assessing the
functional redundancy may serve as a proxy of the resilience capacity of coastal
ecosystems, under the current scenario of increasing anthropogenic and climatic
pressures (Dissanayake et al. 2018). Different organization levels could be affected
by the loss of species within functional groups, for example, their specific ability to
adapt to environmental changes on different spatial scales and the productivity real-
ized for their community as well as the ecosystems services in which they are
involved (Ehrlich and Walker 1998; Wieters et al. 2012).
The abundance data of soft-bottom subtidal macrobenthic invertebrates from
Fiori (2016) was classified in function of different biological traits in order to ana-
lyze their variability along the Principal Channel. Information on the biological
traits of the different taxa has been sourced from previous analyses made in the
inner zone of the Bahía Blanca Estuary by Bravo (2019), as well as from taxa-­
specific literature or syntheses and suitable databases, e.g., polytraits (Faulwetter
et  al. 2014). Biological traits included feeding modes considering the categories
subsurface deposit-feeder, surface deposit-feeder, omnivore, carnivore, filter-feeder,
herbivore, and bacterial grazer; body size classified as small (<1  cm), medium
(1–5 cm), or large (>5 cm); motility classified as discretely motile, motile, or ses-
sile; calcification classified as non-calcified, heavily calcified, or lightly calcified;
and lifestyle classified as burrower, errant, tube builder, or attached (Table 9.2). The
category “not assigned” was used for classification of taxa of low taxonomic
9  Taxonomic and Functional Assessment of Subtidal Macrobenthic Communities… 231

Table 9.2  Main taxa representing the biological traits with the categories to which they were
assigned
Biological
traits Categories Main taxa
Feeding Subsurface Leodamas verax, Flabelligeridae undet., Aricidea sp., Malletia
mode deposit-­ cumingii, Ennucula puelcha, Nucula semiornata, Tellina petitiana
feeder
Surface Onuphis setosa, Polydora sp., Malacoceros sp., Aphelochaeta sp.,
deposit-­ Axiothella sp., Terebellides totae, Terebella plagiostoma,
feeder Dorvilleidae undet., Melinna uruguayi, Heleobia australis, Adrana
electa, Cyrtograpsus affinis, Cyrtograpsus altimanus, Serolis
elliptica, Heterophoxus videns, Neomysis americana, Arthromysis
magellanica
Omnivore Laeonereis acuta, Glycera americana, Lumbrineris tetraura,
Ninoe brasiliensis, Lumbrineriopsis mucronata, Lepidonotinae
undet., Buccinanops deformis, Pagurus criniticornis, Idotea sp.,
Monocorophium insidiosum, Anoplodactylus sp., Nematoda undet.
Carnivore Syllis spp., Epitonium georgettinum, Pleurobranchaea sp., Olivella
sp.
Filter/ aff. Chone sp., Myxicola sp., Sabellaria nanella, Serpulidae undet.,
suspension-­ Corbula patagonica, Pitar rostratus, Stylatula darwini, Ophiura
feeder sp.
Herbivore Chaetopleura angulata
Bacterial Eteone sp., Barnea sp.
grazer
Body size Small P. cornuta, Syllis spp., Aphelochaeta sp., Aricidea sp., Dorvilleidae
undet., E. georgettinum, E. puelcha, N. semiornata, Eteone sp., H.
australis, H. videns, M. insidiosum, Nematoda undet.
Medium L. verax, L. acuta, G. americana, O. setosa, L. tetraura, L.
mucronata, N. brasiliensis, Malacoceros sp., Flabelligeridae
undet., Axiothella sp., aff. Chone sp., T. totae, T. plagiostoma,
Lepidonotinae undet., M. uruguayi, Myxicola sp., M. cumingii, T.
petitiana, C. patagonica, B. deformis, P. rostratus, Barnea sp., A.
electa, C. angulata, Olivella sp., C. affinis, C. altimanus, P.
criniticornis, Idotea sp., Anoplodactylus sp., N. americana, A.
magellanica
Large Pleurobranchaea sp., S. darwini, Ophiura sp.
Motility Discretely L. verax, Malacoceros sp., Aricidea sp., Aphelochaeta sp.,
motile Axiothella sp., E. puelcha, N. semiornata, M. cumingii, T.
petitiana, C. patagonica, P. rostratus, Barnea sp., A. electa
Motile L. acuta, G. americana, L. tetraura, L. mucronata, N. brasiliensis,
Syllis spp., Eteone sp., Lepidonotinae undet., E. georgettinum,
Pleurobranchaea sp., B. deformis, C. angulata, Olivella sp., H.
australis, C. affinis, C. altimanus, P. criniticornis, Idotea sp., H.
videns, Anoplodactylus sp., N. americana, A. americana, Nemertea
undet., Nematoda, Ophiura sp.
Sessile P. cornuta, O. tenuis, Flabelligeridae undet., aff. Chone sp., S.
nanella, T. totae, T. plagiostoma, Serpulidae undet., M. insidiosum,
S. darwini
(continued)
232 M. E. Bravo et al.

Table 9.2 (continued)
Biological
traits Categories Main taxa
Calcification Non-­ L. verax, L. acuta, G. americana, L. tetraura, L. mucronata, N.
calcified brasiliensis, Syllis spp., Eteone sp., Lepidonotinae undet.,
Malacoceros sp., M. uruguayi, Myxicola sp., Aricidea sp.,
Aphelochaeta sp., Axiothella sp., P. cornuta, O. tenuis,
Flabelligeridae undet., aff. Chone sp., S. nanella, T. totae, T.
plagiostoma, Serpulidae undet., Nemertea undet., Nematoda.,
Pleurobranchaea sp.
Heavily E. georgettinum, E. puelcha, N. semiornata, M. cumingii, T.
calcified petitiana, C. patagonica, P. rostratus, Barnea sp., A. electa, B.
deformis, C. angulatus, Olivella sp., C. affinis, C. altimanus,
Ophiura sp.
Lightly P. criniticornis, Idotea sp., H. videns, M. insidiosum,
calcified Anoplodactylus sp., N. americana, A. magellanica, S. darwini
Lifestyle Burrower L. verax, G. americana, Malacoceros sp., L. tetraura, L.
mucronata, N. brasiliensis, Aricidea sp., Aphelochaeta sp., E.
puelcha, N. semiornata, C. patagonica, P. rostratus, Barnea sp.,
Nematoda undet.
Errant L. acuta, Syllis sp., Eteone sp., Lepidonotinae undet., E.
georgettinum, Pleurobranchaea sp., B. deformis, C. angulata,
Olivella sp., H. australis, C. affinis, C. altimanus, P. criniticornis,
Idotea sp., H. videns, Ampithoe sp., Anoplodactylus sp., N.
americana, A. magellanica, Nemertea undet., Ophiura sp.
Tube builder O. setosa, P. cornuta, M. uruguayi, Flabelligeridae undet.,
Axiothella sp., aff. Chone sp., Myxicola sp., S. nanella, T. totae, T.
plagiostoma, Serpulidae undet., M. insidiosum
Attached B. rodriguezii, S. darwini

resolution and/or a lack of taxa-specific literature. The percentage of a given bio-


logical trait category was calculated for each station along the Principal Channel,
and the results are shown in Fig. 9.5.
The characteristics that were equally represented in the macrobenthic organisms
along the Principal Channel were the deposit-feeding diet and the absence of calci-
fied structures, while the remaining biological traits showed differences between the
sites (Fig. 9.5). There were marked differences between the inner zone (stations 1,
2, and 3) and the outer zone (stations 4, 5, and 6). For example, the functional struc-
ture of the inner zone is dominated by small, discretely mobile, and burrowing
organisms, whereas the relative importance of medium-sized, mobile organisms
increases towards the outer zone (Fig.  9.5). The feeding guilds of benthic fauna
from the Bahía Blanca Estuary are mainly represented by subsurface- and surface
deposit-feeders, omnivores, filter/suspension-feeders, and carnivores. The greatest
trophic diversity is found at stations 1 and 6, which are the innermost and the most
external ones. Other biological traits (body size, motility, calcification, lifestyle)
show greater diversity and functional evenness from the middle zone (stations 3 and
4) to the outer zone (stations 5 and 6).
9  Taxonomic and Functional Assessment of Subtidal Macrobenthic Communities… 233

Fig. 9.5  Percentage of a given trait category at a given station, data from Fiori (2016). (a) Feeding
mode: subsurface deposit-feeder (SSD); surface deposit-feeder (SD); omnivore (O); carnivore (C);
filter feeder (F); herbivore (H); bacterial grazer (B); not assigned (NA). (b) Body size: small (S);
medium (M); large (L); not assigned (NA). (c) Motility: discretely motile (D); motile (M); sessile
(S); not assigned (NA). (d) Calcification: non-calcified (N); heavily calcified (H); lightly calcified
(L); not assigned (NA). (e) Lifestyle: burrower (B); errant (E); tube builder (T); attached (S)

Although the mechanisms and quantification of the effects of environmental


parameters, such as salinity, sediment grain size, organic matter content, and hydro-
dynamics, need to be better understood, it is already clear that they play an impor-
tant role in setting up both the taxonomic and the functional composition of benthic
communities. Some considerations below are still hypothetical and serve as illustra-
tions of ecological implications, but they are already supported by several studies
and may serve as potential research targets.
As well as the species-specific relationships of preference/tolerance with each of
the environmental variables that generate changes in community structure in taxo-
nomic terms, they also do so in functional terms. However, dominant diets did not
vary significantly along the gradient even with the replacement of dominant species
234 M. E. Bravo et al.

(Nicolea sp. and Leodamas verax in the inner zone, Terebellides totae in the middle
zone, and Axiothella sp. and Corbula patagonica in the outer zone). Nevertheless, a
greater representation of large body-sized organisms with a filter/suspension feed-
ing mode could be observed, such as Stylatula darwini in the internal zone and
Ophiura sp. and several bivalves in the external zone. Also, in the external zone,
there would be a greater representation of omnivorous and carnivorous organisms of
low abundance but larger body size, like some crustaceans, such as Cyrtograpsus
altimanus, and the polychaetes Glycera americana. On the other hand, as for the
taxonomic analysis, a greater functional diversity is observed in the euhaline zone.
The inner zone of the Bahía Blanca Estuary has relative low energy (Ginsberg
and Aliotta 2011) which may benefit the benthic fauna that are able to feed on
highly concentrated organic matter in fine grain size sediments (surface and subsur-
face deposit-feeders, burrowing organisms), in comparison to other functional
groups. At low flow rates, drag forces are reduced together with mixing rates (Nikora
2010), which coincide with a high content of organic matter and stability of sub-
strate, favoring infaunal organisms characterized by burrowing and deposit-feeding
habits (van der Wal et al. 2017a). On the other hand, the outer zone of the Bahía
Blanca Estuary has relative high energy, reflected in the presence of the largest sub-
tidal sand dunes of the estuary migrating at mean rates of 43  m  year −1 (Minor
Salvatierra et al. 2015). These strong bottom currents in the outer zone of the estu-
ary are associated with greater abundance of omnivores and carnivores, as well as
the higher mobility and body sizes of the benthic fauna. Strong bottom currents can
have large drag effects, even removing and destroying some animals, while high
mixing rates provide food and an efflux of wastes (Nikora 2010). These drag effects
may explain why the proportion of mobile and/or errant organisms increases with
increasing current velocities relative to that of burrowers. Also deposit-feeding
organisms are in low abundance in high-energy regions as the drag effects may flush
them out (Nikora 2010; van der Wal et al. 2017a), whereas the high mixing rates can
supply enough energy to sustain a higher diversity of feeding guilds (Nikora 2010;
van der Wal et al. 2017a). However, the patterns observed only represent momen-
tary conditions in the Bahía Blanca Estuary as hydrodynamic patterns in estuaries
change constantly, associated with both natural and anthropogenic drivers, such as
dredging activities (van der Wal et al. 2017a). It is expected that the distribution of
the ecological patterns linked with hydrodynamic characteristics will change
together with the morphology and hydrodynamics in the estuary.
Biological traits are proxies for understanding how organisms play crucial roles
in marine ecosystems (Dissanayake et al. 2018). In estuaries, different ecosystem
functions are regulated by biophysical processes mediated by benthic organisms
such as bioturbation, bioirrigation, and biogeochemical processes (van der Wal
et al. 2017a). Some macrobenthic organisms with biological traits such as burrow-
ing habits, deposit-feeding strategies, or simply having highly calcified bodies may
modify the physical structure of their own habitats, becoming ecosystem engineers
(Kristensen et al. 2014; Tait et al. 2020). These organisms add environmental com-
plexity, contributing disproportionally to ecosystem functions, and may buffer
anthropogenic stressors (Xie et al. 2018; Tian et al. 2019; Tait et al. 2020). Rather
9  Taxonomic and Functional Assessment of Subtidal Macrobenthic Communities… 235

than species richness per se, the presence, abundance, and biomass of ecosystem
engineers are the main drivers of ecosystem function (Tait et al. 2020). In this sense,
identification of the biological traits of benthic macrofauna involved in ecosystem
functions (e.g., bioturbation, habitat-formers) may help to improve the environmen-
tal management of estuaries.
Burrowing and deposit-feeding are biological traits associated with bioturbation
(Kristensen et al. 2014; Dissanayake et al. 2018). These biological traits are domi-
nant in the muddy sediments in the inner zone of the Principal Channel of the Bahía
Blanca Estuary and may change the oxygenation (redox potential), resuspension,
cycling of organic matter, nutrients, and pollutants (Dean 2008). The changes in
transport and biogeochemistry and, consequently, in the exchange of solutes
between the water-sediment interphase are related to burrowing organisms through
the construction and maintenance of burrows, as well as the ingestion and defeca-
tion of particles. Ventilation leads to rapid transport of solutes and is mediated by
the water currents generated by the respiration and feeding of animals (Kristensen
et al. 2014). Consequently, subsurface organic particles are transferred to the sur-
face oxide sediments, and vice versa, accelerating the degradation of organic matter
at the surface and favoring its resuspension by erosion (Kristensen et  al. 2014).
These functional groups are some of the first in the ecological succession (Dean
2008) to colonize environments enriched with organic matter, such as those found
at sites affected by sewage discharge in the Bahía Blanca Estuary (Fiori et al. 2020).
Metal concentrations are positively correlated with fine grain size sediments in
the Bahía Blanca Estuary (Marcovecchio et  al. 2010; Fernández Severini et  al.
2018). The inner zone of the estuary is particularly sensitive to the accumulation of
pollutants, driven by a combination of high anthropogenic pressure and a low rate
of water renewal. Although abiotic conditions are favorable for the precipitation of
contaminants on the seabed, a high load of bioturbator organisms favors their resus-
pension into the water column (Kristensen et al. 2014; Xie et al. 2018; Tian et al.
2019). High rates of bioturbation carried out by burrowers and deposit-feeding ben-
thic fauna in the inner sector of the estuary are expected to affect the dynamics of
biogeochemical processes, as well as promote release, resuspension, and bioavail-
ability of pollutant substances (Xie et  al. 2018; Tian et  al. 2019). In addition, as
subtidal macrobenthos constitute key prey items for fish of commercial interest,
such as Micropogonias furnieri and Mustelus schmitti, among others (Lopez Cazorla
2004), they may be driving the bioaccumulation of pollutants and even affecting
human health (Kristensen et al. 2014; Barletta et al. 2019; Fiori et al. 2020).
Dominance of bioturbators observed at stations 1 and 2 along the Principal
Channel can exclude some infaunal species due to changes in substrate conditions.
On the one hand, they can make the substrate more unstable (enhancing erosion)
and therefore exclude some infaunal species (such as discrete mobility or sessile
fauna). On the other hand, the enhanced resuspension of particles may exclude
fauna with filter/suspension feeding habits (Dean 2008; Kristensen et al. 2014). For
example, some burrowing polychaetes, such as Arenicola marina, dramatically
affect the species composition at the local scale through vigorous sediment mixing,
strong disturbance, and competition for food (Kristensen et al. 2014). It is possible
236 M. E. Bravo et al.

that similar effects are associated with the high load of bioturbator organisms at sta-
tions in the inner zone of the Bahía Blanca Estuary.
The resuspension of nutrients to the water column by burrowing species may be
important in the inner zone of the Bahía Blanca Estuary, where otherwise low veloc-
ity bottom water currents promote their deposition (Carbone et  al. 2016). The
reverse state can be promoted by suspension-feeders that cause particle deposition
and clear water (Kristensen et al. 2014). Burrowing and deposit-feeders are func-
tional groups with higher tolerance to high loads of nutrients and metals than sus-
pension feeders (Ellis et al. 2017). This may explain the low densities of suspension
feeders in sectors with high food availability but that are polluted and or eutrophi-
cated (Carbone et al. 2016) such as stations 2 and 4 in the Principal Channel.
The ecosystem engineering roles mentioned above are known as allogenic engi-
neering due to physical alteration of the environment mechanically or chemically
(Tait et al. 2020), whereas physical alteration of the environment by the bodies of
benthic organisms, such as habitat building, is known as autogenic engineering (Tait
et al. 2020). Heavily calcified and tube-building benthic organisms have biological
traits typical of autogenic engineering organisms (Tait et al. 2020). These traits are
mainly found in sandy sediments in the external part of the Bahía Blanca Estuary,
and they enhance the settlement of sessile epifauna, such as bryozoans and hydro-
zoans. However, they are not represented by gregarious organisms, so this would
not lead to any significant alterations in the three-dimensional configuration of the
subtidal environment as may occur with some mollusks (Nikora 2010).

9.5  Acoustic Approach in Subtidal Benthic Habitats

Analysis of large-scale variation in the ecological patterns of macrobenthic com-


munities would benefit of including acoustic classification of the seafloor. These
acoustic techniques are widely used by marine geologists and geophysicists to map
morphological features of the seabed with high spatial coverage in a short time
(Anderson et al. 2008). Several studies find a good fit between seabed acoustic clas-
sification and ecological patterns of benthic communities and are therefore increas-
ingly used for benthic habitat mapping (Anderson et al. 2008; van der Wal et al.
2017a, b; Mestdagh et al. 2020).
In the Bahía Blanca Estuary, acoustic anomalies derived from the presence of
hydrocarbon deposits (shallow gas) on the seafloor were found to be associated with
local changes in macrobenthic communities (see Box 9.1; Bravo et al. 2018, Bravo
et al. 2020). It is expected that larger local changes in the ecological patterns than
those described in this chapter were found associated with patchiness of mor-
phosedimentary features of the estuarine seabed. Secondary channels may have a
combination of muddy sediment patches alternating with sandy sediment patches
with or without sand dunes of different dimensions (Ginsberg et al. 2009). It is pos-
sible that these patches behave as different benthic habitats because of their differ-
ent textures and morphodynamics. In the estuary, sand dunes are usually small in
9  Taxonomic and Functional Assessment of Subtidal Macrobenthic Communities… 237

the inner zone (Gómez et al. 1996) and in secondary channels (Ginsberg et al. 2009)
and large in the external zone (Ginsberg and Aliotta 2011; Minor Salvatierra et al.
2015). These dunes provide good information about the dynamics of the bottom
currents (Ginsberg and Aliotta 2011; Minor Salvatierra et al. 2015), also affecting
the benthic communities of macroinvertebrates (Nikora 2010; Van der Wal et  al.
2017b). Other sectors, mainly in dredged sectors and at the intersection of channels
with intensified currents, may have patches with rocky outcrops hosting sessile ben-
thic macrofauna (Aliotta and Lizasoain 2004; Ginsberg et al. 2009). There is also a
wide range of bathymetric variations in the subtidal of the estuary, and benthic com-
munities usually respond to this variable. It is common to observe deep-holes in
areas of channel intersection, reaching relative relief of 2–17 m from the bottom of
the channels (Ginsberg and Perillo 1999). Overall, this highlights the intersections
of tidal channels as morphodynamically complex subtidal environments, with the
potential to host biodiversity hot spots (Van der Wal et al. 2017a). At the same time,
it highlights the potential impact of the deposition of dredging material at channel
intersections, as occurred during 2013 at the mouths of secondary channels on the
Principal Channel (Ginsberg et al. 2014).

9.6  Impacts and Conservation Issues

The impacts of human activities are cumulative and affect all parts of the oceans at
different scales, but the greatest impacts have been found in coastal and continental
shelf environments (Harris and Hughes 2012). Estuaries are among the most
impacted coastal ecosystems, being affected by a wide range of anthropogenic
activities as they serve as ports and waterways to urban and industrial centers (Saiz-­
Salinas and González-Oreja 2000). Kennish et al. (2014) identified a wide range of
anthropogenic factors that degrade and damage estuary ecosystems in the face of
climate change. These include nutrient loading and eutrophication, wastewater and
organic waste, habitat loss and alteration, chemical pollutants, sediment particle
inputs, overfishing, intensive aquaculture, introduced/invasive species, and altered
hydrological regimes. Human activities can have a cumulative impact on the struc-
ture, function, and ecological health of estuaries (Kennish  et  al. 2014). Affected
estuaries are expected to become less resilient to climate variability, most likely as
a result of the loss of biological and functional diversity (Dolbeth et al. 2007).
In the Bahía Blanca Estuary, the main anthropogenic pressures are related to
physical (dredging and deposition of dredging material), chemical (untreated urban
and industrial effluent discharge), and biological (introduction of exotic species)
impacts. Since benthic macroinvertebrates have a rapid and specific response to dif-
ferent types of stress, they are excellent indicators of environmental quality (Pearson
and Rosenberg 1978; Guidetti et al. 2000; Hampel et al. 2009; Dauvin et al. 2012).
Understanding the interaction of benthos and environmental disturbances could be
useful for identifying areas of high environmental susceptibility by facilitating man-
agement (Muniz et al. 2013; Dissanayake et al. 2018). This is even more important
238 M. E. Bravo et al.

considering the ecological role of the benthic fauna in relation to other biological
groups in the pelagic fraction.
The Bahía Blanca Estuary is the most important deepwater port system in
Argentina (see Chap. 2). This requires periodic maintenance of the Principal
Channel by dredging and the deposition of dredging material in intertidal areas for
the filling and expansion of the port complex (Ginsberg et al. 2014). These tasks
have been carried out for more than 30 years, with the most important work done in
1989, 1991, and 2013. Both dredging and the deposition of dredging material have
led to drastic changes in the morphology and sedimentology of the estuarine system
(Ginsberg et al. 2014). Most of the 2013 dredging deposits were made at the mouth
of secondary channels. This disrupts the natural hydrodynamics and balance
between erosion and deposition in these channels that normally function as a hydro-
dynamic network (Ginsberg et  al. 2014). Such changes are expected to alter the
ecological patterns of large-scale benthic communities. At the local level, stabiliza-
tion of dredging material deposits is associated with changes in the grain size of
shallow subtidal sediments (higher representation of coarse sands, gravels, and
poorer selection) related to lower specific richness and biomass, as well as changes
in the species composition of macrobenthic communities (Fiori et al. 2020). They
have also been associated with increased turbidity in the water column and higher
concentrations of metals (chromium, nickel, iron, and lead; Fiori et  al. 2020). In
addition to physically altering the benthic habitat, dredging could promote resus-
pension and/or translocation of contaminating substances trapped on the seafloor
(La Colla et al. 2018).
Through the monitoring program that evaluates the environmental quality of the
estuary (see Chap. 19), pilot sampling has been carried out to learn about the eco-
logical status of the subtidal macrobenthic communities in channels affected by
industrial wastewater and petrochemical discharges (Fiori 2018). This study repre-
sented the first integrated approach to explore the impact of anthropogenic activities
on the macrobenthic assemblages in the estuary. Preliminary results showed that the
sediments of these channels contain metals, polycyclic aromatic hydrocarbons
(PAH), hydrocarbon-degrading bacteria, and the fecal coliform bacteria Escherichia
coli (Fiori 2018). As expected, biological descriptors of the benthic community in
channels affected by anthropogenic effluents indicate that the community was
impoverished in terms of abundance and diversity. In the channel where the indus-
trial effluents are discharged, three species were numerically dominant: the poly-
chaetes Laeonereis acuta and Malacoceros sp., and the clam Tagelus plebeius (Fiori
2018). These polychaetes are deposit-feeding organisms, and the clam is a
suspension-­feeding organism (Holland and Dean 1977). In addition, T. plebeius is a
deep-digging species that inhabits permanent caves (up to 75 cm deep) that show
vertical movements in each tidal cycle (Holland and Dean 1977). These species can
affect the remobilization of some chemical compounds by their feeding mode or
burrowing activity, increasing the release of contaminants from the sediment to the
water column (Schaller 2014) or moving contaminants associated with suspended
matter from the water column to the sediment (Klerks et al. 1997). These changes
in the environmental distribution of contaminants will affect contaminant levels in
other organisms (Klerks 2018).
9  Taxonomic and Functional Assessment of Subtidal Macrobenthic Communities… 239

The area closest to the wastewater effluent discharge, where sediments were
anoxic with a high concentration of cadmium, lead, zinc, and copper, was domi-
nated by Laeonereis acuta (Fiori et al. 2020). Polychaetes can respond to exposure
to high concentrations of metals and toxins by forming metal granules, increasing
the secretion of mucus or accumulation in specific areas of the body followed by
exocytosis (Viarengo and Nott 1993; Gibbs et  al. 2000; Berthet et  al. 2003;
Mouneyrac et  al. 2003). According to toxicological tests, L. acuta exhibits some
tolerance to cadmium exposure. This seems to be related to a high concentration of
antioxidant enzymes that prevents damage caused by oxidative stress or formation
of insoluble metal granules in its tissues for detoxification (Sandrini et  al. 2006,
2008; Carricavur et al. 2018). Weis et al. (2017) showed that L. acuta is a species
tolerant of synergisms of diffuse contaminants that occupies a niche space that is
uncertain for other species more sensitive to contamination. The author found that
the effects of contaminants on molecular levels occur earlier, which may encourage
decision-making by managers and stakeholders in the use of preventive practices
before damage occurs at the population or community level (Weis et  al. 2017).
Individuals of L. acuta in contaminated Brazilian estuaries were two to three times
more exposed than those in undeveloped estuaries, suggesting that they may become
more attractive prey to consumers and favor bottom-up transfer of pollution (bio-
magnification) along the estuarine food web (Weis et al. 2017).
Another threat to local biodiversity associated with port activities is the introduc-
tion of non-native species that can become invasive. Scientists and policymakers are
increasingly considering invasive species as a major threat to the marine biodiver-
sity in coastal environments (Occhipinti Ambrogi et al. 2011). The effects of inva-
sion have been associated with decline in populations of indigenous species (Kappel
2005), changes in the food web (Oguz et al. 2008), and alterations to the structure
and functioning of entire ecosystems (Vilà et al. 2011, Simberloff et al. 2013). The
presence of non-native marine species in the Bahía Blanca Estuary has been docu-
mented from 1970 to the present, mainly associated with artificial hard substrate in
the ports (bryozoans, Lichtschein and Bastida 1980; tunicates and crustaceans
anemones, Molina et al. 2009; bivalves, Dos Santos and Fiori 2010). Many of these
species became dominant in the biofouling communities, even replacing native
biota. It has been proposed that accidental introduction through ship ballast water is
the most likely route of entry of these species into the ecosystem. In contrast, the
non-native species in the soft-bottom benthic communities of the Bahía Blanca
Estuary are unknown, since the inventory is still being completed and only a frac-
tion of the fauna has been identified at the species level (Fiori 2018). As already
mentioned by other authors, the main source of uncertainty regarding biological
invasions in the Southwest Atlantic is the poor documentation of marine biota
(Orensanz et al. 2002).
The continental shelf adjacent to the Bahía Blanca Estuary, known as “El
Rincón,” concentrates demersal fisheries, and so it has been studied more intensely
to characterize the feeding use of the seafloor by benthivorous fish of commercial
interest (Bremec and Giberto 2017). However, in the Bahía Blanca Estuary, the
capture of fish of commercial interest was historically carried out by artisanal
240 M. E. Bravo et al.

fishermen, with little political-economic influence in comparison to that exerted by


the petrochemical industry and port complex. This is reflected in the fact that both
the fish stock and the artisanal fishermen are strongly threatened by the growing
anthropogenic pressure in the area (Truchet et al. 2019; Speake et al. 2020). In turn,
this threatens the natural, historical, and cultural heritage of the Bahía Blanca
Estuary and associated urban centers (Speake et al. 2020). The combination of sci-
entific research and the knowledge of artisanal fishermen could be the key for the
correct development of management policies (Truchet et  al. 2019; Speake et  al.
2020). Research on the subtidal macrobenthic species could determine the quantity
and quality of food that supports species of commercial interest, such as the white-
mouth croaker (Micropogonias furnieri), narrownose shark (Mustelus schmitti),
rays, and stripped weakfish (Cynoscion guatucupa) (Lopez Cazorla 2004; Bremec
and Giberto 2017). This would not only allow the analysis of socioeconomic vari-
ables related to the quantity of fish stocks, but also it has health implications, espe-
cially considering that some of these species evidenced pollutant bioaccumulation
of which the bioavailability is linked to their benthic preys (Oliva et al. 2017).
Assessing the dynamics of pollutants in relation to the macrobenthic community
allows an ecosystem approach, which would include some ecosystem functions that
are affected or used by humans. The results discussed in this chapter indicate that
there are sectors within the estuary that are more sensitive than others in terms of the
biological features of the species. The inner zone of the estuary would be particu-
larly sensitive to pollution because of its low energy, shallow depth (high benthic-­
pelagic exchange), and dominance of bioturbating organisms. These organisms
favor bioremediation of the environment but also the trophic transference and mag-
nification of the pollutants. This would imply environmental and health sensitivity
considering that this area is a feeding, spawning, and breeding area for species of
conservation and fishing interest.

9.7  Closing Remarks

The sampling effort conducted over the last decade increased the number of species
identified along the Principal Channel and smaller channels affected by human
activities (see Chap. 19). Despite this progress, much research remains to be done
to complete the inventory of the biodiversity of the subtidal macrobenthic commu-
nities. Most of the estuary remains unexplored, and some species may have not been
discovered yet because they require further taxonomic research. For example, sev-
eral species of jellyfish have been reported in the study area, but their polyps have
never been found in benthic studies (Dutto et  al. 2017). This also indicates that
although many species in the Bahía Blanca Estuary are meroplanktonic, ecological
research on plankton and benthos has been conducted separately. At this point, the
analytical integration of the planktonic fraction would allow a holistic understand-
ing of the ecological processes involving the subtidal macrobenthos, as with infor-
mation regarding the marine geology of the area.
9  Taxonomic and Functional Assessment of Subtidal Macrobenthic Communities… 241

Box 9.1: Evaluation of the Effect of Shallow Gas on Benthic


Communities in the Bahía Blanca Estuary
In several coastal ecosystems of the world, the increase in the sea level that
occurred during Quaternary period generated the burial of continental organic
matter by fine grain size marine sediments. This created an anaerobic environ-
ment where organic matter decomposition mediated by microorganisms
formed methane gas. When methane gas concentrations exceeded the diffu-
sion capacity, formed bubbles trapped in their interstitial space due to the low
permeability of sediments. After a critic concentration (approximately 1%)
and thanks to the difference in acoustic impedance generated by the bubbles,
the gas deposits can be detected by seismo-acoustic methods, which are the
most commonly used in shallow gas mapping.
Currently, the process by which vertical migration of methane gas deposits
occurs remains not fully understood, but it is related with the relative perme-
ability of sediment layers located above and the physical processes affecting
them (e.g., erosion, deposition; Weschenfelder et al. 2016). This biogenic gas,
cumulated at different distances to the seabed surface, can migrate until be in
touch with the seabed surface or even escape through seabed to the water
column and to the atmosphere. There is an increasing interest in understand-
ing the roles played by the interaction of the seabed methane gas deposits with
the surrounding environment (biotic and abiotic). This significance of bio-
genic shallow gas is associated with both their practical and immediate impli-
cations. For example, gas can be considered as a potential energy resource but
also can impede engineering settlements by reducing shear strength of the
seabed (Fleischer et al. 2001). Additionally, recent findings suggest that we
are just beginning to understand its role in global carbon cycle and other bio-
geochemical cycles. For example, at local scale they are associated with sig-
nificant alterations in diagenetic and redox processes in seabed sediments
(Ramírez-Pérez et  al. 2015; Ramírez-Pérez and De Blas 2017). At global
scale some authors consider them as important sources of atmospheric meth-
ane emissions for climate change (Karisiddaiah and Veerayya 1994; Borges
et al. 2016). However, due to logistical limitations, quantitative measurements
of methane emissions from the seabed are scarce (García Gil et  al. 2011).
Research needs are becoming more important considering a possible increase
in methane emissions from the seabed associated with rising water tempera-
tures (Borges et  al. 2016) and expansion of hypoxic areas which promote
methanogenesis on buried organic carbon (Oppo et al. 2020) and can act in a
synergetic way.
Marine methane gas can release to the water column from deep, thermo-
genic sources at continental slopes. These gas seeps can act as energy source,
nursery, and refuge for local and surrounding benthic fauna (Grupe et  al.
2015; Grey 2016; Levin et al. 2016). Moreover, the biota that lives associated
with methane seepages is estimated to consume up to 80% of the total
242 M. E. Bravo et al.

methane before it reaches the atmosphere (Boetius and Wenzhöfer 2013).


Although the link between fauna and methane seems to be stronger at water
depths greater than 200  m, in coastal systems, hard substrate derived from
methane provides environmental complexity and suitability for sessile fauna
(Jensen et al. 1992). Nevertheless, when gas is trapped in sediments, its eco-
logical effect remains poorly understood. Most studies focused on more
noticeable effects of thermogenic gas seeps, while few studies exploring the
biological effect of shallow biogenic gas (e.g., Judd et al. 2002) have consid-
ered it as despicable. Nevertheless, its biological effect has received only iso-
lated, sparse attention at shallow depths, and recent studies made in the Bahía
Blanca Estuary indicate that these effects may had been underestimated
(Bravo et al. 2018, Bravo et al. in press).
Biogenic shallow gas deposits have been found widely distributed in sev-
eral coastal systems of South America, mainly in Brazil and Argentina
(Weschenfelder et  al. 2016). To date, in Argentina, its presence has been
reported in the Río de La Plata Estuary, the San Matías Gulf, the Beagle
Channel, and the Bahía Blanca Estuary (Weschenfelder et al. 2016). In the
Bahía Blanca Estuary, gas deposits are located at different depths to the sea-
bed covering extensive areas from the inner to the external zones of the
Principal Channel, even extending its distribution to the neighboring conti-
nental shelf “El Rincón” (Andreoli 2018; Bravo et al. 2018). The gas deposits
in the Bahía Blanca Estuary showed a tendency to be closer to the surface of
the seabed or even to be in contact with it towards the innermost zone (Bravo
et al. 2018). Wide sectors of the seabed surface of the subtidal area are char-
acterized by gas-bearing sediments. Additionally, many of these sectors are in
the area where the port complex is located (Bravo et al. 2018).
The wide coverage of gas-bearing sediments in the surface of the seabed in
the Bahía Blanca Estuary gives an excellent opportunity to evaluate its envi-
ronmental effects. The ecological effect of gas presence in soft-bottom seabed
at shallow water depths was studied by Bravo et al. (2018) in the Bahía Blanca
Estuary by comparing a gas site with a control site, both selected based on
seismic evidences. These sites were adjacent to each other. Each sample loca-
tion was selected based on the analysis of a seismic survey profile. This way,
it was established with high precision that both sites only differed in gas pres-
ence, while sharing water depths (2–3 m below the reference level), and geo-
morphological and sedimentological features between them (Bravo et  al.
2018; Bravo et al. in press).
In the Bahía Blanca Estuary, methane gas affects the physical and chemical
properties of seabed sediments (Bravo et  al. 2018; Bravo et  al. in press).
Despite sedimentological features commonly associated with gas when it is
actively seeping, such as carbonate cement precipitation and local sediment
lithification (e.g., Torry Bay, Scotland; Judd et al. 2002), were not observed in
the Bahía Blanca Estuary, gas-bearing sediments were characterized by
9  Taxonomic and Functional Assessment of Subtidal Macrobenthic Communities… 243

having lower shear strength and higher organic matter content than those from
a control site (Bravo et  al. 2018). The lower shear strength was coincident
with findings from other estuarine environments such as in Chesapeake Bay
(USA, Hill et al. 1992). Analyzing the seabed areas covered by gas-bearing
sediments as a soft-bottom benthic habitat, they behave as an unstable sub-
strate with more possibilities of resuspension (Bravo et al. 2018). Its organic
matter enrichment is a common feature of gas-bearing sediments, characteris-
tic shared with other estuaries and bays such as in Chesapeake Bay (USA, Hill
et al. 1992), San Simón Bay (Ría de Vigo et al. 2011), and Skagerrak (Western
Baltic Sea, Laier and Jensen 2007). Some authors had attributed the high
amount of organic matter in the gas-bearing sediments to the use of methane
as an alternative carbon source by the microbial consortium so that its organic
carbon oxidative capacity was exceeded (Hill et al. 1992). This would imply
that the vertical migration of buried methane, promoted by seawater warming,
may add organic carbon on the coastal seafloor (Bravo et al. in press). Without
considering the effect of methane per se, low shear strength and organic
enrichment of gas-bearing sediments may change structural and functional
configurations of faunistic groups associated to the soft-bottom seabed
(Bravo 2019).
There were identified a total of 48 taxa inhabiting gas and control sites,
which were represented differently (Czekanovski-Dice-Sorensen
index = 0.52) in terms of abundance. The differences between sites were also
observed in terms of Shannon diversity and evenness. Total abundance was
lower, while diversity and evenness were higher at gas site than at control site.
The dominant species in gas site, representing 66.7% of total abundance, was
the polychaete Aricidea sp., while in control site, the amphipod Monocorophium
insidiosum dominated the community, representing 72% of total abundance.
In comparisons of biomass, cnidarians were dominant at gas site, while at
control site annelids were the highest biomass group. Total biomass was
higher at the gas site, and this was explained by the contribution of the sea pen
Stylatula darwini which was absent at the control site. Until now the biology
of this species has been poorly studied. It is interesting to carry out research
on the possible preference for gas-bearing sediments by S. darwini, mostly
considering that some other species of sea pens have been reported associated
with gas seeps at continental slopes (Barrie et  al. 2011; Bigham 2016;
Palomino et al. 2016).
Shallow gas appears to affect abundance, composition, and distribution of
benthic species in the Bahía Blanca Estuary. In general, the benthic commu-
nity at gas site was dominated by polychaetes followed by crustaceans,
whereas at control site this relationship was reversed. This could be due to a
differential sensitivity of benthic organisms to the presence of gas in sedi-
ments. Previous studies attributed the macrobenthic community structure to
organic matter enrichment as a result of sewage discharges even at sites
244 M. E. Bravo et al.

distant from these disturbing sources (Elías 1987). Based on findings from
Bravo et al. (2018), higher organic matter content at gas sites appears to be
more related to methane than to sewage effluents.
Considering the expected fauna-sediment relations for the study area, char-
acterized by sandy-mud sediments with high organic matter content, the dom-
inant species should be deposit-feeders (Sanders 1958; Fauchald and Jumars
1979; Rhoads and Germano 1982). Nevertheless, this trend was only found at
control site, while dominant feeding guilds at gas site were suspension-feed-
ers, carnivores, and scavengers followed by surface deposit-feeders. At gas
site dominates the Paraonidae Aricidea sp. whose feeding guild varies between
species from deposit-feeder, carnivore, to suspension-feeder (Fauchald and
Jumars 1979). In contrast, control site is dominated by a detritivore,
Monocorophium insidiosum (Guerra-García et al. 2014). Some other authors
had found organisms belonging to family Corophiidae avoiding gas, consider-
ing them as sensitive to sulfide-rich sediments (Meadows et al. 1981; Judd
et al. 2002). As the methanogenesis process usually follows the hydrogen sul-
fide genesis (Judd 2004), it is expected that this gas could inhibit M. insidio-
sum in methane gas-bearing sediments in the Bahía Blanca Estuary.
Benthic communities at gas site showed a clearly different taxonomic
structure, a markedly lower total abundance of organisms but higher total bio-
mass. Some of the species from surroundig areas seems to avoid gas-bearing
sediments, which may be related with high content of organic matter, low
shear stress, and/or methane and hydrogen sulfide gas toxicity, whereas the
species found in gas site may be obtaining benefits such as exclusion of com-
petitors and/or energy source. For what we know, gas-bearing sediments in
subtidal seabed affect taxonomic structure of benthic communities inhabiting
them, but whether this leads to changes in their ecological functions should be
analyzed in future studies. Thus, the presence of biogenic shallow gas on the
seabed surface may behave as an important modelling factor for the benthic
habitat, making it therefore necessary to take shallow gas distribution into
account for researches on distribution of benthic communities. Understanding
its ecological effects will be favored by future studies in other coastal sectors
where shallow gas covers the surface of the seabed.

References

Aliotta S, Lizasoain G (2004) Tipos de fondos y su caracterización geológica por métodos sis-
moacústicos. In: Piccolo MC, Hoffmeyer MS (eds) Ecosistema del Estuario de Bahía Blanca.
Instituto Argentino de Oceanografía, Bahía Blanca, Argentina, pp 51–59
Alves AT, Petsch DK, Barros F (2020) Drivers of benthic metacommunity structure along tropical
estuaries. SC Rep 10(1):1–12
Anderson JT, Van Holliday D, Kloser R et al (2008) Acoustic seabed classification: current prac-
tice and future directions. ICES J Mar Sci 65(6):1004–1011
9  Taxonomic and Functional Assessment of Subtidal Macrobenthic Communities… 245

Andreoli A (2018) Análisis de los rasgos morfosedimentológicos de fondo en un sector de la


Plataforma Continental al sur de la provincia de Buenos Aires. Tesis de Licenciatura,
Universidad Nacional del Sur
Barrie JV, Cook S, Conway KW (2011) Cold seeps and benthic habitat on the Pacific margin of
Canada. Cont Shelf Res 31(2):S85–S92
Barletta M, Lima AR, Costa MF (2019) Distribution, sources and consequences of nutrients,
Persistent organic pollutants, metals and microplastics in South American estuaries. Sci Total
Environ 651:1199–1218
Berthet B, Mouneyrac C, Amiard JC, Amiard-Triquet C, Berthelot Y, Le Hen A, Smith BD
(2003) Accumulation and soluble binding of cadmium, copper, and zinc in the polychaete
Hediste diversicolor from coastal sites with different trace metal bioavailabilities. Archives of
Environmental Contamination and Toxicology 45(4):468–478
Bigham K (2016) Interpretation of the relationship between benthic fauna, geologic distributions,
and methane seeps at Southern Hydrate Ridge, Oregon continental margin
Borges AV, Champenois W, Gypens N et al (2016) Massive marine methane emissions from near-­
shore shallow coastal areas. Sci Rep 6:27908
Boetius A, Wenzhöfer F (2013) Seafloor oxygen consumption fuelled by methane from cold seeps.
Nat Geosci 6(9):725–734
Bravo ME (2019) Efecto de los depósitos de gas somero del fondo marino en las comunidades
macrobentónicas del estuario de Bahía Blanca. Doctoral thesis, Universidad Nacional del Sur
(UNS). Bahía Blanca (Argentina)
Bravo ME, Aliotta S, Fiori SM et al (2018) Distribution, vertical position and ecological implica-
tions of shallow gas in Bahía Blanca Estuary (Argentina). Estuar Coast Shelf Sci 202:222–231
Bravo ME, Fiori SM, Aliotta S et al (2020) Assessment of the impacts of shallow gas on taxonomic
structure of benthic communities in Bahía Blanca Estuary (Argentina). Estuar Coast Shelf Sci.
in press
Breine NT, De Backer A, Van Colen C et al (2018) Structural and functional diversity of soft-­bottom
macrobenthic communities in the Southern North Sea. Estuar Coast Shelf Sci 214:173–184
Bremec CS (1990) Macrobentos del área de Bahía Blanca (Argentina). distribución espacial de la
fauna. Bol. Inst. Oceanogr 38(2):99–110
Bremec CS, Elías R, Giberto DA (2017) El Rincón. In: Bremec CS, Giberto DA (eds) Comunidades
bentónicas en regiones de interés pesquero en Argentina. Instituto Nacional de Investigación y
Desarrollo Pesquero, Mar del Plata, pp 40–46
Bremec CS, Giberto DA (2017) Comunidades bentónicas en regiones de interés pesquero en
Argentina. Instituto Nacional de Investigación y Desarrollo Pesquero, Mar del Plata, 131 p
Carbone ME, Spetter CV, Marcovecchio JE (2016) Seasonal and spatial variability of macronutri-
ents and Chlorophyll a based on GIS in the South American estuary (Bahía Blanca, Argentina).
Environ Earth Sci 75(9):736
Carricavur AD, Boudet LC, Romero MB, Polizzi P, Marcovecchio JE, Gerpe M (2018)
Toxicological responses of Laeonereis acuta (Polychaeta, Nereididae) after acute, subchronic
and chronic exposure to cadmium. Ecotoxicology and environmental safety 149:217–224.
Chainho P, Costa JL, Chaves ML et  al (2006) Seasonal and spatial patterns of distribution of
subtidal benthic invertebrate communities in the Mondego River, Portugal—a poikilohaline
estuary. In: Marine Biodiversity. Springer, Dordrecht, pp 59–74
Dauer DM, Rodi AJ, Ranasinghe JA (1992) Effects of low dissolved oxygen events on the macro-
benthos of the lower Chesapeake Bay. Estuaries 15(3):384–391
Dauvin JC, Alizier S, Rolet C et al (2012). Response of different benthic indices to diverse human
pressures. Ecol Indic 12(1):143–153
Dauvin JC, Ruellet T (2007) Polychaete/amphipod ratio revisited. Mar Pollut Bull 55(1–6):215–224
Dean HK (2008) The use of polychaetes (Annelida) as indicator species of marine pollution: a
review. Rev Biol Trop 56(4):11–38
246 M. E. Bravo et al.

Diez MJ, Romero MC, Obenat S et al (2009) Distribución de Invertebrados Bentónicos en el Canal
Beagle, Argentina. Distribution of Benthic Invertebrates in the Beagle Channel, Argentina.
Anales del Instituto de la Patagonia 37(2):29–40
Dissanayake NG, Frid CL, Drylie TP et al (2018) Ecological functioning of mudflats: global analy-
sis reveals both regional differences and widespread conservation of functioning. Mar Ecol
Progr Ser 604:1–20
Dolbeth M, Cardoso PG, Ferreira SM et al (2007) Anthropogenic and natural disturbance effects
on a macrobenthic estuarine community over a 10-year period. Mar Pollut Bull 54(5):576–585
Dos Santos EP, Fiori SM (2010) Primer registro sobre la presencia de Crassostrea gigas (Thumberg,
1793) en el Estuario de Bahía Blanca,(Argentina)
Dutto MS, Genzano GN, Schiariti A et al (2017) Medusae and ctenophores from the Bahía Blanca
Estuary and neighboring inner shelf (Southwest Atlantic Ocean, Argentina). Mar Biodiver Rec
10(1):14
Ehrlich P, Walker B (1998) Rivets and redundancy. Bioscience 48(5):387–388
Elias R (1987) Estudio inventarial y ecológico del macrobentos de la Bahía Blanca. Doctoral the-
sis. Universidad Nacional de La Plata, Facultad de Ciencias Naturales y Museo. 262p
Elias R (1992) Quantitative benthic community structure in Blanca Bay and its relationship with
organic enrichment. Mar Ecol 13(3):189–201
Elías R (1995) The subtidal macroinfauna from soft-bottom substrate of the Blanca Bay
(Argentine). Thalassas 11:73–86
Elias R, Bremec CS, Giberto D et  al (2001) Estudio faunístico de las comunidades bentónicas
infaunales de El Rincón. Resultados de la campaña CC 14/00. Inf Téc Int DNI-INIDEP N°
8(/2001):11
Elias R, Ieno EN (1993) La comunidad de Laeonereis acuta Treadwell, 1923 (Polychaeta:
Nereididae) en la región interna de la Bahía Blanca. Iheringia, Ser Zool 75:3–13
Elias R, Iribarne OO, Bremec CS et al (2004) Comunidades bentónicas de fondos blandos. In:
Piccolo MC, Hoffmeyer MS (eds) Ecosistema del Estuario de Bahía Blanca. Instituto Argentino
de Oceanografía, Bahía Blanca, Argentina, pp 179–190
Ellis JI, Clark D, Atalah J et  al (2017) Multiple stressor effects on marine infauna: responses
of estuarine taxa and functional traits to sedimentation, nutrient and metal loading. Sci Rep
7:12013. https://doi.org/10.1038/s41598-­017-­12323-­5
Farías NE, Obenat S, Goya AB (2015) Outbreak of a neurotoxic side-gilled sea slug
(Pleurobranchaea sp.) in Argentinian coasts. New Zeal J Zool 42(1):51–56
Fauchald K, Jumars PA (1979) The diet of worms: a study of polychaete feeding guilds. Oceanogr
Mar Biol Annu Rev 17:193–284
Faulwetter MS, Markantonatou MV, Pavloudi MC et al (2014) Polytraits: a database on biological
traits of marine polychaetes. Biodivers Data J 2:e1024
Fernández Severini MD, Carbone ME, Villagran DM et al (2018) Toxic metals in a highly urban-
ized industry-impacted estuary (Bahia Blanca Estuary, Argentina): spatio-temporal analysis
based on GIS. Environ Earth Sci 77:393. https://doi.org/10.1007/s12665-­018-­7565-­5
Fiori SM (2014) Capítulo IV Comunidades bentónicas. In: Programa de Monitoreo de la calidad
ambiental del Estuario de Bahía Blanca. http://www.bahiablanca.gov.ar/areas-­de-­gobierno/
medio-­ambiente/comité-­tecnico-­ejecutivo/informes-­medio-­ambientales
Fiori SM (2016) Capítulo IV Comunidades bentónicas. In: Programa de Monitoreo de la calidad
ambiental del Estuario de Bahía Blanca. http://www.bahiablanca.gov.ar/areas-­de-­gobierno/
medio-­ambiente/comité-­tecnico-­ejecutivo/informes-­medio-­ambientales
Fiori SM (2018) Capítulo IV Comunidades bentónicas. In: Programa de Monitoreo de la calidad
ambiental del Estuario de Bahía Blanca. http://www.bahiablanca.gov.ar/areas-­de-­gobierno/
medio-­ambiente/comité-­tecnico-­ejecutivo/informes-­medio-­ambientales
Fiori SM, Bravo ME, Elias Ret AL (2020) Effects of sewage effluente on the subtidal macto-
benthic assemblage in a urban estuary (Argentina). Ecol Austral 1(30):134–145. https://doi.
org/10.25260/EA.20.30.1.0.954
9  Taxonomic and Functional Assessment of Subtidal Macrobenthic Communities… 247

Fleischer P, Orsi T, Richardson M et  al (2001) Distribution of free gas in marine sediments: a
global overview. Geo-Mar Lett 21(2):103–122
García-Gil S, de Blas E, Martínez-Carreño N et al (2011) Characterisation and preliminary quanti-
fication of the methane reservoir in a coastal sedimentary source: San Simón Bay, Ría de Vigo,
NW Spain. Estuar Coast Shelf Sci 91(2):232–242
Gibbs PE, Burt GR, Pascoe PL, Llewellyn CA, Ryan KP (2000) Zinc, copper and chlorophyll-
derivatives in the polychaete Owenia fusiformis. Journal of the Marine Biological Association
of the United Kingdom 80(2):235–248
Giberto DA, Bremec CS, Acha E et al (2004) Large-scale spatial patterns of benthic assemblages
in the SW Atlantic: the Rıo de la Plata estuary and adjacent shelf waters. Estuar Coast Shelf
Sci 61(1):1–13
Gimenez J, Lasta M, Bigatti G (2005) Exploitation of the volute snail Zidona dufresnei in Argentine
waters, Southwestern Atlantic Ocean. J Shellfish Res 24(4):1135–1140
Ginsberg SS, Aliotta S (2011) Sediment transport circulation pattern through Mesotidal channels
system. In: Ginsberg SS (ed) Sediment Transport. InTech, Vienna, Austria, pp 275–294
Ginsberg SS, Aliotta S, Lizasoain GO (2009) Sistema interconectado de canales de marea del
estuario de Bahía Blanca, Argentina: evaluación de la circulación de sedimento como carga de
fondo por métodos acústicos. Lat Am J Aquat Res 7(2):231–245
Ginsberg SS, Aliotta S, Vecchi L et al (2014) Vinculación entre el canal de acceso a los puertos de
Bahía Blanca y elsistema interconectado de canales de marea en relación a las tareas dedragado.
XIV Reunión Argentina de Sedimentología, La Plata, Asamblea Argentina de Sedimentología,
Asociación Argentina de Sedimentología
Ginsberg SS, Perillo G (1999) Deep-scour holes at tidal channel junctions, Bahia Blanca Estuary,
Argentina. Mar Geol 160(1–2):171–182
Gómez EA, Ginsberg SS, Perillo GM (1996) Geomorfología y sedimentología de la zona inte-
rior del Canal Principal del Estuario de Bahía Blanca. Revista de la Asociación Argentina de
Sedimentología 3(2):55–61
Gray JS (2002) Species richness of marine soft sediments. Mar Ecol Progr Ser 244:285–297
Grey J (2016) The incredible lightness of being methane-fuelled: stable isotopes reveal alternative
energy pathways in aquatic ecosystems and beyond. Front Ecol Evol 4:8
Grupe BM, Krach ML, Pasulka AL et al (2015) Methane seep ecosystem functions and services
from a recently discovered southern California seep. Mar Ecol 36:91–108
Guerra-García JM, De Figueroa JT, Navarro-Barranco C et  al (2014) Dietary analysis of the
marine Amphipoda (Crustacea: Peracarida) from the Iberian Peninsula. J Sea Res 85:508–517
Guidetti P, Modena M, La Mesa G, Vacchi M (2000) Composition, abundance and stratification of
macrobenthos in the marine area impacted by tar aggregates derived from the Haven oil spill
(Ligurian Sea, Italy). Marine Pollution Bulletin 40(12):1161–1166
Guinder VA, López-Abbate MC, Berasategui AA et al (2015) Influence of the winter phytoplank-
ton bloom on the settled material in a temperate shallow estuary. Oceanologia 57(1):50–60
Guinder VA, Molinero JC, Popovich CA et  al (2012) Dominance of the planktonic diatom
Thalassiosira minima in recent summers in the Bahía Blanca Estuary, Argentina. J Plankton
Res 34(11):995–1000
Güller M, Zelaya DG (2017) A hot-spot of biodiversity in Northern Patagonia, Argentina. Biodivers
Conserv 26(14):3329–3342
Hampel H, Elliott M, Cattrijsse A (2009) Macrofaunal communities in the habitats of intertidal
marshes along the salinity gradient of the Schelde estuary. Estuar Coast Shelf Sci 84(1):45–53
Harris PT, Hughes MG (2012) Predicted benthic disturbance regimes on the Australian continental
shelf: a modelling approach. Mar Ecol Progr Ser 449:13–25
Hill JM, Halka JP, Conkwright R et al (1992) Distribution and effects of shallow gas on bulk estua-
rine sediment properties. Cont Shelf Res 12(10):1219–1229
Holland AF, Dean JM (1977) The biology of the stout razor clam Tagelus plebeius: I. Animal-­
sediment relationships, feeding mechanism, and community biology. Chesap Sci 18(1):58–66
248 M. E. Bravo et al.

Jensen P, Aagaard I, Burke RA Jr et al (1992) Bubbling reefs' in the Kattegat: submarine land-
scapes of carbonate-cemented rocks support a diverse ecosystem at methane seeps. Mar Ecol
Progr Ser 83:102–112
Judd AG, Sim R, Kingston P et al (2002) Gas seepage on an intertidal site: Torry Bay, Firth of
Forth, Scotland. Cont Shelf Res 22(16):2317–2331
Judd AG (2004) Natural seabed gas seeps as sources of atmospheric methane. Environ Geol
46(8):988–996
Kaminsky J, Varisco M, Fernández M et al (2018) Spatial analysis of benthic functional biodiver-
sity in San Jorge Gulf, Argentina. Oceanography 31(4):104–112
Kappel CV (2005) Losing pieces of the puzzle: threats to marine, estuarine, and diadromous spe-
cies. Frontiers in Ecology and the Environment 3(5):275–282
Karisiddaiah SM, Veerayya M (1994) Methane-bearing shallow gas-charged sediments in the east-
ern Arabian Sea: a probable source for greenhouse gas. Cont Shelf Res 14(12):1361–1370
Kennish MJ, Brush MJ, Moore KA (2014) Drivers of change in shallow coastal photic systems: an
introduction to a special issue. Estuaries and coasts 37(1):3–19
Klerks PL, Leberg PL, Lance RF, McMillin DJ, Means JC (1997) Lack of development of pollut-
ant-resistance or genetic differentiation in darter gobies (Gobionellus boleosoma) inhabiting a
produced-water discharge site. Marine Environmental Research 44(4):377–395
Klerks PL, Kascak A, Cazan AM, Adhikary ND, Chistoserdov A, Shaik A, Louka FR (2018)
Effects of the razor clam Tagelus plebeius on the fate of petroleum hydrocarbons: A mesocosm
experiment. Archives of environmental contamination and toxicology 75(2):306–315
Kristensen E, Delefosse M, Quintana CO et al (2014) Influence of benthic macrofauna community
shifts on ecosystem functioning in shallow estuaries. Front Mar Sci 1:41
La Colla NS, Botté SE, Negrin VL et  al (2018) Influence of human-induced pressures on dis-
solved and particulate metal concentrations in a South American estuary. Environ Monit Assess
190(9):532
Laier T, Jensen JB (2007) Shallow gas depth-contour map of the Skagerrak-western Baltic Sea
region. Geo-Mar Lett 27(2–4):127–141
Lana P, Christofoletti R, Gusmão JB Jr et  al (2018) Benthic Estuarine Assemblages of the
Southeastern Brazil Marine Ecoregion (SBME). In: Brazilian Estuaries. Springer, Cham,
pp 117–175
Levin LA, Baco AR, Bowden DA et al (2016) Hydrothermal vents and methane seeps: rethinking
the sphere of influence. Front Mar Sci 3:72
Levin LA, Mendoza GF, Grupe BM (2017) Methane seepage effects on biodiversity and biological
traits of macrofauna inhabiting authigenic carbonates. Deep-Sea Res PT II 137:26–41
Lichtschein de Bastida V, Bastida R (1980) Los briozoos de las comunidades incrustantes de puer-
tos argentinos
Lopez Cazorla A (2004) Peces. El Ecosistema del Estuario de Bahía Blanca. In: Piccolo MC,
Hoffmeyer MS (eds) Ecosistema del Estuario de Bahía Blanca. Instituto Argentino de
Oceanografía, Bahía Blanca, Argentina, pp 191–201
Marcovecchio J, Botté S, Severini MF et al (2010) Geochemical control of heavy metal concentra-
tions and distribution within Bahia Blanca Estuary (Argentina). Aquat Geochem 16:251–266
McNabb P, Selwood AI, Munday R et al (2010) Detection of tetrodotoxin from the grey side-gilled
sea slug Pleurobranchaea maculata, and associated dog neurotoxicosis on beaches adjacent to
the Hauraki Gulf, Auckland, New Zealand. Toxicon 56:466–473
Meadows PS, Deans EA, Anderson JG (1981) Responses of Corophium volutator to sediment
sulphide. J Mar Biol Assoc UK 61:739e748
Mestdagh S, Amiri-Simkooei A, van der Reijden KJ et  al (2020) Linking the morphology and
ecology of subtidal soft-bottom marine benthic habitats: a novel multiscale approach. Estuar
Coast Shelf Sci 106687
Minor Salvatierra MS, Aliotta S, Ginsberg SS (2015) Morphology and dynamics of large subtidal
dunes in Bahia Blanca estuary, Argentina. Geomorphology 246:168–177
9  Taxonomic and Functional Assessment of Subtidal Macrobenthic Communities… 249

Molina LM, Valinas MS, Pratolongo PD, Elias R, Perillo GM (2009) First record of the sea anem-
one Diadumene lineata (Verrill 1871) associated to Spartina alterniflora roots and stems, in
marshes at the Bahia Blanca estuary, Argentina. Biological Invasions 11(2):409–416
Morsan E (2007) Spatial pattern, harvesting and management of the artisanal fishery for pur-
ple clam (Amiantis purpurata) in Patagonia (Argentina). Ocean & Coastal Management
50(5–6):481–497
Mouneyrac C, Mastain O, Amiard JC, Amiard-Triquet C, Beaunier P, Jeantet AY, Rainbow PS
(2003) Trace-metal detoxification and tolerance of the estuarine worm Hediste diversicolor
chronically exposed in their environment. Marine Biology 143(4):731–744
Muniz P, Lana P, Venturini N et al (2013) Un manual de protocolos para evaluar la contaminación
marina por efluentes domésticos. No. 577.727 M3.131 p
Muniz P, Venturini N (2015) Macrobenthic communities in a temperate urban estuary of high
dominance and low diversity: Montevideo Bay (Uruguay). CICIMAR Oceánides 30(1):9–20
Narvarte M, González R, Filippo P (2007) Artisanal mollusk fisheries in San Matías Gulf
(Patagonia, Argentina): an appraisal of the factors contributing to unsustainability. Fish Res
87(1):68–76
Nikora V (2010) Hydrodynamics of aquatic ecosystems: an interface between ecology, biome-
chanics and environmental fluid mechanics. River Res Appl 26(4):367–384
Noguchi T, Arakawa O (2008) Tetrodotoxin – distribution and accumulation in aquatic organisms.
and cases of human intoxication Mar Drugs 6:220–242
Oguz T, Fach B, Salihoglu B (2008) Invasion dynamics of the alien ctenophore Mnemiopsis
leidyi and its impact on anchovy collapse in the Black Sea. Journal of plankton research
30(12):1385–1397
Oliva AL, La Colla NS, Arias AH et al (2017) Distribution and human health risk assessment of
PAHs in four fish species from a SW Atlantic estuary. Environ Sci Pollut R 24(23):18979–18990
Occhipinti-Ambrogi A, Marchini A, Cantone G, Castelli A, Chimenz C, Cormaci M, Piraino S
(2011) Alien species along the Italian coasts: an overview. Biological invasions 13(1):215–237
Oppo D, De Siena L, Kemp DB (2020) A record of seafloor methane seepage across the last 150
million years. Sci Rep 10(1):1–12
Orensanz JML, Schwindt E, Pastorino G, Bortolus A, Casas G, Darrigran G, Vallarino EA (2002)
No longer the pristine confines of the world ocean: a survey of exotic marine species in the
southwestern Atlantic. Biological Invasions 4(1):115–143
Palomino D, López-González N, Vázquez JT et al (2016) Multidisciplinary study of mud volca-
noes and diapirs and their relationship to seepages and bottom currents in the Gulf of Cádiz
continental slope (northeastern sector). Mar Geol 378:196–212
Pearson TH, Rosenberg R (1978) Macrobenthic succession in relation to organic enrichment and
pollution of the marine environment. Oceanogr Mar Biol Ann Rev 16:229–311
Quiroga E, Ortiz P, González-Saldías R et  al (2016) Seasonal benthic patterns in a glacial
Patagonian fjord: the role of suspended sediment and terrestrial organic matter. Mar Ecol Prog
Ser 561:31–50
Ramírez-Pérez AM, De Blas E, García-Gil S (2015) Redox processes in pore water of anoxic sedi-
ments with shallow gas. Sci Total Environ 538:317–326
Ramírez-Pérez AM, de Blas E (2017) Iron reactivity in anoxic sediments in the Ría de Vigo (NW
Spain). Chemosphere 174:8–19
Rhoads DC, Germano JD (1982) Characterization of organism-sediment relations using sediment
profile imaging: an efficient method of remote ecological monitoring of the seafloor (Remots™
System). Mar Ecol Prog Ser 8:115–128
Rouse GW, Goffredi SK, Vrijenhoek RC (2004) Osedax: bone-eating marine worms with dwarf
males. Sci 305(5684):668–671
Saiz-Salinas JI, González-Oreja JA (2000) Stress in estuarine communities: lessons from the
highly-impacted Bilbao estuary (Spain). J Aquat Ecosyst Stress Recovery 7(1):43–55
Sandrini JZ, Regoli F, Fattorini D, Notti A, Inácio AF, Linde‐Arias AR, Monserrat JM (2006) Short‐
term responses to cadmium exposure in the estuarine polychaete Laeonereis acuta (Polychaeta,
250 M. E. Bravo et al.

Nereididae): Subcellular distribution and oxidative stress generation. Environmental


Toxicology and Chemistry: An International Journal 25(5):1337–1344
Sandrini JZ, Lima JV, Regoli F, Fattorini D, Notti A, Marins LF, Monserrat JM (2008) Antioxidant
responses in the nereidid Laeonereis acuta (Annelida, Polychaeta) after cadmium exposure.
Ecotoxicology and environmental safety 70(1):115–120
Sanders HL (1958) Benthic studies in Buzzards Bay. I. Animal-sediment relationships. Limnol
Oceanogr 3:245–258
Scarabino F, Zaffaroni JC, Clavijo C et al (2006) Bivalvos marinos y estuarinos de la costa uru-
guaya: faunística, distribución, taxonomía y conservación. En: Menafra R, Rodríguez-Gallego
L, Scarabino F et al (eds) Bases para la conservación y el manejo de la costa uruguaya. Fauna
Silvestre Uruguaya
Schaller J (2014) Bioturbation/bioirrigation by Chironomus plumosus as main factor controlling
elemental remobilization from aquatic sediments?. Chemosphere 107:336–343
Schiariti A, Berasategui AD, Giberto DA et al (2006) Living in the front: Neomysis americana
(Mysidacea) in the Río de la Plata estuary, Argentina-Uruguay. Mar Biol 149(3):483–489
Simberloff D, Martin JL, Genovesi P, Maris V, Wardle DA, Aronson J, Vila M (2013) Impacts
of biological invasions: what’s what and the way forward. Trends in ecology & evolution
28(1):58–66
Soria G, Orensanz JL, Morsán EM, Parma AM et al (2016) Scallops biology, fisheries, and man-
agement in Argentina. Dev Aquac Fish Sci 40:1019–1046
Speake MA, Carbone ME, Spetter CV (2020) Análisis del sistema socio-ecológico del estuario
Bahía Blanca (Argentina) y su impacto en los servicios ecosistémicos y el bienestar humano
Tait LW, Lohrer AM, Townsend M et  al (2020) Invasive ecosystem engineers threaten benthic
nitrogen cycling by altering native infaunal and biofouling communities. Sci Rep 10(1):1–11
Teske PR, Wooldridge TH (2003) What limits the distribution of subtidal macrobenthos in per-
manently open and temporarily open/closed South African estuaries? Salinity vs. sediment
particle size. Estuar Coast Shelf Sci 57(1–2):225–238
Tian S, Tong Y, Hou Y (2019) The effect of bioturbation by polychaete Perinereis aibuhitensis
on release and distribution of buried hydrocarbon pollutants in coastal muddy sediment. Mar
Pollut Bull 149:110487
Truchet DM, Noceti MB, Villagrán DM et al (2019) Fishers’ ecological knowledge about marine
pollution: what can FEK contribute to ecological and conservation studies of a Southwestern
Atlantic Estuary? J Ethnobiol 39(4):584–606
van der Linden P, Patrício J, Marchini A et  al (2012) A biological trait approach to assess the
functional composition of subtidal benthic communities in an estuarine ecosystem. Ecol Indic
20:121–133
van der Wal D, Lambert GI, Ysebaert T et al (2017a) Hydrodynamic conditioning of diversity and
functional traits in subtidal estuarine macrozoobenthic communities. Estuar Coast Shelf Sci
197:80–92
van der Wal D, Ysebaert T, Herman PM (2017b) Response of intertidal benthic macrofauna to
migrating megaripples and hydrodynamics. Mar Ecol Progr Ser 585:17–30
Viarengo A, Nott JA (1993) Mechanisms of heavy metal cation homeostasis in marine inver-
tebrates. Comparative Biochemistry and Physiology Part C: Comparative Pharmacology
104(3):355–372
Vilà M, Espinar JL, Hejda M, Hulme PE, Jarošík V, Maron JL, Pyšek P (2011) Ecological impacts
of invasive alien plants: a meta‐analysis of their effects on species, communities and ecosys-
tems. Ecology letters 14(7):702–708
Weschenfelder J, Klein AH, Green AN et al (2016) The control of palaeo-topography in the preser-
vation of shallow gas accumulation: examples from Brazil, Argentina and South Africa. Estuar
Coast Shelf Sci 172:93–107
Weis WA, Soares CHL, de Quadros DPC, Scheneider M, Pagliosa PR (2017) Urbanization effects
on different biological organization levels of an estuarine polychaete tolerant to pollution.
Ecological Indicators 73:698–707
9  Taxonomic and Functional Assessment of Subtidal Macrobenthic Communities… 251

Wieters EA, McQuaid C, Palomo G et al (2012) Biogeographical boundaries, functional group
structure and diversity of rocky shore communities along the Argentinean coast. PLoS
One 7(11)
Wilson WH (1990) Competition and predation in marine soft-sediment communities. Annu Rev
Ecol Evol S 21(1):221–241
Xie M, Wang N, Gaillard JF et al (2018) Interplay between flow and bioturbation enhances metal
efflux from low-permeability sediments. J Hazard Mat 341:304–312
Ysebaert T, Herman PM (2002) Spatial and temporal variation in benthic macrofauna and rela-
tionships with environmental variables in an estuarine, intertidal soft-sediment environment.
Marine Ecology Progress Series 244:105–124
Zaidman P, Kroeck MA, Van der Molen S, Williams G, Villalobos LG, Kissner EO, Morsan E
(2016) Local scale variation in the reproductive pattern of the southern geoduck, Panopea
abbreviata (Bivalvia: Hiatellidae), in Patagonia. Revista de biología marina y oceanografía
51(2):359–371
Chapter 10
Shrimps and Prawns

Patricia Marta Cervellini and Jorge Omar Pierini

10.1  Introduction

Marine shrimp are decapod crustaceans, distinguished from other crustaceans by its
thoracic appendages. While the first three pairs (maxillipeds) are modified for
feeding, the remaining five pairs (pereopods) are the walking legs, hence the name
Decapoda or “ten-legs.” They belong to the Suborder Dendrobranchiata, defined by
branched gills, prominent hinges on the pleon, larvae hatching as nauplii or
protozoeae, and the presence of a petasma in males (Tavares and Martin 2010;
Tavares et al. 2009). In its adult stage, they present a body segmented into 19 somites
or parts, divided into an anterior region or “head” covered by a rigid cephalothorax
(the carapace) and a posterior region, the articulated abdomen (pleon), that represents
the edible body part. The cephalothorax projects into a compressed tip or “face,”
with dorsal teeth, and, in some species, ventral teeth are also present.
The Suborder Dendrobranchiata is further divided into the Superfamilies
Penaeoidea and Sergestoidea. The Superfamily Penaeoidea is characterized by the
first three pairs of pereopods being chelate (ending in pincers) and of similar size
and shape (Pérez Farfante 1988). This group includes a little more than 350 species
of commercial interest; among them, about 100 species comprise most of the annual

P. M. Cervellini ()
Instituto de Ciencias Biológicas y Biomédicas del Sur, Universidad Nacional del Sur,
CONICET, INBIOSUR, Bahía Blanca, Argentina
Departamento de Biología, Bioquímica y Farmacia. Universidad Nacional del Sur.,
Bahía Blanca, Argentina
e-mail: pcervell@uns.edu.ar
J. O. Pierini
Instituto Argentino de Oceanografía (IADO-CONICET-UNS), Bahía Blanca, Argentina
Comisión de Investigaciones Científicas de la Provincia de Buenos Aires (CIC),
Bahía Blanca, Argentina

© Springer Nature Switzerland AG 2021 253


S. M. Fiori, P. D. Pratolongo (eds.), The Bahía Blanca Estuary,
https://doi.org/10.1007/978-3-030-66486-2_10
254 P. M. Cervellini and J. O. Pierini

world shrimp catches (Fransen 2014). The Superfamily Sergestoidea can be sepa-
rated from the penaeoids by the lack or reduction in size of the fourth and fifth pairs
of pereopods and by having a smaller number of gills. These shrimps occur in com-
mercial catches; however, because of their small size and soft bodies, they are of
little or no economic value (Pérez Farfante 1988).
Penaeoid and sergestoid shrimps are widely distributed, occurring in marine,
estuarine, and freshwater environments. Most marine species occupy shallow or
moderately deep waters, but some are found at depths of almost 5700 m. Although
many shrimps are pelagic, most are benthic, living on a large variety of bottoms
including mud, sand, rock, fragments of shells, or mixtures of these materials. In
addition, some species frequent coral reefs, and a few others live in association with
other invertebrates (Pérez Farfante 1988; Fransen 2014). They are animals that feed
on decaying particulate organic matter and various invertebrates that they capture
with their multiple appendices located on the head and thorax. The legs in the
abdomen (pleopods) are lined with fine silks that are used to swim, as well as
removing bottom sediments to get buried, especially during daylight hours
(Boschi 2004).
These shrimps are dioecious, and the external structures of the genital system are
the major dimorphic features. The male has two pairs of modified abdominal
appendages on the first and second abdominal segments (the petasma and appendix
masculina) that deliver sperm to the female’s external receptacle (the thelycum)
located between the bases of the fifth walking legs (Bailey-Brock and Moss 1992).
The thelycum can be closed (spermatic mass is internally placed on thelycum plates)
or opened (spermatic mass is exposed on thelycum region) (Dall et al. 1990). They
reproduce through copulating males and females, which release fertilized eggs that
are left free in the water, with exceptions such as certain small shrimps from the
Sergestoidea superfamily, which usually incubate eggs within their own bodies. The
reproduction occurs in the sea (spring to autumn), and depending on the species,
after fertilization the female stores the sperm (packs or spermatophore) in the
seminal receptacles. The embryo develops inside the egg, and free larvae are born
12–24 h after fertilization, depending on the temperature (Scelzo 2016).
The planktonic larva, called nauplius (pl. nauplii), is characterized by having
three pairs of appendices (antennules, antennas, and jaws) and an eye. Nauplius
feed on their own reserves (yolk) until they transform into zoea larvae (protozoeae)
that begin to feed on phytoplankton by filtration. Protozoeae are further transformed
into mysis larvae of carnivorous habits, which acquire adult morphology during
their evolution. There is a time span of approximately 20 days between the first and
last mysis larvae stages, before reaching the postlarval stage that also spans 20 days.
Postlarvae have adult morphology and modify their planktonic life form (life in the
water column) to a benthic behavior (on the substrate), acquiring after a few days
the appearance of adults and sexual maturity. The entire cycle lasts from 30 to 45
days. In the natural environment, postlarvae migrate from breeding areas to the
coast and frequently enter estuaries or brackish lagoons to feed (Iorio et al. 1990;
Scelzo 2016).
10  Shrimps and Prawns 255

10.2  Shrimps Present in the Bahía Blanca Estuary

Because of their economic value and abundance in estuarine and coastal ecosys-
tems, shrimps are widely studied, and the literature about them is extensive. In the
next paragraphs, we will provide updated information on the biological traits and
life cycles of three species of shrimps present in the Bahía Blanca Estuary, which
were selected based on their ecological and economic relevance. Two of them
belong to the Superfamily Penaeoidea, Pleoticus muelleri (Family Solenoceridae)
and Artemesia longinaris (Family Penaeidae), both species targets of artisanal and
industrial fisheries along the Argentine coast; the third species belongs to the
Superfamily Sergestoidea, Peisos petrunkevitchi (Family Sergestidae), which
represents an important link within local food webs.

10.2.1  The Argentine Red Shrimp Pleoticus muelleri

The Argentine red shrimp Pleoticus muelleri (Spence Bate, 1888) is a robust species
characterized by a short rostrum, with a series of teeth in its superior margin (Boschi
1963) and a typically intense pink color in their exoskeleton (Fig  10.1a). This
species is endemic to the southwestern Atlantic (Spivak et  al. 2019), distributed

Fig 10.1  Species of shrimps exploited in Bahía Blanca Estuary (a) Argentine red shrimp Pleoticus
muelleri; (b) Argentine stiletto shrimp Artemesia longinaris. (Photos by Sandra Fiori (a) and Juan
Manuel Alvarez (b))
256 P. M. Cervellini and J. O. Pierini

along South American coastal waters from Rio de Janeiro, Brazil, to Santa Cruz,
Argentina (Angelescu and Boschi 1959; Boschi 1963; Boschi and Gavio 2005) and
frequently found between 3 and 130 m depth (D’Incao 1999; Boschi et al. 1992).
There is a marked sexual dimorphism in this species, as well as large variations in
size and weight along its latitudinal range of occurrence.
In the Gulf of San Jorge, Argentina, the mean carapace length is 39 mm for males
and 43 mm for females. Further north, in subtropical coasts of southeastern Brazil,
mean carapace lengths are smaller (34.7 mm for females and 27.7 mm for males)
(Castilho et al. 2008). The highest adult abundance occurs in the Gulf of San Jorge,
in Patagonian waters, where there is the largest reproductive or stock population.
Minor stock populations locate offshore the Bahía Blanca Estuary and in front of
Mar del Plata (Buenos Aires Province) (Boschi 1986; Boschi and Selzo 1967, 1969,
1971; Cervellini and Mallo 1991; Iorio et al. 1996; Scelzo 2016).
The Argentine red shrimp thrives in water temperatures between 6°C and 23°C
and salinities from 31.5 to 33.5. Because of these wide ranges, populations show
different reproductive behaviors along its distributional range. While reproduction
is a continuous process in tropical areas, it shows strong seasonal oscillations in
higher latitudes (Boschi 1997; Costa et  al. 2004; Castilho et  al. 2008). In the
Argentine Sea, the reproductive season begins in late spring, and spawning peaks in
mid-summer, decreases during fall, and finally disappears in winter (Boschi 1986).
This behavior is supported by the presence-absence of mature, impregnated females
and larval stages in plankton samples. The spawning capacity of this species ranges
from 129,000 to 477,000 eggs per female (Macchi et al. 1992), and larval densities
rise up to 334 protozoa and 85 mysis per 100  m3 (Iorio et  al. 1996). Along the
Argentine coast, larvae of Pleoticus muelleri move large distances, between 120 and
300 nautical miles (about 220 and 550  km, respectively), transported by coastal
currents (Boschi 1989).
The Argentine red shrimp fishery is the main crustacean fishery in Argentina and
one of the most important marine resources in the southwestern Atlantic. Most of
the reported annual landings come from fishing grounds in the Patagonian region.
However, there is still a substantial contribution of the artisanal fleet along Buenos
Aires Province (Garza et al 2017)

10.2.2  The Argentine Stiletto Shrimp Artemesia longinaris

The Argentine stiletto shrimp Artemesia longinaris Spence Bate, 1888 is character-
ized by a long and sinuous rostrum (Boschi 1963). A. longinaris has a smaller size
than Pleoticus muelleri, although there are size variations across its geographical
range. In Buenos Aires Province, in Mar del Plata Port, males and females reach up
to 24 and 29 mm, respectively (Castilho et al. 2008). Adults are transparent with
pigmented spots, due to the presence of chromatophores (Gavio and Boschi 2016),
allowing blending and going unnoticed on sandy bottoms (Fig 10.1b). A. longinaris
is endemic to the southwestern Atlantic and distributed from Rio de Janeiro, Brazil,
10  Shrimps and Prawns 257

to Rawson, Chubut Province, Argentina (D´Incao 1999). The species is infra- and
circalittoral (Boschi et al. 1992; D’Incao 1999; Carvalho-Batista et al. 2011), plenti-
ful in shallow waters up to 20 m depth (Scelzo et al. 2002).
Populations of A. longinaris present phenotypic variations throughout the distri-
bution range. The body size and the mean size at sexual maturity increase with the
latitude, from São Paulo (Brazil) to Mar del Plata (Argentina), but decrease with
latitude from Rio de Janeiro to São Paulo. There are also latitudinal differences in
the reproductive period, which tends to be continuous in lower latitudes and seasonal
in higher latitudes (Carvalho-Batista et al. 2014). These variations may result from
reproductive adaptations to environmental factors, mainly temperature, nutrient
supply, and subsequent plankton production. There are also intrinsic physiological
limitations that constrain the reproductive behavior (Castilho et al. 2007). Despite
this phenotypic plasticity, there is no genetic differentiation among A. longinaris
populations through its entire geographical distribution. The genetic homogeneity is
maintained by larval dispersal and their high migratory capacity, which ensures
gene flow among populations (Carvalho-Batista et al. 2014).
Artemesia longinaris plays a critical role in trophic food webs of the southwest-
ern Atlantic. This shrimp inhabits soft bottoms composed of sand, silt, and clay.
Diet studies revealed that sand is the major component on stomach contents, and
sand grains are covered by films of bacteria, which are the base of shrimp nutrition
(Gavio and Boschi 2004). In turn, A. longinaris is a valuable food item for different
species of fish and other invertebrates. Shrimps have higher caloric content than
polychaetes, gastropods, and echinoderms (Thayer et al. 1973; Capitoli et al. 1994).
Because of its high biomass and energetic quality, a large number of carnivorous
fishes in the coastal waters of Buenos Aires Province prey upon these shrimps,
which constitute their main food source (Boschi 1963; Gavio and Boschi 2016). In
recent years, however, this species has become a common target of both artisanal
and industrial fisheries. The increasing fishing pressure and the need for management
strategies push knowledge about reproductive traits in this short-life cycle species.

10.2.3  The White Shrimp Peisos petrunkevitchi

The white shrimp Peisos petrunkevitchi Burkenroad, 1945 is a small shrimp species
endemic to the southwestern Atlantic from Rio de Janeiro, Brazil, through Uruguay
to Chubut, Argentina (Boschi et  al. 1992; D’Incao and Martins 2000). Females
reach a maximum of 45 mm, and the males being smaller than the females reach
between 10 and 30 mm (Boschi and Cousseau 2004). White shrimps present a short
rostrum, with two dorsal teeth (Costa et al. 2003). It is a typical species of coastal
ecosystems, especially at larval stages.
The presence, distribution, and abundance of P. petrunkevitchi larvae in plankton
samples were studied by Boschi and Scelzo (1969). The entire reproductive cycle
takes place in coastal waters, at less than 30 nm from the shoreline. The spawning
period lasts from October to December, as revealed by the appearance of eggs,
258 P. M. Cervellini and J. O. Pierini

nauplius, and protozoea in plankton samples (Mallo and Boschi 1982; Mallo and
Cervellini 1988; Cervellini and Mallo 1991). Larval development from free eggs
includes four nauplii, five elaphocaris (protozoea), one acanthosoma, and four
mastigopus (postlarvae) stages. Larval development until the first postlarva takes
45–48 days at temperatures of 18 ± 2°C (Mallo 1986). All larval and postlarval
stages of P. petrunkevitchi are active planktonic filter-feeders (Mallo and
Boschi 1982).
White shrimps is an active planktonic filter at all stages of development (Mallo
and Boschi 1982), feeding on diatoms and detritus. In turn, they represent an
important food resource for other marine animals such as fish, birds, and other
crustaceans (Omori 1974). Because of its high abundance and biomass, this is a key
species in food webs of coastal ecosystems. It has also been described as prey for
large whales, expanding the current knowledge on the possible trophic roles of this
species (Bortolotto et al. 2016). In the Bahía Blanca Estuary, P. petrunkevitchi was
reported as an important food item for juveniles of the striped weakfish Cynoscion
guatucupa (Sardiña and López Cazorla 2005a), the whitemouth croaker
Micropogonias furnieri (Sardiña and López Cazorla 2005b), and the Jenyns’ sprat
Ramnogaster arcuata (López Cazorla et al. 2011).

10.3  Migration of Shrimp Larvae and Postlarvae

Shrimps and prawns migrate from coastal to deep waters with trophic and repro-
ductive purposes (Lindley 1987; Wehrtmann 1989; Morgan 1990, 1992). The
mechanisms responsible for larval migration from the spawning areas towards
estuaries and lagoons vary according to the larval location and developmental
stage. However, the immigration of larvae through the estuarine inlets is regulated
by the tidal currents, and these become the main mechanism for the recruitment
of a variety of fishes and shrimps (Rothlisberg et al. 1995; Blanton et al. 1999;
Forward and Tankersley 2001). Several aspects of migration process have been
well documented. The number of larvae is generally greater during the flood tide
than during the ebb tide (DeLancey et  al. 1994; Rothlisberg et  al. 1995; Burke
et al. 1998; Jager and Mulder 1999). However, the number of larvae that migrate
from coastal areas into estuaries aided by flood tide transport depends on the num-
ber that arrives at the inlet of the estuary (Blanton et  al. 1999; Forward and
Tankersley 2001). A second aspect is that many larvae of estuarine-dependent
species concentrate outside the inlets in an optimal position to insure the best
transport during the next flood cycle (Young and Carpenter 1977; Calderón–Pérez
and Poli 1987; Poli and Calderón–Pérez 1987; Rothlisberg et al. 1995; Blanton
et al. 1999; Condie et al. 1999).
The spatial and temporal distribution of shrimp larvae and postlarvae in the
Bahía Blanca Estuary has been studied for more than two decades (Cervellini and
Mallo 1991; Cervellini 1992, 2001). This study has been carried out obtaining
samples of mesozooplankton monthly and during periods of active reproduction
10  Shrimps and Prawns 259

of the species, biweekly. During spring and summer (October and February),
shrimps move offshore to the estuary of Monte Hermoso and migrate to deep
waters 15 km from the coast, where salinity conditions are more stable (Fig. 10.2).
The females spawn, outside the limits of the estuary, thousands of eggs which are
fertilized by males. The eggs are derived in several larval stages and even change
their way of feeding (from phyto- to zooplanktophagous). The first stages feed on
microscopic algae (phytoplankton), but then they become zooplanktophagous
(microcrustaceans) and ingest detritus, and the postlarvae grow until they become
juveniles. That’s when they enter the estuary and swarm in the many canals, where
they find abundant food. The larvae and postlarvae stages of Peisos petrunkevitchi
have been captured outside the estuary (Cervellini and Mallo 1991). From the
results obtained from periodic sampling in the area, we can conclude that larvae
and postlarvae stages of Argentine red shrimp P. muelleri and A. longinaris are
exclusively marine; they do not migrate to coastal lagoons and/or estuaries as
other penaeid families, thus being distinguished from most commercial subtropi-
cal and tropical shrimps.

Fig. 10.2  Simulation of dispersal of Penaeidae larvae made with MOHDIS model. References:
Localization of larvae in green (a) starting point 00:00 hs 05/09/2003, (b) 16:30 hs 05/09/2003, (c)
16:30 hs 06/09/2003, (d) 16:30 hs 07/09/2003
260 P. M. Cervellini and J. O. Pierini

10.3.1  A Hydrodynamic Model

The spatial and temporal variability of physical, biological, and chemical processes
observed at the Bahía Blanca Estuary suggests that a more detailed study of the
processes controlling the dynamics of the area cannot be done based on field data
alone (Campuzano et al. 2014; Angeletti et al. 2018; Pierini et al. 2019 ). Several
measurement efforts would be necessary, and the associated costs would be high. A
possible solution for the lack of data is the use of numerical models as sophisticated
tools of interpolation and extrapolation of field data, both at spatial and temporal
domains. These tools allow the study of different scenarios, forecasting the
subsequent environmental answer as well as the determination of exchange fluxes
between the estuary and the ocean.
Then to investigate the possible migration of larvae and postlarvae shrimp from
penetrated, the estuary of Bahía Blanca was capture and the data obtained from
abundance (n° ind/m3) was incorporate into a numerical model. Samples of plankton
were collected at 13 stations located along the Principal Channel of the Bahía
Blanca Estuary. Data were collected from the monthly survey with Coast Guard GC
75 of the Argentine Naval Prefecture, which covered the mouth of the estuary. The
larvae were captured with a 50-cm-diameter plankton net that is 1.90 m long and
has 224 mm pore opening. Horizontal and oblique horizontal trawls last 10 minutes
at two knots. Samples were fixed with 4% formaldehyde. A flowmeter was used to
estimate the volume filtered by the network, and the number of larvae was converted
to ind/100 m3. Physical-chemical parameters were taken. The average abundance
was indicated throughout the year and the spatial distribution.
Numerical models have been implemented in the Bahía Blanca Estuary, includ-
ing studies of hydrodynamic characteristics (Pierini 2007; Rueda et al. 2013; Pierini
et al. 2012; Campuzano et al. 2014), analyses of potential effects of the wastewater
discharge system (Pierini et al. 2012, 2018), a study of the suspension sediment flow
(Campuzano et al. 2008), a study of larval dispersion and retention of crustacean
species (Cuesta 2010; Miguel 2010), and a study of marsh erosion due to the
dynamic interaction between the crab Neohelice granulata and the halophyte marsh
plant Sarcocornia perennis (Minkoff et al. 2006; Angeletti et al. 2018).
Given the large horizontal dimensions (70 km) relative to the vertical (10 m) in
the estuary, the vertical velocities and accelerations are small relative to the
horizontal components. Thus, the processes generating quantity of movement and
transport occur at a different scale in the vertical and horizontal directions. Due to
this fact, the circulation on these domains is mainly horizontal which implies that
vertical accelerations can be ignored when compared with the gravity effect.
Therefore, the vertical equation of motion may be replaced by the hydrostatic
pressure approximation (Campuzano et al. 2014; Pierini et al. 2019).
The MOHID model was used to simulate the hydrodynamics 2D vertically aver-
aged domain model with a horizontal resolution of 0.01° covering from coordinates
−61.41W, −39.38S to the inner Bahía Blanca Estuary. The model was calibrated
and validated by Campuzano et al. (2014). To adequately represent the study areas,
10  Shrimps and Prawns 261

we used a modified version of the high-density bathymetry (50 × 50 m grid) devel-


oped by Pierini (2007) with bathymetric data used to compose the model domain
coming from two sources: the GEBCO digital atlas, a one-minute global bathymet-
ric grid database, and data from the Bahía Blanca Port Management Consortium
(CGPBB) with a waterline obtained from the evaluation of six sets of Landsat 5 TM
and Landsat 7 ETM data resulting in a high-density bathymetry (Pierini 2007).
Monthly average values of discharge from the main tributaries of the estuary (Sauce
Chico River and Napostá Grande Creek) were based on monitored data (Pierini
2007; Campuzano et al. 2014). The tide data were provided by the Bahía Blanca
Port Management Consortium. Wind intensity (vector module) and wind direction
were provided by the CGPBB and by the private company Oiltanking Ebytem,
located close to Villa del Mar (6 km) (Fig. 1.4, Chap. 1). The waves were considered
a fetch-based wave generation model to predict the significant wave height and
wave period inside the Bahía Blanca Estuary. The fetch wave model uses as input
the fetch distances, water depth, wind velocity, and direction in each grid cell. The
fetch distances were calculated for 16 directions. As wave propagation is not con-
sidered explicitly, this model is only adequate for areas where the ocean does not
directly affect the waves. This is the case of our study area due to the geometry of
the inlet channel, where only local wind waves are generated.
The MOHID model initially was calibrated and validated in the study area and
incorporated with the information obtained at one of the sampling sites and later
coupled to Lagrangian (particle-tracking) module, including larvae behavior. The
MOHID model calculates the movement of particles that simulate larvae dynamics.
In this case, it was implemented to adequately represent the larval migration of
shrimp Penaeoidea. The model tracked the trajectories of larvae and was forced
with tide, creek discharges (Napostá and Sauce Chico), and local wind conditions
which occurred in April 2004, in order to capture a range of environmental
variabilities experienced by larvae and postlarvae. We examined whether larval or
postlarval behavior and tidal conditions could influence dispersal distance. The grid
used in the model is 50 × 50 m. The number of particles released in Puerto Coronel
Rosales zone was proportional to the density of larvae and the area of the grid size.
Several generalizable results emerge from the present Lagrangian simulation model
of larval dispersal. First, predictions of larvae dispersal can be developed from
knowledge of the mean and fluctuating hydrodynamic components and wind
conditions, coupled with basic information about larvae life history characteristics.
Second, expected dispersal scales range from 1 to 100  km depending on these
characteristics (Fig. 6.5). In the study and Lagrangian results shown that the larvae
and postlarvae of shrimp Penaeoidea were recorded only at the mouth of the estuary
of Bahía Blanca and distributed throughout north shore of the Bahía Blanca Estuary
where the abiotic factors, mainly salinity, are more constant resembling those of the
adjacent continental shelf (Cervellini 1988; Cervellini and Mallo 1991; Cervellini
1992; Perillo and Piccolo 1999; Delgado et al. 2017).
It is known that larval stages, particularly of penaeoid shrimp, require stable sea
conditions, since small variations affect larval survival (García and Le Reste 1986).
Larvae are found near the Faro buoy, at the height of Monte Hermoso, with certainty
262 P. M. Cervellini and J. O. Pierini

that the adults leave the estuary to spawn, where a very complex biological cycle
begins and the new specimens enter the estuary as juveniles, to gain weight. Such
habitat would offer shelter and abundance of food at early stages of the species.
They eat microscopic organisms, such as plankton.
Lastly, future integrative research on dispersal patterns, ocean currents, and lar-
val behaviors will be necessary to determine the relative contribution of oceanogra-
phy and behavior to realized dispersal patterns.

10.4  Shrimp Fisheries in Argentina

The Argentine red shrimp, Pleoticus muelleri, is a target of coastal fisheries along
its entire distributional range. In Brazil and Uruguay, the species is a resource of
minor importance, but in Argentina it is the commercially most important crustacean,
and its fishery is among the 25 major shrimp fisheries in the world (Poli and
Calderon-Perez 1987). Based on biological and fisheries data, including reproduction
areas and seasons, growth, recruitment, and fisheries concentration, three
independent populations have been recognized in Argentine waters for P. muelleri:
(1) Mar del Plata Port and neighboring areas up to Querandí Lighthouse (25 miles
north from Mar del Plata Port), (2) Southern Buenos Aires Province including the
Bahía Blanca Estuary, and (3) the Patagonian region (mainly Chubut Province).
From the several past decades, landings of this species have been highly variable,
but in the last years their catches have consistently increased (FAO 2016; SAGPyA
2017). The fluctuations in abundance of P. muelleri are related to changing
environmental and oceanographic conditions, which could cause variable mortalities
in early life stages. In consequence, there is considerable uncertainty in the
sustainable yields (de la Garza et al. 2017). This species is currently captured up to
the maximum advisable levels. Landings in 2016 were equivalent to 25% of the
total landings of all the marine/continental species combined in Argentina (de la
Garza et al. 2017).
Globally, bycatch, or the incidental catch of nontarget organisms, is a major
issue in shrimp trawl fisheries, which may have detrimental effects on marine
resources and wildlife (Gillet 2008; Kelleher 2008). More than a hundred spe-
cies are caught by shrimp trawlers in the Patagonian region, and the Argentine
hake, Merluccius hubbsi, is the most frequent and abundant species in these
incidental catches (Bertuche 1999; Pettovello 1999; Gongora et al. 2009). Hake
is the most important demersal species for the commercial fishing in Argentina
(Cordo 2005). Incidental catches have been also recognized as a significant
source of seabird mortality and one of the main threats to seabirds at sea (Croxall
et al. 2012). The Patagonian shrimp fishery has a negative effect on populations
of the Magellanic penguin, Spheniscus magellanicus, due to direct incidental
mortality and also by overlapping between penguin diet (anchovy and hake) and
bycatch species (Gandini et al. 1999; Marinao and Yorio 2014). Between 2010
10  Shrimps and Prawns 263

and 2017, 70–100% of the red shrimp production was fished within San Jorge
Gulf; currently, the area is closed for protection. The capture of Argentine red
shrimp in the period between 2004 and 2016 is shown in Fig 10.3.
In recent years, the Argentine stiletto shrimp, Artemesia longinaris, has become
an important target of the commercial fisheries as a consequence of the decline of
more traditional and profitable marine shrimps. Artisanal fisheries occur along its
entire distribution, and industrial fisheries are mainly concentrated in southern
Brazil and Argentina (D’Incao et al. 2002). Catches are mostly commercialized in
the internal market and show large seasonal and inter-annual fluctuations, leading
to efforts through a commercially sustainable aquaculture, capable of providing a
continuous supply. It is captured in appreciable concentrations over a narrow
coastal strip that does not exceed 5 miles (depths between 2 and 30 m). Recent
studies suggest that it is possible to consider A. longinaris as a metapopulation in
which there are more stable subpopulations, with high densities throughout the
year, such as those from southern Brazil to Argentina and from Macaé (Boschi
1969). These high-density subpopulations may serve as sources of new individu-
als for the less-stable sink populations, like those from São Paulo State. Source
populations would be critical for the implementation of management strategies,
such as the creation of protected areas and the implementation of off-­season peri-
ods, directed to sustain fisheries along the entire range of distribution (Carvalho-
Batista et al. 2014).

Fig 10.3  Evolution of landings registered for the Argentine red shrimp, Pleoticus muelleri,
between 2004 and 2016 for all ports in Argentina. (Adapted Secretary of Fisheries and Aquaculture.
Ministry of Agriculture, Livestock and Fisheries. Argentina)
264 P. M. Cervellini and J. O. Pierini

10.4.1  A
 rtisanal Shrimp Fisheries in the Bahía
Blanca Estuary

In the southern coast of ​​Buenos Aires Province, Pleoticus muelleri and Artemesia
longinaris are target species of small-scale artisanal fisheries. Artisanal fisheries are
defined as those who extract fishing resources for their commercialization, using
small-scale, low-technology, and low-capital fishing practices (Defeo and Castilla
2005). Along the Bahía Blanca Estuary and Anegada Bay (Fig 1.2, Chap. 1), several
sites were identified as fishing zones, but 75% of the total catches of both species
come from three areas: the Principal Channel (45%) and Vieja Channel (11%) in the
Bahía Blanca Estuary and Riacho Azul (14%) located in Anegada Bay (Suquele and
Colautti 2005). The main landing ports are Ingeniero White and Coronel Rosales in
the Bahía Blanca Estuary (Figs. 2.2 and 2.3, Chap. 2).
The artisanal fleet is composed of small boats (up to15 m length) locally known
as flota amarilla (yellow fleet) with limited autonomy, and each boat may be
accompanied by several smaller boats (Fig.  10.4). Instead of the typical bottom
trawling, shrimp fishers in the Bahía Blanca Estuary make use of tidal currents.
Fishing nets like those used for trawling are deployed at high tide or ebb tide, each
net attached to a pair of anchors. Fishing operations are carried through complete
flood tides or ebb tides, and shrimps, along with incidental fishes and crustaceans
dragged by bottom tidal currents, get trapped into the net. Each fishing boat carries
about 10 nets, along with 20 anchors, and the fishing maneuver is all manual labor
carried by fishers on board the small accompanying boats (Bertuche et al. 1998).
Nets are revised twice during each tide to take the catch. The management and
control of the nets are carried out by the canoes driven by an outboard motor or oars.
The boats are generally in number of six. The catch of each net is transported to the
main vessel, which remains fixed in one place throughout the fishing operation. On
board, fishermen manually separate P. muelleri and A. longinaris from and
accompanying fauna and select those shrimps that are commercial in size. In
addition, the crustaceans that will be used for sale are processed (boiled, salted, or
sulfided) or placed in a cool place to ensure they arrive in good condition at port.
The choice of specimen preservation will depend on the remaining time and
temperature for landing. In the Bahía Blanca Estuary, there are two fishing seasons:
one in summer (January to May) and the other in winter (August to spring). In the
former nets are located in the middle of water column; in the latter nets are anchored
over the bottom.
Shrimp abundance in the Bahía Blanca Estuary has strong seasonal and inter-­
annual oscillations for both species (Cervellini and Piccolo 2007). The migratory
breeder behavior would be related with seasonal landing fluctuations. In the past,
outside the estuary, there was an important fishing activity of trawlers of greater
magnitude and autonomy that come from northern Buenos Aires Province, such as
Mar del Plata Port, that could impact negatively over the local fisheries. Since 2006
fishing with trawls is prohibited in El Rincón area, outside the Bahía Blanca Estuary;
however, it has not been evaluated whether this management measure has favored
10  Shrimps and Prawns 265

Fig. 10.4 (a) Artisanal shrimp fishing boat, (b) fishermen select the catches (c) bycatch. (Photo
by: (a) Sandra Fiori, (b and c) Gabriela Blasina)

artisanal shrimp and prawn fisheries into the estuary. On the other hand, the inter-­
annual variability seems to be related with climatic conditions. It was observed that
when the average annual temperature exceeded 16 °C, in the following year or in the
same year, the catches of the resource decreased significantly (Fig.  10.5). The
maximum catches in the Bahía Blanca Estuary are registered every 3–5  years
(Cervellini and Piccolo 2007).
Since 1999, one of the rules that must be obeyed by those who practice artisanal
or commercial fishing is the obligation to submit to the enforcement authority in the
Delegation that has the Secretary of Fisheries Activities of the Province of Buenos
Aires, in the Port of Ingeniero White; they must declare to the authority the catches
made in each tide. When the information contained in the parts corresponding to a
given period of time, provides valuable information, allow understand the dynamics
of the fishery or implement management measures that tend to achieve sustainable
exploitation of resources.
266 P. M. Cervellini and J. O. Pierini

600 16.50

500 16.00

400 15.50
Total Landing (tn)

Temperature (°C)
300 15.00

200 14.50

100 14.00

0 13.50
1992 1993 1994 1995 1996 1997 1998 1999 2000 2001 2002 2003 2004 2005
YEARS

Artemesia longinaris Pleoticus muelleri Temperature (°C)

Fig. 10.5  Annual catches of Pleoticus muelleri and Artemesia longinaris and average annual tem-
perature in the Bahía Blanca Estuary during 1992 and 2005. (Adapted from Cervellini and
Piccolo 2007)

10.5  Conclusions

There are many factors that affect the presence of shrimp in the Bahía Blanca
Estuary, both natural and anthropogenic. Both species have very different
fluctuations in annual catch volumes. The dynamics of the fishery in the estuary is
markedly seasonal and is related to the biological characteristics of the species. It is
recommended to monitor the status of these populations taking into account two
factors: the cycle study life and the dynamics of juveniles and adults. It is also
recommended to control the fishing activities that take place in the outside area of​​
the estuary. In this way, exploitation will be optimized through the creation of
scientific-technical bases for sustainable development of the fisheries of these two
resources so valuable to the community of Bahía Blanca. As we have anticipated in
previous paragraphs, eggs and larval stages of crustaceans penaeoid are an important
part of the zooplankton. The abundance is of great interest in the economy of the
ecosystem and allows to know the fauna composition of the zooplankton, especially
of those members that behave like meroplankton, the distribution and abundance of
them becomes important within the plankton for being indicators of water bodies,
for being food for another. Finally, we want to point out that one of the great
advantages of using a mathematical model for integrated coastal zone management
is that datasets collected in different periods and sampling intervals can be integrated
into a single tool to reproduce and analyze the processes taking place in the water
body. The hydrodynamic module of the MOHID model has been validated and
10  Shrimps and Prawns 267

calibrated for the entire estuary, but it is important to note that there is only
oceanographic data about the Principal Channel and poor data acquisition for the
rest of the estuary. From the performed simulations, it could be concluded that
horizontal dispersion of larvae is only at the outer part (mouth) of the Bahía Blanca
Estuary and has a significant importance over horizontal distribution throughout the
northern shore of the estuary where currents and abiotic factors, mainly salinity, are
more constant. This conclusion was obtained by model results, taking into account
the sampling period of the larvae, where Lagrangian particles develop their
dispersion like passive particles that move and extend over this coastal sector where
they were also affected by the atmospheric (wind) and oceanographic parameters
(currents, waves, etc.). The combination of data analysis and numerical modelling
allows a more comprehensive conceptual model of the Bahía Blanca Estuary
hydrodynamics that could aid in decision-making of local managers. One of the
greatest advantages of using modelling tools in integrated coastal zone management
is that datasets collected at different periods and sampling intervals could be
consolidated in a tool to reproduce periodical phenomena, i.e., tides and currents,
and to analyze different system descriptors such as tidal or current distribution,
larval or postlarval spatial distribution, etc. Numerical modelling also allows the
isolation and discrimination of single processes and can aid with data accuracy and
consistency – important factors to consider if we want to analyze larval retention
and nursery areas and in the future observe their influence in fishing. Multidisciplinary
analysis to date has not been carried out in our study area which is vital for a correct
coastal management.

References

Angelescu VA, Boschi EE (1959) Estudio biológico pesquero del langostino de Mar del Plata en
conexión con la operación nivel medio. Servicio Hidrografía Naval, Público H-1017, 137 p
Angeletti S, Pierini JO, Cervellini PM (2018) Suspended sediment contribution resulting from
bioturbation in intertidal sites of a SW Atlantic mesotidal estuary: data analysis and numerical
modelling. Sci Mar 82(4):245–256
Bailey-Brock JH, Moss SM (1992) Penaeid taxonomy, biology and zoogeography. In: Fast AW,
Lester LJ (eds) Marine shrimp culture: principles and practices. Developments in aquaculture
and fisheries science, vol 23. Elsevier Science Publisher BV, Dordrecht
Bertuche D (1999) Sobre la captura incidental de la merluza en la pesquería del langostino
patagónico Inf. Tec. INIDEP N° 76/99, 5pp
Bertuche D, Fischbach C, Boccanfuso J (1998) Síntesis de los resultados obtenidos en el moni-
toreo del recurso langostino realizado en el sector costero de la Pcia. De Bs. As, al sur de Puerto
Ingeniero White. Inf. Téc. Inst. INIDEP N°129
Blanton J, Werner FE, Kapolnai A et al (1999) Wind-generated transport of fictitious passive larvae
into shallow tidal estuaries. Fish Oceanogra 8(2):210–223
Bortolotto GA, Kolesnikovas CKM, Freire AS et al (2016) Young humpback whale Megaptera
novaeangliae feeding in Santa Catarina coastal waters, Southern Brazil, and a ship strike
report. Mar Biodivers Rec 9(1):29
Boschi EE (1963) Los camarones comerciales de la familia Penaeidae de la costa Atlántica de
América del Sur. Clave para el reconocimiento de las especies y datos bioecológicos. Boletín
del Instituto de Biología Marina (3):1–40
268 P. M. Cervellini and J. O. Pierini

Boschi EE (1969) Estudio Biológico Pesquero del camarón Artemesia longinaris Bate de Mar del
Plata. Bol Inst Biol Mar N° 18:1–4
Boschi EE (1986) La pesquería del langostino del litoral patagónico. Cuadernos de Redes 20:20–26
Boschi EE (1989) Biología pesquera del langostino del litoral patagónico de Argentina (Pleoticus
muelleri). Ser Contrib INIDEP Mar del Plata 646:1–71
Boschi EE (1997) Las pesquerías de crustáceos decápodos en el litoral de la República Argentina.
Invest Mar Valparaíso 25:19–40
Boschi EE (2004) Camarones y especies afines. In: Boschi EE, Cousseau MB (eds) La vida entre
mareas, vegetales y animales de las costas de Mar del Plata, INIDEP, Argentina, pp 205–212
Boschi EE, Gavio MA (2005) On the distribution of decapod crustaceans from the Magellan
Biogeographic Province and the Antarctic region. Sci Mar 69(S2):195–200
Boschi EE, Scelzo MA (1967) Campaña de Pesca Exploratorio Camaronera en el Litoral de la
Provincia de Buenos Aires, 16-24 de Febrero de 1967, Proy Des Pesq (FAO), Publ N°2:1–22
Boschi EE, Scelzo MA (1969) Nuevas Campañas Exploratorias Camaroneras en el Litoral
Argentino, con Referencias al Plancton de la Región. Proy Des Pesq (FAO), Publ N°16:1–31
Boschi EE, Scelzo MA (1971) Últimos resultados de las investigaciones sobre los peneidos com-
erciales en la Argentina (Marcaciones, Campañas camarones y cultivos). CARPAS/5/Doc
Téc 4:1–17
Boschi EE, Cousseau MB (2004) La vida entre mareas: Vegetales y animales de las costas de Mar
del Plata, Argentina, 1 ed. INIDEP, Mar del Plata
Boschi EE, Fischbach CE, Iorio MI (1992) Catálogo ilustrado de los crustáceos estomatópodos y
decápodos marinos de Argentina
Burke JS, Ueno M, Tanaka Y et al (1998) The influence of environmental factors on early life his-
tory patterns of flounders. J Sea Res 40(1):19–32
Burkenroad MD (1945) A new sergestid shrimp (Peisos petrunkevitchii, n. gen., n. sp.), with
remarks on its relationships. Trans Connecticut Acad Arts Sci 36:553–591
Calderón-Pérez JA, Poli CR (1987) A physical approach to the postlarval Penaeus immigration
mechanism in a Mexican coastal lagoon (Crustacea: Decapoda: Penaeidae). Anales del Instituto
de Ciencias del Mar y Limnología. Universidad Nacional Autónoma de México 14(1):147–156
Campuzano FJ, Pierini JO, Leitão PC (2008) Hydrodynamics and sediments in Bahia Blanca estu-
ary: data analysis and modelling. In: Neves R, Baretta J, Mateus M (eds) Perspectives on inte-
grated coastal zone Management in South America. IST Press, Lisbon, pp 483– 503
Campuzano FJ, Pierini JO, Leitão PC (2014) Characterization of the Bahía Blanca estuary by data
analysis and numerical modelling. J Mar Syst 129:415–424. https://doi.org/10.1016/10.1016/j.
jmarsys.2013.09.001
Capitoli R, Bager A, Ruffino M (1994) Contribuçao ao conhecimento das relaçoes tróficas
bentônico-demersais nos fundos de pesca do camarao Artemesia longinaris Bate, na regiao da
Barra da Lagoa dos Patos, RS, Brasil. Nauplius 2:53–74
Carvalho-Batista A, Simões SM, Lopes M et  al (2011) Ecological distribution of the shrimp
Pleoticus muelleri (Bate, 1888) and Artemesia longinaris Bate, 1888 (Decapoda, Penaeoidea)
in the southeastern Brazilian littoral. Nauplius 19:135–143
Carvalho-Batista A, Negri M, Pileggi LG et al (2014) Inferring population connectivity across the
range of distribution of the stiletto shrimp Artemesia longinaris Spence Bate, 1888 (Decapoda,
Penaeidae) from DNA barcoding: implications for fishery management. ZooKeys 457:271
Castilho AL, Gavio MA, Costa RC et al (2007) Latitudinal variation in population structure and
reproductive pattern of the endemic south american shrimp Artemesia longinaris (Decapoda:
Penaeoidea). J Crustacean Biol 27(4):548–552
Castilho AL, Da Costa R, Fransozo A et  al (2008) Reproduction and recruitment of the South
American red shrimp, Pleoticus muelleri (Crustacea: Solenoceridae), from the southeastern
coast of Brazil. Mar Biol Res 4(5):361–368. https://doi.org/10.1080/17451000802029536
Cervellini PM (1988) Las larvas y postlarvas de los crustáceos Decápoda en el estuario de Bahía
Blanca. Variaciones estacionales y su relación con los factores ambientales. PhD thesis,
Universidad Nacional del Sur, Bahía Blanca, Argentina, 246 pp
10  Shrimps and Prawns 269

Cervellini PM (1992) Seasonal and spatial distribution of decapod Crustacea larvae in the Bahía
Blanca estuary, Argentina. J Aqua Trop 7:35–46
Cervellini PM (2001) Variabilidad en la abundancia y retención de larvas de crustáceos decápodos
en el estuario de Bahía Blanca, Provincia de Buenos Aires, Argentina. Invest Mar Valparaíso
29(2):25–33
Cervellini PM, Mallo J (1991) Distribución espacial de estadios larvales de Penaeidae en Bahía
Blanca, Argentina. Biol Pesq 20:5–11
Cervellini PM, Piccolo MC (2007) Variación anual de la pesca del langostino y camarón en el
estuario de Bahía Blanca. Geoacta 32:111–118
Condie SA, Lonerang NR, Die DJ (1999) Modeling the recruitment of tiger prawns Penaeus escu-
lentus and P. semisulcatus to nursery grounds in the Gulf of Carpentaria, northern Australia:
Implications for assessing stock-recruitment relationships. Mari Ecol Progr Series 178:55–68
Cordo H (2005) Evaluación del estado del efectivo sur de 41°S de la Merluza (Merluccius hubbsi)
y estimación de la captura biológicamente aceptable correspondiente al año 2005. Instituto
Nacional de Investigación y Desarrollo Pesquero (INIDEP), Mar del Plata
Costa RCD, Fransozo A, Melo GAS, Freire FADM (2003) Chave ilustrada para identificação dos
camarões Dendrobranchiata do litoral norte do estado de São Paulo, Brasil. Biota Neotropica
3(1):1–12
Costa RC, Fransozo A, Pinheiro AP (2004) Ecological distribution of the shrimp Pleoticus muelleri
(Bate, 1888) (Decapoda: Penaeoidea) in southeastern Brazil. Hydrobiologia 529(1-3):195–203
Croxall JP, Butchart SHM, Lascelles B et al (2012) Seabird conservation status, threats and prior-
ity actions: a global assessment. Bird Conservation International 22:1–34
Cuesta AC (2010) Composición, abundancia estacional y dispersión horizontal de especies de
mesozooplancton en la parte media del estuario de Bahía Blanca (Puerto Rosales). Tesis de
Licenciatura en Ciencias Biológicas de la Universidad Nacional del Sur, 50 pp
D’Incao F (1999) Subordem Dendrobranchiata (camarões marinhos). In: Buckup L, Bond
Buckup G (eds) Os Crustaceos do Rio Grande do Sul. Editora da Universidade, Porto Alegre,
pp 275–299
D’Incao F, Martins STS (2000) Brazilian species of the genera Acetes H. Milne Edwards, 1830 and
Peisos Burkenroad, 1945 (Decapoda: Sergestidae). J Crustacean Biol 20(5):78–86
Dall W, Hill J, Rothlisberg PC et al (1990) The biology of Penaeidae. Adv Mar Biol 27
De la Garza J, Moriondo Danovaro P, Fernández, M (2017) An overview of the argentine red
shrimp (Pleoticus muelleri, Decapoda, Solenoceridae) fishery in Argentina. Biology, fishing,
management and ecological interactions. Mar del Plata: Instituto Nacional de Investigaciony
Desarrollo Pesquero INIDEP. Contribution N° 2080
Defeo O, Castilla JC (2005) More than one bag for the world fishery crisis and keys for co-man-
agement successes in selected artisanal Latin American shellfisheries. Rev Fish Biol Fisher
15:265–283
DeLancey LB, Jenkins JE, Whitaker JD (1994) Result of long term, seasonal sampling for Penaeus
postlarvae at Breach inlet, South Carolina. Fish Bull 92:633–640
Delgado AL, Ferrelli F, Piccolo MC et  al (2017) Influence of Physical-Biological Variability
and Legal Regulations on Fisheries Resources in the Southern Coastal Zone of Buenos Aires
Province. Argentina Argentina Anuário do Instituto de Geociências – UFRJ 40:5
FAO (2016) El estado mundial de la pesca y la acuicultura 2016. Contribución a la seguridad ali-
mentaria y la nutrición para todos. Roma, 224 p
Forward RB Jr, Tankersley RA (2001) Selective tidal stream transport of marine animals. Oceanogr
Mar Biol. An Annual Review 39:305–353
Fransen CHJM (2014) Shrimps and prawns. The living marine resources of the Eastern Central
Atlantic 1:37–196
Gandini PA, Frere E, Pettovello AD et al (1999) Inter-action between Magellanic penguins and
shrimp fisheries in Patagonia, Argentina. Condor 101:783–789
García S, Le Reste L (1986) Ciclos vitales, dinámica, explotación y ordenación de las poblaciones
de camarones peneidos costeros. FAO Doc. Téc. Pesca 203:180
270 P. M. Cervellini and J. O. Pierini

Gavio MA, Boschi EE (2004) Biology of the shrimp Artemesia longinaris Bate, 1888 (Crustacea:
Decapoda: Penaeidae) from Mar del Plata coast, Argentina. Nauplius 12(2):83–94
Gavio MA, Boschi EE (2016) Historia de vida del camarón de Mar del Plata Artemesia longinaris
(Decapoda, Penaeoidea, Penaeidae). In: Boschi EE (ed) Los crustáceos de interés pesquero
y otras especies relevantes en los ecosistemas marinos. El Mar Argentino y sus recursos pes-
queros, 6. Instituto Nacional de Investigación y Desarrollo Pesquero INIDEP, Mar del Plata,
pp 51–58
Gillet R (2008) Global study of shrimp fisheries. Fisheries technical paper no. 475. Food and
Agriculture Organization of the United Nations, Rome, pp 331
Góngora ME, Bovcon ND, Cochia PD (2009) Ictiofauna capturada incidentalmente en la pes-
quería de langostino patagónico Pleoticus muelleri Bate, 1888. Rev Biol Mar Oceanogr
44(3):583–593
Incao, Fernando & Valentini, Helio & Rodrigues, Luiz. (2002). Avaliação da pesca de camarões
nas regiões Sudeste e Sul do Brasil. Atlântica. 24. 103–116
Iorio MI, Scelzo MA, Boschi EE (1990) Larval and post-larval development of the shrimp
Pleoticus muelleri, Bate, 1888 (Crustacea, Decapoda, Solenoceridae). Sci Mar 54(4):329–341
Iorio MI, Macchi GJ, Fischbach CE et al (1996) Estudios sobre la dinámica reproductiva del lan-
gostino Pleoticus muelleri en el área de Bahía Blanca (Provincia. De Buenos Aires, Argentina).
Frente Marit 15:111–118
Jager Z, Mulder J (1999) Transport velocity of flounder larvae (Platichthys flesus L.) in the Dollard
(Ems Estuary). Estuar Coast Shelf Sci 49(3):327–346
Kelleher K (2008) Descartes en la pesca de captura marina mundial. FAO Documento Técnico de
Pesca 470:1–147
Lindley JA (1987) Continuous Plankton Records: the geographical distribution and seasonal cycles
of decapod crustacean larvae and pelagic post-larvae in the north-eastern Atlantic Ocean and
the North Sea, 1981-3. J Mar Biol Ass UK 67:145–167
López Cazorla A, Pettigrosso RE, Tejera L et al (2011) Diet and food selection by Ramnogaster
arcuata (Osteichthyes, Clupeidae). J Fish Biol 78:2052–2060
Macchi GJ, Iorio MI, Christiansen HE (1992) Aspectos del desove y fecundidad del langos-
tino Pleoticus muelleri (Bate, 1888) (Crustacea, Decapoda, Solenoceridae). Rev Biol Mar
27(1):43–58
Mallo JC (1986) Desarrollo larval y cultivo en laboratorio del camarón argentino Peisos petrunkev-
itchi Burkenroad, 1945 (Crustacea, Decapoda, Sergestidae). Biología Pesquera 15:316
Mallo JC, Boschi EE (1982) Contribución al conocimiento del ciclo vital del camarón Peisos
petrunkevitchi de la región de Mar del Plata, Argentina (Crustacea, Decapoda, Sergestidae).
Physis, Buenos Aires A 41:85–98
Mallo J, Cervellini PM (1988) Distribution and abundance of larvae and postlarvae of Artemesia
longinaris, Pleoticus muelleri and Peisos petrunkevitchi (Crustacea, Decapoda; Penaeidae) in
the coastal waters in the Bahía Blanca Bay, Argentina. Aquacult Trop 3:1–9
Marinao CJ, Yorio PM (2014) Fishery discards and incidental mortality of seabirds attending
coastal shrimp trawlers at Isla Escondida, Patagonia, Argentina Wilson Ornithological Society
2011–2012
Miguel C (2010) Variación estacional de las larvas de Decápoda en la parte interna del estuario de
Bahía Blanca (Puerto Cuatreros) y dispersión horizontal de las zoeas de Neohelice granulata
(=Chasmagnathus granulatus) (Decápoda, Varunidae) en relación con la circulación del agua.
Tesis de grado de la Universidad Nacional del Sur, 40 pp
Minkoff DR, Escapa M, Ferramola FE et al (2006) Effects of crab-halophytic plant interactions
on creek growth in a S.W. Atlantic Salt Marsh: A cellular automata model. Estuar Coast Shelf
Sci 69:403–413
Morgan SG (1990) Impact of planktivorous fishes of dispersal, hatching, and morphology of estua-
rine crab larvae. Ecology 71:1639–1652
Morgan SG (1992) Predation by planktonic and benthic invertebrates on larvae of estuarine crabs.
J Exp Mar Biol Ecol 163:91–110
Omori M (1974) The biology of pelagic shrimps in the ocean. Adv Mar Biol 12:233–324
10  Shrimps and Prawns 271

Pérez Farfante I (1988) Illustrated key to penaeoid shrimps of commerce in the Americas. NOAA
Technical Report NMFS 64
Perillo GME, Piccolo MC (1999) Geomorphological, physical characteristics of Bahía Blanca
estuary, Argentina. The Argentina estuaries: a review. In: Perillo GM, Piccolo MC, Pino
Quivira M (eds) Estuaries of South America: their geomorphology and dynamics. Springer,
Berlin, pp 195–216
Pettovello A (1999) By-catch in the Patagonian red shrimp (Pleoticus muelleri) fishery
Pierini JO (2007) Circulation and transport in coastal areas of the Bay estuary Blanca, PhD thesis,
University of Buenos Aires, Buenos Aires, 225 pp. (in Spanish)
Pierini JO, Lovallo M, Telesca L et al (2012) Investigating prediction performance of an artificial
neural network and a numerical model of the tidal signal at Puerto Belgrano, Bahía Blanca
estuary (Argentina) Acta Geophysica 61(6) Dec 2013:1522–1537. https://doi.org/10.2478/
s11600-­012-­0093-­x
Pierini JO, Campuzano FJ, Leitão CP et  al (2018) Atmospheric influence over the residence
time in the Bahia Blanca Estuary, Argentina. Thalass Int J Mar Sci. https://doi.org/10.1007/
s41208-­018-­0120-­z
Pierini JO, Campuzano FJ, Leitão PC et al (2019) Atmospheric influence over the time residence
in the Bahía Blanca estuary. Thalass Int J Mar Sci 35(1):275–286. https://doi.org/10.1007/
s41208-­018-­0120-­z
Poli CR, Calderón-Pérez JA (1987) Effect of hydrological changes at the mouth of the Baluarte
River on the immigration of postlarvae from Penaeus vannamei (Boone) and P. stylirostris
(Stimpson) to Huizache-Caimanero lagoon system, Sinaloa, Mexico (Crustacea: Seafood
Watch 2018)
Rothlisberg PC, Church JA, Fandry B (1995) A mechanism for near-shore concentration and estua-
rine recruitment of post-larval Penaeus plebejus Hess (Decapoda, Penaeidae). Estuar Coast
Shelf Sci 40(2):115–138
Rueda JG, Otero-Díaz LJ, Pierini JO (2013) Caracterización hidrodinámica en un estuario tropi-
cal de América del Sur con régimen microtidal mixto (Bahía de Cartagena, Colombia) Boletín
científico. CIOH 31:159–174
SAGPyA (2017) Resolución 304-2017. Secretaría de Agricultura, Ganadería, Pesca y Alimentación.
Sanidad Animal
Sardiña P, Lopez Cazorla A (2005a) Feeding habits of the juvenile striped weakfish, Cynoscion
guatucupa Cuvier 1830, in Bahía Blanca estuary (Argentina): seasonal and ontogenetic
changes. Hydrobiologia 532:23–30
Sardiña P, Lopez Cazorla A (2005b) Trophic ecology of the of whitemouth croaker, Micropogonias
furnieri (Pisces: Sciaenidae), in south-western Atlantic waters. J Mar Biol Assoc UK
85:405–413
Scelzo MA (2016) Biología reproductiva del langostino y del camarón de las aguas marinas argen-
tinas. In: Boschi EE (ed) Los crustáceos de interés pesquero y otras especies relevantes en
los ecosistemas marinos. El Mar Argentino y sus recursos pesqueros, 6. Instituto Nacional de
Investigación y Desarrollo Pesquero INIDEP, Mar del Plata, pp 71–88
Scelzo MA, Martínez Arca J, Lucero NM (2002) Diversity, density and biomass of the macro-
fauna component of the “shrimp-shrimp” fishing funds, in front of Mar del Plata, Argentina
(1998–1999). National Institute for Fisheries Research and Development, Journal of Fisheries
Research and Development 15:43–65
Spivak ED, Farías NE, Ocampo EH et al (2019) Annotated catalogue and bibliography of marine
and estuarine shrimps, lobsters, crabs and their allies (Crustacea: Decapoda) of Argentina and
Uruguay (Southwestern Atlantic Ocean). Frente Marítimo 26:1–164
Suquele P, Colautti D (2005) Pesca artesanal y comercial en los puertos de la Ría de Bahía Blanca,
Buenos Aires. Análisis de las partes pesqueras correspondientes a los años 2000, 2001, 2002,
2003 y 2004. Dirección de Asuntos Agrícolas, 13 pp
Tavares C, Martin J (2010) Suborder Dendrobranchiata Bate, 1888. In: Schram FR, von Vaupel
Klein JC, Forest J, Charmantier-Daures M (eds) Treatise on Zoology – Anatomy, Taxonomy,
Biology – The Crustacea, Decapoda, Volume 9 Part A Eucarida: Euphausiacea, Amphionidacea,
and Decapoda (partim), vol 9A, pp 99–164
272 P. M. Cervellini and J. O. Pierini

Tavares C, Serejo C, Martin J (2009) A preliminary phylogenetic analysis of the Dendrobranchiata


based on morphological characters. In: Martin JW, Crandall KA, Felder DL (eds) Decapod
crustacean Phylogenetics. Crustacean Issues 18:255–273
Thayer GW, Schaaf WE, Angelovic JW et  al (1973) Caloric measurements of some estuarine
organisms. Fish B-NOAA 71(1):289–296
Wehrtmann IS (1989) Seasonal occurrence and abundance of caridean shrimp larvae at Helgoland.
German Bight Helgoländer Meeresuntersuchungen 43(1):87–112
Young PC, Carpenter SM (1977) Recruitment of postlarval penaeid prawns to nursery areas in the
Moreton Bay, Queensland. Mar Freshwater Res 28:745–773
Chapter 11
Ecology and Biology of Fish Assemblages

Juan Manuel Molina, Gabriela Blasina, and Andrea Lopez Cazorla

11.1  Estuaries as Key Habitat for Fish

Estuaries are important habitats for several species of fish, due to the high levels of pri-
mary production which supports highly diverse and abundant prey. Estuaries also pro-
vide nursery and feeding grounds for many species (Elliott et al. 2007). In these regions
environmental gradients are large and generate a unique combination of biotic and abi-
otic factors (Day et al. 1989). The major biotic and abiotic factors which determine the
distribution and abundance of fish in estuaries are shown in Fig. 11.1. These factors are
not independent but interact directly and indirectly with the fish species that inhabit
estuaries. Hydrographic conditions directly influence mouth condition, estuarine water
temperature, salinity, turbidity, and dissolved oxygen concentrations and indirectly
affect habitat diversity, productivity, fish recruitment, food availability, and competition
(Whitfield 1999). The high productivity of estuaries has often been identified as the
main reason why fish are attracted to these areas in such large numbers, which is
explained by the large food webs these environments support. Food, especially detritus
and benthic invertebrates, is abundant in most estuarine systems. However, the avail-
ability of a particular food type is likely to show marked fluctuations over time and
space, particularly in response to environmental changes which characterize all types of
systems on the subcontinent (Potter et al. 2015). Species that are broadly tolerant of
biotic and abiotic variability are at a considerable advantage over those fish species that
cannot survive such fluctuations, because the former are able to occupy a food-rich
environment from which many potential competitors are excluded (Whitfield 1999).
Bahía Blanca Estuary thrives with fish life. This estuary is a complex ecosystem
encompassing several islands, salt marshes, mudflats, and large tidal flats which

J. M. Molina () · G. Blasina · A. L. Cazorla


Departamento de Biología, Bioquímica y Farmacia, Universidad Nacional del Sur,
Bahía Blanca, Argentina
Instituto Argentino de Oceanografia (IADO-CONICET/UNS), Bahía Blanca, Argentina

© Springer Nature Switzerland AG 2021 275


S. M. Fiori, P. D. Pratolongo (eds.), The Bahía Blanca Estuary,
https://doi.org/10.1007/978-3-030-66486-2_11
276 J. M. Molina et al.

Fig. 11.1  Major abiotic (blue) and biotic (green) factors that influence the distribution and abun-
dance of fish populations in estuaries. (Modified from Day et al. 1989)

generate marked environmental gradients thus promoting a high diversity of bio-


logical systems in a relatively small area (see Chaps. 1 and 3 of this book). This
diverse ecotone provides several different types of habitats for fish. A world-scale
review by Blaber (2002) states that the number of fish species in subtropical and
tropical estuaries is much greater than in temperate regions: at least 100 species,
with some reaching more than 200. Pasquaud et al. (2015) evaluated the latitude as
potential factor determining fish species richness in estuaries, and they also identi-
fied a fish species richness distribution in relationship with latitude, with the lowest
values of richness at 40° S. Yañez-Arancibia et al. (1980) recorded 121 species in
tropical estuaries and lagoons of the Caribbean Sea. Atlantic areas appear relatively
richness-poor regarding the worldwide average species richness of tropical estuar-
ies (Simier et al. 2006). In the southwestern Atlantic temperate regions, the estua-
rine ichthyofauna is composed of ~110 species in dos Patos Lagoon (32° S), South
Brazil (Chao et al. 1985); 119 fish species inhabit the Río de la Plata estuary (34° S)
(Jaureguizar et al. 2016); and a total of 28 species were registered in Mar Chiquita
coastal lagoon (37° S) in Argentina (González Castro et al. 2009). The number of
fish species (32; Table 11.1) registered in Bahía Blanca Estuary (39° S) is consistent
with the widely held view that latitude plays a critical role influencing richness, with
tropical areas being more diverse than temperate ones.
More than 25 different mechanisms have been suggested for generating system-
atic latitudinal patterns in biodiversity, commonly emphasizing reasons as to why
the tropics are highly specious. These include explanations based on chance, histori-
cal perturbation, environmental stability, habitat heterogeneity, productivity, and
interspecific interactions. In these explanations, mean annual temperature is com-
monly used as a proxy of the energy of the system (Gaston 2000), i.e., the systems
with a higher temperature would have higher energy, although this is not always the
case. Among the different hypotheses that have been proposed in order to explain
gradients in species richness, the species richness-energy relationship hypothesis
received the greatest support, mostly from studies of terrestrial or freshwater organ-
isms. For marine organism, a few studies addressed this framework, and their results
are not consensual. In estuaries, the species richness-energy relationship has not
11  Ecology and Biology of Fish Assemblages 277

Table 11.1  Fish species presence in the inner, middle, and external zones of Bahía Blanca Estuary
in the two periods of sampling: 1979–1983 and 2017–2018
Inner zone Middle zone External zone
Species 1979–1983 2017–2018 1979–1983 2017–2018 1979–1983
Galeorhinus galeus ●
Mustelus schmitti ● ● ● ●
Notorhynchus cepedianus ● ●
Squatina guggenheim ●
Sympterygia acuta ● ● ●
Sympterygia bonapartii ● ● ● ●
Myliobatis goodei ● ● ● ● ●
Conger orbignianus ● ● ● ●
Brevoortia aurea ● ● ● ● ●
Ramnogaster arcuata ● ● ● ● ●
Anchoa marinii ● ●
Lycengraulis sp. ● ● ● ●
Genidens barbus ● ●
Macruronus magellanicus ● ●
Porichthys porosissimus ● ● ● ● ●
Mugil liza ● ● ●
Odontesthes argentinensis ● ● ● ● ●
Acanthistius patachonicus ● ●
Dules auriga ●
Pomatomus saltatrix ● ● ● ● ●
Parona signata ● ● ●
Cynoscion guatucupa ● ● ● ● ●
Macrodon ancylodon ● ●
Menticirrhus americanus ● ●
Micropogonias furnieri ● ● ● ● ●
Pogonias cromis ● ●
Umbrina canosai ●
Percophis brasiliensis ●
Stromateus brasiliensis ● ● ● ● ●
Paralichthys patagonicus ●
Paralichthys orbignyanus ● ● ● ● ●
Oncopterus darwinii ●
Symphurus jenynsi ● ● ●
Total species 12 16 21 23 29

been considered in former studies probably due to the complexity and the variability
of these systems (Pasquaud et al. 2015).
Bahía Blanca Estuary is a shallow temperate estuarine system that produces
hypersaline conditions, where the salt concentrations in the inner portion of the
estuary often exceed those of the inner continental shelf (average salinity is 33 ppt
even though values as low as 17.3 ppt and as high as 41.9 ppt have been registered;
278 J. M. Molina et al.

Freije et  al. 2008). The estuary can be divided in three zones, depending on the
composition of fish species that make use of the environment: inner (from the head
of estuary to Ingeniero White), middle (from Ingeniero White to Puerto Rosales),
and external (from Puerto Rosales to the mouth of the estuary) (see references in
Fig. 2.1; Chap. 2). The inner zone has abundant intertidal habitats that are harsh and
variable, forcing fish to either perform tidal migrations or be exposed to the natu-
rally variable abiotic conditions. This zone also gathers most of the freshwater out-
puts of the tributary river meaning salinity can vary greatly. The inner and middle
zones are where most of the human activity takes place. The last zone is abundant
in islands and channels of varying depth. Some of these islands and channels are
populated by important salt marshes. This large mudflat contracts in low tide, forc-
ing fish to the deeper channels. The changes in temperature in this area are the wid-
est of the three zones. The external zone is not subjected to changes in environmental
parameters as much as the other two and resembles oceanic conditions the most.

11.2  Distribution and Composition of Fish Assemblage

In 1979 Lopez Cazorla started surveying the fish species composition in Bahía
Blanca Estuary. In these surveys, which were carried out until 1983, fishing was
performed by the local artisanal fishermen, who used gear specific for their fishing
needs, whether it is shrimp nets, channel closure nets, or gill nets (Fig. 11.2). These
surveys covered the inner, middle, and external zones of Bahía Blanca Estuary.
Between 2017 and 2018, our research team started conducting seasonal surveys on
the estuary. Fishing was performed using two of the fishing gears used by fishermen
(shrimp nets and channel closure nets, Fig. 11.2 a and b, respectively), albeit with
lower fishing effort.
In this estuary, fishermen use these types of nets taking advantage of the strong
flow of the tidal regime. Shrimp nets are usually set on high tide, and low tide flow
forces fish into the net bag. Gill nets are placed in between tides, as fish moving in
and out of the inner portions of the estuary are more likely to get entangled in them.
Channel closure nets are placed on the inner channels on high tide, and as water
level lowers during low tide, fish are pushed against the net, and the fishermen close
the net and collect them. Demersal and demersal-benthic fish are most abundantly
captured using these methods.
In the former surveys, 45 sampling trips were performed, while in the recent
surveys 8 sampling trips were carried out. In both surveys fish were identified, and
several measurements were taken (Table  11.1). Of the 32 species of fish found
inhabiting the waters of Bahía Blanca Estuary, 7 were chondrichthyans and 25
osteichthyans.
From the initial surveys conducted in Bahía Blanca Estuary, summarized in
Table 11.1, it is possible to conclude that there is a progression of species richness,
from the inner zone (12 species) to the middle (21 species) and the external zones
(29 species). The zoning pattern described at the start of this chapter can explain, up
11  Ecology and Biology of Fish Assemblages 279

Fig. 11.2  Net types


employed by artisanal
fishermen of Bahía Blanca
Estuary, also used for the
scientific samplings. (a)
Shrimp nets, (b) channel
closure nets, and (c) gill
nets. (Modified from
Lopez Cazorla 2004)

to certain extent, these findings. The inner zone is challenging for fish given its wide
range of water salinity. Anadromous fish, such as Myliobatis goodei and Mugil liza,
have physiological adaptations that allow them to exploit the resources of this zone.
Fish species with tolerance to changes in temperature and salinity can live or move
in and out of the middle zone. Most chondrichthyans have less tolerance to abiotic
stressors in comparison with bony fish; hence, they occur on the middle and external
zones, with only M. goodei moving deeper into the inner zone. We have found
exclusively marine fish as well, like Dules auriga or Macruronus magellanicus. The
inner and middle zones of the estuary are common fishing grounds for the two study
periods; hence, a comparison between both is only possible for those areas.
Table  11.1 shows such comparison. In general, the more recent surveys found a
greater number of species, especially considering those found in the inner zone,
even though the fishing effort in 1979–1983 was considerably larger. In that period
a total of 45 sampling trips were made, 14 in the inner zone (31.11%), 26 in the
middle zone (57.78%), and 5  in the external zone (11.11%). In comparison, the
2017–2018 period was composed of 8 field samplings, 4 in the inner zone and 4 in
280 J. M. Molina et al.

the middle zone. Despite the comparatively low number of samples done in the
second period, the increase in the number of species found is significant. Two spe-
cies, in particular, appeared in the recent surveys that are noteworthy: Genidens
barbus and Macruronus magellanicus. A notable absence in the recent samplings is
the silverside mullet Mugil liza. Another fact increasing the number of species in the
recent surveys of the inner zone is the occurrence of species that were previously
registered only up to the middle zone, like Conger orbignianus, M. schmitti, and
S. bonapartii. These species seem to have moved to the inner areas in recent years.
In estuaries, salinity is an important factor influencing fish abundance and distri-
bution. The often abrupt changes in salinity common to estuaries can cause consid-
erable physiological demands on fishes. Although fishes living in estuaries are
adapted to salinity fluctuations, individual response to salinity stress varies by spe-
cies and scale. Low diversity and richness of fishes observed in hypersaline systems
have been attributed to the osmoregulatory stress that fishes withstand (Whitfield
2016). As expected, also in Bahía Blanca Estuary, richness values decrease as the
amplitude of salinity fluctuation increases. Inner zone species were captured in
salinity ranges of 25–36, while external zone species were captured in salinity
ranges of 30–36.

11.3  S
 patial, Seasonal, and Long-Term Variations in the Fish
Assemblage of Bahía Blanca Estuary

Knowledge on the spatial and seasonal dynamics of fish species, especially in tem-
perate coastal areas, is usually incomplete (García-Charton and Pérez-Ruzafa 2001;
Topping et al. 2006). This lack of knowledge is an issue for effective and sustainable
management of fish populations and their exploitation (McCormick and Choat
1987; Blyth-Skyrme et  al. 2006). The seasonal data obtained in Bahía Blanca
Estuary from the most recent set of surveys suggest that winter and spring contained
the highest species richness, 20 and 17 species, respectively (Table 11.2). Conversely,
autumn and summer had the lowest, 12 and 7, respectively. In the previous survey,
however, richness was highest in autumn and summer (26 and 23 species, respec-
tively), while it was lower in winter and spring (18 and 20 species, respectively).
Some of the migratory species appear all year round in both periods, such as
C. guatucupa, M. furnieri, and M. schmitti, although their peeks of abundance may
vary according to the species. The occurrence of these species all year round was
due to the presence of their juveniles, which remain in the estuary, while the adults
migrate elsewhere. Resident species, like Ramnogaster arcuata, are present all
year round.
For both periods, there is a clear differentiation between summer and autumn
samples in the one hand and winter and spring samples on the other. While no clear
pattern of species replacement between periods is evident in our results, these two
season groups coincide with the periods of warmer and colder temperatures
11  Ecology and Biology of Fish Assemblages 281

Table 11.2  Fish species presence in autumn (A), winter (W), spring (Sp), and summer (Su) in
Bahía Blanca Estuary in the two periods of sampling: 1979–1983 and 2017–2018
1979–1983 2017–2018
Species A W Sp Su A W Sp Su
Galeorhinus galeus ● ● ●
Mustelus schmitti ● ● ● ● ● ● ● ●
Notorhynchus cepedianus ● ● ● ●
Squatina guggenheim ● ● ● ●
Sympterygia acuta ● ● ● ●
Sympterygia bonapartii ● ● ● ● ● ●
Myliobatis goodei ● ● ● ● ● ●
Conger orbignianus ● ● ● ●
Brevoortia aurea ● ● ● ● ● ● ●
Ramnogaster arcuata ● ● ● ● ● ● ●
Anchoa marinii ● ● ●
Lycengraulis sp. ● ● ● ● ●
Genidens barbus ● ●
Macruronus magellanicus ●
Porichthys porosissimus ● ● ● ● ● ● ●
Mugil liza ● ● ● ●
Odontesthes argentinensis ● ● ● ● ● ● ●
Acanthistius patachonicus ● ● ●
Dules auriga ● ●
Pomatomus saltatrix ● ● ● ● ●
Parona signata ● ● ● ● ● ●
Cynoscion guatucupa ● ● ● ● ● ● ● ●
Macrodon ancylodon ● ● ● ●
Menticirrhus americanus ● ●
Micropogonias furnieri ● ● ● ● ● ● ● ●
Pogonias cromis ● ●
Umbrina canosai ●
Percophis brasiliensis ● ●
Stromateus brasiliensis ● ● ● ● ● ●
Paralichthys patagonicus ●
Paralichthys orbignyanus ● ● ● ● ● ●
Oncopterus darwinii ●
Symphurus jenynsi ● ● ● ●
Total 26 20 18 23 12 20 17 7

respectively, indicating that water temperature must play an important role in the
seasonal species turnover. Being a shallow estuary means that the waters of Bahía
Blanca Estuary heat up and cool off quicker than those of the open sea in front of it,
as the heat retention within the estuary is poor. This desynchronized heating and
cooling of the estuarine waters and the adjacent sea is important for species that
employ an optimum temperature strategy to increase fitness, for example (Elisio
282 J. M. Molina et al.

et al. 2017). Many chondrichthyan species have been shown to exhibit this behavior,
in which the individuals perform small-scale migrations to feed in high-temperature
areas, and then return to colder areas to digest (Neer et al. 2007). This journey may
involve swimming several kilometers and may take hours to complete. Seasonal
dynamics such as this is thought to be responsible for the differences in species
composition (Wonton 1992) and might be responsible for the species composition
differences observed in Bahía Blanca Estuary.
Further spatial structure in the fish community of Bahía Blanca Estuary was
found in the 1979–1983 surveys, using hierarchical classification procedures
(Sneath and Sockal 1973) applied to the matrix of similarity indexes between sam-
pling sites. Sampling sites close to shore were grouped together, some of them
located on the internal zone and the southwestern quadrant of Bahía Blanca Estuary.
A second separated cluster was formed by sites that were close to the main channel.
Further subdivisions of these two groups had much smaller mean similarity indexes.
The robustness of the discrimination between the two groups is supported by the
total (100%) coincidence of the groups using the two similarity indexes. As
explained at the first section of this chapter, depth plays an important role in struc-
turing fish communities. Additionally, habitat resources such as food and shelter
and environmental stressors vary greatly between close-shore intertidal environ-
ments and deeper subtidal zones. The main dredged portion of the inner zone of the
navigation channel might represent an artificial deep subtidal environment, promot-
ing species composition differences with the adjacent shallow intertidal shore habi-
tats as described by Carbines and Cole (2009) in a similar estuary in New Zealand.

11.4  S
 pecies Spotlight: Biological Description of Some
of the Fish Species of Bahía Blanca Estuary

11.4.1  Mustelus schmitti (Springer, 1939)

Locally called “gatuzo,” the narrownose smoothhound Mustelus schmitti (Fig. 11.3a)


is a small shark of the Triakidae family, attaining a maximum total length of 110 cm.
This shark is endemic to the Southwest Atlantic Ocean, from the south of Brazil to
the Argentinean Patagonia (22° S to 47°45’ S), dwelling from coastal waters to up
to 120 m of depth (Menni 1985). This shark is known to migrate seasonally in large
numbers between wintering grounds in southern Brazil and summer grounds in
Argentina (Vooren 1997) and also performs seasonal ingresses to estuaries, pro-
tected bays, and gulfs (Lopez Cazorla 2004; Chiaramonte and Pettovello 2000;
Colautti et al. 2010). M. schmitti is one of the most studied sharks of Argentina:
presently its reproduction, food habits, age, and growth and other of its biological
processes have been described (Menni 1985; Chiaramonte and Pettovello 2000;
Sidders et al. 2005; Segura and Milessi 2009; Colautti et al. 2010; Molina and Lopez
11  Ecology and Biology of Fish Assemblages 283

Fig. 11.3  Fish species spotlight of Bahía Blanca Estuary. (a) Mustelus schmitti, (b) Myliobatis
goodei, (c) Brevoortia aurea, (d) Ramnogaster arcuata, (e) Porichthys porosissimus, (f)
Odontesthes argentinensis, (g) Cynoscion guatucupa, (h) Micropogonias furnieri, (i) Pogonias
cromis, and (j) Paralichthys orbignyanus. (Photos a, b, c, f, g, h, i, and j modified from Cousseau
MB and Rosso JJ (2020) in Peces Argentina, edited by Vázquez Mazzini Editores, Ciudad de
Buenos Aires-Argentina. Photos by Gabriela Blasina d and e)
284 J. M. Molina et al.

Cazorla 2011; Molina et al. 2017). This species feeds mainly on crustaceans when
close to the coast and on fish as it migrates to deeper waters. Seasonal and ontoge-
netic differences in diet composition have been described, with polychaetes being
more important in the colder months of the year, while decapods become the main
prey item in the warmer months (Molina and Lopez Cazorla 2011). Larger nar-
rownose smoothhound sharks feed on larger crabs and fish, while neonates and
juveniles prey on a greater variety of crustaceans and polychaetes. The species
reaches sexual maturation at approximately 5 years old, with the males maturing
faster than the females and also attaining less weight and length (Molina et  al.
2017). The species migrates to the nursery areas of Bahía Blanca Estuary to give
birth in spring-summer and then mates and leaves as temperature drops by the
beginning of autumn. It gives birth to up to six pups, the size and number depending
on the size and age of the female. The maximum age determined for this species is
20 years old (Molina et al. 2017).
The narrownose smoothhound is the most exploited elasmobranch species in
Argentina, Brazil, and Uruguay, with an important percentage of the capture des-
tined to exportation to England and China (Molina and Lopez Cazorla 2011). This
shark is caught by both industrial and artisanal fishing fleets. The exploitation of this
species throughout its distribution range led to recent declines in its populations
despite maximum permitted catch regulations established by Argentina and Uruguay
(Molina and Lopez Cazorla 2011). The narrownose smoothhound is currently con-
sidered endangered in these two countries by the IUCN (Massa et al. 2010).

11.4.2  Myliobatis goodei (Garman, 1885)

The southern eagle ray, locally called “chucho,” Myliobatis goodei (Fig. 11.3b) is a
large stingray of the Myliobatidae family, reaching a meter in disc width. This spe-
cies distributes from south California (35°N) to the south of Argentina (40°S). This
chondrichthyan is diadromous, tolerating a wide range of salinity. It enters estuarine
waters to give birth (Refi 1975). Two very similar species cohabit with M. goodei,
M. ridens (Ruocco et al. 2012) and M. freminvillii (Aguiar et al. 2004).
M. goodei migrates to bays and estuaries during the warm months of spring and
summer, to improve neonates and juveniles’ access to food and shelter, providing
effective protection against predators and optimum conditions for development
(Castro 1993; Simpfendorfer and Mildward 1993). This migrating behavior of
M. goodei was also reported by Molina and Lopez Cazorla (2015), Jaureguizar et al.
(2003b), and Lopez Cazorla (1987) in Anegada Bay, Río de la Plata, and Bahía
Blanca Estuary, respectively.
Molina and Lopez Cazorla (2015) inferred that the mating and spawning season
for M. goodei may occur in summer, although the authors lack a complete yearly
series of gonadosomatic index values (GSI) to accurately demonstrate this. The
presence of mature males and pregnant females with highly developed embryos in
the uterus in summer, and recently born pups, would strengthen this hypothesis and
11  Ecology and Biology of Fish Assemblages 285

also imply that they became pregnant immediately after parturition (Hamlett 1999).
M. goodei in the study area behave as generalist feeders, with a uniform diet com-
posed mainly of bivalves. Trophic level of M. goodei in Anegada Bay (3.2) charac-
terizes it as a secondary consumer (Molina and Lopez Cazorla 2015).
This species is captured as bycatch but retained and sold internally as well as
exported in significant amounts. M. goodei is assigned as Data Deficient by the
IUCN, given the possible population threats it faces (Stehmann 2009).

11.4.3  Brevoortia aurea (Spix and Agassiz, 1829)

Locally called “saraca,” the Brazilian menhaden Brevoortia aurea (Fig. 11.3c) is a


planktonic clupeid that reaches up to 40 cm in total length. B. aurea is a euryhaline
fish, distributed from Salvador de Bahía in Brazil to San Matías Gulf (41°S) in
Argentina. This species is abundant in estuarine waters. It has been estimated that it
lives up to 12 years and the females reach larger sizes. Adults (larger than 20 cm TL)
are captured inside Bahía Blanca Estuary in the spring and summer months, while
only juveniles remain during autumn and winter (Lopez Cazorla 1985). While
B. aurea is not targeted by any specific fishing fleet, it is an important bycatch com-
ponent in coastal industrial and artisanal fisheries. The IUCN has evaluated the
conservation status of this species as Least Concern with a stable population trend
(Di Dario et al. 2017).

11.4.4  Ramnogaster arcuata (Jenyns, 1842)

Jenyns’ sprat Ramnogaster arcuata (Fig.  11.3d), locally called “saraquita,” is a


pelagic fish species of the Clupeidae family. It has a wide distribution in coastal
areas in the Southwestern Atlantic Ocean, from southern Brazil (estuary of the Patos
Lagoon) to Tierra del Fuego in southern Argentina (Lopez Cazorla et al. 2011). It is
confined to the external areas of rivers and coastal lagoon mouths, which are char-
acterized by moderate salinity ranges. It is an estuarine-resident species that com-
pletes the entirety of its life cycle within estuaries (Garcia and Vieira 2001). The
species reaches a total length of 130 mm and present a relatively short life span with
a maximum age registered of 3 years. Sexual maturity is reached at 76 mm total
length (1  year), and spawning season begins in spring (Lopez Cazorla and
Sidorkewicj 2009). According to its trophic habits, it has been classified as a zoo-
planktivorous feeder, and its diet composition exhibited monthly variability in the
main prey items consumed (Lopez Cazorla et al. 2011).
Ramnogaster arcuata is a main functional component of the ecosystem of Bahía
Blanca Estuary, where it is not only one of the most abundant species but also one
of the most commonly caught fish. It also represents a key food item for C. guatu-
cupa and P. orbignyanus, two of the most economically important fish species in the
286 J. M. Molina et al.

area. The coastal habits and short life span of R. arcuata make it an excellent organ-
ism to be considered as a bio-indicator of aquatic environmental health (Lopez
Cazorla and Sidorkewicj 2009; Ronda et  al. 2019). The IUCN has evaluated the
conservation status of this species as Least Concern (Di Dario et al. 2017).

11.4.5  Porichthys porosissimus (Cuvier, 1829)

Porichthys porosissimus (Fig. 11.3e) is a species of batrachoid known as Atlantic


midshipman and locally as “lucerna.” The reported maximum length for this species
is 32 cm TL. It is an abundant species, caught as bycatch throughout its distribution
range, from Rı́o de Janeiro, Brazil, to Golfo de San Matías (41°S), Argentina. This
species inhabits coastal waters from 30 to 200  m deep (Cousseau and Perrotta
2013). Other closely related species exhibit complex mating behaviors, which
include nest building, sound- and bioluminescence-mediated courtship, and paren-
tal care (Tsujii et al. 1972). Photophores, a special type of skin cells that produce
bioluminescence, are present in P. porosissimus and would likely play a homolo-
gous role in mating and communication between individuals. Mature individuals of
this species croak when disturbed, meaning they are capable of producing sound,
much like their northern counterparts.
Shrimp trawling and traps capture this species as bycatch, but only one study
exists on the biology of this species (Vianna et al. 2000). In Bahía Blanca Estuary,
it is captured mainly in the external zone, all year long. The largest adults were pres-
ent during spring, while the smallest juveniles were found in summer. Size range
was from 9 to 31 cm TL, and the most frequent sizes were 17–31 cm TL (Lopez
Cazorla 1987). The IUCN has yet to evaluate the conservation status of this species.

11.4.6  Odontesthes argentinensis (Valenciennes, 1835)

Locally called “pejerrey,” Odontesthes argentinensis (Fig.  11.3f) is a large-sized


silverside that is widely distributed along the Atlantic Ocean coast between Sao
Paulo, Brazil, and Rawson in Argentina (García 1987; Dyer 2000). It is a planktonic
species that feeds on zooplankton, mainly crustaceans. This fish reaches 48 cm in
total length, the males being larger than the females (Molina 2013). O. argentinen-
sis inhabits shallow coastal waters, and juveniles have been found to be abundant in
the surf zones of sand beaches. As for other atherinids, this fish shows a great phe-
notypic plasticity that allows its adaptation to different environments (Bamber and
Henderson 1988) involving a wide range of salinities. This allows this species to
inhabit estuaries and inshore waters, where it likely migrates during late spring and
summer to reproduce (Cousseau and Perrotta 2013; Beheregaray and Levy 2000;
Bemvenuti 2005). According to gonad ripening takes place between September and
November with a peak in October, in the nearby location of Anegada Bay. Llompart
11  Ecology and Biology of Fish Assemblages 287

et al. (2013) also describe age and growth of this species, which attains a maximum
of 7 years, growing quickly in the first 2.
Odontesthes argentinensis in Bahía Blanca Estuary behave similarly, with the
main captures being done in the external zone using gill nets and channel closure
nets. Bahía Blanca Estuary represents a breeding area for this species, and spawning
occurs from late August to November (Lopez Cazorla 2004). The commercial
importance of this species is limited to the Argentinean market and is targeted by
coastal artisanal fleets (Cousseau and Perrotta 2013). The IUCN has yet to evaluate
the conservation status of this species.

11.4.7  Cynoscion guatucupa (Cuvier 1829)

The striped weakfish, Cynoscion guatucupa (Fig. 11.3g), is locally called “pesca-


dilla de red.” It is a pelagic fish species, which has a wide geographical distribution,
extending from Río de Janeiro (22°S) in Brazil to San Matías Gulf (43°S) in
Argentina (Cousseau and Perrotta 2013). This fish presents dietary shifts during
ontogeny. It feeds from pelagic to benthic crustaceans on its early stages (mysids,
sergestids, shrimps) and eventually increases progressively in ichthyophagy (mainly
an increase consumption of alevins and young fish) as it develops into adulthood. It
present dietary seasonal differences could be due to changes in abundance and
availability of its prey species in the environment (Lopez Cazorla 1996; Sardiña and
Lopez Cazorla 2005a). C. guatucupa performs seasonal migrations, swimming
northwards between autumn and spring, leaving the fishing grounds of Uruguay and
Argentina to move to the coast of Brazil, only to return to the south in summer
(Villwock de Miranda and Haimovici 2007). Lopez Cazorla (1996) reports the
influence of changes in temperature and salinity as triggers for the spawning move-
ments of C. guatucupa. Spawning occurs outside of the estuary from spring to mid-
autumn (Cassia 1986; Lopez Cazorla 2000), and the fry is pushed into the estuary
by tidal movements. Small juveniles recruited from late spring move to deeper
waters (25–50 m) in late autumn, when they reach a mean total length of 9.8 cm
(age 0+). They remain there for the next 1–2 years before joining the adult stock’s
seasonal movements (Haimovici et  al. 1996; Lopez Cazorla 2000; Sardiña and
Lopez Cazorla 2005a). The total length of adult fish ranges from 34 to 63 cm, and
the ages range from 3 to 23 years (Lopez Cazorla 2000; Ruarte and Sáez 2008).
The IUCN has yet to evaluate the conservation status of this species.

11.4.8  Micropogonias furnieri (Desmarest 1823)

The whitemouth croaker Micropogonias furnieri (Fig. 11.3h), locally called “cor-


vina rubia,” is a demersal fish of the family Sciaenidae, inhabiting coastal waters up
to 60  m deep. It is a euryhaline fish distributed widely in marine and estuarine
288 J. M. Molina et al.

systems of the eastern American coast from the Gulf of Mexico (20°20’N) to “El
Rincón” (41°S) in Argentina (Carozza et al. 2004). The maximum recorded size for
the species is 74 cm in TL, reaching sexual maturity at 33 cm of TL, which corre-
sponds to 4 or 5 years of age. The reproductive period of M. furnieri is very long and
extends from spring to summer (Macchi et al. 2003). Spawning occurs in highly
saline coastal waters, and subsequently, larval M. furnieri enter coastal estuaries
during winter months. In Bahía Blanca Estuary, species reproduction occurs in El
Rincón area during spring, with the subsequent drift of eggs and larvae into estuary.
Juveniles with sizes of 2 and 18 cm total length (Lt) remain inside the estuary from
early summer to winter and then leave the region. At late spring, entrance to the
estuary of individuals in the adult state with a size range from 30 to 72 cm Lt begins,
and they remain in the area until autumn (Lopez Cazorla 2004). Young-of-year
(YOY) and adult M. furnieri utilize estuarine habitats for feeding and growth
(Jaureguizar et al. 2003a; Lopez Cazorla 2004). The species has been identified as a
generalist feeder, and its stomach contents largely reflect seasonal changes in prey
availability, meaning it has a broad dietary niche width (Mendoza-Carranza and
Vieira 2008). Previous studies in estuarine habitats have documented ontogenetic
changes in diet. YOY individuals rely heavily on polychaetes in their diets but also
consume other food items such as chaetognaths, copepods, and amphipods.
Evidence indicates that M. furnieri changes its feeding habits as it gets larger, rely-
ing more heavily on organisms such as mysids and fish. Adults have been described
as opportunistic bottom-feeders that eat decapod crustaceans, such as crabs and
shrimps, polychaetes, and, occasionally, small fishes (Sardiña and Lopez Cazorla
2005b; Giberto et al. 2007; Blasina et al. 2016).
Micropogonias furnieri is one of the most abundant demersal fishes of South
American estuaries and an important component of artisanal and coastal industrial
fisheries in Brazil, Uruguay, and Argentina (Carozza et al. 2004). M. furnieri fishery
in the Río de la Plata estuary is mainly artisanal, with fish being caught mostly in
winter in Samborombón Bay area of Argentina and during spring and summer in
Santa Lucia area of Uruguay (Jaureguizar et al. 2003a, b). In Bahía Blanca Estuary,
it was the second most important fish resource, captured in spring and summer. It
reached 16% of the commercial landings between 1972 and 1992, although between
1994 and 1996 landings strongly decreased and values as low as 2% were reported
(Lopez Cazorla 2004). The IUCN has evaluated the conservation status of M. furni-
eri as Least Concern, although a decreasing population trend is mentioned (Aguilera
et al. 2015).

11.4.9  Pogonias cromis (Linnaeus, 1766)

The black drum Pogonias cromis (Fig. 11.3i), locally called “corvina negra,” is a
demersal coastal fish distributed along the western Atlantic Ocean from
Massachusetts, USA, to south of Buenos Aires Province in Argentina. It is an
estuarine-­
dependent species and the largest sciaenid found in the estuarine
11  Ecology and Biology of Fish Assemblages 289

environments of Argentina. It reaches up to 120 cm in TL and is sexually mature at


the end of the second year of life, at 28.5–33  cm of TL (Cousseau and Perrotta
2013). Spawning takes place in regions associated with estuaries, inside or outside
of them, and mainly in spring. P. cromis juveniles live in estuarine areas as they can
tolerate a wide range of salinities and water temperatures, eventually moving to
offshore marine waters when they reach the adult stage. Adults are usually common
in shallow coastal and estuarine waters and occasionally occur further from the
coast (Macchi et al. 2002; Rubio et al. 2018).
Juvenile and adult P. cromis exploit a variety of benthic food resources and can
use their strong pharyngeal teeth to crush the shells of mollusks and crustaceans
(Blasina et al. 2016; Rubio et al. 2018). Gut content analyses have identified signifi-
cant seasonal differences in the diet composition and trophic niche breadth. Because
of this P. cromis has been classified as an opportunist predator (Blasina et al. 2010).
Pogonias cromis is the target of an important recreational and commercial fish-
ery in the Gulf of Mexico, and it is commercially harvested in inshore waters of
Samborombón Bay, a semi-enclosed region inside the Río de la Plata estuary in
Argentina. Fishing effort occurs mainly between late winter and summer, but is
especially high during spring, when P. cromis forms large schools in shallow waters.
This behavior contributes to intensified commercial and recreational activity on this
species (Macchi et  al. 2002; Rubio et  al. 2018). P. cromis is classified as Least
Concern status by the IUCN, but a decreasing population trend is mentioned (Chao
et al. 2015). In Bahía Blanca Estuary, in the period 1979–1983, this species was
registered only in three opportunities, all in the same winter month (June) and in
Ingeniero White (the middle zone) (Lopez Cazorla 1987). In 2017–2018 this spe-
cies was captured also in the middle zone, but during summer (Table 11.1 and 11.2).

11.4.10  Paralichthys orbignyanus (Jenyns, 1842)

The flounder, locally called “lenguado,” Paralichthys orbignyanus (Fig. 11.3j) is a


commercially important species generally found in the shallow waters from Rı́o de
Janeiro (22° S) southwards in Brazil to San Matías Gulf (41° S) in Argentina
(Cousseau and Perrotta 2013). Two other species of the genus Paralichthys live in
Argentinean waters, P. isosceles and P. patagonicus (Lopez Cazorla 2005). P. orbig-
nyanus is a typical benthonic fish with a wide temporal and spatial distribution in
Bahía Blanca Estuary. Two other species of the genus Paralichthys live in
Argentinean waters P. isosceles and P. patagonicus (Lopez Cazorla 2005).
The maximum ages recorded for both sexes corresponded to 6 years in males and
7 years in females, respectively, suggesting that P. orbignyanus is not a long-living
species, and the females are longer and heavier than males (Lopez Cazorla 2005).
Similar data have been published for P. adspersus by Escobar (1995), who aged this
species and found they live a maximum of 6 years and reach up to 74 cm in total
length. In addition, Dı́az de Astarloa and Munroe (1998) observed that the longest
TL for P. orbignyanus was 61 cm for males and 103 cm for females, respectively,
290 J. M. Molina et al.

although they made no reference to age. Females are longer and heavier than males.
Larger size in females could be indicative of a life history strategy supportive of
increasing egg production (Masuda et  al. 2000). The growth difference between
females and males was also observed in P. adspersus females which exhibit a length
significantly larger than males (Escobar 1995). The length growth registered in
P. orbignyanus males and females in Bahía Blanca Estuary was significantly higher
than that of P. isosceles reported by Fabré and Cousseau (1990).
Paralichthys orbignyanus has an active growth period in summer and interrupts
its growth in winter. Spawning occurs in the period extending from November to
January (spring-summer), and the eggs and larvae of this species are found in
January and February (summer) in the area next to the estuary mouth (Lopez
Cazorla 2005). This suggests that spawning occurs out of the estuary, as with other
species of bony fish, such as C. guatucupa (Lopez Cazorla 1996, 2000). A similar
behavior has been described for other Pleuronectidae (Kareius bicoloratus) which
spawn off the coasts, at depths ranging from 20 to 50 m. Once larvae reach 10–15 mm
in total length, they approach the coast, migrating to nursery grounds (Malloy et al.
1996). The IUCN has yet to evaluate the conservation status of this species.

11.5  Fish Habitat Uses

The life of fish in estuaries is conditioned by the abundance of food and variations
in the abiotic parameters of the water (Elliott et  al. 2007). Fish species found in
estuaries use these systems in a variety of ways, and this usage can change at differ-
ent life stages. Ecological characteristics of fish species found in estuaries can be
divided into three main functional aspects: (1) the use fish make of the estuary dur-
ing their life cycle, (2) reproductive characteristics, and (3) feeding preferences and
strategies. Elliott et al. (2007) name the three functional groups as “estuarine use
functional group,” “reproductive mode functional group,” and “feeding mode func-
tional group” respectively.

11.6  Estuarine Use Functional Group

Many species spawn in marine waters and enter estuaries for variable periods, while
others complete their life cycle within the estuary, and yet others employ the estuary
as a feeding area (Potter et  al. 2015). Thus, fish assemblages include estuarine-­
resident species, freshwater and marine species that typically use estuaries at a spe-
cific life stage, as well as migratory diadromous species (Elliott et al. 2007). Each
of these categories is considered to contain two or more functional guilds that
represent characteristics associated with the spawning, feeding, and/or refuge loca-
tions, which in some cases involve migratory movements between estuaries and
other ecosystems (Whitfield 2016).
11  Ecology and Biology of Fish Assemblages 291

Guild approach categorization of fishes was proposed by Elliott et al. (2007) and
refined by Potter et al. (2015). Two fish guilds are dominant in Bahía Blanca Estuary
(Table 11.3): marine estuarine-opportunists and estuarine-residents; they are repre-
sented in estuaries by different life stages and are associated with different food
chains (see Feeding mode functional group section). Marine-estuarine opportunists
are predominantly juvenile fish making use of this ecosystem as a nursery area.

Table 11.3  Species frequency of occurrence in Bahía Blanca Estuary. R rare species (up to 33%
of occurrence), C common species (34–66% of occurrence), F frequent species (more than 66% of
occurrence)
Species Resident Migrant Straggler
Galeorhinus galeus R
Mustelus schmitti F
Notorhynchus cepedianus R
Squatina guggenheim R
Sympterygia acuta F
Sympterygia bonapartii F
Myliobatis goodei F
Conger orbignianus F
Brevoortia aurea F
Ramnogaster arcuata F
Anchoa marinii F
Lycengraulis sp. C
Genidens barbus R
Macruronus magellanicus R
Porichthys porosissimus F
Mugil liza C
Odontesthes argentinensis F
Acanthistius patachonicus R
Dules auriga R
Pomatomus saltatrix C
Parona signata C
Cynoscion guatucupa F
Macrodon ancylodon R
Menticirrhus americanus R
Micropogonias furnieri F
Pogonias cromis R
Umbrina canosai R
Percophis brasiliensis R
Stromateus brasiliensis R
Paralichthys patagonicus R
Paralichthys orbignyanus F
Oncopterus darwinii R
Symphurus jenynsi C
Total species 14 17 2
292 J. M. Molina et al.

These fish species regularly enter estuaries in substantial numbers but use, to varying
degrees, coastal marine waters as alternative nursery areas. In small-scale studies,
some authors have pointed out that estuary mouth width was the most important
variable explaining a significant part of the variability in fish species richness
(Nicolas et al. 2010; Pasquaud et al. 2015). Estuaries with large mouths can attract
numerous brackish water species, as well as marine-estuarine opportunist fish spe-
cies (Martinho et al. 2009; Vinagre et al. 2009). Estuarine-resident guilds are com-
posed by species with populations in which the individuals complete their life cycle
within the estuary. While a number of the marine estuarine-opportunist species have
economic importance for the recreational and local artisanal fishermen, none of the
small resident species, which are a highly productive component in this estuary, are
utilized. In addition, a number of the marine straggler species are frequently regis-
tered. These species enter estuaries sporadically and in low numbers and are most
common in zones where salinity typically does not decline far below approximately
33 ups. Due to hypersaline conditions in Bahía Blanca Estuary, no freshwater fish
species has been registered (Table 11.3).
In Bahía Blanca Estuary, the biology and life history of each species condition
the use they make out of the resources available. Resident fish species (i.e., estuarine-­
resident) here, for example, have a remarkable tolerance to environmental varia-
tions, while migrant species (i.e., marine estuarine-opportunists) exhibit behaviors
that allow them to exploit the high productivity of the intertidal ecotone and leave
the area when the conditions become unfavorable. The latter species have a com-
paratively low tolerance to shifts in abiotic variables.

11.6.1  Resident Species of Bahía Blanca Estuary

In Bahía Blanca Estuary, there are resident chondrichthyans and osteichthyans.


Skates of the genus Sympterygia lay eggs all year round and are commonplace all
along the waters of the estuary. The most ubiquitous species is Sympterygia
bonapartii, the shortnose south Atlantic skate. This species lays eggs protected with
a fibrous black capsule and with four tendrils. The eggs are placed around sub-
merged vegetation or debris so that the tendril holds them in place.
The absence of natural hard substrate and the relative scarcity of submerged
macroalgae offer little shelter for resident reef fish; however, species of genera
Dules and Acanthistius, both associated with rocky bottoms and reefs, have been
found to occur within the waters of the estuary. Soft sediments, however, are ideal
for soles and flounders. In Bahía Blanca Estuary, there are four species of flounders
and one of soles. Of these species, only the flounder Paralichthys orbignyanus
occurs frequently and with considerable abundances, in so as to become a targeted
species of the local artisanal fishermen.
The high turbidity of the estuary means that ambush predators like the southern
Atlantic midshipman (Porichthys porosissimus) have no problem procuring food.
Indeed this species is very abundant in the estuary, where it performs seasonal
11  Ecology and Biology of Fish Assemblages 293

migrations. It enters from the external area of the estuary in spring and summer to
mate and care for their young. This species is caught abundantly in shrimp nets from
October to February during ebb tide, suggesting it uses the currents to move in and
out of the inner part of the estuary on a daily basis. By autumn/winter it is already
unlikely to fish any in the inner zone of the estuary. A possible explanation is the
drop in water temperature which in the inner, and shallower, section of the estuary
is much more intense and sharp. Changes in water salinity also offer another pos-
sible explanation, while it fluctuated sharply in the inner zone, values in the external
and middle areas remained relatively constant (Lopez Cazorla 2004), so perhaps the
distribution of this fish is due to a behavioral escape from the fluctuating conditions,
of temperature and salinity, of the inner zone. Prey availability may also be a factor
contributing to this species’ distribution. Its main prey, the prawn Peisos petrunkev-
itchi, spawns by the end of winter, prompting an abundance peak during spring and
summer. In the late summer months, spawners concentrate in the outer part of the
estuary (Mallo and Cervellini 1988), representing a valuable protein and energy
source for P. porosissimus, who might follow their seasonal movement patterns.
Little else is known about this species, as there are no specific studies on it yet.

11.6.2  Migrant Species of Bahía Blanca Estuary

In Bahía Blanca Estuary, the increases in water temperature and salinity during the
warmer months of spring are thought to trigger migratory movements of certain
species of fish. However, little is known about the drivers of migration movements
in the migrant species of Argentina. It is theorized that some species utilize the estu-
ary as feeding ground, others as a nursery for their young, as mating area, or spawn-
ing waters. Regardless of the driver, several fish species shoal into the estuary in
different times of the year and then leave.
The high productivity, the availability of refuge, and favorable conditions in
spring and early summer in Bahía Blanca Estuary seem to be a reasonable explana-
tion as to what draws migrating fish species to these waters. Sciaenids like Cynoscion
guatucupa and Micropogonias furnieri migrate to the estuary and spawn before
entering estuarine waters for feeding. Adults of C. guatucupa presented two abun-
dance peaks in the estuary: one in early autumn and a more important peak in early
spring. On the other hand, the higher abundance of M. furnieri adults into the estu-
ary has been registered during spring and summer (Lopez Cazorla 2004). Given that
nutrient load increases around winter, with a consequent increase in phytoplankton
biomass, it is not surprising that these sciaenids have a bimodal spawning behavior.
This behavior might favor the larvae produced in the early spawning event with
more food availability, at the expense of lower growing temperatures. Larger juve-
niles of these species, preying on copepods and brachyuran larvae, are captured at
the end of spring and throughout the summer, autumn, and winter, exploiting the
zooplankton biomass explosion that follows the peaks of phytoplankton.
294 J. M. Molina et al.

The abundant invertebrate assemblages also represent an outstanding food source


for benthic predators like M. furnieri. This species uses the intertidal during high
tide to prey on polychaetes and crabs among the “cangrejales.” Local fishermen
report they can see the nuzzling marks of M. furnieri in the mud of the “cangrejal”
during low tide. Although this species has several specializations in its mouth to
feed on hard-shelled prey, like crabs, it will opt for other benthonic prey if they are
abundant, exhibiting an opportunistic feeding behavior.
Chondrichthyans also migrate into the estuary to give birth and lay eggs. Adults
of the triakid Mustelus schmitti can be found in the waters of the estuary from the
end of winter up to the beginning of summer. The juveniles can be found from sum-
mer to early winter. When this species enters the estuary, it shifts the diet to con-
sume almost exclusively crabs. Females migrate into the inner parts of the estuary
to give birth to 2–8 pups and then mate with the males. Stingrays of the genus
Myliobatis also move into the estuary in spring to give birth and mate, utilizing the
abundant populations of polychaetes and crabs to load up on energy reserves. They
can be found up until April (autumn). Two species are known to occur here, M. goo-
dei and M. ridens, which were thought to be one species until recently (Ruocco
et al. 2012).
Large sharks like the seven-gill shark (Notorynchus cepedianus), copper shark
(Carcharhinus brachyurus), and sand tiger shark (Carcharias taurus) also utilize
estuarine waters as a nursery area during the warm months of spring and summer
and also as a hunting ground for both fish and pinnipeds. Studies on these species in
Bahía Blanca Estuary are lacking; the only available reference is the presence of
N. cepedianus (Lopez Cazorla 1987).

11.6.3  Straggler Species

This category is represented by fish species that occur “accidentally” in estuaries;


they generally occupy it for only very short periods of time and in limited areas. In
Bahía Blanca Estuary, two species have been registered that come from other
regions and that visit the estuary to use its favorable conditions opportunistically:
Genidens barbus and Macruronus magellanicus. These findings represent the first
record of both species of fish in Bahía Blanca Estuary. G. barbus is an anadromous
species inhabiting estuaries and the marine continental shelf from Bahía in Brazil
(17° 00′ S) to San Blas Bay in Argentina (40° 32′ S) (Avigliano and Volpedo 2015).
On the other hand, M. magellanicus is distributed on intermediate platform of the
Argentine Sea and in the gulfs of San Jorge and San Matías. On the platform its
distribution is closely related to the Malvinas current, and it has been recorded at
temperatures that varied from 3 to13° C (Cousseau and Perrotta 2013). Possible
explanations for these irregular records are diverse; in some cases they could be due
to atypical abiotic conditions, such as the extraordinary incursion of marine waters
or the occurrence of adverse climatic conditions in the area outside the estuary, or it
could even be due to an intrusion when following their preys.
11  Ecology and Biology of Fish Assemblages 295

11.7  Reproductive Mode Functional Group

The spawning features and the degree of parental care are required to define repro-
ductive modes in fishes (DeMartini and Sikkel 2006; Elliott et al. 2007). Fish spe-
cies are first divided into oviparous and viviparous, according to the maternal
investment in individual offspring (DeMartini and Sikkel 2006). Viviparous species
present internal fertilization and live-bearing of young with a broad range of post-­
fertilization provisioning, from no (strictly lecithotrophic viviparity) with live-­
bearing of young provisioned entirely by ovum yolk to extensive provisioning
beyond the nutrition provided by ovum yolk (matrotrophic viviparity). On the other
extreme are the oviparous species with lecithotrophic maternal provisioning (lim-
ited to the yolking of ovarian oocytes prior to fertilization) and external fertilization.
Oviparous species are distinguished on the basis of their egg characteristics, mode
of release, and the degree of parental care provided to eggs (Franco et al. 2008).
These reproductive modes determine offspring survival; according to the optimiza-
tion theory, parental care implies a greater investment on offspring (therefore, larger
individual offspring) at the expense of the number of offspring in which it is per-
formed. Within of viviparous species category, the most extreme example of parental
care is matrotrophic viviparity (DeMartini and Sikkel 2006).
According to reproductive modes described in Jaureguizar et al. (2016), different
reproductive strategies occur in estuarine and migrant assemblages of Bahía Blanca
Estuary. The 78% of the species are oviparous (62·4% producing pelagic eggs, 6·2%
eggs that settle on the substratum and adhesive eggs 9·4% and), followed by 15·6%
of viviparous and 6·4% of ovoviviparous species. Of the 13 estuarine-resident spe-
cies registered, most (61·5%) produce pelagic eggs, spawning within the estuary or
in its influence area (e.g., Ramnogaster arcuata, Oncopterus darwinii and Porichthys
porosissimus). Species that produce adhesive eggs that are able to attach to substrata
and the vegetation are second in importance (23·5%; e.g., Odontesthes argentinen-
sis), and finally two species (15%) were ovoviviparous (Sympterygia acuta and
S. bonapartii). The reproductive strategy of marine migrant fish was similar to that
of the estuarine-residents, as 58·8% of these species spawn pelagic eggs (e.g.,
Micropogonias furnieri, Cynoscion guatucupa, and Brevoortia aurea), followed by
viviparous species (29·4%; e.g., Mustelus schmitti and Myliobatis goodei). There
was only one ovoviviparous species whose male carry the eggs in their mouth
(Genidens barbus).

11.8  Feeding Mode Functional Group

Trophic ecology studies seek to identify the feeding habits of species through the
analysis of the major items consumed. Knowledge on the diets of species is one of
the basic requirements for a closer examination of the relationships between
296 J. M. Molina et al.

organisms in a given ecosystem. A very close relationship exists between the quan-
tity, quality, and availability of food and the distribution and abundance of consumer
organisms (Dantas et al. 2013; Campos et al. 2015). The structures of fish assem-
blages that use the shallow areas of estuaries are strongly influenced by trophic
relationships (Blasina et al. 2016). Knowledge on the structure of the trophic web
allows the description of the energy flow in an ecosystem and the understanding of
the ecological relationships among organisms (Dantas et al. 2013).
Although opportunism is a widely reported feeding strategy used by estuary-­
associated fish (Gerking 1994; Elliott et al. 2007), intrinsic factors such as morpho-
logical and behavioral constraints set the boundaries on what food items can be
taken from the environment, thus affecting the individual’s ability to obtain certain
prey. Extrinsic interactions (of a species or an individual with both the environment
and other community members) will also influence the diet of said individual (Elliott
et al. 2002; Horn and Ferry-Graham 2006). For example, the foraging range of the
fish will affect which prey are encountered and can be potentially included in the
diet, while mouth adaptations and morphology will determine which among the
potential prey are ultimately consumed. The trophic categories from literature were
revised by Franco et al. (2008), and feeding mode functional groups were identified
by combining information on predominant diet and feeding location. The trophic
groups, indicating the main types of food exploited by fish within estuarine environ-
ments and the estuarine compartments (e.g., pelagic, benthic) where these resources
are taken, are:
• Microbenthivores: feeds mainly on benthic, epibenthic, and hyperbenthic fauna,
with prey size <1 cm.
• Macrobenthivores: feeds mainly on benthic, epibenthic, and hyperbenthic fauna,
with prey size >1 cm.
• Planktivores: feeds predominantly on zooplankton and occasionally on phyto-
plankton in the water column, mainly by filter feeding.
• Hyperbenthivores/zooplanktivores: feeds just over the bottom, predominantly
either on smaller mobile invertebrates living over the bottom or zooplankton;
diverse prey capture mechanisms (ram, suction, or manipulation).
• Hyperbenthivores/piscivores: feeds just over the bottom, predominantly either
on larger mobile invertebrates living over the bottom or fish; diverse prey capture
mechanisms (ram, suction, or manipulation).
• Detritivores: feeds on all the small organisms in or on the surface layer of the
substratum (e.g., benthic algae such as diatoms, microfauna, and, to a lesser
extent, smaller meiofauna) and associated organic matter (usually of plant ori-
gin); ingests relatively large volumes of sand or mud (by suction mechanisms);
digests the food material and passes out the inorganic particles.
• Herbivores: grazes predominantly on living macroalgal and macrophyte
material.
• Omnivores: ingests both plant and animal material by feeding mainly on macro-
phytes, periphyton, epifauna, and filamentous algae.
11  Ecology and Biology of Fish Assemblages 297

11.8.1  Feeding Habits of the Fish of Bahía Blanca Estuary

With the objective to comparatively describe and analyze the trophic spectrum of
the most common species and their interrelationship, Lopez Cazorla (1987) studied
the stomach content of 1035 specimens belonging to 7 species of fish, between 1980
and 1982. The species studied were Sympterygia bonapartii, Mustelus schmitti,
Porichthys porosissimus, Odontesthes argentinensis, Micropogonias furnieri,
Cynoscion guatucupa, and Paralichthys orbignyanus. Her results indicate that
Neohelice granulata is the most important food source for almost all the species
studied. The remainder of the dietary items presented great differences in the pro-
portions consumed by each of the species. Sympterygia bonapartii feeds mainly on
benthic decapodic crustaceans, predominately peneids and brachyurans. The diet of
Mustelus schmitti was found to be composed mainly of benthonic decapod crusta-
ceans, polychaetes, and young fish. Porichthys porosissimus fed almost exclusively
on Peisos petrunkevitchi with a small percentage of misidaceans. Odontesthes
argentinensis showed a diet consisting principally of Neohelice granulata, gastro-
pods, misidaceans, and amphipods. Micropogonias furnieri consumed N. granulata
and P. petrunkevitchi as their most common prey. Cynoscion guatucupa presented a
diet which consisted mainly of Pleoticus muelleri, Artemesia longinaris,
P. petrunkevitchi, and young fish of the Ramnogaster arcuata, Brevoortia aurea,
and O. argentinensis. Paralichthys orbignyanus preys on the following fish species:
O. argentinensis, B. aurea, R. arcuata, Parona signata, C. guatucupa, Pomatomus
saltatrix, P. orbignyanus, and Lycengraulis olidus (Fig. 11.4).
Crabs and polychaetes constitute the principal or most important food for ben-
thonic and demersal fish, while Decapoda Natantia such as A. longinaris, P. muel-
leri, and P. petrunkevitchi were the principal food of demerso-planktonic fish.
Although fish select certain types of prey, depending on their size and habitat, con-
sumption of prey depends, above all, on the availability and the community struc-
ture of the prey.
In this regard, a brief description of the community structure of prey can help to
interpret these findings. In Bahía Blanca Estuary, an adequate supply of food for the
diverse life stages of the fish communities depends on a sequential abundance of
progressively bigger prey, from autumn to late summer. Nutrient abundance
increases by the end of summer, reaching a peak in the autumn months (between
April and June) (see Chap. 3). This increase in the nutrient load produces a trophic
cascade, triggering blooms in the planktonic fractions. The main phytoplankton
bloom occurs between June (winter) and October (spring), followed by the meso-
zooplankton explosion in November (spring). November is the month when misci-
daceans and brachyuran larvae, important prey items in the diet of several fish
species of the estuary, register their maximum abundance. Teleost larvae of demer-
sal species, like Cynoscion guatucupa, feed on small planktonic crustaceans like
Acartia tonsa and then shift their diets to larger prey like miscidacean Arthromysis
magellanica. Peneid crustaceans, like Peisos petrunkevitchi, Artemesia longinaris,
and Pleoticus muelleri, occur mainly between January (summer) and June and are
298 J. M. Molina et al.

Fig. 11.4  Frequency of occurrence of major prey items in the stomach contents of the studied fish
species of Bahía Blanca Estuary

the food of larger juveniles, which eventually prey on other fish. Pups of chondrich-
thyan are too large to feed on small planktonic prey, exploiting the abundance of
larger crustaceans by the end of spring and mid-summer. Resident species like
Ramnogaster arcuata breed in spring, when abundant food is available for the adults
but also for the larvae and juveniles.
The benthic fish species that feed on reptant crustaceans and polychaetes do not
experience such a pronounced seasonality as that of the plankton. The intertidal
ecotone in Bahía Blanca Estuary is dominated by euryhaline vascular plants, like
Sarcocornia perennis and Spartina alterniflora. These plants dominate the mud flats
that cover most of the intertidal environment of the estuary. This habitat is charac-
terized by fast-changing temperatures, oxygen-poor substrate, and lack of refuge.
However, the association between the burrowing crab Neohelice granulata and the
salt marsh plant S. perennis creates a unique habitat for other invertebrates, as it
promotes sediment oxygenation and nutrient turnover (Parodi 2004). This particular
type of habitat is called “cangrejal,” in Spanish meaning “land of crabs.” Polychaetes
are particularly benefited, and their abundances in sites with cangrejal are signifi-
cantly higher than in sites without cangrejal. This bolstered abundance of crabs and
polychaetes might explain why these are the most predominant and important prey
11  Ecology and Biology of Fish Assemblages 299

items for the fish species present in Bahía Blanca Estuary. Cangrejal sites could be
essential for the thriving of the whole trophic network of the estuary (Elías
et al. 2004).

Box 11.1 Artisanal Fishery in Bahía Blanca Estuary


It is widely recognized that small-scale fisheries play an important role in
providing food and livelihood to people, contributing to poverty reduction and
sustainable development in several places around the globe (FAO 2005). In
particular, developing countries greatly benefit from this type of fishery, as
they constitute the main source of both food and income for people living
along the coast (Blaber et al. 2000). Artisanal fishermen communities develop
an intricate relationship with the marine environment and the species that
constitute their sustenance, which greatly aids the success of conservation and
management practices.
The artisanal fishery in Bahía Blanca Estuary has been carried out since the
beginning of 1900. Fishermen employ a combination of fishing gear through-
out the year, consisting of shrimp nets, channel closures, and gill nets. Each
type of net is employed to target a particular species or group of species and
is performed in different times of the year to increase yield and reduce bycatch
of unwanted species. Traditionally a family business, fishermen formed the
Cooperativa Pesquera Whitense (White’s Fishery Cooperative), a cooperative
organization, between 1945 and 1999. The cooperative had a fish processing
plant and handled the marketing of the fishermen catch, allowing a better
income for the families of the fishermen and a regulatory frame for the activ-
ity. However, this organization closed in 1999 due to the collapse of the arti-
sanal fisheries of several species (Lopez Cazorla et al. 2014).
The Collapse of a Fishery
Between 1972 and 1992, catches of Cynoscion guatucupa reached 50% of the
total annual landings. However, at the end of the 1990s, catches dropped to
15%. In 2004 the Argentinean Government implemented fishing closures in
El Rincón area, as a management measure to control the increasing landings
and the decrease in biomass of C. guatucupa and several other commercial
species (Carozza and Fernández Aráoz 2009). But the closure came too late to
save the artisanal fishermen of Bahía Blanca Estuary; between 2000 and 2004,
the artisanal fishery in the south of Buenos Aires Province collapsed.
In Lopez Cazorla et al. (2014), we explore the causes of the collapse of the
fishery within the estuary. A cursory look at the reported landings suggested
that the greatly increased fishing pressure from industrial vessels operating
outside the estuary had depleted the stocks of C. guatucupa. The annual

(continued)
300 J. M. Molina et al.

Box 11.1  (continued)


commercial landings of the Argentinean fleet reached 5000  t in the early
1970s. After that, landings increased sharply to 20,000–48,000 t in the decade
between 1995 and 2004 (Villwock de Miranda and Haimovici 2007). From
1992 to 1998, the number of industrial vessels targeting striped weakfish at
the northern continental shelf of Argentina doubled, and the amount of effort
measured in fishing hours quadruplicated (Ruarte et al. 2000). For C. guatu-
cupa, the first scientific results pinpointing the decrease in the yields of the
artisanal fleet fishing this species in Bahía Blanca Estuary were presented by
Lopez Cazorla (2004). Carozza et al. (2004) mentioned that since 2000, there
was an increase in landings of several coastal species at El Rincón area, espe-
cially during the reproductive season of most of them. Additionally, Aubone
et  al. (2006) mentioned that from 1995 to 2006, biomass of C. guatucupa
stocks south to the 39° S was severely depleted.
In Fig. 11.5, we plotted landings, effort, and yield (CPUE) for both fleets
from 1992 to 2009. Landings in Bahía Blanca Estuary seem to increase
steadily since 2004, whereas landings in El Rincón area increased to higher
values between 1994 and 2002. Since 2001, landings dropped to lower values
in El Rincón area, while they started to increase in Bahía Blanca Estuary.
Effort applied in Bahía Blanca Estuary decreased from 1992 to 2000; from
2001 up to 2004, it remained low and constant, and since 2005 a slight recov-
ery was observed. Effort applied on El Rincón area increased in 1996–2001
period from the lower values of 1992–1995. From 2002, effort dropped con-
siderably, remaining low until 2003, although a slightly positive trend can be
observed until 2009. In Bahía Blanca Estuary, there was a steady drop in
yield, from 1992 to 1994, and this decrease continues until 1998. On 1999,
yield peaked at 265 kg/day and then dropped again until 2003. Starting on
2004, there was a net increase of the yield. Yields in El Rincón area increase
since 1993, and up to 2001, around the same average values. Yield between
2002 and 2009 presented an important increase.
Our evidence suggests that the landings of the artisanal fleet operating in
Bahía Blanca Estuary were affected by the increased fishing pressure exerted
by the industrial fishing fleet of El Rincón area. Effective management of this
fishery needs to be implemented to attain sustainability. While the fishing
closures in El Rincón area provided certain respite for the weakfish popula-
tions, they are not enough to rebuild the stock of this species. Cynoscion
guatucupa stocks need to be recovered before a sustainable fishery of this
species can be implemented (Lopez Cazorla et al. 2014).
11  Ecology and Biology of Fish Assemblages 301

Fig. 11.5  Reported landings (a), effort (b), CPUE (c), and model estimates for Bahía Blanca
Estuary (BBE) and El Rincón area (ERA) for Cynoscion guatucupa in 1992–2009. Data for Bahía
Blanca Estuary is presented with boxes and a gray line for the model estimates. Data for El Rincón
area is presented with triangles and a black line for the model estimates. (Modified from Lopez
Cazorla et al. 2014)
302 J. M. Molina et al.

References

Aguiar AA, Gallo V, Valentin JL (2004) Using the size independent discriminant analysis to distin-
guish the species of Myliobatis Cuvier (Batoidea: Myliobatidae) from Brazil. Zootaxa 464:1–7
Aguilera Socorro O, Fredou FL, Haimovici M et  al (2015) The IUCN Red List of Threatened
Species 2015: Micropogonias furnieri. https://doi.org/10.2305/IUCN.UK.2015-­4.RLTS.
T195076A49232972.en. Accessed 30 Aug 2019
Aubone A, Ruarte C, Di Marco E (2006) Un modelo matricial estructurado por estadios de tallas
para la pescadilla de red (Cynoscion guatucupa) al sur de los 39°S, en el periodo 1995-2005.
Inf Téc Ases y Transf DNI-INIDEP 20:16
Avigliano E, Volpedo AV (2015) New records of anadromous catfish Genidens barbus (Lacépéde,
1803) in the Paraná Delta (South America): evidence of extension in the migration corridor?
Mar Biodivers Rec 8:e23. https://doi.org/10.1017/S175526721400147X
Bamber RN, Henderson PA (1988) Pre-adaptative plasticity in atherinids and the estuarine seat of
teleost evolution. J Fish Biol 33:17–23
Beheregaray LB, Levy J (2000) Population genetics of the silverside Odontesthes argentinensis
(Teleostei, Atherinopsidae): evidence for speciation in an estuary of southern Brazil. Copeia
2000:441–447
Bemvenuti MA (2005) Osteologia comparada entre as espécies de peixes-rei Odontesthes Everman
& Kendal (Osteichthyes, Atherinpasidae) do sistema lagunar Patos-Mirim, no extremo sul do
Brasil. Rev Bras Zool 22:293–305
Blaber SJM (2002) Fish in hot water: the challenges facing fish and fisheries research in tropical
estuaries. J Fish Biol 61:1–20
Blaber SJM, Cyrus DP, Albaret JJ et al (2000) Effects of fishing on the structure and functioning
of estuarine and nearshore ecosystems. ICES J Mar Sci 57:590–602
Blasina GE, Barbini SA, Díaz de Astarloa JM (2010) Trophic ecology of the black drum, Pogonias
cromis (Sciaenidae) in Mar Chiquita coastal lagoon (Argentina). J Appl Ichthyol 26:528–534
Blasina GE, Molina JM, Lopez Cazorla A et  al (2016) Relationship between morphology and
trophic segregation in four closely related sympatric fish species (Teleostei, Sciaenidae). C R
Biol 339:498–506
Blyth-Skyrme RE, Kaiser MJ, Hiddink JG et al (2006) Conservation benefits of temperate marine
protected areas: variation among fish species. Conserv Biol 20:811–820
Campos D, Silva A, Sales N et al (2015) Trophic relationships among fish assemblages on a mud-
flat within a Brazilian Marine protected area. Braz J Oceanogr 63:429–442
Carbines G, Cole RG (2009) Using a remote drift underwater video (DUV) to examine dredge
impacts on demersal fishes and benthic habitat complexity in Foveaux Strait, Southern New
Zealand. Fish Res 96:230–237
Carozza C, Fernandez Araoz N (2009) Análisis de la actividad de la flota en el área de “El Rincón”
dirigida al variado costero durante el período 2000-2008 y situación de los principales recursos
pesqueros. Inf Téc INIDEP 23:p18
Carozza C, Lasta C, Ruarte C et al (2004) Corvina rubia (Micropogonias furnieri). In: Sánchez
RP, Bezzi SI, Boschi EE (eds) El Mar Argentino y sus recursos pesqueros. Tomo 4: Los peces
marinos de interés pesquero. Caracterización biológica y evaluación del estado de explotación.
INIDEP, Mar del Plata, pp 255–270
Cassia MC (1986) Reproducción y fecundidad de la pescadilla de red (Cynoscion striatus). Publ
Com Téc Mix Fr Mar 1:191–203
Castro JI (1993) The shark nursery of bulls bay, South Carolina, with a review of the shark nurser-
ies of the southeastern coast of the United States. Environ Biol Fish 38:37–48
Chao LH, Pereira LE, Vieira JP (1985) Estuarine fish community of the dos Patos Lagoon, Brazil.
A baseline study. In: Yañez-Arancibia A (ed) Fish community ecology in estuaries and coastal
lagoons: towards an ecosystem integration. UNAM Press, Mexico, pp 429–450
11  Ecology and Biology of Fish Assemblages 303

Chao L, Vieira JP, Brick Peres M et al (2015) The IUCN Red List of Threatened Species 2015:
Pogonias cromis. https://doi.org/10.2305/IUCN.UK.2015-­2.RLTS.T193269A49230598.en.
Accessed 30 Aug 2019
Chiaramonte GE, Pettovello AD (2000) The biology of Mustelus schmitti in southern Patagonia,
Argentina. J Fish Biol 57:930–942
Colautti D, Baigun C, Lopez Cazorla A et al (2010) Population biology and fishery characteristics
of Smoothhound Mustelus schmitti in Anegada Bay, Argentina. Fish Res 106:351–357
Cousseau MB, Perrotta RG (2013) Peces marinos de Argentina: biología, distribución, pesca. 4th.
ed. Instituto Nacional de Investigación y Desarrollo Pesquero INIDEP, Mar del Plata
Cousseau MB, Pequeño G, Mabragaña E, Lucifora LO, Martínez P, Giussi A (2020) The
Magellanic Province and its fish fauna (South America): Several provinces or one? J Biogeogr
47: 220– 234. https://doi.org/10.1111/jbi.13735
Dantas DV, Barletta M, Ramos JA et al (2013) Seasonal diet shifts and overlap between two sym-
patric catfishes in an estuarine nursery. Estuaries 36:237–256
Day JW, Hall CAS, Kemp WM, Yáñez-Arancibia A (1989) Estuarine Ecology. Wiley, New York
DeMartini EE, Sikkel PC (2006) Reproduction. In: Allen LG, Pondella DJ, Horn MH (eds) The
ecology of marine fishes—California and adjacent waters. University of California Press, San
Francisco, pp 483–452
Di Dario F, Williams JT, Palla H (2017) The IUCN Red List of Threatened Species 2017: Ramnogaster
arcuata. https://doi.org/10.2305/IUCN.UK.2017-­3.RLTS.T98822415A98886225.en.
Accessed 30 Aug 2019
Díaz de Astarloa JM, Munroe TA (1998) Systematics, distribution and ecology of commercially
important paralichthyid flounders occurring in Argentinean-Uruguayan waters (Paralichthys,
Paralichthyidae): an overview. J Sea Res 39:1–9
Dyer BS (2000) Revisión sistemática de los pejerreyes de Chile (Teleostei, Atheriniformes). Est
Oceanol 19:99–127
Elías R, Iribarne O, Bremec C et al (2004) Comunidades bentónicas de fondos blandos. In: Píccolo
MC, Hoffmeyer MS (eds) El ecosistema del estuario de Bahía Blanca. Instituto Argentino de
Oceanografía, Bahía Blanca, pp 179–190
Elisio M, Colonello JH, Cortés F et al (2017) Aggregations and reproductive events of the nar-
rownose smooth-hound shark (Mustelus schmitti) in relation to temperature and depth in
coastal waters of the South-Western Atlantic Ocean (38–42°S). Mar Freshw Res 68:732–742
Elliott M, Hemingway KL, Costello MJ et al (2002) Links between fish and other trophic levels.
In: Elliott M, Hemingway K (eds) Fishes in estuaries. Wiley, New York, pp p124–p216
Elliott M, Whitfield AK, Potter IC et al (2007) The guild approach to categorizing estuarine fish
assemblages: a global review. Fish Fish 8:241–268
Escobar BEA (1995) Dimorfismo sexual, crecimiento y fecundidad del lenguado común
(Paralichthys adspersus) de la costa central del Perú. Thesis to obtain the Fishing Engineering
degree, Facultad de Pesquería, Universidad Nacional Agraria La Molina Perú, p 63
Fabré NN, Cousseau MB (1990) Sobre la determinación de la edad y el crecimiento del lenguado
Paralichthys isosceles aplicando retrocálculo. Rev Brasil Biol 50:345–354
FAO (2005) Increasing the contribution of small-scale fisheries to poverty alleviation and food
security, FAO Technical Guidelines for Responsible Fisheries No 10. FAO, Rome, p 79
Franco A, Elliott M, Franzoi P, Torricelli P (2008) Life strategies of fishes in European estuaries:
the functional guild approach. Mar Ecol Prog Ser 354:219–228
Freije RH, Spetter CV, Marcovecchio JE et al (2008) Water chemistry and nutrients of the Bahía
Blanca Estuary. In: Neves R, Baretta J, Mateus M (eds) Perspectives on integrated coastal zone
management in South America. IST Press, Lisboa, pp 241–254
García ML (1987) Contribución al conocimiento sistemático y biológico de los Atherinidae del
Mar Argentino. PhD Thesis. Universidad Nacional de La Plata, La Plata
Garcia AM, Vieira JP (2001) O Aumento da diversidade de peixes no estuario da Lagoa dos Patos
durante o episodio El Niño 1997–1998. Atlantica 23:133–152
304 J. M. Molina et al.

García-Charton JA, Pérez-Ruzafa A (2001) Spatial pattern and the habitat structure of a
Mediterranean rocky reef fish local assemblage. Mar Biol 138:917–934
Gaston KJ (2000) Global patterns in biodiversity. Nature 405:220–227
Gerking SD (1994) Feeding ecology of fish. Academic Press, California
Giberto DA, Bremec CS, Acha EM (2007) Feeding of the whitemouth croaker Micropogonias
furnieri (Sciaenidae; Pisces) in the estuary of the Rio de la Plata and adjacent uruguayan
coastal waters. Atlantica 29:75–84
González Castro M, Díaz de Astarloa JM, Cousseau MB et al (2009) Fish composition in a South-­
Western Atlantic temperate coastal lagoon: spatial–temporal variation and relationships with
environmental variables. J Mar Biol Assoc U K 89:593–604
Haimovici M, Martins AS, Vieira PC (1996) Distribuicao e abundancia de peixes teleósteos demer-
sais sobre a plataformacontinental do sul do Brasil. Rev Bras Biol 56:27–50
Hamlett WC (1999) Sharks, skates, and rays: the biology of elasmobranch fishes. The Johns
Hopkins University Press, Maryland
Horn MH, Ferry-Graham LA (2006) Feeding mechanisms and trophic interactions. In: Allen LG,
Pondella DJ, Horn MH (eds) The ecology of marine fishes: California and adjacent waters.
University of California Press, Berkeley, pp 387–410
Jaureguizar A, Bava J, Carozza CR et  al (2003a) Distribution of the whitemouth croaker
(Micropogonias furnieri) in relation to environmental factors at the Río de la Plata estuary,
South America. Mar Ecol Prog Ser 255:271–282
Jaureguizar A, Menni R, Bremec AC et al (2003b) Fish assemblage and environmental patterns in
the Río de la Plata estuary. Estuar Coast Shelf Sci 56:921–933
Jaureguizar A, Solari A, Cortés F et al (2016) Fish diversity in the Río de la Plata and adjacent
waters: an overview of environmental influences on its spatial and temporal structure. J Fish
Biol 89:569–600
Llompart FM, Colautti DC, Maiztegui T et al (2013) Biological traits and growth patterns of pejer-
rey Odontesthes argentinensis. J Fish Biol 82:458–474
Lopez Cazorla A (1985) Edad, crecimiento y comportamiento migratorio de Brevoortia aurea
(Agassiz, 1829) (Osteichthyes, Clupeidae) de Bahía Blanca (Argentina). Investig Pesq
49:297–314
Lopez Cazorla AC (1987) Contribución al conocimiento de la ictofauna marina del área de Bahía
Blanca. PhD thesis, Universidad Nacional del Sur, Bahía Blanca
Lopez Cazorla A (1996) The food of Cynoscion striatus (Cuvier) (Pisces: Sciaenidae) in the Bahía
Blanca area, Argentina. Fish Res 28:371–379
Lopez Cazorla A (2000) Age structure of the population of weakfish Cynoscion guatucupa
(Cuvier) in the Bahıa Blanca waters, Argentina. Fish Res 46:279–286
Lopez Cazorla A (2004) Peces del estuario de Bahía Blanca. In: Píccolo M, Hoffmeyer M (eds) El
ecosistema del estuario de Bahía Blanca. Instituto Argentina de Oceanografía, Bahía Blanca,
pp 191–201
Lopez Cazorla A (2005) On the age and growth of flounder Paralichthys orbignyanus (Jenyns,
1842) in Bahía Blanca estuary, Argentina. Hydrobiologia 537:81–87
Lopez Cazorla AL, Sidorkewicj N (2009) Some biological parameters of Jenyns’ sprat Ramnogaster
arcuata (Pisces: Clupeidae) in South-Western Atlantic waters. Mar Biodivers Rec 2:127
Lopez Cazorla A, Pettigrosso R, Tejera L et al (2011) Diet and food selection by Ramnogaster
arcuata (Osteichthyes, Clupeidae). J Fish Biol 78:2052–2066
Lopez Cazorla A, Molina JM, Ruarte C (2014) The artisanal fishery of Cynoscion guatucupa in
Argentina: exploring the possible causes of the collapse in Bahía Blanca estuary. J Sea Res
88:29–35
Macchi GJ, Acha EM, Lasta CA (2002) Reproduction of black drum (Pogonias cromis) in the Rio
de la Plata estuary, Argentina. Fish Res 59:83–92
Macchi GJ, Acha EM, Militelli MI (2003) Seasonal egg production of whitemouth croaker
(Micropogonias furnieri) in the Río de la Plata estuary, Argentina–Uruguay. Fish Bull
101:332–342
11  Ecology and Biology of Fish Assemblages 305

Mallo P, Cervellini P (1988) Distribution and abundance of larvae and postlarvae of Artemesia
longinaris, Pleoticus muelleri and Peisos petrunkevitchi (Crustacea: Decapoda: Penaeida) in
the coastal waters of the Bahia Blanca Bay, Argentina. J Aquacult Trop 3:1–9
Malloy KD, Yamashita Y, Yamada H et al (1996) Spatial and temporal patterns of juvenile stone
flounder Kareius bicoloratus growth rates during and after settlement. Mar Ecol Prog Ser
131:49–59
Martinho F, Dolbeth M, Viegas I et al (2009) Environmental effects on the recruitment variability
of nursery species. Estuar Coast Shelf Sci 83:460–468
Massa A, Hozbor N, Chiaramonte GE et  al (2010) Mustelus schmitti. The IUCN Red List of
Threatened Species. https://doi.org/10.2305/IUCN.UK.2006.RLTS.T60203A12318268.en.
Accessed 22 Oct 2019
Masuda Y, Ozawa T, Onoue O, Hamada T (2000) Age and growth of the flathead, Platycephalus
indicus, from the coastal waters of West Kyushu, Japan. Fish Res 46:113–121
McCormick MI, Choat JH (1987) Estimating total abundance of a large temperate-reef fish using
visual strip-transects. Mar Biol 96:469–478
Mendoza-Carranza M, Vieira J (2008) Whitemouth croaker Micropogonias furnieri (Desmarest,
1823) feeding strategies across four southern Brazilian estuaries. Aquat Ecol 42:83–93
Menni RC (1985) Distribución y biología de Squalus acanthias, Mustelus schmitti y Galeorhinus
vitaminicus (Chondrichthyes) en el mar argentino en agosto-setiembre de 1978. Rev Mus La
Plata 13:151–182
Molina JM (2013) La comunidad íctica de Bahía Anegada: estructura, composición, dinámica
estacional y aspectos biológicos. PhD thesis, Universidad Nacional del Sur, Bahía Blanca
Molina JM, Cazorla AL (2015) Biology of Myliobatis goodei (Springer, 1939), a widely distrib-
uted eagle ray, caught in northern Patagonia. J Sea Res 95:106–114
Molina JM, Lopez Cazorla A (2011) Trophic ecology of Mustelus schmitti (Springer, 1939) in a
nursery area of northern Patagonia. J Sea Res 65:381–389
Molina JM, Blasina GE, Lopez Cazorla AC (2017) Age and growth of the highly exploited nar-
rownose smooth-hound (Mustelus schmitti ) (Pisces: Elasmobranchii). Fish Bull 115:365–379
Neer JA, Rose KA, Cortés E (2007) Simulating the effects of temperature on individual and popu-
lation growth of Rhinoptera bonasus: a coupled bioenergetics and matrix modeling approach.
Mar Ecol Prog Ser 329:211–223
Nicolas D, Lobry J, Le Pape O, Boet P (2010) Functional diversity in European estuaries: relat-
ing the composition of fish assemblages to the abiotic environment. Estuar Coast Shelf Sci
88:329–338
Parodi E (2004) Marismas y algas bentónicas. In: Píccolo M, Hoffmeyer M (eds) El ecosistema
del estuario de Bahía Blanca. Instituto Argentina de Oceanografía, Bahía Blanca, pp 101–107
Pasquaud S, Vasconcelos RP, França S et al (2015) Worldwide patterns of fish biodiversity in estu-
aries: effect of global vs. local factors. Estuar Coast Shelf Sci 154:122–128
Potter IC, Tweedley JR, Elliott M, Whitfield AK (2015) The ways in which fish use estuaries: a
refinement and expansion of the guild approach. Fish Fish 16:230–239
Refi SM (1975) Myliobatidae y Dasyatidae del litoral bonaerense de la República Argentina y
estudio comparado del mixopterigio (Chondrichthyes, Myliobatoidea). Physis 34:121–136
Ronda AC, Oliva AL, Arias AH et al (2019) Biomarker responses to polycyclic aromatic hydro-
carbons in the native fish Ramnogaster arcuata, South America. Int J Environ Res 13:77–89
Ruarte CO, Sáez MB (2008) Estudio preliminar sobre la estructura de edades y el crecimiento de
la pescadilla de red (Cynoscion guatucupa, Pisces, Sciaenidae) en el área sur de la Provincia de
Buenos Aires. Rev Invest Des Pesq 19:37–44
Ruarte C, Lasta C, Carozza C (2000) Pescadilla de Red (Cynoscion guatucupa). In: Bezzi SI,
Akselman R, Boschi EE (eds) Síntesis del estado de las pesquerías marítimas argentinas y de la
Cuenca del Plata, Años 1997–1998, con la actualización de, vol 1999. INIDEP, Mar del Plata,
pp 65–74
Rubio KS, Ajemian M, Stunz GW et al (2018) Dietary composition of black drum Pogonias cromis
in a hypersaline estuary reflects water quality and prey availability. J Fish Biol 93:250–262
306 J. M. Molina et al.

Ruocco NL, Lucifora LO, Díaz de Astarloa JM et  al (2012) Morphology and DNA barcoding
reveal a new species of eagle ray from the Southwestern Atlantic: Myliobatis ridens sp. nov.
(Chondrichthyes: Myliobatiformes: Myliobatidae). Zool Stu 51:862–873
Sardiña P, Lopez Cazorla AC (2005a) Feeding habits of the juvenile striped weakfish, Cynoscion
guatucupa Cuvier 1830, in Bahía Blanca estuary (Argentina): seasonal and ontogenetic
changes. Hydrobiologia 532:23–38
Sardiña P, Lopez Cazorla A (2005b) Trophic ecology of the whitemouth croaker, Micropogonias
furnieri (Pisces: Sciaenidae), in South-Western Atlantic waters. J Mar Biol Assoc U K
85:405–413
Segura AM, Milessi AC (2009) Biological and reproductive characteristics of the Patagonian
smoothhound Mustelus schmitti (Chondrichthyes, Triakidae) as documented from an artisanal
fishery in Uruguay. J Appl Ichthyol 25:78–82
Sidders MA, Tamini LL, Perez JE (2005) Reproductive biology of Mustelus schmitti springer,
1939 (Condrichtyes, Triakidae) in Puerto Quequen Buenos Aires Province. Rev Mus Argent
Cienc Nat 7:89–101
Simier M, Laurent C, Ecoutin JM et al (2006) The Gambia River estuary: a reference point for
estuarine fish assemblages studies in West Africa. Est Coast Shelf Sci 69:615–628
Simpfendorfer CA, Milward NE (1993) Utilisation of a tropical bay as a nursery area by sharks of
the families Carcharhinidae and Sphyrnidae. Environ Biol Fish 37:337–345
Sneath PH, Sokal RR (1973) Numerical taxonomy: the principles and practice of numerical clas-
sification, 1st edn. W. H. Freeman, San Francisco
Stehmann M (2009) Myliobatis goodei. The IUCN Red List of Threatened Species. https://doi.
org/10.2305/IUCN.UK.2009-­2.RLTS.T161436A5423507.en. Accessed 22 Oct 2019
Topping DT, Lowe CG, Caselle JE (2006) Site fidelity and seasonal movement patterns of adult
California sheephead Semicossyphus pulcher (Labridae): an acoustic monitoring study. Mar
Ecol Prog Ser 326:257–267
Tsuji FI, Barnes AT, Case JF (1972) Bioluminescence in the marine teleost, Porichthys notatus, and
its induction in a non-luminous form by Cypridina luciferin (Ostracoda). Nature 237:515–516
Vianna M, Tomas AR, Verani JR (2000) Aspects of the biology of the Atlantic Midshipman,
Porichthys porosissimus (Teleostei, Batrachoididae): an important by-catch species of shrimp
trawling off southern Brazil. Rev Brasil Oceanog 48:131–140
Villwock de Miranda L, Haimovici M (2007) Changes in the population structure, growth and
mortality of striped weakfish Cynoscion guatucupa (Sciaenidae, Teleostei) of southern Brazil
between 1976 and 2002. Hydrobiologia 589:69–78
Vinagre C, Santos FD, Cabral HN et al (2009) Impact of climate and hydrology on juvenile fish
recruitment towards estuarine nursery grounds in the context of climate change. Estuar Coast
Shelf Sci 85:479–486
Vooren CM (1997) Demersal elasmobranchs. In: Seeliger U, Odebrecht C, Castello JP (eds)
Environment and biota of the Patos Lagoon Estuary. Springer-Verlag, Berlin, pp 141–146
Whitfield AK (1999) Ichthyofaunal assemblages in estuaries: a South African case study. Rev Fish
Biol Fisher 9:151–186
Whitfield AK (2016) Biomass and productivity of fishes in estuaries: a South African case study.
J Fish Biol 89:1917–1930
Wonton RJ (1992) Fish ecology. Blackie and Son, London
Yañez-Arancibia A, Linares AFA, Day JW (1980) Fish community structure and function in
Terminos Lagoon, a tropical estuary in the southern Gulf of Mexico. In: Kennedy VS (ed)
Estuaries perspectives. Academic Press, New York, pp 465–482
Chapter 12
Bahía Blanca Estuary and the Importance
of Wetlands for the Conservation of Sea
Turtles

Victoria Massola, Laura Prosdocimi, Cristina Suldrup,


and Juan Facundo Sosa

Resumen
Las tortugas marinas, símbolo de longevidad, paciencia, fertilidad y libertad, son
reptiles que aparecieron en la tierra hace unos 200 millones de años. Hay siete espe-
cies agrupadas en dos familias Cheloniidae y Dermochelyidae. Estos animales están
presentes en todos los océanos y mares del mundo, excepto en las regiones polares.
Las tortugas marinas son especies migratorias, que se mueve entre sus áreas de
anidación y alimentación. Desde finales de primavera hasta principios de otoño es
frecuente la aparición de tres especies de tortugas marinas en el litoral de la provin-
cia de Buenos Aires, Argentina. Estas son la tortuga cabezona (Caretta caretta), la
tortuga laúd (Dermochelys coriacea) y la tortuga verde (Chelonia mydas). El

V. Massola ()
Fundación para la Asistencia de Animales Marinos y Educación Ambiental FRAAM,
Lokeren, Belgium
Reserva Natural Provincial Bahía Blanca – Organismo Provincial Para el Desarrollo
Sostenible OPDS, Provincia de Buenos Aires, Bahía Blanca, Argentina
Programa Regional para la Investigación y Conservación de Tortugas Marinas en,
Buenos Aires, Argentina
L. Prosdocimi
Programa Regional para la Investigación y Conservación de Tortugas Marinas en,
Buenos Aires, Argentina
Dirección de Planificación y Gestión de Pesquerías, Subsecretaría de Pesca y Acuicultura,
Ministerio de Agricultura, Ganadería y Pesca, Buenos Aires, Argentina
C. Suldrup · J. F. Sosa
Fundación para la Asistencia de Animales Marinos y Educación Ambiental FRAAM,
Lokeren, Belgium
Programa Regional para la Investigación y Conservación de Tortugas Marinas en,
Buenos Aires, Argentina

© Springer Nature Switzerland AG 2021 307


S. M. Fiori, P. D. Pratolongo (eds.), The Bahía Blanca Estuary,
https://doi.org/10.1007/978-3-030-66486-2_12
308 V. Massola et al.

estuario de Bahía Blanca es un extenso humedal marino costero y es una importante


área de alimentación, en el Atlántico sudoccidental, para la tortuga verde durante su
etapa de desarrollo juvenil, principalmente debido a su dieta omnívora. En esta área
costera es un importante centro urbano, industrial, portuario y de pesca artesanal. El
objetivo de este capítulo es actualizar la información sobre el valor de los hume-
dales del estuario de Bahía Blanca como una importante área de alimentación para
las tortugas marinas, principalmente la especie de tortuga verde. Además, se presen-
tarán algunos proyectos de investigación locales y los esfuerzos de conservación y
educación ambiental que han estado en progreso desde 2003 hasta la actualidad.
Palabras Claves
Tortugas marinas, áreas de alimentación, captura incidental, especies en peligro de
extinción

12.1  Introduction

Sea turtles belong to the reptile class; they are vertebrates that have been evolving
since approximately 120 million years ago, a successful group of animals that has
survived from periods of stable times to significant environmental change, adapting
to marine life. There are seven species of sea turtles that are distributed in all the
seas and oceans of the world, except in Antarctica. They swim in tropical, subtropi-
cal, and temperate waters. They are grouped into two families: Cheloniidae on one
hand has six species whose shell has obvious shields – hawksbill turtle (Eretmochelys
imbricata), green turtle (Chelonia mydas), loggerhead turtle (Caretta caretta),
Kemp’s ridley turtle (Lepidochelys kempii), olive ridley turtle (Lepidochelys olivá-
cea), and flatback turtle (Natator depressus). Dermochelyidae on the other hand has
a single species and shell without shields: leatherback turtle (Dermochelys coria-
cea). Throughout their life cycle, they occupy diverse ecosystems, moving from
hundreds to thousands of kilometers between the nesting, reproduction, feeding,
and developing areas, using beach, coastal zone, neritic, pelagic, and demersal envi-
ronments. Some species use large “blue corridors” swimming across national and
international coasts; therefore, they are considered highly migratory reptiles
(Bolten 2003).
They play a highly important ecological role. In the food plot, they serve as prey
and predators alike, contributing to the healthy maintenance of the structure and
dynamics of the marine ecosystem. In adulthood, the green turtle becomes an herbi-
vore, feeding on seaweed and seagrass. Through its feces, it contributes nitrogen to
the environment, thus facilitating the regrowth of pastures and maintaining the
replacement of nutrients. They are also considered indicators of the health of the sea
and coasts (Gonzalez Carman et al. 2011).
In some ancestral cultures and folklore of many Asian and Western countries,
they represent a symbol of longevity, calm, and patience – capacities and powers
12  Bahía Blanca Estuary and the Importance of Wetlands for the Conservation of Sea… 309

attributed to an animal considered mythical, sacred, and full of wisdom (Frazier


2003, 2005).
Despite being surrounded by so much spirituality, they have been intensely
exploited since ancient times, as early as the sixteenth century, until late in the twen-
tieth century for meat and egg consumption (Seminoff 2004; IUCN 2014), putting
many populations of different species at severe risk (Broderick et al. 2001, 2006;
Mc Clenachan et al. 2006).
On nesting beaches, where females spend a small part of their lives, they face
real estate projects, looting of eggs, exploitation of meat, skin and shell trade by
humans, attacks by domestic pets and other introduced animals, and presence of
solid urban waste near nesting areas, as well as global warming that affects both
sands where eggs mature and seawater. However, in recent years protection mea-
sures have been adopted on nesting beaches and other mitigation measures in the
water, for example, to reduce interaction with fisheries, like the mandatory use of
turtle-excluding devices (TED); the instruments are still insufficient for the protec-
tion and to ensure the viability of their populations.
Of the seven species of sea turtles, five can be found on the IUCN Red List of
Threatened Species as either “endangered” or “critically endangered.” The olive
ridley turtle, hawksbill turtle, and leatherback turtle are listed as “critically endan-
gered,” the loggerhead and green turtle as “endangered,” Kemp’s ridley turtle as
“vulnerable,” and the flatback turtle as “insufficient data,” which means that its con-
servation status is not clear due to the lack of data gathered (IUCN 2014;
Bolten 2003).
Throughout their complex life cycle, postmodernity has exposed them to differ-
ent threats including longlines, bycatch in coastal artisanal fisheries such as pelagic
commercial fishing, boat accidents, oil spills, and discharge of other chemical sub-
stances, as well as solid urban waste pollution mainly plastic. In the last 100 years,
threats of anthropic origin sea turtles must face have increased exponentially, com-
promising the viability of their population. For these reasons, the measures and
actions taken for the protection and conservation of these wonderful marine reptiles
must be applied cross-border based on the wide range of action and movement cor-
ridors used by sea turtles.
In the face of the unfavorable panorama of the Americas in the 1990s, a coopera-
tion strategy was adopted between governmental and nongovernmental organiza-
tions and research groups aiming to coordinate conservation actions within the
framework of a regional strategy. In 1996 in Caracas, Venezuela, the Inter-American
Convention for the Protection and Conservation of Sea Turtles (CIT) was imple-
mented – an intergovernmental treaty that promotes the protection, conservation,
and recovery of sea turtle populations and the habitats they depend on. This was
based on the most reliable data available and considered the environmental, socio-
economic, and cultural characteristics of the States Parties (ANNEX I – ACTA CFP
No. 31/2018). The ILC was launched in 2001 as a space for dialogue and action in
310 V. Massola et al.

the Americas to achieve concerted management of the species (Marco Solano,


Secretary Pro Tempore Inter-American Convention for the Protection and
Conservation of Sea Turtles 2004).

12.2  Sea Turtles in Argentina

Until 2003 it was thought that the presence of sea turtles in Argentina was peculiar
and accidental. There were few reports of beach-stranded animals with sanitary
complications, accidentally bycaught by fishery boats, and in case of being dead,
they are turned into local museum collections (Freiberg 1938; 1942, 1945; Gallardo
1977; Frazier 1984). This lack of information, plus the temperate coastal waters,
leads to the belief that the presence of these marine reptiles was rare and that the few
animals found here were disoriented or had been wrongly dragged by ocean cur-
rents (González Carman et al. 2011).
Since the conformation of the Programa Regional de Investigación y Conservación
de Tortugas Marinas de Argentina (PRICTMA) (Regional Program of Investigation
and Conservation of Argentine Sea Turtles) in 2003, protocols were established for
systematic procedures to collect information about the presence of sea turtles in our
country.
Different kinds of organizations began to work, led by this Program, by setting
consensual objectives and strategies in order to optimize and enhance their technical
and logistic capacities for the investigation and conservation of sea turtles.
At present, the PRICTMA is composed of nine organizations, distributed along
the Argentine coast: the Ecoparque Aquarium in Buenos Aires; Peyú Project;
Aquamarina; Sea World Foundation (at San Bernardo del Tuyú); ECOFAM Project;
Mar del Plata Aquarium Foundation; Cooperative Association of the Multiple Use
Natural Reserve Bahía Blanca, Bahía Falsa, Bahía Verde; the Marine and Fishery
Biology Institute Alm. Storni; and Natural Patagonia Foundation (Fig. 12.1).
Thanks to the work strategies, of which the key point was the integration and
cooperation between the different institutions in a logistic and academic level, some
of the ruled objectives were reached in short time, in order to answer some basic
questions about biology and about the different sea turtles visiting our waters. As a
result of the working strategies, there also emerged two main tools: Marking
Program and Satellite Monitoring Program for Sea Turtles.
In 2018, the National Action Plan for the Conservation of Sea Turtles (PAN
Tortugas) was established, which has two programs: the National Action Program to
reduce the interaction of sea turtles with fisheries in the Argentine Republic
(Resolution of the Federal Fisheries Council No. 14/2018) and the National Action
Program to reduce the interaction of sea turtles with marine debris in the Argentine
Republic (Resolution of the Federal Environment Council No. 317/2015).
Within the framework of this plan, the Multiple Use Reserve and the FRAAM
foundation are aligned on the objectives and conservation strategies for these marine
reptiles, as well as for species of mammals, birds, and chondrichthyans.
12  Bahía Blanca Estuary and the Importance of Wetlands for the Conservation of Sea… 311

Fig. 12.1  Institutions that are part of the Programa Regional de Investigación y Conservación de
Tortugas Marinas de Argentina (PRICTMA). The map shows the distribution of the institutions:
(1) Ecoparque; (2) Peyú Project; (3) Sea World Foundation; (4) Aquamarina; (5) Mar del Plata
Aquarium Foundation; (6) Ecofam; (7) Multiple Use Reserve Bahía Blanca, Bahía Falsa, Bahía
Verde (OPDS); (8) Almirante Storni Institute; (9) Natural Patagonia Foundation

12.3  Biology of the Species Found in Bahía Blanca

From the monitoring of the fishing localities and beach sectors, it was determined
that, effectively, presence of sea turtles was not random; on the contrary, a stage in
the life of three sea turtle species out of the seven species from the world happens in
our sea. These species are the green turtle (Chelonia mydas), loggerhead turtle
(Caretta caretta), and the leatherback turtle (Dermochelys coriacea). Recently, two
specimens of hawksbill turtle (Eretmochelys imbricata) were found: the first one
was by an accidental fishermen bycatch at San Bernardo’s beach, and the second
was found dead at Bahía Blanca’s estuary, being the first records of this species in
Argentina. Genetics studies made from them showed that they came from Brazilian
nesting zones (Prosdocimi et al. 2014a, b).
Green turtle (Chelonia mydas): the green turtle is the biggest in the family of
Cheloniidae (Pritchardand and Mortimer 1999). Generally, it shows an oval shell
dorsoventrally flattened with five vertebral shields, four pairs of costal shields, and
four pairs of inframarginal shields. It is characterized for having a couple frontal
scales and four pairs of post-orbital scales. The upper jaw has a slightly denticulate
border, while the inferior jaw has sharper denticulate border. Each fin shows one
nail, though it may have two (Fig. 12.2) (Pritchardand and Mortimer 1999). Like the
312 V. Massola et al.

Fig. 12.2  Morphological characteristics of the three species of sea turtles present in the Argentine
Sea. Modified and reprinted with permission from Wyneken, J. 2001. The Anatomy of Sea Turtles.
NOAA Tech. Memo. NMFS-SEFSC-470

other species of the family, the hatchlings’ shell color is mostly black or dark gray
with a whitish plastron. As they grow, the shell changes to dark coffee or olive
green. Colors in adults can vary from stained to striped brown, gray, black, or green
tones (Fig. 12.2). The common name of the species came from the green color of the
corporal fat (Pritchardand and Mortimer 1999) (Fig. 12.2).
Loggerhead turtle (Caretta caretta) has five pairs of costal shields and five cen-
tral and two supracaudals. Another distinctive characteristic is that they show two
pairs of prefrontal scales, three pairs of post-orbital scales, and a characteristic red-­
coffee dorsoventral coloration in subadults and adults. Also present are three or in
rare occasions four inframarginal scales with no pores. The nape scale is in contact
with the first two costal scales, different from the green and hawksbill turtles (Dodd
1988). The presence of asymmetric anomalies in the scale disposition is very fre-
quent and can lead to mistakes in the specific determination (Rivilla et al. 2005).
The head is proportionally bigger compared to the other species, hence the charac-
teristic name. The frontal fins are relatively shorter in comparison with the other
species of the family, showing two nails in each fin (Pritchard and Mortimer 1999)
(Fig. 12.2).
12  Bahía Blanca Estuary and the Importance of Wetlands for the Conservation of Sea… 313

Leatherback turtle (Dermochelys coriacea): Unlike the others species of sea


turtles, this species has no keratinized shell. It was thought that the shell is covered
by a thick skin, recovering a mosaic of dermic ossicles connected by cartilage
(Wyneken 2001). The shell has a markedly tapered form, with seven prominent
ridges that enhance swimming (Wyneken 2003). The head is triangular with two
conspicuous maxillary cusps covered by skin (Pritchard and Mortimer, 1999). A
distinctive sign of the specie is a rose stain in the back of the head that is used for
photo identification (McDonald and Dutton 1996) (Fig. 12.2).
Each one of these species registered in Argentina is found in different stages of
their life. The green turtles, for example, are in their early youth, recently raised
from the oceanic environment to the neritic one. Their weight range is from 3 to
10 kg, with a shell length between 30 and 50 cm. Nevertheless the loggerhead tur-
tles have a wider size range, between 45 and 90 cm, and a weight range between 15
and 40 kg, both young and adults. The leatherback turtles are subadults and adults
with length up to 180  cm and more than 200  kg weight (González Carman
et al. 2011).
The hawksbill turtles found correspond to young individuals (Prosdocimi et al.
2014a, b). The presence of sea turtles in our coast is seasonal, mainly during sum-
mer and fall. Most of the records are concentrated in Bahía Samborombón, Cabo
San Antonio, and Bahía Blanca Estuary, though there are notices of fluvial raids in
the La Plata River, the Paraná River, and the Uruguay River and also their presence
in the colder waters of the north of Patagonia, as in San Matías, San José, and Golfo
Nuevo (González Carman et al. 2011, 2012, 2016a, b).
The majority of the registers of sea turtles were made in the estuarial areas of
Buenos Aires coast. Those areas are characterized for being highly productive and
offer a great biomass formed by jellyfish, mollusks, and other organisms which are
part of the sea turtles’ food. This, added to the information obtained by the analysis
of the stomach contents, allows to affirm that Argentine coastal waters are a feeding
and development zones for these sea species. Nevertheless, thanks to the informa-
tion obtained by the Marking Plan and the Satellite Following Program, we know
that our waters are integrated to a great feeding zone, that is, the South Occidental
Atlantic, also involving coastal and oceanic waters from Uruguay and Brazil. By the
end of autumn, the sea turtles that were feeding all along Argentine coasts move to
warmer waters in the south of Brazil, where they remain feeding. In some cases,
they return to the same place they fed in our coast (López Mendilaharsu et al. 2009;
Fossette et al. 2010; Gonzalez Carman et al. 2012, 2016a, b).
The origin of the sea turtles that arrived in Argentine waters and specifically in
Bahía Blanca’s zone has been recently unveiled through poblational genetic studies.
Most of the green turtles come from nesting beaches in Ascension Island, a small
volcanic island placed in front of the African coast, and, in a lesser percentage, from
nesting beaches of Suriname and Trindade Island (Brazil) (Prosdocimi et al. 2012).
In the case of leatherback turtles, the recovering of metal marks and genetics stud-
ies, let us identify its origin as nesting beaches in Occidental Africa (Gabon) (Billes
et al. 2006; Prosdocimi et al. 2014b). Lastly, for the case of the loggerhead turtle, all
the individuals came from Brazilian populations (Prosdocimi et al. 2015).
314 V. Massola et al.

12.4  Identified Conservation Problems

The highly productive areas chosen by the sea turtles to feed and grow are also the
ones chosen by independent and industrial fishing fleets from Argentina and
Uruguay. The accidental bycatches of the three species (green turtle, loggerhead
turtle, and leatherback turtle) have been registered mainly in gillnets and trawls,
while in the shrimp nets only green turtles have been caught (Gonzalez Carman
et al. 2011).
In a pilot study made at the port from San Clemente del Tuyú, Buenos Aires
coast, it was estimated that an average of 100 green turtles are caught in gillnets,
most of them were found dead in gillnets from suffocation (Albareda et al. 2007a).
Although the fishing arts that interact with sea turtles in Argentina have been identi-
fied, currently there is no other information regarding the capture rates and the
impact that this threat has on their populations.
On the other hand, in the stomach contents of green turtles analyzed in the last
5 years from the accidental capture in the Bahía de Samborombón, there are remains
of anthropogenic residues, mostly plastic bags (Albareda et al. 2007b). These have
also been found in live sea turtles that were in rehabilitation, who had defecated the
residues for several days. Plastic bags and fecal matter are part of a kind of solid
“framework” in the intestine, which can end up causing intestinal obstruction that
could damage the mucosa and alter its normal functionality. A gas-filled intestine,
unable to evacuate regularly, acts as a “life jacket,” not allowing the sea turtles to
dive, move, escape from predators, or feed. This causes a slow deterioration of their
physical condition, finally leading to a long and agonizing death. In our latitude this
chronical deterioration could affect their normal migration to warmer waters, failing
to escape on time from the low winter temperatures of Buenos Aires coast and
finally being dragged to the beach in a lethargic status, victims of a hypothermic
deadly picture (Albareda et al. 2007b).
In addition to sea turtle populations, the accidental bycaught, and the ingestion
of anthropic residues, we must add conservation problems shared with Uruguay and
Brazil (Bugoni et al. 2001; Laporta et al. 2006) and the need to coordinate region-
ally the actions for conservation taken by each country.

12.5  Conservation Actions

Scientific research in the last 6 years has allowed us to understand the basic aspects
of the biology of sea turtles in Argentina and, in that way, the main conservation
problems, through the development of interdisciplinary and multidisciplinary activ-
ities, such as poblational genetics studies, satellite monitoring, marking plan, health
assessment, feeding studies, and accidental bycatch assessment. The insertion of
young investigators inside the national academic circuit (such as the Scientific and
Technical Investigation National Council, Buenos Aires University, Marine and
12  Bahía Blanca Estuary and the Importance of Wetlands for the Conservation of Sea… 315

Fishery Biology Institute Almirante Storni and the Fishery Investigation and
Developing National Institute) to develop some investigation lines of PRICTMA
guaranteed the ability to find and to consolidate management activities. On the other
hand, thanks to the increasingly active participation of independent fishermen and
joint effort with other research projects, such as the Franciscana Dolphin
Conservation Program led by Aquamarina, different mitigation measures are cur-
rently being tested to reduce bycatch of these dolphins and sea turtles with gillnets.
The FRAAM foundation is a nonprofit entity (registration ID: PJ 40843) which
works with veterinary professionals, biologists, and environmental educators. The
main function of the foundation is the primary care mainly of sea turtles. In addition
to that, the training of undergraduate and graduate students (volunteers), environ-
mental education directed to general public, and students of all levels and related
careers, as well as research support, are all goals of the foundation.

12.5.1  S
 ea Turtles in the Southwestern Atlantic: Red ASO
and PRICTMA – Start of Research

A very valuable instrument that emerged in 2003 is the creation of the ASO,
Southwestern Atlantic network. The ASO interacts with almost all fisheries in the
Exclusive Economic Zones of Brazil, Uruguay, and part of Argentina and the adja-
cent international waters of the Atlantic Ocean between 5° N to 45° S and 20° to 65°
W. This network has allowed the exchange of experiences between researchers and
institutions in the region, as well as support to develop combined initiatives, thus
strengthening actions for the conservation of sea turtles (Domingo et  al. 2006).
Regarding the conservation of biodiversity and in particular sea turtles, Argentina,
Brazil, and Uruguay signed, approved, and ratified, through respective national
laws, some of the following International Conventions:
• Convention on Wetlands of International Importance (RAMSAR
CONVENTION 1971).
• Convention on International Trade in Endangered Species of Wild Fauna and
Flora (CITES), Washington, DC, 1973.
• Convention on the Conservation of Migratory Species of Wild Animals (CMS),
Bonn, 1979.
• United Nations Convention on the Law of the Sea (CONVEMAR), (New
York, 1982).
• Convention on Biological Diversity (Rio de Janeiro, 1992).
• Inter-American Convention for the Protection and Conservation of Sea Turtles
(IAC) (San José, 2001).
In 2010, Argentina approved the National Law 26.600, through which it adheres
to the Inter-American Convention for the Protection and Conservation of Sea Turtles
(CIT). In June 2011, the national government deposited its instrument of
316 V. Massola et al.

ratification. This meant among other measures to assume the commitment of protec-
tion and conservation methods for these species in our latitude, through the promo-
tion of scientific research, environmental education, and the dissemination of
information, also promoting participatory work and training of independent fisher-
men to minimize incidental catches, retention, damage, and killing of sea turtles
during fishing activities. The National Ministry of Environment and Sustainable
Development (SGA y DS) is the CIT enforcement authority. Going further in the
conservation strategies, in 2014 the work began on the preparation of the National
Action Plan for the Conservation of Sea Turtles in the Argentine Republic
(PAN-TM). In November of the same year, a workshop was carried out with the
participation of national and provincial government agencies, members of the sci-
entific sector, and nongovernmental organizations. Since then they work on the
issue together. Representatives of the National State, the provinces of Buenos Aires
and Río Negro, and the Autonomous City of Buenos Aires participated in the prepa-
ration of the PAN-TM, since the sea turtles reach the marine coast of the Province
of Río Negro from Río de la Plata. Research efforts were mainly invested in two
National Action Programs: Reduce the Interaction of Sea Turtles with Marine
Wastes (Approved by Resolution COFEMA 317/2015) and Reduce the Interaction
of Sea Turtles with Fisheries (Approved by Resolution 14/2018 Federal Fisheries
Council).

12.5.2  O
 ur Work in the South of Buenos Aires Province:
Wetlands of Bahía Blanca Estuary

In the estuary of Bahía Blanca, the multiple use reserve Bahía Blanca, Bahía Falsa,
Bahía Verde (RNBB) is established, and adjacent to the protected area in the coastal
town of Villa del Mar, the Foundation for Reception and Marine Animal Attendance
and Environmental Education (FRAAM), both institutions are part of
PRICTMA. Since 2003, action has been taken by different groups of people: out-
reach campaigns directed to the independent fishermen communities in the area,
formal environmental education activities, articulating with educational establish-
ments of all levels of education, and nonformal education directed to the gen-
eral public.
Using different communication strategies and incorporating other knowledge,
adding the valuable empirical knowledge of fishermen and the general community,
the conservation work has focused on raising public awareness of the seasonal pres-
ence of sea turtles in the waters of the estuary of Bahía Blanca, recognizing and
identifying problems of anthropic origin that different species face in our latitude,
and from a holistic approach seeking solutions that minimize conflict with local
actors. The varied range of activities has generated a gradual yet positive response
from the different actors, allowing favorable results in the first scientific research
tasks carried out in this coastal sector of Buenos Aires Province.
12  Bahía Blanca Estuary and the Importance of Wetlands for the Conservation of Sea… 317

Actions:
• Initial survey of the problem
• Environmental education workshops for students and teachers
• Publicity talks
• Training for veterinarians and voluntary university students
• Interviews with fishermen
• General information talks to fishermen
• Training for fishermen
On our coasts, the direct observation of sea turtles is difficult mainly due to two
factors: the turbidity of estuarial waters and also the animals’ own behavior, which
emerges briefly to breathe. To avoid this problem, an important project was launched
through the Satellite Monitoring Program. The placement of satellite transmitters to
several animals in four coastal locations, Magdalena, San Clemente del Tuyú, and
Villa del Mar – estuary of Bahía Blanca – in Buenos Aires Province and San Antonio
Oeste in Río Negro Province, has provided valuable information. The satellite mon-
itoring of green turtles, loggerhead, and leatherback turtles allowed us to determine,
for example, that Río de la Plata as the southern area of ​​Buenos Aires Province and
in the North of Rio Negro Province are areas of feeding, intensely used by these
species. It was also known that sea turtles frequent the fishing areas of the coastal
fleets of Buenos Aires Province and Uruguay. The common areas that juvenile sea
turtles use and where fishing activities are carried out are also contaminated by solid
garbage that comes from cities like Buenos Aires and Montevideo.
Throughout these years, constant monitoring efforts made it possible to deter-
mine that part of the Argentine continental shelf is the southernmost feeding and
development zone of the Southwestern Atlantic for at least three species: green
turtles, loggerhead turtles, and leatherback turtles (González Carman et al. 2011).
Most of the bycatches have been of green turtles, small-sized individuals with an
early youth development stage. The two hawksbill turtle individuals found corre-
spond to juvenile animals (Prosdocimi et al. 2014a).
Regarding the distribution in Argentina, it is conditioned by the temperature of
the sea.
Therefore, the waters overlying the northern sector of the Argentine continental
shelf would be their main habitat. There the temperature of the sea ranges between
18° and 23 ° C in summer and averages 8 ° C in winter, unlike the waters overlying
the southern sector of the platform that are colder. The presence of sea turtles, in
turn, is registered from the end of spring until the beginning of autumn, showing a
marked seasonal presence probably governed by the low water temperature of
8–10 °C in winter (González Carman et al. 2011, 2012).
A practical tool to gain quick insight into the behavior of marine turtles is satel-
lite tracking. With these the migratory routes and foraging habitats can be estimated
with accuracy (Godley et al. 2008).
From February 2008 to March 2009, nine green turtles (recovered from entan-
glement on the coast of Argentina) were instrumented with satellite tracking devices
318 V. Massola et al.

Fig. 12.3  Juvenile green turtle equipped with satellite tracker being released in Bahía Blanca
Estuary, carried out by the Bahía Blanca Nature Reserve staff, park rangers, and PRICTMA staff
(OPDS 2009)

(Fig.  12.3), six of them in Bahía Blanca Estuary (Villa del Mar and El Rincón)
(González Carman et al. 2012).
During summer and fall, tracked juveniles were located in the coastal waters of
Argentina and Uruguay. In winter sea turtles migrated to warmer waters in southern
Brazil. Migration started during the fall, as evidenced by an increase in mean move-
ment rate and the northern position of the 20 ° C isotherm. In spring, most sea turtles
remained in the waters of Brazil and Río de la Plata (Gonzalez Carman et al. 2012).
The use of this technique allowed to know the most intense foraging season in
Bahía Blanca Estuary (El Rincón, Villa del Mar), in the coast of Buenos Aires, and
in the La Plata River, an area of 11,282  km, during summer months and early
autumn. While in winter and spring period, for the same purpose, sea turtles used
the coasts of Uruguay and Brazil (Gonzalez Carman et al. 2012).

12.5.2.1  S
 ummary of Actions for the Protection and Conservation
of Marine Turtles in the Wetlands of Bahía Blanca Estuary

Bahía Blanca – OPDS Natural Reserve, Environmental Education Program and the
Foundation for Reception and Marine Animal Attendance FRAAM, Villa del Mar.
Our sustained work without interruption since 2003 to date has focused its efforts
on training and divulgation. It has been designed and directed to different actors:
independent fishermen’s community (artisanal); educational community at all lev-
els of education; connecting with academic institutions such as the South National
University and government organizations such as Coronel Rosales Municipality and
12  Bahía Blanca Estuary and the Importance of Wetlands for the Conservation of Sea… 319

Fig. 12.4  Training in primary assistance with sea turtles for Prefectura Naval Argentina (Coast
Guard) staff and lifeguards of Villa del Mar and Arroyo Pareja (Coronel Rosales), carried out by
staff of the FRAAM Foundation

its Honorable Deliberative Council, Bahía Blanca Municipality (Fig. 12.4); perma-


nent support of the mass media for all the activities carried out:
• 2003–2015: We started training aimed at artisanal fishermen from the towns of
Ing. White, Villa del Mar, and Coronel Rosales. We continue to interact today.
• 2004–2019: Uninterruptedly, from the Formal Environmental Education
Program, we began to work on the topic Marine Turtles from the Initial Level to
the Tertiary Level.
• From 2012 to date: World Sea Turtle Day Remembrance – traveling photo shows,
plastic workshops, puppet theater, oral exhibitions by different specialists.
• 2012, 2013, and 2014: Beach campaigns in Monte Hermoso, Pehuén Co, and
Villa del Mar during the summer season to inform tourists about the biology,
ecology, and behavior of sea turtles on our coasts and particularly about the nega-
tive impacts they face as a consequence of inorganic waste left on the beach and
in the sea.
• 2012, 2013, and 2014: Puppet Theater “S.O.S Who Can Help Us” – the story
tells the problems faced by a green turtle and a mermaid in the sea and the beach
because of anthropogenic waste. The work ends with songs and solutions shared
in dialogue with the public.
• 2013: Inauguration of “Your Garbage Affects Us” Poster. Awareness and infor-
mation about the environmental problems faced by the sea turtles; the poster was
installed in Villa del Mar in a very busy sector, on a public access beach, during
the development of the 5th edition of the Wetlands Festival. Proposed at the
Education for Action Workshop, ASO Mar del Plata 2009.
320 V. Massola et al.

• 2013: Approval of the first Municipal Ordinance No. 3408 “Emblematic Species
of Coronel Rosales District.” In the list of animals designated, the incorporation
of the three species of sea turtles frequent on the southwest coasts of Buenos
Aires was contemplated with special interest.
• 2013–2015: Video presentation – “Los Amigos de Las Tortugas” tells the story
of the life of an independent fisherman and his connection with sea turtles. It was
exhibited in the Cultural Historic Center of the South National University of
Bahía Blanca, in Union y Fuerza hall in Punta Alta City, during the 5th, 6th, and
7th edition of the Wetlands Festival in Villa del Mar.
• 2013–2014: Second and third presentation at the Sustainable House of Culture
Hall (UNS) with Sea Turtles, made of recycled material, the approval for the
installation of the Clean Beaches Campaign Poster, located in the seaside of Villa
del Mar Town, result of the 2009 ASO Workshop, being the second installed in
the southwest of Buenos Aires Province.
• 2014: Participation to the first preparatory workshop of the National Sea Turtle
Conservation Plan  – the FRAAM Foundation was present during the 2-day
workshop (26th and 27th of November of 2014), which took place at the head-
quarters of the National Ministry of Environment and Sustainable Development,
the first preparatory workshop of the National Action Plan for the Conservation
of Sea Turtles (PAN-TM) of the Argentine Republic. In the context of this work-
shop, we participate in the preparation of the National Action Program to reduce
the interaction of sea turtles with marine debris in Argentina.
• 2014: Training of veterinarians and volunteer students in different endoparasite
sampling and analyzing techniques.
• 2015: Signed Act Agreement for the non-use of plastic bags  – in the general
framework of the commitment assumed for the defense and protection of the
environment, on February 2015 an agreement was signed between Bahía Blanca
Municipality, Bahía Ambiental SAPEM, FRAAM Foundation, South
Conservation Association TELLUS, and large supermarket companies
(Cooperativa Obrera, Walmart Argentina, Carrefour, Burgos, and Super Vea).
This agreement has the purpose of eliminating plastic bags from supermarkets.
• 2015: Necropsy and Research Techniques Workshop in Sea Turtles. National
Action Program to reduce the interaction of sea turtles with marine debris in the
Argentine Republic.
• 2015: Educational and Outreach Strategies Workshop for the reduction of marine
litter in the area of the Río de la Plata and Buenos Aires Coast. National Action
Program to reduce the interaction of sea turtles with marine debris in the
Argentine Republic.
• 2017 Marine Garbage Workshop: Tools for a Better Impact on Public Policies
and Cultural Change. National Action Program to reduce the interaction of sea
turtles with marine debris in the Argentine Republic.
• 2019: First follow-up workshop on the National Action Plan for the Conservation
of Sea Turtles (PAN Tortugas): ST Program – Fisheries and ST Program – Marine
Residues.
12  Bahía Blanca Estuary and the Importance of Wetlands for the Conservation of Sea… 321

12.5.2.2  Veterinary Actions

The veterinary staff has a protocol to follow in which different types of evaluations
are carried out before the animal enters the Foundation if necessary. As previously
mentioned, over the years, several training sessions have been carried out for differ-
ent groups of people: university students, volunteers, and independent fishermen,
among others. These fishermen, being the first ones to get in contact with sea turtles,
will follow the instructions received, in addition to those that the corresponding
personnel give them for the evaluation and management of the animal at the place.
In some cases the sea turtle could be half drowned, they will make the first han-
dlings by putting the animal in an oblique angle position, head down, in a soft and
protected place, in order to prevent more injuries.
This coastal area is a migratory route for animals such as penguins, sea turtles,
whales, fishes, and birds at different stages of their life cycle, from breeding areas
to feeding places, and due to this it is possible to find a sea turtle (or other animals)
stranded on the beach. In these cases, the action protocol is the same: the state of the
animal must be evaluated as a first measure, and then the corresponding personnel
should be contacted as soon as possible  – FRAAM Foundation, Rangers, and
Prefectura Naval Argentina (Coast Guard). Part of the protocol of action in these
cases is to keep people at a good distance from the animal, avoid interacting with it
in any way (feed, touch, etc.), to reduce stress. If the veterinary team is not in the
area, the mentioned staff present at the scene will make an assessment of the ani-
mal’s condition, if possible while on the telephone to the veterinarians, who will
give instructions to verify if the animal presents some type of obvious injury or
unusual behavior. If any type of anomaly is found at the anatomical or ethological
level, it might be decided to transfer the animal to the Foundation where it will be
given assistance with the necessary additional elements and corresponding equip-
ment for the species in question.
The animal species that most frequently enter the Foundation are the green turtle
(Chelonia mydas), the Magellanic penguin (Spheniscus magellanicus), and the two-­
haired sea lion (Arctocephalus australis). Other less frequent species are the log-
gerhead turtle (Caretta caretta) with two recorded assists and the one-haired sea
lion (Otaria flavescens). There are other species, much less frequent, such as the
Antartic fur seal (Arctocephalus gazella) and the Sub Antartic fur sea (Arctocephalus
tropicalis).
Among the diagnoses made to the admitted animals, different factors are taken
into account: if the injury requires surgery, if the animal must be placed in a dry area
in case it has suffered drowning, or if it is very active, it is placed directly in a pool
to observe its movements, reactions, and responses to the different stimuli adminis-
tered. If the reaction is normal, we stimulate food intake, corresponding to the stage
of the animal in question.
In this area, the sea turtles caught incidentally are at the juvenile stage of their
cycle in which they have a non-selective diet, so they were fed with algae and
arthropods.
322 V. Massola et al.

When the sea turtle’s responses are positive according to its behavior, it is
released in the area where it was incidentally captured.
Before the release of these animals, the corresponding biometric data were taken:
the length (straight and curved) and width (straight and curved) of the carapace and
plastron and the length and width of the head, tail, and the four fins are measured
(Workbook ASO 2004); the weight is measured upon entering the Foundation and
upon release (Fig. 12.5). Blood samples are taken from the sinus occipital venous
with a 25/8 needle and a 5 cc syringe. In addition to the above measurements, a
micro-blood count can be made to see the proportion of blood cells by counting the
percentage of proteins and blood staining to see abnormalities, as well as a small
skin extract for gender determination.
All this biometric data is sent for scientific studies, for example, obtaining popu-
lation data (to determine the origin). In addition to the above, the type of food
administered, the one chosen by the animal itself, the quantity, the frequency, and
whether this food was alive or dead are also measured and recorded. Once the turtle
is ready to be released, it is marked with two metal rings, one in each fin (depending
on the species), which consists a non-corrosive alloy printed with an alphanumeric
identification, which will serve in the event that there is a recapture of the animal
throughout the migratory route, to know where and when it was marked, and so
compare the past biometric data with the current one. All this is developed follow-
ing a national and international protocol. Some animals can be released in groups,
as is the case with penguins. But sea turtles, which are independent and solitary
animals, are released in the place where they were found or captured. Different spe-
cies of birds have also been received, which after examination, if they cannot be
released after recovery, will be sent to specialized centers or will remain in the

Fig. 12.5  Veterinary staff and FRAAM volunteers carrying out the biometrics on a juvenile green
turtle. (Photo by V Massola)
12  Bahía Blanca Estuary and the Importance of Wetlands for the Conservation of Sea… 323

Foundation facilities, to teach visitors about the problems of trying to domesticate


wild animals. Any animal recovered and showing no remaining sequelae is released.
Those who unfortunately do not survive (either because of the type of injuries they
have had or because they did not arrive on time at the Foundation), their bodies are
used for educational and scientific purposes.

Acknowledgments  We would like to thank to the artisanal fishermen of Villa del Mar, especially
to the memory of Mario Delgado, Coronel Rosales, and Ingeniero White; also the personnel of the
Prefectura Naval Argentina (Coast Guard), Bahia Blanca Delegation, Coronel Rosales, and Monte
Hermoso; SIPA Station (Rescue, Fire, and Environmental Protection); Oscar Liberman AIku,
Carolina Parodi, the staff and park rangers of the Natural Reserve Bahía Blanca, Bahía Falsa,
Bahía Verde and the FRAAM Foundation Volunteers.

References

Albareda DA, Bordino P, Prosdocimi L et al (2007a) Captura accidental de tortuga verde (Chelonia
mydas) en la pesquería artesanal del sur de la Bahía Samborombón, Buenos Aires, Argentina.
En: III Jornadas de Conservación e Investigación de Tortugas Marinas en el Atlántico Sur
Occidental. Piriápolis, Uruguay. 37 p
Albareda D, Álvarez K, Iglesias M et al (2007b) ¡No te comas mi basura! Evidencia de ingestión
de desechos antrópicos en tortugas verdes (Chelonia mydas) del estuario del Río de la Plata y
la Bahía de Samboronbóm – Pcia. de Buenos Aires – Argentina. En: Libro de resúmenes de la
III Jornadas de Investigación y Conservación de Tortugas Marinas del Atlántico Sur Occidental
(ASO), Piriápolis, Uruguay, pp 39–40
Bailey H, Shillinger G, Palacios D et al (2008) Identifying and comparing phases of movement by
leatherback turtles using state-space models. J Exp Mar Biol Ecol 356:128–135
Billes A, Fretey J, Verhage B et al (2006) First evidence of leatherback movement from Africa to
South America. Marine Turtle Newsletter, IUCN/SSC MTSG News 111:13–14
Bolten AB (2003) Variation in sea turtle life history patterns: neritic versus oceanic developmental
stages. In: Lutz PL, Musick JA, Wyneken J (eds) The biology of sea turtles, vol 2. CRC Press,
Boca Raton, pp 243–457
Breed GA, Jonsen ID, Myers RA et al (2009) Sex-specific, seasonal foraging tactics of adult grey
seals (Halichoerus grypus) revealed by statespace analysis. Ecology 90(11):3209–3221
Broderick AC, Godley BJ, Hays GC (2001) Trophic status drives inter- annual variability in nest-
ing numbers of marine turtles. Proc R Soc Lond B 268:1481–1487
Broderick AC, Frauenstein R, Glen F, Hays GC et al (2006) Are green turtles globally endangered?
Glob Ecol Biogeogr 15:21–26
Bugoni L, Krause L, Petry MV (2001) Marine debris and human impacts on sea turtles in southern
Brazil. Mar Pollut Bull 41(12):1330–1334
Carpenter JW (2006) 3rd Edition. Formulary for Exotic Animals
Dodd CK Jr (1988) Synopsis of the biological data on the loggerhead sea turtle Caretta caretta
(Linnaeus, 1758). US Fish Wild Serv Biol Rep 88(14):1–110
Domingo A, Bugoni L, Prosdocimi L, et al (2006) The impact generated by fisheries on Sea Turtles
in the Southwestern Atlantic. WWF Progama Marino para Latinoamérica y el Caribe, San José,
Costa Rica
Fossette S, Girard C, López-Mendilaharsu M, Miller P et al (2010) Atlantic leatherback migratory
paths and temporary residence areas. PLoS One 5(11):e13908. https://doi.org/10.1371/Journal.
pone.0013908.
324 V. Massola et al.

Frazier J (1984) Las tortugas marinas en el Océano Atlántico Sur Occidental. La Plat: Asociación
Herpetologíaca Argentina 22 p. ISSN 0326-5528
Frazier J (2003) Prehistoric and ancient historic interactions between humans and marine turtles.
In: Lutz PL, Musick JA, Wyneken J (eds) The biology of sea turtles, vol II. CRC Press, Boca
Raton, pp 1–38
Frazier J (ed) (2005) Marine turtles as flagships. MAST (Maritime Stud) 3(4):1–303
Freiberg MA (1938) Catálogo sistemático y descriptivo de las tortugas argentinas. Memorias del
Museo de Entre Ríos, Zoología 9:1–25
Freiberg MA (1942) La tortuga laúd Dermochelys coriacea (L.) frente a las costas argentinas.
Physis 19(52):263–265
Freiberg MA (1945) Observaciones sobre las tortugas de mar que se encuentran frente a las costas
argentinas. Physis 20(55):50–53
Gallardo JM (1977) Reptiles de los alrededores de Buenos Aires. Buenos Aires Editoral
EUDEBA:213
Gallo BMG, Macedo S, Giffoni BB et al (2006) Sea turtle conservation in Ubatuba, South Eastern
Brazil, a feeding area with incidental capture in coastal fisheries. Chelonian Conserv Biol
5(1):93–101
Godley BJ, Blumenthal JM, Broderick AC et al (2008) Satellite tracking of sea turtles: where have
we been and where do we go next? Endanger Species Res 4:3–22
González Carman V, Álvarez K, Prosdocimi L et al (2011) Temperate SW Atlantic: a feeding and
developmental hábitat for endangered sea turtles. Mar Biol Res 7:500–508
González Carman V, Falabella V, Maxwell S et al (2012) Revisiting the ontogenetic shift para-
digm: the case of juvenile green turtles in the SW Atlantic. J Exp Mar Biol Ecol 429:64–72
González Carman V, Mandiola A, Alemany A et al (2016a) Distribution of mega faunal species
in the Southwestern Atlantic: key ecological areas and opportunities for marine conservation.
ICES J Mar Sci. https://doi.org/10.1093/icesjms/fsw019
González Carman V, Bruno I, Maxwell S et  al (2016b) Habitat use, site fidelity and conserva-
tion opportunities for juvenile loggerhead sea turtles in the Río de la Plata, Argentina. Mar
Biol 163:20
Hart KM, Fujisaki I (2010) Satellite tracking reveals habitat use by juvenile green sea turtles
Chelonia mydas in the Everglades, Florida, USA. Endanger Species Res 11:221–232
Hart KM, Lamont MM, Fujisaki I (2011) Common coastal foraging areas for loggerheads in the
Gulf of Mexico: opportunities for marine conservation. Biol Conserv 145(1):185–194
IUCN (2014) IUCN Red List of Threatened Species. Version 2014.2. http://www.iucnredlist.org/.
Accessed 13 Mar 2020
Jonsen ID, Myers RA, James MC (2007) Identifying leatherback turtle foraging behaviour from
satellite telemetry using a switching state-space model. Mar Ecol Prog Ser 337:255–264
Laporta M, Miller P, Ríos M et al (2006) Conservación y manejo de tortugas marinas en la zona
costera uruguaya. In: Menafra R, Rodríguez-Gallego L, Scarabino F et al (eds) Bases para la
conservación y manejo de la costa uruguaya. Vida Silvestre Uruguay, Montevideo, pp 259–269
López-Mendilaharsu M, Rocha CFD, Miller P et al (2009) Insight son leatherback turtle move-
ments and high use areas in the Southwest Atlantic Ocean. J Exp Mar Biol Ecol 378:31–39.
https://doi.org/10.1016/j.jembe.2009.07.010
Luschi P, Hays GC, Del Seppia C et al (1998) The navigational feats of green sea turtles migrating
from Ascension Island investigated by satellite telemetry. Proc R Soc Lond B 265:2279–2284
Makowski C, Seminoff JA, Salmon M (2006) Home range and habitat use of juvenile Atlantic
green turtles (Chelonia mydas L.) on shallow reef habitats in Palm Beach, Florida, USA. Mar
Biol 148:1167–1179
Maxwell SM, Breed GA, Nickel BA et al (2011) Using satellite tracking to optimize protection of
long-lived marine species: Olive Ridley sea turtle conservation in Central Africa. PLoS One
6(5):e19905
Mc Clenachan L, Jackson JBC, Newman MJH (2006) Conservation implications of historic sea
turtle nesting beach loss. Front Ecol Environ 4(6):290–296
12  Bahía Blanca Estuary and the Importance of Wetlands for the Conservation of Sea… 325

McClellan CM, Read AJ (2009) Confronting the gauntlet: understanding incidental capture of
green turtles through fine-scale movement studies. Endanger Species Res 10:165–179
McDonald DL, Dutton PH (1996) Use of PIT tags and photo identification to revise remigration
estimates of leatherback turtles (Dermochelys coriacea) eating in St. Croix, U.S. Virgin Island,
1979-1995. Chel Cons Biol 2(2):148–152
Patterson TA, Thomas L, Wilcox C et al (2008) Statespace models of individual animal movement.
Trends Ecol Evol 23(2):87–94
Piola AR, Matano RP (2001) Brazil and Falklands (Malvinas) currents. In: Steele JH, Thorpe SA,
Turekian KK (eds) Encyclopedia of ocean sciences. Academic, London, pp 340–349
Polovina JJ, Kobayashi DM, Seki MP et al (2000) Turtles on the edge: movement of loggerhead
turtles (Caretta caretta) along oceanic fronts, spanning longline fishing grounds in the central
North Pacific, 1997–1998. Fish Oceanogr 9:71–82
Pritchard PCH, Mortimer JA (1999) Taxonomy, external morphology, and species identification.
In: Eckert KL, Bjorndal KA, Abreu-Grobois FA et al (eds) Research and management tech-
niques for the conservation of sea turtles. IUCN/SSC Marine Turtle Specialist Group. 235 p
Prosdocimi L, González Carman V, Albareda D et al (2012) Genetic composition of green turtle
feeding grounds in coastal waters of Argentina based on mitochondrial DNA. J Exp Mar Biol
Ecol 412:37–45
Prosdocimi L, Bruno I, Diaz L, González-Carman V, Albareda DA, Remis MI (2014a) Southernmost
reports of the Hawksbill Sea turtle, Eretmochelys imbricata (Linnaeus, 1766), in Argentina.
Herpetol Rev 45(1):1–5
Prosdocimi L, Dutton PH, Albareda DA et al (2014b) Origin and genetic diversity of leatherbacks
(Dermochelys coriacea) at argentine foraging grounds. J Exp Mar Biol Ecol 458:13–19
Prosdocimi L, Bugoni L, Albareda DA et al (2015) Are stocks of immature loggerhead sea turtles
always mixed? J Exp Mar Biol Ecol 466:85–91
Rivilla JC, Alís S, Alís L et al (2005) Ejemplar de tortuga boba (Caretta caretta) con anomalías
morfológicas en los escudos del espaldar. Bol Asoc Herpetol Esp 15:98–100
Seaman DE, Powell RA (1996) An evaluation of the accuracy of kernel density estimators for
home range analysis. Ecology 77(7):2075–2085
Seminoff JA (2004) Global status assessment green turtle (Chelonia mydas). Marine Turtle
Specialist Group. The World Conservation Union IUCN, Gland
Seminoff JA, Resendiz A, Nichols WJ (2002) Home range of green turtles Chelonia mydas at a
coastal foraging area in the Gulf of California. Mexico Mar Ecol Prog Ser 242:253–265
Work Booklet ASO (2004) Data notes standardization. Project, Tamar
Worton BJ (1989) Kernel methods for estimating the utilization distribution in homerange studies.
Ecology 70(1):164–168
Wyneken J (2001) The anatomy of Sea Turtles. U. S. Department of Commerce NOAA Technical
Memorandum NMFS-SEFSC-470. 172 p
Wyneken J (2003) In: Lutz P, Musick JA, Wyneken J (eds) The external, musculoskeletal and
neuro-anatomy of sea turtles: the biology of sea turtles (II). CRC Press, Boca Raton, pp 39–77
Chapter 13
Shorebirds and Seabirds’ Ecology
and Conservation

Natalia S. Martínez-Curci, Germán O. García, Leandro Marbán,


Pía Simonetti, and Sergio M. Zalba

13.1  Introduction

Shorebirds and seabirds belonging to the order Charadriiformes are probably some
of the most conspicuous animals inhabiting Bahía Blanca Estuary. This extremely
diverse order of birds is second only to passerines in number of families and species
(Lovette 2016). Having a global distribution that ranges from the Arctic to the
Antarctic, they can be found in almost any terrestrial habitat type, as well as in the
sea. Consistently, they exhibit remarkably varied breeding strategies and migration

N. S. Martínez-Curci ()
Grupo Vertebrados, Instituto de Investigaciones Marinas y Costeras, Universidad Nacional de
Mar del Plata, Consejo Nacional de Investigaciones Científicas y Técnicas (CONICET),
Buenos Aires, Argentina
Coastal Solutions Fellows Program, Cornell Lab of Ornithology, Cornell University,
Ithaca, NY, USA
Bird Ecology Lab, Instituto de Ciencias Marinas y Limnológicas, Universidad Austral de
Chile, Valdivia, Chile
e-mail: nsm85@cornell.edu
G. O. García
Grupo Vertebrados, Instituto de Investigaciones Marinas y Costeras, Universidad Nacional de
Mar del Plata, Consejo Nacional de Investigaciones Científicas y Técnicas (CONICET),
Buenos Aires, Argentina
L. Marbán · S. M. Zalba
GEKKO, Grupo de Estudios en Conservación y Manejo, Departamento de Biología,
Bioquímica y Farmacia, Universidad Nacional del Sur, Bahía Blanca, Argentina
Consejo Nacional de Investigaciones Científicas y Técnicas (CONICET),
Buenos Aires, Argentina
P. Simonetti
Instituto Argentino de Oceanografía (IADO), Universidad Nacional del Sur, CONICET,
Buenos Aires, Argentina

© Springer Nature Switzerland AG 2021 327


S. M. Fiori, P. D. Pratolongo (eds.), The Bahía Blanca Estuary,
https://doi.org/10.1007/978-3-030-66486-2_13
328 N. S. Martínez-Curci et al.

systems. Although the monophyly of Charadriiformes is well established, the rela-


tionship of the families within this order has been controversial. According to the
most recent evidence (Paton et al. 2003; Thomas et al. 2004; Hackett et al. 2008),
the order consists of three major groups: Charadrii (includes plovers, oystercatch-
ers, avocets, stilts, thick-knees, and sheathbills); Scolopaci (includes sandpipers,
seedsnipes, jacanas, and painted-snipes); and Lari (includes skuas, skimmers, gulls,
and terns). The former two groups constitute what are commonly known as shore-
birds, while the latter consists of seabirds. This chapter is organized into three sec-
tions in which we will describe some relevant ecological aspects of shorebirds
(Sect. 13.2) and seabirds (Sect. 13.3) with emphasis on the species of Bahía Blanca
Estuary. Finally, we will analyze their main conservation challenges both at global
and local contexts (Sect. 13.4).

13.2  Shorebirds

Shorebirds, also called waders, are small- to medium-sized birds closely associated
with open environments, especially wetlands that inhabit all continents. Most of
them spend at least part of their lives on tidal flats at estuarine or marine shorelines,
while others use inland habitats such as grasslands, rivers, lakes, and lagoons. As a
consequence of their cosmopolitan distribution and habitat associations, these birds
exhibit diverse morphological adaptations that include marked variations in body
size and bill morphologies. For example, the white-rumped sandpiper (Calidris fus-
cicollis), one of the smallest shorebird species in Bahía Blanca Estuary, is 15–18 cm
in length and 40–60  g in mass and has a short, fairly straight bill (Fig.  13.1a;
Parmelee 2020). At the opposite end, the Hudsonian godwit (Limosa haemastica),
one of the largest shorebird species, is 36–42 cm in length and 200–350 g in mass
and has a long, upturned bill (Fig. 13.1b; Walker et al. 2020). These morphological
variations allow shorebirds to exploit different trophic resources while coexisting in
large numbers at key sites along their flyways (Martínez-Curci et al. 2015). Flyways
are routes, used by several species, that encompass the full range of breeding, stop-
ping, staging, and non-breeding areas occupied by a population during the annual
cycle (Boere and Stroud 2006). As we will see, Bahía Blanca Estuary constitutes
one of the most important sites identified within the Atlantic Americas Flyway.
There are about 215 recognized shorebird species in the world, distributed
unevenly among 14 families (Colwell 2010). Records in Bahía Blanca account for
37 species in 9 families (Table 13.1). The largest number of species is hosted by
Charadriidae and Scolopacidae. These two families show quite marked morphologi-
cal and behavioral differences. Members of Charadriidae (i.e., plovers and allies)
have relatively large heads with short bills and large eyes (Fig. 13.1c) that facilitate
visual foraging for what they use a “run-stop-peck” strategy (Colwell and Haigh
2019) that is considered the ancestral condition in shorebirds (Barbosa and Moreno
1999). On the other hand, Scolopacidae members (i.e., sandpipers and allies) have
long bills relative to their body size that are varied in shapes (e.g., upturned in
13  Shorebirds and Seabirds’ Ecology and Conservation 329

Fig. 13.1  Some of the most common shorebirds in Bahía Blanca Estuary: (a) white-rumped sand-
piper (Calidris fuscicollis); (b) Hudsonian godwits (Limosa haemastica) with basic (left) and alter-
nate plumage (right); (c) two-banded plover (Charadrius falklandicus); (d) red knots (Calidris
canutus) with basic, intermediate, and alternate plumage. (Photos by Natalia S. Martínez-Curci (a,
b, d), Lautaro Rodríguez Astorino (c))

godwits, straight in knots, decurved in curlews, wedge-shaped in turnstones;


Figs.  13.1a,b,d). Most members of this family feed on buried prey using tactile
strategies (Barbosa and Moreno 1999) that in many cases rely on mechanoreceptors
located at the tip of the bill (Piersma et al. 1998); others use their brush-like tongues
to feed on biofilm (Elner et al. 2005).
Another notable difference between Scolopacidae and Charadriidae is related to
their breeding areas and migration strategies. Most sandpipers are northern-latitude
breeders that in many cases undertake amazing migrations to the southern extreme
of the globe. Most plovers, however, breed on temperate and tropical latitudes and
are less prone to migrate long distances (Colwell and Haigh 2019). These patterns
are reflected in the two migratory currents (norther vs. austral migrants), with oppo-
site chronologies, that converge in Bahía Blanca Estuary. Northern migrants (i.e.,
birds that breed in North America during the boreal summer/austral winter and then
migrate to the southern hemisphere) are the most abundant shorebirds in Bahía
Table 13.1  List of Charadriiformes recorded in Bahía Blanca Estuary (for the Charadrii and Scolopaci suborders, we followed Martínez-Curci and Petracci
330

(2016, Petracci and Delhey 2005); for Lari we followed Petracci and Sotelo (2013); Petracci and Delhey (2005) and authors’ experience). The migratory pattern
indicates the following: summer visitor (Sv), winter visitor (Wv), northern migrant (NM), austral migrant (AM), or resident (R). The probability of observation
(Prob Obs) refers to the frequency with which species can be observed within the estuary: abundant (A: frequent observations of large abundances), common (C:
frequent observations), uncommon (U: infrequent but mostly annual observations), occasional (O: very infrequent observations, which generally do not occur
every year). Conservation status (Cons Stat) shows the categorization of the species according to national (Nat) and international (Int) criteria. For the national
categorization, we followed MAyDS and Aves Argentinas (2017) keeping the acronyms of the original document published in Spanish: critically endangered
(EC), endangered (EN), threatened (AM), vulnerable (VU), and not threatened (NA). For the international categorization, we followed IUCN (2020): near
threatened (NT) and low concern (LC). We also followed IUCN (2020) for population trends (Pop Trend): increasing (I), stable (S), decreasing (D), unknown (U)
Migratory Prob Cons Stat Pop
pattern Obs Nat Int Trend
Charadrii
Charadriidae
American golden plover Pluvialis dominica Sv, NM C NA LC D
Grey plover Pluvialis squatarola Sv, NM O NA LC U
Tawny-­throated dotterel Oreopholus ruficollis Wv, AM C NA LC D
Southern lapwing Vanellus chilensis R C NA LC I
Semipalmated plover Charadrius semipalmatus Sv, NM C NA LC S
Collared plover Charadrius collaris R U NA LC D
Two-­banded plover Charadrius falklandicus Wv, AM A NA LC S
Rufous-­chested plover Charadrius modestus Wv, AM U NA LC U
Haematopodidae
American oystercatcher Haematopus palliatus R A NA LC S
Blackish oystercatcher Haematopus ater Wv, AM O NA LC U
Magellanic oystercatcher Haematopus leucopodus Wv, AM V NA LC S
Recurvirostridae
Black-­necked stilt Himantopus mexicanus R A NA LC U
Chionidae
Snowy sheathbill Chionis albus Wv, AM O NA LC S
N. S. Martínez-Curci et al.

Pluvianellidae
Magellanic plover Pluvianellus socialis Wv, AM O EN NT S
Migratory Prob Cons Stat Pop
pattern Obs Nat Int Trend
Charadrii
Charadriidae
Scolopaci
Scolopacidae
Upland sandpiper Bartramia longicauda Sv, NM O VU LC I
Whimbrel Numenius phaeopus Sv, NM U NA LC D
Hudsonian godwit Limosa haemastica Sv, NM A NA LC D
Ruddy turnstone Arenaria interpres Sv, NM U NA LC D
Red knot Calidris canutus Sv, NM C EC NT D
Sanderling Calidris alba Sv, NM U NA LC U
White-­rumped sandpiper Calidris fuscicollis Sv, NM A NA LC D
Baird’s sandpiper Calidris bairdii Sv, NM C NA LC S
Pectoral sandpiper Calidris melanotos Sv, NM U NA LC S
Stilt sandpiper Calidris himantopus Sv, NM O NA LC I
Buff-­breasted sandpiper Calidris subruficollis Sv, NM U AM NT D
13  Shorebirds and Seabirds’ Ecology and Conservation

Short-­billed Dowitcher Limnodromus griseus Sv, NM V NA LC D


South American snipe Gallinago paraguaiae R U NA LC S
Wilson’s phalarope Phalaropus tricolor Sv, NM U NA LC I
Spotted sandpiper Actitis macularius Sv, NM O NA LC D
Greater yellowlegs Tringa melanoleuca Sv, NM C NA LC S
Willet Tringa semipalmata Sv, NM O NA LC S
Lesser yellowlegs Tringa flavipes Sv, NM C NA LC D
Thinocoridae
Least seedsnipe Thinocorus rumicivorus R O NA LC S
(continued)
331
Table 13.1 (continued)
332

Migratory Prob Cons Stat Pop


pattern Obs Nat Int Trend
Charadrii
Charadriidae
Jacanidae
Wattled jacana Jacana jacana R O NA LC S
Rostratulidae
South American painted-­snipe Nycticryphes semicollaris R U NA LC D
Lari
Stercorariidae
Parasitic jaeger Stercorarius parasiticus Sv, AM O NA LC S
Rynchopidae
Black skimmer Rynchops niger Sv, NM C NA LC D
Laridae
Brown-­hooded gull Chroicocephalus maculipennis R C NA LC I
Gray-­hooded gull Chroicocephalus cirrocephalus R U NA LC S
Olrog’s gull Larus atlanticus R A VU NT S
Kelp gull Larus dominicanus R A NA LC I
Gull-­billed tern Gelochelidon nilotica R C NA LC D
South American tern Sterna hirundinacea Wv, AM U NA LC D
Snowy-­crowned tern Sterna trudeaui R A NA LC S
Sandwich tern Thalasseus sandvicensis R U NA LC S
Royal tern Thalasseus maximus R U NA LC S
N. S. Martínez-Curci et al.
13  Shorebirds and Seabirds’ Ecology and Conservation 333

Blanca Estuary. This group dominates the assemblage from austral spring to late
summer/early fall. Some species, such as the white-rumped sandpiper (Fig. 13.1a),
which is the most abundant shorebird in the area, spend the whole non-breeding
season at Bahía Blanca. However, its abundance may be higher during the passage
of individuals that spend the non-breeding season in southern latitudes (Petracci and
Sotelo 2013, Belenguer et al. 1992). Other species, like the red knot (Calidris canu-
tus; Fig.  13.1d), use the area for fueling during northward migration and can be
found almost exclusively from March to May (Petracci and Sotelo 2013). With the
arrival of the autumn, and during the winter, the assemblage becomes dominated by
austral migrants (i.e., birds that breed in southern South America, mainly in
Patagonia, during the austral spring and summer and then migrate north reaching
southern Brazil). Among austral migrants, the two-banded plover (Charadrius falk-
landicus; Fig. 13.1c) is the most abundant species with some individuals remaining
in the area throughout the year (Petracci and Sotelo 2013). Residents (i.e., birds that
remain in the same area year-round), such as the American oystercatcher
(Haematopus palliatus; Fig.  13.5a, in Box 13.2) or the black-necked stilt
(Himantopus mexicanus), represent a smaller fraction of the assemblage (Martínez-­
Curci and Petracci 2016).

13.2.1  Migration

Perhaps one of the most fascinating aspects of shorebirds is their ability to migrate.
Over 60% of shorebird species undertake short- to long-distance migrations
(Colwell 2010). Several members make some of the most extreme migratory flights
in the world exceeding 30,000  km per year. Migration annually takes shorebirds
from breeding to non-breeding grounds where they may spend as much as 8 to
10 months of the year (Colwell 2010). From an evolutionary point of view, migra-
tion strategies may arise if the benefits of moving seasonally are greater than the
benefits of staying in one region (Lack 1968). However, to take advantage of the
world’s seasonality, birds must perform energetically demanding flights.
Different species have developed different migration strategies, undertaking
journeys of varying length and duration (Piersma 1987). Most species travel in
many short periods of migratory flight interspersed with short periods of refueling
(that strategy is called “hop”). Others may do some long-distance flights inter-
spersed with resting and fueling periods (“skip”). Finally, in the most extreme cases,
shorebirds can travel in one very long migratory flight (“jump”). Following the lat-
ter strategy, some of the species that inhabit Bahía Blanca Estuary make nonstop
endurance flights of about 8000 km in 6 days in the case of red knots (Niles et al.
2010) and more than 10,000 km in 7 days in the case of Hudsonian godwits (Senner
et al. 2014).
To accomplish these long-distance migratory flights, shorebirds rely on energy
and metabolic water that is mainly stored subcutaneously in the form of lipids but
also as proteins in flight muscles (Jenni and Jenni-Eiermann 1998; Jenni-Eiermann
334 N. S. Martínez-Curci et al.

et al. 2002). Immediately before starting the trip, they deposit huge amounts of tri-
glycerides at the same time they hypertrophy their flight muscles without any
noticeable increase in exercise (Lindström and Piersma 1993). In this way, migra-
tory shorebirds are capable of doubling their mass prior to migration (Piersma and
Gill 1998) for which they spend considerable time foraging while increasing their
consumption rate in a period of hyperphagia (Burger and Olla 1984).

13.2.2  Dependence on Key Sites

To achieve high intake rates and fat loads, shorebirds depend on a limited number
of key sites that serve as staging areas (i.e., locations where birds fuel up in a pause
during migration), stopover sites (i.e., locations where migrant birds take a short
break to rest a feed), and non-breeding grounds (i.e., locations where migrant birds
spend most of the non-breeding season). Most of these sites are wetlands that pro-
vide large amounts of high-quality food. Shorebirds’ diet consist mainly of inverte-
brates; in Buenos Aires province, some of their major prey items are polychaetes,
crabs, mollusks, and insects (Iribarne and Martínez 1999; Petracci 2002; Ieno et al.
2004; Ribeiro et  al. 2004; Martínez-Curci et  al. 2015; Martínez-Curci and
Petracci 2016).
Several shorebird species exhibit philopatry to the mentioned key sites, as well
as aggregative behaviors during the non-breeding season and migration. This ten-
dency to return in large flocks to the same stopping, staging, and non-breeding areas
confers several benefits, such as familiarity with resources and conditions or
decreased predation risk (Shuter et al. 2011). However, since significant percent-
ages of several populations are concentrated in a limited number of sites, it also
makes shorebirds more susceptible to habitat loss or degradation (Myers 1983;
Shuter et  al. 2011). About 50% of shorebirds with known population trends are
declining, and the conservation of this highly migratory species requires interna-
tional cooperation. In pursuit of this, Bahía Blanca Estuary has joined the Western
Hemisphere Shorebird Reserve Network (WHSRN), an international effort aimed
to conserve shorebirds and their habitats across the Americas through action at a
network of key sites. Holding more than 20,000 shorebirds per year and more than
1% of the biogeographical populations of American oystercatcher (Fig. 13.5 in Box
13.2), two-banded plover (Fig.  13.2c), Hudsonian godwit (Fig.  13.2b), red knot
(Fig.13.1d), and white-rumped sandpiper (Fig. 13.2a), Bahía Blanca Estuary was
designated as a site of regional importance within the WHSRN in 2016 (https://
whsrn.org/whsrn_sites/estuario-­de-­la-­bahia-­blanca/).
13  Shorebirds and Seabirds’ Ecology and Conservation 335

Fig. 13.2  Representative seabirds of the order Charadriiformes in Bahia Blanca Estuary: (a) black
skimmers (Rynchops niger); (b) flock of terns, South American tern (Sterna hirundinacea; left),
sandwich terns (Thalasseus sandvicensis; back), and snowy-crowned terns (Sterna trudeaui;
front); (c) Olrog’s gulls (Larus atlanticus). (Photos by Hugo Gribman (a), Claudio Rodríguez (b),
Pablo Fernández (c))

13.3  Seabirds

Seabirds are defined as those birds living in and making their living from the marine
environment, which includes coastal areas, estuaries, wetlands, and oceanic islands
(Schreiber and Burger 2001). Within Charadriiformes, seabirds are grouped in the
suborder Lari that comprises about 127 species. In Bahía Blanca Estuary, this group
is mostly represented by gulls and terns (see Table  13.1). From a morphological
perspective, Lari constitutes an heterogeneous group. In Bahia Blanca these birds
range from small size (e.g., South American tern Sterna hirundinacea, 170–220 g in
weight and has an 84–86 cm wingspan; Gochfeld et al. 2020) to large species (e.g.,
kelp gull Larus dominicanus, 810–1335 g in weight and has a 128–142 cm wing-
span, Burger et al. 2020). They are also heterogeneous from an ecological point of
view. Some members feed on a few types of prey so they are considered specialists,
while others are able to use a wide range of resources being considered generalists.
Furthermore, there are resident species but also trans-equatorial migrants that travel
large distances across several oceanographic systems.
336 N. S. Martínez-Curci et al.

Seabirds are equally at home on land, in the air, and in the water, changing from
one to the other, often daily. Such flexibility requires unique morphological and
physiological adaptations to an environment that has also exerted selective forces on
their behavior, ecology, and demography. Bills, feet, and body shapes show innu-
merable adaptations to various lifestyles, allowing seabirds to swim and dive using
webbed feet or even their wings to propel themselves in the water. Although all
seabirds use their bills to capture and handle food, they exhibit adaptations for dif-
ferent types of feeding. An emblematic example is that of skimmers, which are
represented in Bahía Blanca by the black skimmer (Rynchops niger; Fig  13.2a).
They have a specialized bill in which the lower mandible is compressed laterally
and is longer than the upper mandible. This provides a greater surface area that,
along with their flight mechanics, allows skimmers to catch fish by efficiently skim-
ming the water surface with their lower mandible as they search for prey with tactile
cues (Zusi 1996). Adaptations for feeding in marine environments are not just
restricted to bills. By feeding in the sea, these birds also must deal with high physi-
ological loads of salt. To accomplish this, they rely on salt glands (i.e., organs for
excreting excess salts) that are found in shallow depressions around the orbit, above
the eye, which help to cope with a diet overloaded with salt (Schmidt-Nielsen 1960).
In addition, just like shorebirds, seabirds have the capacity to cope with fattening
periods as a physiological adaptation to migration.

13.3.1  Foraging

Seabirds are at the higher trophic levels of the marine food web. When breeding,
they must return to their nest after every fishing trip to carry food for their chicks, a
pattern referred by ecologist as central-place foraging (Orians and Pearson 1979).
During the non-breeding season, social foraging is widespread, and it usually
involves several species (Thiebault et al. 2014). Foraging activity can occur during
the day or night, and presumably these different foraging habits have evolved in
response to the behavior of their preys. Seabird diet consists mainly of fish, crusta-
ceans, and/or mollusks. However, some species have incorporated anthropogenic
items in their trophic spectrum. In Argentina, for example, coastal gulls such as the
Olrog’s gull (Larus atlanticus) and the kelp gull have been registered preying on
resources facilitated by fishery activities (Berón et al. 2013; Marinao et al. 2019).
To obtain food, different foraging strategies, generally linked to morphological
and/or physiological features, are used. Gulls fed by methods that include picking
up items from the ground, surface dipping, jump-plunging, and other forms of dip-
ping. In terns (Fig. 13.2b), plunge-diving, diving-to-surface, dipping, and hawking
(to catch insects) are well-documented feeding techniques (Cabot and Nisbet 2013).
Although most seabirds catch live prey, scavenging is a feeding method employed
by a small proportion of seabirds, especially gulls and skuas (Furness et al. 2007).
Many scavenging species have increased dramatically in numbers, and these
increases have often been attributed to the feeding opportunities presented over
13  Shorebirds and Seabirds’ Ecology and Conservation 337

many decades by fishery waste (Oro et al. 2013). Kleptoparasitism (i.e., parasitic
interaction in which one animal steals food from another; Rothschild and Clay
1952) is another very special technique for acquiring prey. Skuas have often been
considered to be highly evolved as specialized kleptoparasites, that is, they may use
kleptoparasitism for all or most of their energy acquisition (Brockmann and Barnard
1979). Specialist kleptoparasites exhibit adaptations that are apparently absent in
opportunist species (e.g., ability to detect and attack hosts carrying food concealed
in the proventriculus, to sustain prolonged aerial chases, and to adapt their breeding
cycle to match that of their host; Furness 1987). Within Bahía Blanca Estuary, there
are some records where the Arctic skua (Stercorarius parasiticus) has been observed
parasitizing on sandwich terns (Thalasseus sandvicensis; Fig. 13.2b) (Petracci and
Sotelo 2013). Unlike skuas, gulls and terns may resort to food theft opportunisti-
cally, using a range of foraging tactics. In these cases, kleptoparasitism is context-­
dependent and often occurs during periods of low availability of primary food
sources (García et al. 2010). Even in opportunistic kleptoparasites, some individuals
may be more specialized as kleptoparasites than others, and this behavior may be
related to reproductive performance (García et al. 2011, 2013). Recent studies on
terns showed that in individuals specialized as kleptoparasites, the energetic reward
of a kleptoparasitic event is sex-specifically related to the tactic used by the parasite,
showing signs of within-individual improvement with age in both sexes (García
et al. 2020).

13.3.2  Breeding

Most seabird species are social and nest in colonies that can include hundreds, thou-
sands, tens of thousands, or even hundreds of thousands of pairs (Schreiber and
Burger 2001). Almost all the species included in this chapter nest seasonally, being
at breeding colonies for 3–5 months. After breeding, most of them migrate or dis-
perse to non-breeding areas where they spend more than half the year. Migration
patterns can include a variety of strategies, durations, and distances covered. The
dominant mating system is monogamous with cooperation of both parents for a suc-
cessful incubation and chick rearing. During breeding, there is some differentiation
of sex roles. In terns, for example, males establish the nest territory and take the
leading role in defending it and feed their mates during the period of egg formation
and egg-laying. Some species show, as well, high mate fidelity. Courtship feeding
can occur early in the season, as a way for females to assess how good food provid-
ers their partners might be. It is also important to allow proper egg formation during
the pre-incubation stage, relieving females from the acquisition of food that con-
sume time and energy. A nest of varying structure is usually built to protect the eggs.
Some are placed on cliffs and islands, away from predators, while others can be
placed in cavities to prevent attacks from aerial predators. Seabirds grouped as
Charadriiformes lay more than one egg; mechanisms related to egg-size asymmetry
and laying/hatching asynchrony may lead to reduced survival of the last chicks if
338 N. S. Martínez-Curci et al.

the environmental variables are not favorable. Chick-rearing period is about


25–30 days.
Among the seabirds of Bahía Blanca Estuary (Table 13.1), only two species of
gulls reproduce regularly in the area (Yorio et al. 1998). One of them is Olrog’s gull
(Fig. 13.2c), an endemic species to the Atlantic coast of southern South America,
which is one of the few Larus species in the world with adverse conservation status
(see Sect. 13.3). There are scant identified Olrog’s gull breeding sites, and most of
the breeding population is concentrated in Bahía Blanca Estuary. The largest colony
reported to date is located at “Islote de la Gaviota Cangrejera” Natural Reserve
(Delhey et al. 2001a; Yorio et al. 2005, Petracci et al. 2008). After the breeding sea-
son, this species disperses along the Atlantic coast reaching southern Brazil.
Although the species remains throughout the year in the estuary and is considered a
local resident (Table 13.1), recent evidence indicates that the population that breeds
in Bahía Blanca Estuary may have a partial migration behavior as observed in the
San Blas bay population (Copello et al. 2020). Olrog’s gull is considered one of the
few gulls specialized in feeding on crabs (Escalante 1966; Herrera et  al. 2005;
Petracci et al. 2004). However, they can occasionally feed on other items (Copello
and Favero 2001; Petracci et al. 2007; Berón et al. 2013); recent studies conducted
in non-breeding areas try to understand the ways in which they access to novel food
sources (García et al. 2019; Castano et al. 2020). The other species that breeds in the
area is the kelp gull. It has generalist habits, an extremely large population size, and
a breeding distribution that spans the coasts and islands of much of the southern
hemisphere (including South America, Africa, New Zealand, Australia, Subantarctic
Islands, and Antarctic Peninsula; Burger and Gochfeld 1996a, b). Although there is
no published information on the number of breeding pairs inhabiting Bahía Blanca
Estuary, it is known that the Argentine population has increasing trends. It has been
argued that, in some coastal areas, population growth would be related to the use of
anthropogenic food subsidies (Lisnizer et al. 2011).

13.4  Conservation

13.4.1  Main Threats

The last centuries have been marked by a rapid loss of biodiversity (Pimm et al.
2014; Ceballos et  al. 2015), being habitat loss and degradation one of the main
identified causes (Sala et al. 2000; Hoekstra et al. 2005). This biodiversity crisis not
only involves extinctions but also a decrease in the number of individuals in local
populations, which can result in changes in the composition of communities and the
function of ecosystems (Dirzo et al. 2014; Rosenberg et al. 2019). Accordingly, both
shorebird and seabird species have been experiencing drastic population declines
(Croxall et al. 2012; Colwell 2010; Rosenberg et al. 2019).
13  Shorebirds and Seabirds’ Ecology and Conservation 339

Decreasing trends are caused by a variety of threats including climate change


among the most important on a global scale (Lascelles et al. 2016; van Gils et al.
2016; Kubelka et al. 2018). The effects of this threat are particularly relevant to spe-
cies inhabiting coastal habitats such as estuaries and breeding in high latitudes,
where climate change has already significantly modified the ecosystem. Some of
the most important land-based threats include the presence of invasive alien species,
problematic native species (e.g., those that have become super-abundant), human
disturbance, changes in land use and land cover, habitat deterioration, commercial
and residential development, and hunting (Sutherland et al. 2012a, b). Finally, at the
sea, some of the main problems are associated with bycatch (in gillnet, trawl, and
other fisheries; Paz et al. 2018), pollution (oil spills, chemical contaminants, plastic
and marine debris; Quadri Adrogué et al. 2019), noise (busy shipping lanes, seismic
surveys, and sonar; Pichegru et  al. 2017), prey depletion caused by overfishing,
energy production, and mining (Tasker et al. 2000; Croxall et al. 2012).
Bahía Blanca Estuary has been subject to novel threats and increasing levels of
pollution during the last decades. Global data indicates that although estuarine and
coastal systems provide vital habitats for many plants and animals, as well as a wide
variety of goods and services for millions of people, they are some of the most heav-
ily used and threatened natural systems in the world (Barbier et al. 2011). The prox-
imity to Bahía Blanca Estuary to urban areas, industrial parks, and ports has led to
contamination by coliforms, hydrocarbon derivatives, pesticides, and heavy metals
as some of the major threats (Martínez-Curci and Petracci 2016). In recent years,
problems associated with the expansion of invasive alien species have also become
relevant. One of these cases is the potential threat given by the Pacific oyster
(Magallana gigas). Although this alien mollusk does not seem to negatively affect
the feeding habits of some shorebirds (Escapa et al. 2004), it might cause structural
changes in the ecosystem (Luckenbach 1984; Kelly et al. 1996). On the other hand,
the establishment and expansion of an alien plant species named opposite-leaved
saltwort (Salsola inermis) is threatening the largest breeding colony of Olrog’s gull
at “Islote de la Gaviota Cangrejera” Natural Reserve, causing the displacement of
nests to suboptimal sites.

13.4.2  Shorebirds and Seabirds’ Conservation Status

Among the 46 shorebird and seabird species recorded at Bahía Blanca, 35% have
decreasing population trends, 41% are stable, only 13% are increasing, and the
remaining 11% have unknown trends (Table 13.1). According to the categorization
of the International Union for Conservation of Nature, four of the Charadriiformes
species inhabiting Bahía Blanca Estuary are near threatened (IUCN 2020). These
are three shorebirds, the Magellanic plover (Pluvianellus socialis), the red knot, and
the buff-breasted sandpiper (Calidris subruficollis), and one seabird, Olrog’s gull.
The latter three species are listed in the Appendix I of the Convention on Migratory
340 N. S. Martínez-Curci et al.

Species (CMS), while the Magellanic plover is listed in the Appendix II along with
other families of Charadriiformes.
The Magellanic plover is an austral migrant that might be occasionally observed
in brackish lagoons and salt ponds such as “La Vidriera” (Petracci and Sotelo 2013).
Having a small but stable population estimated in 1500–7000 mature individuals
(BirdLife International 2016), it is considered endangered in Argentina (MAyDS
and Aves Argentinas 2017). The red knot is a northern migrant that is commonly
seen in Bahía Blanca during northward migration (Petracci and Sotelo 2013). It has
a Holarctic breeding distribution with six recognized subspecies of which only
C. canutus rufa reaches the coast of South America up to Tierra del Fuego (Baker
et  al. 2020). Having numbered 67,500 individuals in the mid-1980s, dropped to
around 13,000 by 2011 (Morrison and Ross 1989; Andres et al. 2012), and remained
in the range 10,000–15,000 up to 2020 (R.I.G. Morrison, pers. comm.), the popula-
tion that spends the non-breeding season at Tierra del Fuego is considered critically
endangered in Argentina (MAyDS and Aves Argentinas 2017). The buff-breasted
sandpiper, a northern migrant that mostly spends the non-breeding season in the
Pampas region (McCarty et al. 2020), is considered an uncommon species in Bahía
Blanca. It can be found in saltmarshes dominated by chickenclaws (Sarcocornia
perennis) and short grass areas (Delhey et  al. 2001b; Delhey and Petracci 2004;
Petracci and Sotelo 2013). Its population is estimated in 15,300–56,000 mature
individuals (BirdLife International 2017) being considered endangered in Argentina
(MAyDS and Aves Argentinas 2017). Finally, Olrog’s gull breeds in Bahía Blanca;
its largest colony is located at “Islote de la Gaviota Cangrejera” with about 3500
nests (Yorio et al. 2013). Its population is estimated in 9800–15,600 mature indi-
viduals, and although it has a stable trend (BirdLife International 2018), it is subject
to different impacts within the estuary such as coastal development, unregulated
tourism, recreational fisheries, invasive alien species, and pollution (Yorio et  al.
2013, Boxes 13.1 and 13.2).

13.4.3  Conservation Efforts

Despite the aforementioned threats and species with adverse conservation status, a
positive aspect toward the conservation of Bahía Blanca Estuary and its birds is that
most of the ecosystem is legally protected (Martínez-Curci and Petracci 2016).
There are four protected areas: “Bahía Blanca” Natural Reserve under municipal
administration; “Bahía Blanca, Bahía Falsa, and Bahía Verde” and “Islote de la
gaviota cangrejera” Natural Reserves administered by the System of Natural
Protected Areas of Buenos Aires Province; and “Defensa Baterías – Charles Darwin”
Natural Reserve under the jurisdiction of national parks and the Argentine navy. As
already mentioned, the area was included among the WHSRN key sites. In addition,
“Bahía Blanca, Bahía Flasa, and Bahía Verde” Natural Reserve was recognized as
13  Shorebirds and Seabirds’ Ecology and Conservation 341

Box 13.1 Invasive Coastal Plants and the Conservation of Seabirds


Seabirds are one of the most threatened groups among birds, with almost a
third of them categorized under some degree of threat (i.e., critically endan-
gered, endangered, or vulnerable), and another 11% considered as near threat-
ened (Croxall et  al. 2012). Additionally, nearly half of seabird species are
known or suspected to be declining (Croxall et al. 2012). Additionally, and
given that most of seabirds breed on islands, the rise in sea level projected by
scientists represents a novel threat associated with the loss and retraction of
breeding habitat (Reynolds et al. 2015) and will also produce changes in the
intertidal areas where they feed.
Colonial nesting is widespread among almost all species of seabirds, with
monospecific or mixed reproductive groupings ranging from a few pairs up to
millions of birds breeding together and commonly exhibiting a strong fidelity
to natal sites. Although these high concentrations are temporary, bringing
together high densities of individuals increases their exposure to potential
threats.
Invasive exotic species (IES) are recognized as one of the principal threats
to almost half the species of seabirds, mostly represented by land mammals
that prey on their breeding colonies (Dias et al. 2019). Much less known is the
impact of invasive plants that alter the structure of the breeding habitat.
Natural processes like trampling by breeding birds, the mobilization of mate-
rials for the construction of the nest, and the eutrophication caused by the
deposition of guano produce important physicochemical changes in the soil
(Otero et al. 2018) that can open invasion windows for the establishment of
opportunistic plant species. Their propagules can arrive dispersed by the tide,
by the wind, or by the birds themselves. However, stress factors associated
with coastal habitats, mainly salinity, represent a strong environmental barrier
for the establishment of new species. That is probably the reason why there
are very few records of plants impacting seabirds (Dias et al. 2019) and also
why this relationship has been poorly explored in the literature. However, the
increase in international maritime traffic, with the arrival of increasingly
larger ships, in greater numbers and from a greater diversity of regions, sig-
nificantly increases the chances of arrival of species with the ability to estab-
lish and invade this type of habitats.
Not Just Another Gull
Located on Bahía Blanca Estuary, the “Islote de la gaviota cangrejera” Nature
Reserve (also named “Isla del Puerto,” Fig.  2.2; Chap. 2) hosts the largest
nesting site for Olrog’s gull (Larus atlanticus). The name of the reserve, in
fact, refers to this species that concentrates 70% of the total of its reproductive
population, estimated in 4860–7790 pairs, in the south-east tip of the island
(Yorio et al. 2013). This nesting site covers an area of less than 24 hectares,
and it is just 3 linear kilometers from the Port of Ingeniero White and the
associated petrochemical and industrial hub. Yorio et  al.’s (2013) review
342 N. S. Martínez-Curci et al.

indicates that industries are important source of persistent organochlorine


pesticides, hydrocarbons, and heavy metals in Bahía Blanca Estuary, where
the impact would be even greater due to its accumulation in crabs, the key
prey of Olrog’s gulls on their breeding grounds. Periodic operations of main-
tenance dredging of the channels increase this risk due to resuspension of
polluted sediments on the water (Roberts 2012). The increase in the popula-
tions of the kelp gull (Larus dominicanus), a much more opportunistic species
and capable of taking advantage of food resources associated with human
activity, including domestic waste and fishing discards, seems to be another
factor threatening Olrog’s gull due to eventual competition for breeding areas.
Lesions by sport fishing lines and deaths following the ingestion of discarded
baited hooks have also been reported for the species in its wintering grounds
(Berón and Favero 2009). Although larids are one of the largest and least
threatened groups of seabirds (Croxall et al. 2012), Olrog’s gull has a set of
characteristics that make it a vulnerable species. Its low population size,
restricted distribution range, migratory behavior, and the abovementioned
threats led to its classification as globally near threatened (BirdLife
International 2018), vulnerable in Argentina (MAyDS and Aves Argentinas
2017), and threatened in Uruguay (Azpiroz and Caballero-Sadi 2017).
A Saltmarsh Menace: A Case of Coastal Plant Invasion
Recently, a new menace was detected in the Bahía Blanca Estuary: the estab-
lishment and spread of a novel coastal plant species that colonizes the species
breeding grounds. Salsola inermis (Amaranthaceae) is an annual halophyte
plant native to the Mediterranean, reported in the region of Bahía Blanca eight
decades ago, but that apparently managed to advance on the coastal sector just
in the last few years. It grows into practically monospecific blocks that replace
native shrubs of Sarcocornia perennis and Heterostachys sp., which are lower
and typically grow more scattered, leaving bare ground spaces that disappear
with the advance of S. inermis. Invasive plants preferentially colonize areas
used by gulls for nesting, which are located in sectors slightly higher than the
surrounding environment. As the plants grow, they cover the nesting area with
densities that prevent their use by gulls. Dead and dry plants remain in the
area until the following year, making this interference even more intense
(Fig. 13.3). This change results in the movement of nesting sites toward the
perisphere, increasing nest exposure to extraordinary tides (Fig. 13.4).
Invasion of Coastal Environments: A Special Management Challenge
Managing invasive species in coastal environments offers particular chal-
lenges, including difficulties in detecting population centers and accessing
invaded areas. From this point of view, it is key to identify the vectors and
dispersal routes, as well as the environments with the greatest susceptibility to
being invaded (Zalba et al. 2000).

(continued)
13  Shorebirds and Seabirds’ Ecology and Conservation 343

Fig. 13.3  Traditional nesting areas of Olrog’s gull (Larus atlanticus) colonized by Salsola iner-
mis. Dry specimens from the previous year are observed in the foreground, while bright green
plants below and among them are young specimens from the current season. Photo by
Leandro Marbán

Fig. 13.4  Heat maps representing the spatial distribution of nests in the largest breeding colony of
Olrog’s gull (Larus atlanticus) a before and b after the Salsola inermis invasion in the “Isla del
Puerto” nature reserve. Green polygons indicate the distribution of the invasive species. The draw-
ings below the maps are a schematic representation of the invasion process and their impact.
(Modified from Marbán et al. 2019)
344 N. S. Martínez-Curci et al.

Each plant of S. inermis can produce more than 16,000 seeds (Marbán and
Zalba 2019), and, once it is mature and dry, its aerial part can detach from its
root and roll with the wind, such as many tumbleweeds of the same genus,
expanding its dispersal distance. The discontinuous distribution of this spe-
cies along the coast, with great distances between populations, and the record
of entire mature specimens floating adrift with the currents of water suggest
the hydrochory as a principal way of long-distance dispersal. Additionally,
seeds themselves are capable of staying floating up to a week and able to
germinate even after a long exposure to seawater (Marbán and Zalba 2019).
Hydrodynamic models combining information about tidal currents and the
distribution of the species in the area further support this hypothesis (Marbán,
unpublished data). Once they reach the coast, in order to germinate and suc-
cessfully establish, seeds need to reach the supralittoral zone or areas slightly
above the tidal line, where floods are sporadic or absent. In mean, just like the
sites that Olrog’s gulls use to nest.
Given the high vulnerability of the breeding environments of Olrog’s gulls
to be reached and colonized by Salsola inermis, and considering the impact
that can be projected for this process on the population dynamics and survival
of this endemic gull, in 2017 control actions of the invasive plants were initi-
ated. The area covered by the species was about 3360 square meters, and 6
control campaigns were required with a group of 12 volunteers to remove the
plants growing in the area. Manual removal of specimens took place during
the months of February and March in order to minimize interference with
Olrog’s gulls’ reproductive activities and to remove S. inermis specimens
before they release their seeds. Monitoring of the effects of this intervention
on the breeding activity of Olrog’s gulls only started recently, and no results
are available yet, but the changes in the structure of the breeding habitat are
remarkable.
Beyond the aforementioned restrictions, a set of traits of the invasive spe-
cies and the characteristics of the invasion process in the area allow us to
propose its eventual eradication: the annual life cycle of Salsola and the short
persistence of its seeds in the soil (Marbán and Zalba 2019), the ease of detec-
tion and manual removal, the early stage of the invasion in the area, and the
possibility of modeling the main directions and dispersal intensities of its
propagules according to sea currents. In the medium term, it is recommended
to maintain continuous monitoring actions and rapid control interventions to
keep the breeding area safe from invasion, and to identify and, if possible,
eradicate nearby populations that can serve as sources of propagules. If such
a strategy is not implemented, the persistence of Olrog’s Gulls, and of a set of
other shorebird species in this area, will possibly face their main obstacle.
13  Shorebirds and Seabirds’ Ecology and Conservation 345

Box 13.2 Contribution to the Study of Contamination in Shorebirds and


Seabirds from the Bahía Blanca Estuary
Anthropogenic contaminants from industrial, domestic, and agricultural run-
off are a major harmful issue for shorebirds and seabirds. Estuarine environ-
ments act as filters for the large amount of organic and inorganic compounds
that are transported by rivers, streams, underground drainage, and effluents
from terrestrial systems (Caçador et al 2012; Fu et al. 2013) toward the open
sea. Among these substances, some elements are of particular interest due to
their persistence in the environment, their biogeochemical recycling, and the
risk they pose to living beings. Regarding this, industrial and domestic dis-
charges are one of the most problematic issues since they carry several pollut-
ants, such as heavy metals, microplastics, different persistent organic
compounds (OCs, OPs, PAHs), and other emerging contaminants (Arias et al.
2010; Spetter et al. 2015; Fernández Severini et al. 2019; Villagrán et al. 2019).
As consumers, shorebirds and seabirds are found at most trophic levels of
the food web playing a key role in marine and estuarine ecosystems. Because
they are long-lived and are often at the top of their food chains, they are par-
ticularly vulnerable to a wide range of contaminants associated with the bio-
accumulation of pollutants throughout their whole life (Burger 1993).
Estuarine sediments constitute the most important sink of pollutants in
aquatic ecosystems. In this sense, shorebirds and seabirds may be particularly
vulnerable to these contaminants as they forage primarily on benthic inverte-
brates (White et al. 1993; Braune and Noble 2009).
Trace Metals in Shorebirds and Seabirds from the Bahía Blanca Estuary
Trace metals, although they may have an anthropogenic origin, are also natu-
ral components in the environment and can be found in terrestrial and aquatic
systems, as well as in the atmosphere. Shorebirds and seabirds are exposed to
different nonessential (and thus potentially toxic beyond a threshold) ele-
ments such as silver (Ag), cadmium (Cd), lead (Pb), and mercury (Hg) and
also essential elements that could be toxic at high levels such as arsenic (As),
cobalt (Co), chromium (Cr), copper (Cu), iron (Fe), manganese (Mn), nickel
(Ni), selenium (Se), and zinc (Zn) (Lucia et al. 2014).
Although the bioavailability of metals is associated with physicochemical
factors such as pH, salinity, redox potential, organic matter content, and sedi-
ment grain size (Eggleton and Thomas 2004; Du Laing et al. 2009; Acosta
et al. 2011; Zhao et al. 2013), bioaccumulation is also strongly influenced by
the various accumulation strategies of aquatic organisms. For birds, there are
several factors that influence uptake, accumulation, and biomagnification of
metals, including exposure pathways, species of the metal, and bioavailabil-
ity, as well as a number of host factors, such as trophic status, location, forag-
ing behavior, nutrition, body condition, gender, size, genetic variability, and
age (Burger et al. 2008).
346 N. S. Martínez-Curci et al.

Shorebirds and seabirds acquire metals both through water, which they
drink, and through the diet, which in most cases includes marine invertebrates.
Once the metals enter the bird’s body, they can be accumulated in the tissues
or eliminated by deposition through the feathers during periods of rapid
growth (after molting), by the uropygeal gland and by the salt gland (Burger
and Gochfeld 1996a, b; Burger et al. 1993, 2000). In turn, females can seques-
ter some metals such as Mn, Hg, Cr, Pb, and Cd in egg content and eggshells,
which could put embryonic development at risk.
There is a rich worldwide literature on heavy metals in shorebirds and
seabirds. These works can be categorized according to whether they are labo-
ratory studies, measurements of residues in sick or dead birds, surveys of
metal levels in a given species, and finally, biomonitoring studies, which have
been increasing over the years given that waterbirds are valuable for environ-
mental monitoring because they are long-lived predators and are capable of
integrating pollutant levels over a large area by bioaccumulation (Furness
1993; Burger and Gochfeld 2001; Zhang and Ma 2011; Lucia et  al. 2014;
Saalfeld et al. 2016; Dias et al. 2019).
Although there is an extensive database on the levels of heavy metals in the
Bahía Blanca Estuary (see Chap. 4), there are no sufficient studies of heavy
metals carried out specifically on the marine and coastal birds in this wetland.
The American oystercatcher (Haematopus palliatus; Fig. 13.5) is a resi-
dent shorebird that reproduces and feeds in the Bahía Blanca Estuary (Delhey
and Petracci 2004; Simonetti 2012). The foraging behavior of this species has
been described on a variety of intertidal benthic prey such as crabs, clams,
oysters, limpets, mussels, and polychaetes (Goss-Custard and Durell 1988;
Nol and Humphrey 1994; Bachmann and Martinez 1999; Daleo et al. 2005).
One of its main prey items in the Bahía Blanca Estuary is the burrowing crab
(Neohelice granulata), which is considered a key species within this ecosys-
tem, given that it could play a major role in the transference of pollutants to
higher trophic levels. Previous studies on this benthic invertebrate in the Bahía
Blanca Estuary showed low to medium concentrations of several trace metals
(Ferrer 2001; Beltrame et al. 2009, 2011; Simonetti et al. 2012, 2013, 2018).
Both eggs and eggshells of shorebirds and seabirds have been used in vari-
ous studies on levels of heavy metals because they have certain advantages
over other tissues. Sampling eggs takes less time, is easy to handle, can be
collected with little interference, and its extraction puts less pressure on a
population than that of adults or juveniles, especially if only one egg is
removed from a given clutch (Dev and Bhattacharjee 2010).
As previously mentioned, females can sequester several metals in eggs and
eggshells (Burger and Gotchfeld 2002; Lam et al. 2005; Barata et al. 2010;
Dev and Bhattacharjee 2010; Hashmi et al. 2013), where metal concentrations
represent recent exposure as well as mobilization from stored materials in
females at the time of egg formation (Burger and Gochfeld 1995; Burger and
Gochfeld 1996a, b).

(continued)
13  Shorebirds and Seabirds’ Ecology and Conservation 347

Fig. 13.5 (a) American oystercatcher (Haematopus palliatus), on the tidal plain of the
Bahia Blanca Estuary, (b) clutch during the reproductive season (Photos by Natalia
S. Martínez-Curci (a) Pia Simonetti (b))

One of the few studies made on trace metals in shorebirds from the Bahía
Blanca Estuary was the work of Simonetti et al. (2015). In this work, concen-
tration of Cu, Cd, Cr, Pb, Zn, and Ni were determined in eggshells of American
oystercatchers. Only infertile eggs (i.e., not hatched) and eggs from abandoned
nests were selected to avoid affecting the reproductive success. All metals
showed detected levels, with exceptionally high Cd concentrations (Table 13.2).
Cu, Cr, Zn, and Ni are four essential metals which play an important role in
the metabolism of birds. The values obtained in this study were within the val-
ues described in the literature for different species of birds like egrets, herons,
albatrosses, and passerines (Dauwe et al. 2005; Ikemoto et al. 2005; Dev and
Bhattacharjee 2010; Al-Obaidi et al. 2012; Hashmi et al. 2015; Dolci et al. 2017).
Regarding Cd and Pb, two nonessential metals, although the concentra-
tions obtained for Pb were within the values described in the literature, the Cd
348 N. S. Martínez-Curci et al.

Table 13.2  Trace metal concentrations (μg/g dry weight) in eggshells of American
oystercatcher (Haematopus palliatus), from the Bahía Blanca Estuary
Cu Cd Cr Ni Zn Pb
Range 0.52–2.63 1.70–14.87 0.03–0.84 0.89–7.59 1.02–3.86 2.13–9.79
Std 0.52 3.38 0.03 0.89 1.13 2.33
Median 1.97 11.62 0.78 6.06 2.08 8.14

values found in this study were surprisingly higher (Ayas 2007; Dev and
Bhattacharjee 2010; Hashmi et  al. 2013; Avazpour et  al. 2014; Dolci et  al.
2017; van Aswegen et al. 2019).
La Sala et al. (2011) studied the levels of Hg in feathers in adults and live
and dead chicks of Olrog’s gull (Larus atlanticus), from the Bahia Blanca
Estuary. As for eggs, feathers are another pathway for elimination of metals
that are accumulated during the inter-molt periods (Monteiro and Furness
1995). Moreover, they are useful indicators of heavy metal contamination
because birds sequester heavy metals in their feathers, and the proportion of
the body burden that is in feathers is relatively constant for any metal (Burger
1993). The results of this study showed detectable concentrations of Hg in
both adults and chicks. For breeding adults Hg in feathers ranged between
2.20 and 16.97 μg/g dry weight. According to Burger and Gotchfeld (2002),
these values largely fall within the range that is considered toxic, and there is
a potential risk about long-term effects of Hg pollution on this population. The
values recorded for feathers of dead and living chicks were lower than adults,
ranging between and 1.12–3.00  μg/g dry weight for dead chicks and
0.77–3.02  μg/g dry weight for living chicks. These differences could be
explained by (1) a fairly simultaneous growth of all feather types in pre-fledg-
lings, (2) Hg concentrations in adult feathers reflecting metal input throughout
a vast geographical area and a substantially longer accumulation period, and
(3) a wider array of prey items during the non-breeding season which might
be exposing the adults to higher Hg concentrations from other food resources
such as fish discards (La Sala et al. 2011).
During the chick-rearing period, Olrog’s gull adults prey mainly on the
crabs N. granulata and Cyrtograpsus angulatus, which they feed to their
young. Although there are no data on the concentration of Hg in any of these
invertebrates, as previously mentioned, there is vast information about the
bioaccumulation of heavy metals in N. granulata, which suggests the possi-
bility that this metal is also bioaccumulated in these organisms, thus making
Olrog’s gull adults, and especially small chicks, susceptible to the effects of
Hg (La Sala et al. 2011).
From these results, added to the state of knowledge of the presence of trace
metals in one of the prey items for oystercatchers and gulls, metals that are
13  Shorebirds and Seabirds’ Ecology and Conservation 349

bioavailable in the environment may also be transferred to other seabirds and


shorebird species through food chains. Given the toxicity of these elements, it
seems a potential risk in these species, since many of these metals can have
effects on the development and nervous system of birds.
Microplastics and Other Contaminants
Plastic pollution in marine and estuarine environments is a worldwide con-
cern and has been studied since the 1970s. The durability of the plastic implies
that it is retained for years or centuries, while if it is not exposed to bacterial
activity or UV radiation, degradation does not occur (Lambert et al. 2014).
Large plastic debris, called “macroplastics,” generate a significant environ-
mental impact due to the injury and death of seabirds, mammals, fish, and
reptiles as a result of plastic entanglement and ingestion (Derraik 2002;
Gregory 2009; Lozano and Mouat 2009). Marine animals often mistake these
fragments for food by ingesting them, which can cause damage to internal
organs, digestive tract blockage, and ulcers among other affections (Blais
et al. 2005; Gregory 2009).
In seabirds, the effect of plastic ingestion has been a cause of concern pri-
marily for two reasons, firstly because of the frequency with which seabirds
ingest plastic and secondly because of emerging evidence of impacts on body
condition and transmission of toxic chemicals, which could lead to changes in
mortality or reproduction (Wilcox et al. 2015). In recent years, “microplas-
tics,” referred as tiny plastic particles with a diameter smaller than 5  mm
(Thompson et al. 2004), have become a subject of great concern at the envi-
ronmental level. Because of their composition and relatively large surface
area, microplastic fragments can be vectors for the diffusion of organic con-
taminants. Ingestion of microplastics may therefore be introducing toxins to
the base of the food chain, from where there is potential for bioaccumulation
(Teuten et al. 2009).
Again, although there are a considerable number of studies on plastic pol-
lution in seabirds worldwide, to date there are no published studies for shore-
birds and seabirds in the Bahía Blanca Estuary. The only registry available and
whose investigation is currently under study is that of La Sala (pers. comm.),
where plastic debris were detected in the stomachs of chicks of Olrog’s gull.
Considering the limited information available as well as the possible
adverse effects of these pollutants on shorebirds and seabirds, a comprehen-
sive long-term research covering relevant disciplines is necessary to assess the
effects of this and other pollutants on the health of populations of these spe-
cies in the Bahía Blanca Estuary.
350 N. S. Martínez-Curci et al.

one of the Important Bird and Biodiversity Areas (IBA-BA15) in Argentina (Di
Giacomo 2005), it was included within the Valuable Grassland Areas (VGAs;
Bilenca and Miñarro 2004), and it is also one of the high priority areas for the con-
servation of Nearctic migratory birds in the southern cone grasslands of South
America (Di Giacomo and Parera 2008).
Remarkably, migratory shorebirds and the red knot were declared emblematic
species of Bahía Blanca and Coronel Rosales counties, respectively. The legal
protection of the ecosystem, as well as the recognition of its importance by interna-
tional and local communities, show signs of hope for Bahía Blanca Estuary to
remain a healthy ecosystem in the long term and thus to continue supporting viable
shorebirds and seabirds’ populations.

References

Acosta JA, Jansen B, Kalbitz K et al (2011) Salinity increases mobility of heavy metals in soils.
Chemosphere 85:1318–1324. https://doi.org/10.1016/j.chemosphere.2011.07.046
Al-Obaidi FA, Mehdi BA, Al-Shadeedi SM (2012) Identification of inorganic elements in egg shell
of some wild birds in Baghdad. Adv Appl Sci Res 3(3):1454–1458
Andres BA, Smith PA, Morrison RIG et al (2012) Population estimates of North American shore-
birds. Wader Study Group Bull 119:178–194
Arias AH, Fernández Severini MD, Delucchi F et al (2010) Persistent pollutants monitoring in a
South Atlantic coastal environment. Case study: the Bahía Blanca Estuary. In: Ortiz AC, Griffin
NB (eds) Pollution monitoring. Nova Science Publishers, New York, pp 1–30
Avazpour FG, Bakhtiari AR, Sajjadi MM (2014) Levels of vanadium in tern eggs and sediments of
Shidvar and Bani Farour Islands, Persian gulf. Podoces 9(1)
Ayas Z (2007) Trace element residues in eggshells of grey heron (Ardea cinerea) and black-­
crowned night heron (Nycticorax nycticorax) from Nallihan bird paradise, Ankara-Turkey.
Ecotoxicology 16:347–352
Azpiroz A, Caballero-Sadi D (2017) Gaviota cangrejera (Larus atlanticus). In: Azpiroz AB,
Jiménez S, Alfaro, M (eds.) Libro Rojo de las Aves del Uruguay. Biología y conservación de
las aves en peligro de extinción a nivel nacional. Categorías “Extinto a Nivel Regional”, “En
Peligro Crítico” y “En Peligro”. DINAMA y DINARA, Montevideo.
Bachmann S, Martinez MM (1999) Feeding tactics of the American Oystercatcher (Haematopus
palliatus) on Mar Chiquita Coastal Lagoon, Argentina. Ornitol Neotrop 10:81–84
Baker A, González P, Morrison RIG, Harrington BA (2020) Red Knot (Calidris canutus), ver-
sion 1.0. In: Billerman SM (ed) Birds of the world. Cornell Lab of Ornithology, New York.
Available via 10.2173/bow.redkno.01. Accessed 8 May 2020
Barata C, Fabregat MC, Cotín J et al (2010) Blood biomarkers and contaminant levels in feathers
and eggs to assess environmental hazards in heron nestlings from impacted sites in Ebro basin
(NE Spain). Environ Pollut 158(3):704–710
Barbier EB, Hacker SD, Kennedy C et al (2011) The value of estuarine and coastal ecosystem
services. Ecol Monogr 81:169–193
Barbosa A, Moreno E (1999) Evolution of foraging strategies in shorebirds: an ecomorphological
approach. Auk 116:712–725
Belenguer C, Delhey K, Di Martino S et al (1992) Observaciones de aves playeras migratorias de
Bahia Blanca. Boletín Informativo Grupo Argentino de Limícolas Nro 10
Beltrame MO, De Marco SG, Marcovecchio JE (2009) Influences of sex, habitat, and seasonality
on heavy-metal concentrations in the burrowing crab (Neohelice Granulata) from a coastal
lagoon in Argentina. Arch Environ Contam Toxicol 58(3):746–756
13  Shorebirds and Seabirds’ Ecology and Conservation 351

Beltrame MO, De Marco SG, Marcovecchio JE (2011) The burrowing crab Neohelice granulata as
potential bioindicator of heavy metals in estuarine systems of the Atlantic Coast of Argentina.
Environ Monit Assess 172(1–4):379–389
Berón MP, Favero M (2009) Mortality and injuries of Olrog’s Gull (Larus atlanticus) individuals
associated with sport fishing activities in Mar Chiquita coastal lagoon, Buenos Aires Province.
Hornero 24(2):99–102
Berón MP, Seco Pon JP, García GO et al (2013) The diet of Olrog’s Gull (Larus atlanticus) reveals
an association with fisheries during the non-breeding season. Emu 113:69–76
Bilenca D, Miñarro F (2004) Identificación de Áreas Valiosas de Pastizal (AVPs) en las Pampas
y Campos de Argentina, Uruguay y sur de Brasil. Fundación Vida Silvestre Argentina,
Buenos Aires
BirdLife International (2016) Pluvianellus socialis. In: The IUCN Red List of Threatened Species.
Available via https://doi.org/10.2305/IUCN.UK.2016-­3.RLTS.T22693570A93413261.en.
Accessed 8 May 2020
BirdLife International (2017) Calidris subruficollis. In: The IUCN Red List of Threatened Species.
Available via https://doi.org/10.2305/IUCN.UK.2017-­1.RLTS.T22693447A111804064.
Accessed 8 May 2020
BirdLife International (2018) Larus atlanticus. In: The IUCN Red List of Threatened Species.
Available via https://doi.org/10.2305/IUCN.UK.2018-­2.RLTS.T22694286A132538305.en.
Accessed 8 May 2020
Blais JM, Kimpe LE, McMahon D et al (2005) Arctic seabirds transport marine-derived contami-
nants. Science 309(5733):445–445
Boere GC, Stroud DA (2006) The flyway concept: what it is and what it isn’t. In: Boere GC,
Galbraith CA, Stroud DA (eds) Waterbirds around the world. The stationery office, Edinburgh,
pp 40–47
Braune BM, Noble DG (2009) Environmental contaminants in Canadian shorebirds. Environ
Monit Assess 148(1–4):185–204
Brockmann HJ, Barnard CJ (1979) Kleptoparasitism in birds. Anim Behav 27:487–514
Burger J (1993) Metals in avian feathers: bioindicators of environmental pollution. Rev Environ
Toxicol 5:203–311
Burger J, Gochfeld M (1995) Heavy metal and selenium concentrations in eggs of herring gulls
(Larus argentatus): temporal differences from 1989 to 1994. Arch Environ Contam Toxicol
29(2):192–197
Burger J, Gochfeld M (1996a) Family Laridae (gulls). In: del Hoyo J, Elliott A, Sargatal J (eds)
Handbook of the birds of the world, vol 3. Lynx Edicions, Barcelona, pp 572–623
Burger J, Gochfeld M (1996b) Heavy metal and selenium levels in Franklin's gull (Larus pipixcan)
parents and their eggs. Arch Environ Contam Toxicol 30(4):487–491
Burger J, Gochfeld M (2001) On developing bioindicators for human and ecological health.
Environ Monit Assess 66(1):23–46
Burger J, Gotchfeld M (2002) Effects of chemicals and pollution on seabirds. In: Schreiber EA,
Burger J (eds) Biology of marine birds. CRC Press, Boca Raton, pp 485–525
Burger J, Olla BL (eds) (1984) Shorebirds: migration and foraging behavior. Plenum Press,
New York, pp 73–123
Burger J, Rodgers JA, Gochfeld M (1993) Heavy metal and selenium levels in endangered wood
storks Mycteria americana from nesting colonies in Florida and Costa Rica. Arch Environ
Contam Toxicol 24(4):417–420
Burger J, Trivedi CD, Gochfeld M (2000) Metals in herring and great black-backed gulls from
the New York bight: the role of salt gland in excretion. Environ Monit Assess 64(3):569–581
Burger J, Gochfeld M, Jeitner C et al (2008) Assessment of metals in down feathers of female com-
mon eiders and their eggs from the Aleutians: arsenic, cadmium, chromium, lead, manganese,
mercury, and selenium. Environ Monit Assess 143(1–3):247
Burger J, Gochfeld M, Garcia EFJ et al (2020) Kelp Gull (Larus dominicanus), version 1.0. In: del
Hoyo J, Elliott A, Sargatal J, Christie DA, de Juana E (eds) Birds of the world. Cornell Lab of
Ornithology, New York. Available via 10.2173/bow.kelgul.01. Accessed 8 May 2020
352 N. S. Martínez-Curci et al.

Cabot D, Nisbet I (2013) Tern. Harper Collins, London


Caçador I, Costa JL, Duarte B et al (2012) Macroinvertebrates and fishes as biomonitors of heavy
metal concentration in the Seixal Bay (Tagus estuary): which species perform better? Ecol
Indic 19:184–190
Castano MV, Biondi LM, Zumpano F et al. (2020). Behavioral responses to a novel feeding prob-
lem in the Olrog’s Gull Larus altanticus. Mar Ornithol (in press)
Ceballos G, Ehrlich PR, Barnosky AD et  al (2015) Accelerated modern human–induced spe-
cies losses: entering the sixth mass extinction. Sci Adv 1:e1400253. https://doi.org/10.1126/
sciadv.1400253
Colwell MA (2010) Shorebird ecology, conservation, and management. University of California
Press, California
Colwell MA, Haigh SM (2019) An overview of the world’s plovers. In: Colwell MA, Haigh SM
(eds) The population ecology and conservation of Charadrius plovers. CRC Press Taylor/
Francis Group, London, pp 3–15
Copello S, Favero M (2001) Foraging ecology of Olrog’s Gull Larus atlanticus in Mar Chiquita
Lagoon (Buenos Aires, Argentina): are there age-related differences? Bird Conserv Int
11:175–188
Copello S, Suárez N, Yorio P et al (2020) Distribution of Olrog’s Gull Larus atlanticus from Bahía
San Blas during the non-breeding period: signals of partial migration. Bird Conserv Int:1–10.
https://doi.org/10.1017/S0959270920000234
Croxall JP, Butchart SHM, Lascelles B et al (2012) Seabird conservation status, threats and prior-
ity actions: a global assessment. Bird Conserv Int 22:1–34
Daleo P, Escapa M, Isacch JP et al (2005) Trophic facilitation by the oystercatcher Haematopus
palliatus Temminick on the scavenger snail Buccinanops globulosum Kiener in a Patagonian
bay. J Exp Mar Biol Ecol 325(1):27–34
Dauwe T, Janssens E, Bervoets L et al (2005) Heavy metal concentrations in female laying great
tits (Parus major) and their clutches. Arch Environ Contam Toxicol 49:249–256
Delhey, JKV, Petracci, PF (2004) Aves marinas y costeras. In: Ecosistema del estuario de Bahía
Blanca. Piccolo MC, Hoffmeyer M (eds). Edi UNS, pp 203–220
Delhey JKV, Carrete M, Martínez M (2001a) Diet and feeding behaviour of Olrog’s Gull Larus
atlanticus in Bahía Blanca, Argentina. Ardea 89:319–329
Delhey K, Petracci PF, Pérez CHF (2001b) Observaciones de algunos Charadriiformes en el sur de
la provincia de Buenos Aires, Argentina. Nuestras Aves 42:14–16
Derraik JGB (2002) The pollution of the marine environment by plastic debris: a review. Mar
Pollut Bull 44:842–852
Dev B, Bhattacharjee PC (2010) Heavy metals in egg shells of six species of Ardeidae (Aves) from
Barak Valley, Assam. ADBU J Eng Technol 5(1):48–52
Di Giacomo AS (2005) Áreas importantes para la conservación de las aves en Argentina. Sitios
prioritarios para la conservación de la biodiversidad. Aves Argentinas/Asociación Ornitológica
del Plata, Buenos Aires. Available via https://www.avesargentinas.org.ar/aica. Accessed 8
May 2020
Di Giacomo AS, Parera AF (2008) 20 Áreas prioritarias para la conservación de las aves migra-
torias neárticas en los pastizales del cono sur de Sudamérica. Aves Argentinas / Asociación
Ornitológica del Plata, Buenos Aires
Dias MP, Martin R, Pearmain EJ et  al (2019) Threats to seabirds: a global assessment. Biol
Conserv 237:525–537
Dirzo R, Young HS, Galetti M et al (2014) Defaunation in the Anthropocene. Science 345:401–406.
https://doi.org/10.1126/science.1251817
Dolci NN, Sá F, da Costa ME et al (2017) Trace elements in feathers and eggshells of brown booby
Sula leucogaster in the Marine National Park of Currais Islands, Brazil. Environ Monit Assess
189(10):496
13  Shorebirds and Seabirds’ Ecology and Conservation 353

Du Laing G, Rinklebe J, Vandecasteele B et  al (2009) Trace metal behaviour in estuarine and
riverine floodplain soils and sediments: a review. Sci Tot Environ 407:3972–3985. https://doi.
org/10.1016/j.scitotenv.2008.07.025
Eggleton J, Thomas KV (2004) A review of factors affecting the release and bioavailability of
contaminants during sediment disturbance events. Environ Int 30:973–980. https://doi.
org/10.1016/j.envint.2004.03.001
Elner RW, Beninger PG, Jackson DL et al (2005) Evidence of a new feeding mode in western
sandpiper (Calidris mauri) and dunlin (Calidris alpina) based on bill and tongue morphology
and ultrastructure. Mar Biol 146:1223–1234
Escalante R (1966) Notes on the Uruguayan population of Larus belcheri. Condor 68:507–510
Escapa M, Isacch JP, Daleo P et  al (2004) The distribution and ecological effects of the intro-
duced Pacific Oyster Crassostrea gigas (Thunberg, 1973) in Northern Patagonia. J Shellfish
Res 23:765–772
Fernández Severini MD, Villagrán DM, Buzzi NS et al (2019) Microplastics in oysters (Crassostrea
gigas) and water at the Bahía Blanca Estuary (Southwestern Atlantic): an emerging issue of
global concern. Reg Stud Mar Sci 32:100829. https://doi.org/10.1016/j.rsma.2019.100829
Ferrer LD (2001) Estudio de los diversos metales pesados en sedimentos del estuario de Bahia
Blanca y sus efectos toxicos sobre el cangrejo Chasmagnathus granulata. Doctoral thesis,
Universidad Nacional del Sur, Argentina
Fu J, Tang XL, Zhang J et al (2013) Estuarine modification of dissolved and particulate trace met-
als in major rivers of East-Hainan, China. Cont Shelf Res 57:59–72
Furness RW (1987) Kleptoparasitism in seabirds. In: Croxal JP (ed) Seabirds, feeding biology and
role in marine ecosystem. Cambridge University Press, Cambridge, pp 77–99
Furness RW (1993) Birds as monitors of pollutants. In: Birds as monitors of environmental change.
Springer, Dordrecht, pp 86–143
Furness RW, Edwards AE, Oro D (2007) Influence of management practices and of scaveng-
ing seabirds on availability of fisheries discards to benthic scavengers. Mar Ecol Prog Ser
350:235–244. https://doi.org/10.3354/meps07191
García GO, Favero M, Vassallo AI (2010) Factors affecting kleptoparasitism by gulls in a multi-­
species seabird colony. Condor 112:521–529
García GO, Becker PH, Favero M (2011) Kleptoparasitism during courtship in Sterna hirundo and
its relationship with female reproductive performance. J Ornithol 152:103–110
García GO, Favero M, Becker PH (2013) Intraspecific kleptoparasitism improves chick growth
and reproductive output in common terns Sterna hirundo. Ibis 155:338–347
García GO, Nicoll A, Castano M et al (2019) Evaluation of neophobia in a threatened seabird:
Olrog’s Gull (Larus atlanticus) as a case study. Emu 119:166–175
García GO, Paterlini CA, Favero M et al (2020) Age-, sex- and tactic-specific kleptoparasitic per-
formance in a long-lived seabird. J Ornithol 161:183–188
Gochfeld M, Burger J, de Juana E et al (2020) South American tern (Sterna hirundinacea), ver-
sion 1.0. In: del Hoyo J, Elliott A, Sargatal J et al (eds) Birds of the world. Cornell Lab of
Ornithology, New York. Available via 10.2173/bow.soater1.01. Accessed 8 May 2020
Goss-Custard JD, Durell SLVD (1988) The effect of dominance and feeding method on the intake
rates of oystercatchers, Haematopus ostralegus, feeding on mussels. J Anim Ecol:827–844
Gregory MR (2009) Environmental implications of plastic debris in marine settings: entangle-
ment, ingestion, smothering, hangers-on, hitch-hiking and alien invasions. Philos T R Soc B
364:2013–2025
Hackett SJ, Kimball RT, Reddy S et al (2008) A phylogenomic study of birds reveals their evolu-
tionary history. Science 320:1763–1768
Hashmi MZ, Malik RN, Shahbaz M (2013) Heavy metals in eggshells of cattle egret (Bubulcus
ibis) and little egret (Egretta garzetta) from the Punjab province, Pakistan. Ecotox Environ
Safe 89:158–165
Hashmi MZ, Abbasi NA, Tang X et al (2015) Egg as a biomonitor of heavy metals in soil. In:
Heavy metal contamination of soils. Springer, Cham, pp 127–143
354 N. S. Martínez-Curci et al.

Herrera G, Punta G, Yorio P (2005) Diet specialization of the threatened Olrog’s gull Larus atlan-
ticus during the breeding season at Golfo San Jorge, Argentina. Bird Conserv Int 15:89–97
Hoekstra JM, Boucher TM, Ricketts TH et al (2005) Confronting a biome crisis: global disparities of
habitat loss and protection. Ecol Lett 8:23–29. https://doi.org/10.1111/j.1461-­0248.2004.00686.x
Ieno E, Alemany D, Blanco DE et al (2004) Prey size selection by red knots feeding on mud snails
at Punta rasa (Argentina) during migration. Waterbirds 27:493–498
Ikemoto T, Kunito T, Tanabe S et al (2005) Non-destructive monitoring of trace element levels
in short-tailed albatrosses (Phoebastria albatrus) and black-footed albatrosses (Phoebastria
nigripes) from Torishima Island, Japan, using eggs and blood. Mar Pollut Bull 51:889–895
Iribarne OO, Martínez MM (1999) Predation on the southwestern Atlantic fiddler crab (Uca uru-
guayensis) by migratory shorebirds (Pluvialis dominica, P. squatarolla, Arenaria interpres and
Numenius phaeopus). Estuaries 22:47–54
IUCN (2020) The IUCN Red List of Threatened Species. Version 2020–1. Available via https://
www.iucnredlist.org. Accessed 8 May 2020
Jenni L, Jenni-Eiermann S (1998) Fuel supply and metabolic constraints in migrating birds. J
Avian Biol 29:521–528
Jenni-Eiermann S, Jenni L, Kvist A et al (2002) Fuel use and metabolic response to endurance
exercise: a wind tunnel study of a long-distance migrant shorebird. J Exp Biol 205:2453–2460
Kelly JP, Evens JG, Stallcup RW et al (1996) The effects of aquaculture on habitat use by wintering
shorebirds. California Fish and Game 82:160–174
Kubelka V, Šálek M, Tomkovich P et al (2018) Global pattern of nest predation is disrupted by
climate change in shorebirds. Science 362:680–683. https://doi.org/10.1126/science.aat8695
La Sala LF, Petracci PF, Smits JE et al (2011) Mercury levels and health parameters in the threat-
ened Olrog’s Gull (Larus atlanticus) from Argentina. Environ Monit Asses 181(1–4):1–11
Lack D (1968) Bird migration and natural selection. Oikos 19:1–9
Lam JC, Tanabe S, Lam MH et al (2005) Risk to breeding success of waterbirds by contaminants
in Hong Kong: evidence from trace elements in eggs. Environ Pollut 135(3):481–490
Lambert S, Sinclair C, Boxall A (2014) Occurrence, degradation, and effect of polymer based
materials in the environment. Rev Environ Contam Toxicol 227(227):1–53
Lascelles BG, Taylor PR, Miller MGR et al (2016) Applying global criteria to tracking data to
define important areas for marine conservation. Divers Distrib 22:422–431
Lindström A, Piersma T (1993) Mass changes in migrating birds: the evidence for fat and protein
storage re-examined. Ibis 135:70–78
Lisnizer N, García Borboroglu P, Yorio P (2011) Spatial and temporal variation in population
trends of kelp gulls in northern Patagonia, Argentina. Emu 111:259–267
Lovette IJ (2016) Avian diversity and classification. In: Lovette IJ, Fitzpatrick JW (eds) The
Cornell lab of ornithology handbook of bird biology, 3rd edn. Wiley, West Sussex, pp 7–61
Lozano RL, Mouat J (2009) Marine litter in the North-East Atlantic region: Assessment and priori-
ties for response KIMO international
Lucia M, Bocher P, Chambosse M et al (2014) Trace element accumulation in relation to trophic
niches of shorebirds using intertidal mudflats. J Sea Res 92:134–143
Luckenbach MW (1984) Biogenic structure and foraging by five species of shorebirds (Charadrii).
Estuar Coast Shelf S 196:691–696
Marbán LM, Petracci P, Sotelo M, Zalba S (2019) Plantas vs. gaviotas: colonizando una colonia.
Un problema de invasión. In: Libro de resúmenes de la XVIII.
Marbán LM, Zalba SM (2019) When the seeds go floating in: A salt marsh invasion. Estuar Coast
Shelf S 231:106442. https://doi.org/10.1016/j.ecss.2019.106442.
Marinao C, Suárez N, Yorio P (2019) Trophic interactions between the Kelp Gull (Larus dominica-
nus) and royal and Cayenne terns (Thalasseus maximus maximus and Thalasseus sandvicensis
eurygnathus, respectively) in a human-modified environment. Can J Zool 97:904–913. https://
doi.org/10.1139/cjz-­2019-­0047
Martínez-Curci NS, Petracci P (2016) Aves playeras del litoral costero de la provincia de Buenos
Aires: ecología y conservación. In: Athor J, Celsi CE (eds) La costa atlántica de Buenos Aires:
13  Shorebirds and Seabirds’ Ecology and Conservation 355

naturaleza y patrimonio cultural. Fundación de Historia Natural Felix de Azara, Buenos Aires,
pp 204–233
Martínez-Curci NS, Azpiroz AB, Isacch JP et al (2015) Dietary relationships among Nearctic and
Neotropical migratory shorebirds in a key coastal wetland of South America. Emu 115:326–334
MAyDS (Ministerio de Ambiente y Desarrollo Sostenible), Aves Argentinas (2017) Categorización
de las Aves de la Argentina según su estado de conservación. Available via https://avesar-
gentinas.org.ar/sites/default/files/Categorizacion-­de-­aves-­de-­la-­Argentina.pdf. Accessed 8
May 2020
MAyDS, AA (Ministerio de Ambiente y Desarrollo Sustentable and Aves Argentinas) (2017).
Categorización de las Aves de la Argentina (2015). Informe del Ministerio de Ambiente y
Desarrollo Sustentable de la Nación y de Aves Argentinas, edición electrónica. Buenos Aires,
Argentina
McCarty JP, Wolfenbarger LL, Laredo CD, Pyle P, Lanctot RB (2020) Buff-breasted sandpiper
(Calidris subruficollis), version 1.0. In: Rodewald PG (ed) Birds of the world. Cornell Lab of
Ornithology, Ithaca, New York. Available via 10.2173/bow.bubsan.01. Accessed 8 May 2020
Monteiro LR, Furness RW (1995) Seabirds as monitors of mercury in the marine environment.
Water Air Soil Pollut 80:851–870
Morrison RIG, Ross RK (1989) Atlas of Nearctic shorebirds on the coast of South America.
Canadian Wildlife Service, Ottawa
Myers JP (1983) Conservation of migrating shorebirds: staging areas, geographic bottlenecks, and
regional movements. Am Birds 37:23–25
Niles LJ, Burger J, Porter RR et al (2010) First results using light level geolocators to track red
knots in the Western hemisphere show rapid and long intercontinental flights and new details
of migration pathways. Wader Study Group Bull 117:123–130
Nol E, Humphrey RC (1994) American Oystercatcher (Haematopus palliatus) In: Poolea and Gill
(eds) The birds of North America. Academy of Natural Sciences and American Ornithologists’
Union, Filadelfia y Washington DC, pp 1–24
Orians GH, Pearson NE (1979) On the theory of central place foraging. In: Horn DJ, Mitchell
RD, Stairs GR (eds) Analysis of ecological systems. Ohio State University Press, Columbus,
pp 154–177
Oro D, Genovart M, Tavecchia G et al (2013) Ecological and evolutionary implications of food
subsidies from humans. Ecol Lett 16:1501–1514. https://doi.org/10.1111/ele.12187
Otero XL, De la Peña-Lastra S, Pérez Alberti A et al (2018) Seabird colonies as important global
drivers in the nitrogen and phosphorus cycles. Nat Commun 9:246. https://doi.org/10.1038/
s41467-017-02446-8.
Parmelee DF (2020) White-rumped sandpiper (Calidris fuscicollis), version 1.0. In: Poole
AF, Stettenheim PR, Gill FB (eds) Birds of the world. Cornell Lab of Ornithology, Ithaca,
New  York. Available via https://doi-­org.proxy.library.cornell.edu/10.2173/bow.whrsan.01.
Accessed 8 May 2020
Paton TA, Baker AJ, Growth JG et al (2003) RAG-1 sequences resolve phylogenetic relationships
within Charadriiform birds. Mol Phylogenet Evol 29:268–278
Paz JA, Seco Pon JP, Favero M et  al (2018). Seabird interactions and by-catch in the anchovy
pelagic trawl fishery operating in northern Argentina. Aquat Conserv. https://doi.org/10.1002/
aqc.2907
Petracci PF (2002) Diet of sanderling in Buenos Aires province, Argentina. Waterbirds 25:366–370
Petracci, PF, Delhey JKV (2005) Guía de las aves marinas y costeras de la ría de Bahía Blanca. In:
Petracci PF, Delhey JKV (eds). Bahía Blanca
Petracci PF, Sotelo M (2013) Aves del Estuario de Bahía Blanca: Una herramienta para su cono-
cimiento y conservacion. Grupo Editorial Muelle Sur, Bahía Blanca
Petracci PF, La Sala L, Aguerre G et al (2004) Dieta de la gaviota cocinera (Larus dominicanus)
durante el período reproductivo en el Estuario de Bahía Blanca, Buenos Aires, Argentina. El
Hornero 19(1):23–28
356 N. S. Martínez-Curci et al.

Petracci PF, Delhey KV, Sotelo MR (2007) Hábitos granívoros en la Gaviota Cangrejera (Larus
atlanticus): implicancias sobre su estatus de especialista. Hornero 22:51–54
Petracci PF, Sotelo MR, Díaz L (2008) Nuevo registro de nidificación de la gaviota cangrejera
(Larus atlanticus) en la Reserva Natural Bahía Blanca, Bahía Falsa y Bahía Verde, Buenos
Aires, Argentina. Hornero 23(1):37–40
Pichegru L, Nyengera R, McInnes AM et al (2017) Avoidance of seismic survey activities by pen-
guins. Sci Rep 7:16305. https://doi.org/10.1038/s41598-­017-­16569-­x
Piersma T (1987) Hop, skip or jump? Constraints on migration of arctic waders by feeding, fatten-
ing, and flight speed. Limosa 60:185–194
Piersma T, Gill RE (1998) Guts don’t fly: small digestive organs in obese Bar-tailed godwits. Auk
115:196–203
Piersma T, van Aelst R, Kurk K et al (1998) A new pressure sensory mechanism for prey detection
in birds: the use of principles of seabed dynamics? Proc R Soc Lond B 265:1377–1383
Pimm SL, Jenkins CN, Abell R et al (2014) The biodiversity of species and their rates of extinction,
distribution, and protection. Science 344:1246752. https://doi.org/10.1126/science.1246752
Quadri Adrogué A, Miglioranza KSB, Copello S et al (2019) Pelagic seabirds as biomonitors of
persistent organic pollutants in the Southwestern Atlantic. Mar Pollut Bull 149:110516. https://
doi.org/10.1016/j.marpolbul.2019.110516
Reunión Argentina de Ornitología, Universidad Nacional del Centro de la Provincia de Buenos
Aires, Tandil, 4-6 September 2019. https://doi.org/10.13140/RG.2.2.24074.41923
Reynolds MH, Courtot KN, Berkowitz P et al (2015) Will the Effects of Sea-Level Rise Create
Ecological Traps for Pacific Island Seabirds?. PLOS ONE 10(9):e0136773. https://doi.
org/10.1371/journal.pone.0136773.
Ribeiro PD, Iribarne OO, Navarro D et al (2004) Environmental heterogeneity, spatial segrega-
tion of prey, and the utilization of Southwest Atlantic mudflats by migratory shorebirds. Ibis
146:672–682
Roberts DA (2012) Causes and ecological effects of resuspended contaminated sediments (RCS) in
marine environments. Environ Int, 40:230–243. https://doi.org/10.1016/j.envint.2011.11.013.
Rosenberg KV, Dokter AM, Blancher PJ et al (2019) Decline of the North American avifauna.
Science 366:120–124. https://doi.org/10.1126/science.aaw1313
Rothschild M, Clay T (1952) Fleas, flukes, and cuckoos: a study of bird ectoparasites. Collins Son
and Co, London
Saalfeld DT, Matz AC, McCaffery BJ et al (2016) Inorganic and organic contaminants in Alaskan
shorebird eggs. Environ Monit Assess 188(5):276
Sala OE, Chapin FS III, Armesto JJ et al (2000) Global biodiversity scenarios for the year 2100.
Science 287:1770–1774. https://doi.org/10.1126/science.287.5459.1770
Schmidt-Nielsen K (1960) The salt-secreting gland of marine birds. Circulation 21:955–967
Schreiber EA, Burger J (2001) Seabirds in the marine environment. In: Schreiber EA, Burger J
(eds) Biology of marine birds. CRC Press, Florida, pp 1–15
Senner NR, Hochachka WM, Fox JW et al (2014) An exception to the rule: carry-over effects do
not accumulate in a long-distance migratory bird. PLoS One 9:e86588. https://doi.org/10.1371/
journal.pone.0086588
Shuter JL, Broderick AC, Agenw DJ et  al (2011) Conservation and management of migratory
species. In: Milner-Gulland EJ, Fryxell JM, Sinclair ARE (eds) Animal migration: a synthesis.
Oxford University Press Inc, New York, pp 172–206
Simonetti P (2012) El ostrero pardo, Haematopus palliatus, en el Estuario de Bahía Blanca:
Estudio de la biología reproductiva, uso del hábitat y el potencial condicionamiento ambien-
tal como consecuencia del impacto antrópico Doctoral Thesis Universidad Nacional del Sur,
Bahía Blanca, p. 142
Simonetti P, Botté SE, Fiori SM et al (2012) Heavy-metal concentrations in soft tissues of the bur-
rowing crab Neohelice granulata in Bahía Blanca Estuary, Argentina. Arch Environ Contam
Toxicol 62:243–253
13  Shorebirds and Seabirds’ Ecology and Conservation 357

Simonetti P, Botté SE, Fiori SM et al (2013) Burrowing crab (Neohelice granulata) as a potential
bioindicator of heavy metals in the Bahía Blanca Estuary, Argentina. Arch Environ Contam
Toxicol 64:110–118
Simonetti P, Botté SE, Marcovecchio JE (2015) Exceptionally high Cd levels and other trace ele-
ments in eggshells of American oystercatcher (Haematopus palliatus) from the Bahía Blanca
Estuary, Argentina. Mar Pollut Bull, 100(1):495–500
Simonetti P, Botté SE, Marcovecchio JE (2018) Heavy metal bioconcentration factors in the bur-
rowing crab Neohelice granulata of a temperate ecosystem in South America: Bahía Blanca
estuary, Argentina. Environmental Science and Pollution Research, 25(34):34652–34660
Spetter CV, Buzzi NS, Fernández EM et  al (2015) Assessment of the physicochemical condi-
tions sediments in a polluted tidal flat colonized by microbial mats in Bahía Blanca Estuary
(Argentina). Mar Pollut Bull 9:491–505. https://doi.org/10.1016/j.marpolbul.2014.10.008
Sutherland WJ, Alves JA, Amano T et al (2012a) A horizon scanning assessment of current and
potential future threats to migratory shorebirds. Ibis 154:663–679
Sutherland WJ, Mitchell R, Prior SV (2012b) The role of ‘conservation evidence’ in improving
conservation management. Conservation Evidence 9(11):1–2
Tasker ML, Camphuysen CJ, Cooper J et al (2000) The impacts of fishing on marine birds. ICES
J Mar Sci 57:531–547. https://doi.org/10.1006/jmsc.2000.0714
Teuten EL, Saquing JM, Knappe DRU et al (2009) Transport and release of chemicals from plastic
to the environment and to wildlife. Philos T R Soc B 36:2027–2045
Thiebault A, Mullers RHE, Pistorius P et al (2014) Local enhancement in a seabird: reaction dis-
tances and foraging consequence of predator aggregations. Behav Ecol 25:1302–1310
Thomas GH, Wills MA, Székely T (2004) A supertree approach to shorebird phylogeny. BMC
Evol Biol 4:28
Thompson RC, Olsen Y, Mitchell RP et al (2004) Lost at sea: where is all the plastic? Science 304:838
van Aswegen JD, Nel L, Strydom NA, Minnaar K et al (2019) Comparing the metallic elemental
compositions of kelp Gull Larus dominicanus eggs and eggshells from the Swartkops Estuary,
Port Elizabeth, South Africa. Chemosphere 221:533–542
van Gils JA, Lisovski S, Lok T, Meissner W et  al (2016) Body shrinkage due to Arctic warm-
ing reduces red knot fitness in tropical wintering range. Science 352:819–821. https://doi.
org/10.1126/science.aad6351
Villagrán DM, Fernández Severini MD, Biancalana F et al (2019) Bioaccumulation of heavy met-
als in mesozooplankton from a human-impacted south western Atlantic estuary (Argentina). J
Mar Res 77:217–241. https://doi.org/10.1357/002224019826887362
Walker BM, Senner NR, Elphick CS et al (2020) Hudsonian godwit (Limosa haemastica), ver-
sion 1.0. In: Poole AF, Stettenheim PR, Gill FB (eds) Birds of the world. Cornell Lab of
Ornithology, New  York. Available via https://doi-­org.proxy.library.cornell.edu/10.2173/bow.
whrsan.01. Accessed 8 May 2020
White DC, Smith GA, Guckert JB et al (1993) Nearshore benthic marine sediments. In: Antarctic
microbiology. Wiley-Liss, New York, pp 235–240
Wilcox C, Van Sebille E, Hardesty BD (2015) Threat of plastic pollution to seabirds is global,
pervasive, and increasing. Proc Natl Acad Sci 112(38):11899–11904
Yorio P, Rábano D, Rabuffetti F et al (1998) Distribución reproductiva y abundancia de las aves
marinas de la Provincia de Buenos Aires: de Bahía Blanca a Punta Redonda. In: Yorio P, Frere
E, Gandini P et  al (eds) Atlas de la distribución reproductiva de aves marinas en el litoral
patagónico argentino. Fundación Patagonia Natural and Wildlife Conservation Society, Buenos
Aires, pp 19–28
Yorio P, Bertellotti M, García Borboroglu P (2005) Estado poblacional y de conservación de gavi-
otas que se reproducen en el litoral marítimo argentino. Hornero 20:53–74
Yorio P, Petracci P, García Borboroglu P (2013) Current status of the threatened Olrog’s Gull Larus
atlanticus: global population, breeding distribution and threats. Bird Conserv Int 23:477–486
358 N. S. Martínez-Curci et al.

Zalba SM, Sonaglioni MI, Compagnoni CA, Belenguer CJ (2000) Assessing the risk of inva-
sion by an exotic plant in a coastal nature reserve. Biol Conserv 93(2):203–208. https://doi.
org/10.1016/S0006-3207(99)00146-9
Zhang WW, Ma JZ (2011) Waterbirds as bioindicators of wetland heavy metal pollution. Procedia
Environ Sci 10:2769–2774
Zhao S, Feng C, Wang D et al (2013) Salinity increases the mobility of Cd, Cu, Mn, and Pb in
the sediments of Yangtze Estuary: relative role of sediments’ properties and metal speciation.
Chemosphere 91:977–984. https://doi.org/10.1016/j.chemosphere.2013.02.001
Zusi RL (1996) Family rynchopidae (Skimmers). In: del Hoyo J, Elliott A, Sargatal J (eds)
Handbook of the birds of the world, vol 3. Lynx Edicions
Chapter 14
Marine Mammals: Is the Bahía Blanca
Estuary and Its Area of Influence
Important for Their Conservation?

Gisela Giardino, Estela M. Luengos Vidal, Victoria Massola,


M. Agustina Mandiola, Joaquín C. M. Gana, Diego Rodríguez,
and Ricardo Bastida

14.1  Role of Marine Mammals in Marine Ecosystems

Marine mammals include around 134 extant species that are primarily dwelling or
dependent on the ocean for food. This group comprises three mammalian orders:
Sirenia, 4 species (manatees and dugongs); Carnivora, 41 species (polar bears, sea
otters, and pinnipeds: sea lions and fur seals, true seals, and walruses); and Cetacea,
89 species (baleen whales and toothed whales such as dolphins) (Berta et al. 2015).
Marine mammals are not uniformly distributed among the oceanographic areas.
Some of them inhabit deep water, marine shelf areas, and areas of open ocean, and
few species live in freshwater. They live in several habitats and ecosystems from the
Arctic to the Antarctic. Only in the Argentine Sea, 50 species of marine mammals
have been cited so far, some of them are very frequent and well known, while others
are little known and of circumstantial presence (Bastida and Rodríguez 2009).
Marine mammals are considered as top or apex predators whose food habits
are in many cases specialized, feeding on krill, fish, or even birds and mammals.

G. Giardino () · M. A. Mandiola · J. C. M. Gana · D. Rodríguez · R. Bastida


Instituto de Investigaciones Marinas y Costeras (IIMyC), FCEyN, UNMdP – CONICET,
Mar del Plata, Argentina
E. M. Luengos Vidal
INBIOSUR (Instituto de Ciencias Biológicas y Biomédicas del Sur), CONICET-UNS,
Bahía Blanca, Argentina
Grupo de Ecología Comportamental de Mamíferos, Laboratorio de Fisiología Animal,
Departamento de Biología, Bioquímica y Farmacia, Universidad Nacional del Sur,
Bahía Blanca, Argentina
V. Massola
Fundación para la Recepción y Asistencia de Animales Marinos (FRAAM). Sarmiento
esquina Pichincha Villa del Mar S/N°. Reserva Natural de Usos Múltiples, Bahía Blanca,
Bahía Falsa y Bahía Verde, Organismo Provincial para el Desarrollo Sostenible,
Bahía Blanca, Argentina

© Springer Nature Switzerland AG 2021 359


S. M. Fiori, P. D. Pratolongo (eds.), The Bahía Blanca Estuary,
https://doi.org/10.1007/978-3-030-66486-2_14
360 G. Giardino et al.

They are foremost consumers at most trophic levels, from primary producers to
predatory fish and even to other marine mammals. Because of their large body size
and abundance, they are thought to have a great influence on the structure and func-
tion of some marine communities. In addition, as top predators, they can serve as
key indicator species of marine ecosystem dynamics through changes in their abun-
dance, behavior, and health. Consequently, marine mammals could have five main
functions: as consumers, as preys, as vectors of nutrient and material flux, as a par-
ticular and new “habitat,” and as environmental sentinels (Roman et al. 2014).

14.1.1  Marine Mammals as Consumers

Mammals may function as keystone species in some marine communities. As a


result, a serious depletion in their numbers can cause major changes in species sta-
tus, starting a chain of extinctions through the food web. Croll et al. (2006) esti-
mated that 65% (range 53–86%) of the North Pacific Ocean’s primary production
was required to sustain the large whale populations prior to commercial whaling.
The total metabolic rate of a whale is high, as an endotherm, but because of its
immense size, it has a low mass-specific metabolic rate, relative to smaller animals.
The amount of food required to sustain 1 blue whale (Balaenoptera musculus) could
support 7 smaller minke whales (Balaenoptera acutorostrata) or 1500 penguins, but
their collective biomass reach just 50% or 8% of the biomass of a blue whale,
respectively, because of the relatively higher metabolic rates of these smaller ani-
mals (Roman et al. 2014).
If primary production is held constant, reducing whale populations lowers the
potential for marine ecosystems to retain carbon, both in living biomass and in car-
casses that sink to the ocean floor (Pershing et al. 2010). After whale biomass was
removed from the Antarctic system (approximately 84%), it is estimated that
150  million tons of krill go uneaten each year. Crabeater seals (Lobodon carci-
nophaga), Antarctic fur seals (Arctocephalus gazella), leopard seals (Hydrurga lep-
tonyx), and penguins, all krill-eating predators, began to increase, thus moving the
Antarctic marine ecosystem to new equilibrium levels. These species directly ben-
efited from the removal of whales and may now be hindering the recovery of whale
stocks (Trites et al. 1997).
In addition, pinnipeds can affect their ecosystem, influencing the benthic fauna
and community structure. Estes et al. (1998) explain the effect of the abundance of
sea otters (Enhydra lutris) on the structure of kelp communities, and Anderson
(1995) suggested that the overhunting of local sea otter populations could have led
to the extinction of Steller’s sea cows (Hydrodamalis gigas). How are related sea
otters and Steller’s sea cows? Sea otters feed on sea urchins, keeping their number
low, which is essential for the maintenance of the dense kelp beds that, in turn, are
the main food source for sea cows.
14  Marine Mammals: Is the Bahía Blanca Estuary and Its Area of Influence Important… 361

In British Columbia the annual diet of harbor seals (Phoca vitulina) contains
about 4% of salmon and 43% of hake. Contrary to expectations, the harbor seals
may be benefiting salmon because they affect the abundance of hake, which is one
of the largest predators of young salmons (Trites 1997).

14.1.2  Marine Mammals as Prey

Cetaceans and pinnipeds are also eaten by several species. White sharks
(Carcharodon carcharias) have been observed killing and feeding on small odonto-
cetes and scavenging on carcasses of large cetaceans (Cockcroft et al. 1989). Shark
predation attempts on large marine animals are rare, but recent studies suggest that
they may occur more often than previously thought (Bornatowski et al. 2012). There
are several reports of killer whales (Orcinus orca) attacking greater whales. In
Península Valdés (Patagonia), the stranding behavior of orcas to capture both, sea
lions (Otaria flavescens) and elephant seals (Mirounga leonina), is known world-
wide. Also in Patagonia, killer whales were observed feeding on dusky dolphins
(Lagenorhynchus obscurus) and chasing common dolphins (Delphinus delphis)
(Coscarella et al. 2015), while in Brazil and in Península Valdés, they were observed
attacking Southern right whale (Eubalaena australis) calves (Bastida and Rodríguez
2009; Ott et al. 2017).
In Antarctica, groups of killer whales cause a wave to make the seal pups fall
from the floating ice and thus feed on them. In the Buenos Aires Province, killer
whales have been observed attacking franciscanas (Pontoporia blainvillei), bottle-
nose dolphins (Tursiops truncatus), and fur seals (Arctocephalus australis) (Bastida
and Rodríguez 2009). Other indirect evidence that whales are victims of predation
are the high frequency of rake mark scars found on their flukes (Reeves et al. 2006).
Opportunistic terrestrial predators of pinnipeds (especially pups) include wolves,
dogs, foxes, jackals, hyenas, and pumas, in addition to some birds that feed on pin-
niped pups (Weller 2018). At least five species of pinnipeds have been documented
to feed on other pinniped species (Harcourt 1993). The strong tendency of pinni-
peds to haul out on ice or islands limits the impact of terrestrial predators on these
populations. The main prey of polar bears (Ursus maritimus) throughout their range
is the ringed seal (Pusa hispida). Polar bears hunt ringed seals on both fjord and
open sea ice. Other marine predators that can have serious effects on some pinniped
populations include adult male sea lions and leopard seals, killer whales, and sev-
eral species of large sharks. Stomach content analyses indicate that white sharks
prefer pinnipeds or whales to other prey such as birds or sea otters. This selective
preference for marine mammals with extensive lipid stores may be necessary to
meet their elevated muscle temperatures and high growth rates in the cold waters
where their attacks on pinnipeds are concentrated (Ainley et al. 1985).
362 G. Giardino et al.

14.1.3  M
 arine Mammals as Vectors of Nutrient
and Material Flux

Whales facilitate nutrient transfer by releasing fecal plumes near the surface after
feeding at depth, so nutrients are moved from highly productive to less productive
areas. During urine and fecal plume production, nitrogen and iron is released, indi-
rectly increasing prey productivity and abundance. These plumes enhance phyto-
plankton blooms and carbon sequestration (Lavery et al. 2010). Furthermore, during
whale migration nutrients transfer through urea and placentas from areas of high to
low productivity. Whale carcasses sequester carbon to the deep sea, where they
provide a massive pulse of organic enrichment as well as habitat and food for many
endemic species, including chemosynthetic bacteria and invertebrate hosts (Roman
et al. 2014). Similarly, the excrement of pinnipeds contributes to the turnover and
recycle nutrients, mainly in large colonies (Trites 1997). When foraging, whales and
pinnipeds can locally influence the ocean physical environment. Through diving
and surfacing, whales can enhance the upward transport of nutrient-rich deep
waters, as they pass through density gradients during feeding sessions (Dewar et al.
2006). Humpback whales (Megaptera novaeangliae) intentionally disturb sandy
bottoms and shell-hash beds (a mix of mud, sand, and broken shells) to flush sand
lance, a prey, from their burrows (Hain et al. 1995), contributing mechanical energy
to the ocean. In addition, under certain conditions, the bubble nets that humpback
whales make to catch schools of small fish, along with bottom disturbance, play an
important role in the flux of materials, since these activities also can break the ther-
mocline, facilitating the transport of nutrients from colder to warmer layers in the
ocean. Stranded whales can subsidize terrestrial food webs (Chap. 15 in this book).

14.1.4  Marine Mammals as a Particular and New “Habitat”

The ocean floor is rich on detritus due to the contribution made by the carcasses of
the great whales that fall from the surface. Dead whales contribute with proteins and
lipids, yielding massive pulses of organic enrichment to a realm that is often nutri-
ent and energy impoverished. Only one gray whale (Eschrichtius robustus) of
40 tons, provides nearly 2 million g carbon, equivalent to more than 2000 years of
the background carbon flux for an area equivalent to the carcass surface (Smith
2006). Whale falls influence deep-sea environments, by altering local food avail-
ability, providing habitat structure, and supporting diverse biotic assemblages
(Lundsten et al. 2010). Dead bodies of marine mammals undergo ecological succes-
sion from a stage dominated by mobile scavengers that actively consume the soft
tissues, passing through the enrichment-opportunist stage and getting to a sulpho-
philic stage where chemosynthetic bacteria dominate (Smith 2006; Lundsten et al.
2010; Amon et al. 2013). The persistent food-rich conditions and widespread occur-
rence of whale carcasses has led to ecological and evolutionary opportunities on the
14  Marine Mammals: Is the Bahía Blanca Estuary and Its Area of Influence Important… 363

deep-sea floor, in a manner similar to that of hydrothermal vents and cold seeps
(Smith 2006). In the North Pacific, more than 60 macrofaunal species have been
associated with one single whale fall. Regarding whale carcasses stranded on coast-
lines, after death, it is a mechanism for transporting marine biomass to the sea–land
interface. Although the number of stranded whales is small as compared with those
that sink (Smith 2006), carcasses can attract and nourish many terrestrial consumers.

14.1.5  Marine Mammals as Sentinels

As marine mammals have long lives and move over great distances, they can regis-
ter ecological variation across large spatial and long temporal scales. Thus, they can
be considered sentinels of marine ecosystem change. Selecting the appropriate
marine mammal species to use as sentinel of change depends on the ecological
alteration of concern (Moore 2008, 2018). Apex predators as whales, seals, and sea
lions can indicate environmental changes and degradation. For example, if we want
to know broad scale shifts in ecosystems, migratory mysticete whales may be inves-
tigated, whereas polar cetaceans are more useful for assessing the effects of rapid
changes in sea ice conditions and its impacts on food webs in these strongly sea-
sonal ecosystems. On the other hand, coastal dolphins are good indicators of pollut-
ant or disease vectors in nearshore habitat (Moore 2008).
Harwood (2001) says that critical habitat for marine mammals can be defined in
terms of the ecological units that provide safe areas to breed and forage. The layer
of blubber for insulation that marine mammals have is suitable for the accumulation
of lipophilic pollutants. The information obtained from this layer is useful for indi-
cating the prevalence and persistence of pollutants in marine ecosystems (Gil et al.
2006; Panebianco et  al. 2012; Perez-Venegas et  al. 2018; Muñoz-Arnanz et  al.
2019). In addition, profiles of stable carbon isotopes and fatty acids in blubber can
be used to infer the diet of marine mammals, thereby providing evidence of changes
in food webs within marine ecosystem (Budge et al. 2008; de Castro et al. 2016).
Changes in individual body condition can demonstrate shifts in the prey base and
food web structure as well as alterations in pathogen transmission. Indeed, to
explore variability of ecosystem productivity and health, it seems essential to incor-
porate the biology and ecology of marine mammals and other apex predators in
multidiscipline programs of research (Moore 2008). An understanding of the role of
marine mammals in marine ecosystems is important because it provides a context
within which to evaluate the potential impact of their predation on prey populations
and community structure, and the impact of variation in prey populations, due to
harvesting by humans, and environmental change on the dynamics of marine mam-
mals. Mammals may function as keystone species in some marine communities
such as in the famous example about the effect of the abundance of sea otters
(Enhydra lutris) on the structure of kelp communities (Estes et al. 1998).
Indeed, marine mammals can guide human stewardship activities, reflect the
ocean’s role in climate interactions across regions, demonstrate ecosystem
364 G. Giardino et al.

vulnerabilities and health, and thereby lead to ways to enhance human health
(Moore 2008). Moreover, as charismatic megafauna, marine mammals capture the
attention and concern of the public. This capability provides clear opportunities for
education and outreach on oceanic and environmental themes.

14.2  Marine Mammal in the Bahía Blanca Estuary

Stranded marine mammals allow researchers to gather valuable information about


many elusive species. They offer a unique opportunity to learn about a species’ life
history, population structure, occurrence, disease prevalence, and anthropogenic
causes of mortality and to understand fossil assemblages. It is important to consider
that stranded animals are not a random sample of the free-living population as most
of them are unhealthy specimens. On the other hand, the stranding range detection
is a dependent variable of the human effort and the possibilities of access to coastal
sites. Beyond the health status of animals, stranding episodes may also reflect
changes in environmental conditions and diverse human activities, shipping traffic
collisions, oil exploration, etc. (Leeney et al. 2008). Cetaceans may strand alive for
different reasons including behavioral tendencies of particular species to follow a
leader, disorientation caused by geographical anomalies in the earth’s magnetic
field and acoustical “dead zones,” coastal areas where echolocation signals are dis-
torted, anthropogenic factors such as bycatch in fisheries or contaminants, and
infectious diseases (Peltier et al. 2013). Pinnipeds seem to be particularly suscepti-
ble to entanglement in marine debris, perhaps because of their exploratory nature.
Entanglement lesions could develop into infected chronic wounds and the entangled
pinnipeds may live for months or years with a plastic line, or net, cutting into their
skin and muscle tissues (Butterworth 2016). Interaction with fisheries is another
source of animal death or injury. Impacts may be direct consequence of bycatch or
shooting by fishermen, as well as indirect competition for fish resources and
fisheries-­induced changes to ecosystems that cause nutritional stress among pinni-
peds (Kovacs et al. 2012).
The increase of global concern regarding the status of marine mammals pro-
moted government programs for facilitating the public record of stranding episodes
and fostered investigations into the causes of mortality. In some places, there is a
long history of systematic records, for example, in the UK after the ASCOBANS
(Agreement on the Conservation of Small Cetaceans of the Baltic and North Seas)
signature in 1993, a stranding reporting scheme was established along with a
research program to investigate death causes. One of the longest-term stranding
projects is the one carried out by Leeney et al. (2008) that covers a record period of
96 years on the Cornwall coast (UK) and islands of Sicily (Italy). In Argentina there
are some initiatives like Museo Acatushún in Tierra del Fuego, LAMAMA Lab in
Puerto Madryn, Marine Mammal Lab at Mar del Plata University, Fundación
Mundo Marino in San Clemente del Tuyú, Fundación Aquamarina in Pinamar and
FRAAM (Marine Animal Rescue and Assistance Foundation) in Bahía Blanca
14  Marine Mammals: Is the Bahía Blanca Estuary and Its Area of Influence Important… 365

Estuary. All these research groups record systematic data on stranding marine
mammals.
We made a list of marine mammals that could potentially be present in Bahía
Blanca Estuary and surroundings, following Bastida and Rodríguez (2009) and the
recently submitted Red List of Threatened Species for Argentina (SADSN-SAREM
2019). The information about some marine mammals is scarce and incomplete, but
approximately 37 species have their distribution range within the area, 29 cetaceans
and 8 pinnipeds (Table  14.1). According to the IUCN Red List of Threatened
Species (IUCN, International Union for the Conservation of Nature) and Red List
of Threatened Species for Argentina (SADSN-SAREM 2019), almost 39% of the
species are “Data Deficient.” Some Data Deficient species may be well studied and
its biology be well known, but still lack appropriate data on its abundance and/or
distribution. All the pinnipeds in the area are considered Least Concern. Among the
cetaceans of possible occurrence in Bahía Blanca Estuary, 17% of the species
belong to the threatened categories (“Endangered” or “Vulnerable”); any taxon in
those categories faces a high or very high risk of extinction in the wild. Finally, the
remaining marine mammal species belong to the “Data Deficient” (38%) or “Least
Concern” categories (22%). This last category is implying that the population exist
or spreads through its entire range. We also analyzed local information from two
sources Fidalgo (2004) and FRAAM marine mammal database. Fidalgo (2004)
reported 10 different marine mammal species from sightings, entanglements or
stranding episodes, and other 6 species from records than needed confirmation
(Table 14.1).
FRAAM marine mammals’ database has records that come mostly from stranded
animals both alive and dead, or that arrived to the coast in poor health conditions.
These records span from 2004 to 2019  in the geographical area from Balneario
Marisol to Bahía San Blas. FRAAM recorded 71 marine mammals, most of them
pinnipeds (79%), particularly belonging to the Otariidae family (sea lions and fur
seals) (75%) (Fig. 14.1). Four species of pinnipeds were registered in the area, but
the South American sea lion (Otaria flavescens, see species data sheet a) and the
South American fur seal (Arctocephalus australis, see species data sheet b) were the
most frequent (Fig. 14.1). In addition, O. flavescens has an historic assessment in
the area near Punta Lobos, on the Trinidad Island, in the Bahía Blanca Estuary (see
box text in this chapter). Currently the colony is still active, although no studies
have been reported recently. The presence of A. australis is expected due to the cur-
rent population growth (Crespo et  al. 2015; Mandiola 2015). SADSN-SAREM
(2019) has listed both species as “Least Concern” and their main conservation prob-
lem is their interaction with anthropic activities (Romero et al. 2011; Mandiola et al.
2017). Most individuals were juvenile and subadult animals in good health condi-
tion, that leave the water to rest or because they were molting. Another pinniped
recorded in Bahía Blanca Estuary was the Southern elephant seal (Mirounga leo-
nina), which was registered three times, always as solitary subadult specimens. The
presence of this species has also been increasing in other coastal areas of the Buenos
Aires Province and in the harbor colonies of Mar del Plata and Necochea (Bastida
and Rodríguez 2009; Group of Biology, Ecology and Conservation of Marine
Table 14.1  List of potential species of marine mammals surrounding the Bahía Blanca Estuary based on the information of Bastida and Rodríguez (2009) and
366

SADSN-SAREM (2019), national conservation status and Fidalgo (2004)


Cat. FRAAM Fidalgo
Order Group Family Common Name Scientific Name SAREM data (2004)
Carnivora Pinnipeds Otariidae (sea lions and fur South American sea Otaria flavescens LC P P
seals) lion
South American fur Arctocephalus australis LC P P
seal
Antarctic fur seal Arctocephalus gazella LC P
Subantarctic fur seal Arctocephalus tropicalis LC
Phocidae (true seals) Southern elephant seal Mirounga leonina LC P
Weddell seal Leptonychotes weddellii LC
Crabeater seal Lobodon carcinophaga LC
Leopard seal Hydrurga leptonyx LC
Cetacea Mysticeti (baleen Balaenidae Southern right whale Eubalaena australis LC P P
whales)
Neobalaenidae Pygmy right whale Caperea marginata DD
Balaenopteridae Blue whale Balaenoptera musculus EN
Fin whale Balaenoptera physalus EN
Sei whale Balaenoptera borealis EN P
Bryde’s whale Balaenoptera edeni DD
Antarctic minke whale Balaenoptera DD P
bonaerensis
Dwarf minke whale Balaenoptera DD
acutorostrata
Humpback whale Megaptera novaeangliae LC P
Odontoceti (toothed Physeteridae Sperm whale Physeter macrocephalus VU NC
whales)
Kogiidae Pygmy sperm whale Kogia breviceps DD
G. Giardino et al.
Cat. FRAAM Fidalgo
Order Group Family Common Name Scientific Name SAREM data (2004)
Dwarf sperm whale Kogia sima NA P
Pontoporiidae La Plata river dolphin Pontoporia blainvillei VU P P
Delphinidae Common dolphin Delphinus delphis LC P
Lahille’s bottlenose Tursiops truncatus VU P P
dolphin gephyreus
Southern right whale Lissodelphis peronii DD
dolphin
Dusky dolphin Lagenorhynchus LC NC
obscurus
Risso’s dolphin Grampus griseus LC
Killer whale Orcinus orca LC P
False killer whale Pseudorca crassidens DD P
Long-finned pilot Globicephala melas LC P
whale
Phocoenidae Burmeister’s porpoise Phocoena spinipinnis DD P
Spectacled porpoise Phocoena dioptrica LC NC
Ziphiidae Cuvier’s beaked whale Ziphius cavirostris DD P
Southern bottlenose Hyperoodon planifrons DD
whale
Arnoux’s beaked whale Berardius arnuxii DD
Gray’s beaked whale Mesoplodon grayi DD P NC
Héctor’s beaked whale Mesoplodon hectori DD
Strap-toothed whale Mesoplodon layardii DD P
Ref:* discrepancies between Bastida and Rodríguez (2009) and SADSN-SAREM (2019). Categories of conservations status
DD data deficient, LC Least Concern, VU vulnerable, NA without application, P presence, NC need confirmation
14  Marine Mammals: Is the Bahía Blanca Estuary and Its Area of Influence Important… 367
368 G. Giardino et al.

Fig. 14.1  Marine mammals recorded in Bahía Blanca Estuary and surrounding areas by the envi-
ronmental non-governmental organization Marine Animals Assistance and Rescue
Foundation (FRAAM)

Mammals, UNMdP, unpublished data). One young Antarctic fur seal (Arctocephalus
gazella) has also been recorded as a wandering species with occasional records in
Argentine coasts.
The cause of stranding was determined based on a combination of results from
clinical examinations or gross necropsy. Animals, that were sick or injured and pos-
sibly admitted for rehabilitation, were provided supportive treatment and released
afterwards if their condition improved (see Sect. 4 in this chapter). Dead individuals
represented 27% of all pinnipeds, and from those alive, 66% were released after a
short period of rehabilitation. Similar number of cases was registered in each year,
decreasing slightly in autumn.
One of the toothed whales (Odontoceti) more frequently recorded was the small
dolphin Pontoporia blainvillei, (see species data sheet c) mostly adult and subadult
females, followed by males and less frequently newborns. Most of them were found
with skin injuries related to entanglement. In the last 10 years, the information on
the biology and ecology of this dolphin in Argentina has increased exponentially;
14  Marine Mammals: Is the Bahía Blanca Estuary and Its Area of Influence Important… 369

however, updated information about main population parameters, such as mortality,


abundance, and population trends, is missing for a more accurate assessment
(Denuncio et al. 2019). The annual incidental mortality of this species in Argentina
reaches levels between 3.5% and 5.6%, far exceeding the 2% suggested as maxi-
mum sustainable level by the International Whaling Commission (Crespo et  al.
2010). Nowadays this dolphin is considered as Vulnerable according to the last Red
List of Threatened Species for Argentina (Denuncio et al. 2019).
Lahille’s bottlenose dolphins (Tursiops truncatus gephyreus, see species data
sheet d), is another frequent Odontoceti species registered by fishermen or people
that use the estuary in a recreational way, but individuals do not show a stranding
occurrence higher than other cetaceans. According to the last categorization, T. t.
gephyreus is considered an endangered species (EN) (Vermeulen et  al. 2019).
Bottlenose dolphins coming from San Antonio Bay, Río Negro, are genetically iso-
lated from those in Uruguay and southern Brazil and were identified as an evolu-
tionarily significant unit within the Southwestern Atlantic (Fruet et  al. 2014).
Information on bottlenose dolphins in the country is dispersed in time and space
(Vermeulen et al. 2018), but the Argentine population of bottlenose dolphins may
have decreased at an estimated rate of 20% in two generations, and there may be
less than 250 dolphins alive (Vermeulen et al. 2019).
In addition, killer whales (Orcinus orca) are frequently seen alive in the area,
mainly during the southern summer. Other toothed whales occasionally found
stranded and dead in the Bahía Blanca Estuary belong to the Ziphiidae and Kogiidae
families, both of them of pelagic habits.
Baleen whales (Mysticeti) have been also recorded in the area. The first hump-
back whale (Megaptera novaeangliae) was recorded in 2011 (Angeletti et al. 2014),
followed by two whales in 2018 and one in 2019. Humpback whales move annually
from their main breeding grounds in Abrolhos (Brazil) to their temperate and polar
summer feeding areas (Jefferson et al. 2015; Andriolo et al. 2010). Thus, Buenos
Aires Province is in the middle of their journey. Estimations performed in the
Abrolhos archipelago showed that population densities of humpback whales were
increasing at a rate of approximately 7.4% per year and whales were re-occupying
old distribution areas (Zerbini et  al. 2004; Andriolo et  al. 2010). Another baleen
whale very common in the area is the Southern right whale (Eubalaena australis)
observed from April to October. Their records in Buenos Aires Province have
increased gradually since 1970 (Mandiola et al. 2020).
The Bahía Blanca Estuary is located between two main reproductive areas
(Península Valdés Argentina and Santa Catarina-Brazil) of the Southern right whale,
so this site, like the entire Buenos Aires coast, is considered as transit area. Whales
recorded in autumn and winter season could be individuals travelling from their
southern feeding areas to the breeding areas of southern Brazil (Mandiola et  al.
2020). Nevertheless, recent satellite-tracking studies from Golfo San Matías and
Península Valdés show variable individual movement patterns. Most tracked whales
370 G. Giardino et al.

made coastal and offshore migrations to feeding grounds after the breeding season
with no clear displacement pattern (Zerbini et al. 2016, 2018).
Most of the right whale records in the Bahía Blanca Estuary correspond to live
whales’ sightings. Only one Southern right whale was recorded stranded in the area
with clear scars of vessel collision.
Antarctic minke whales (Balaenoptera bonaerensis) were also recorded three
times; the first specimen was a calf which died after a short time, another was found
alive and released (in 2006), and the last one in 2007 was found dead ashore.
Table 14.1 and Fig. 14.1 summarize all marine mammals recorded in the area.
Considering the previous information, basic information about the four species
most likely to be observed in Bahía Blanca Estuary is presented:
• South American sea lion – Otaria flavescens (Shaw, 1800)
• South American fur seal – Arctocephalus australis (Zimmerman, 1783)
• La Plata River dolphin – Pontoporia blainvillei (Gervais and d’ Orbigny, 1844)
• Lahille’s bottlenose dolphin – Tursiops truncatus gephyreus (Montagu, 1821)

 outh American Sea Lion Otaria flavescens


14.2.1  S
(Shaw, 1800)

Common name: Lobo marino de un pelo Sudamericano, León marino (Spanish),


South American sea lion (English)

14.2.1.1  Description

Medium to large size, adult males between 2.1–2.8 m in length and weight around
300–350  kg (Fig.  14.2); females between 1.5–2  m and 170  kg (Bastida and
Rodríguez 2009). It is clearly different from other sea lions because adult males
have an aspect more similar to a lion, and because it is the largest Otariidae species
in the region. They exhibit strong sexual dimorphism, adult males have their neck
and chest region covered with longer, thicker, coarser guard hairs, which can give
the impression of a mane. Females lack mane, and their body is more stylized. The
color is highly variable, ranging from reddish brown lighter to yellowish tones,
especially in females (Bastida and Rodríguez 2009). At birth, pups weigh 10–15 kg
and are 75–85 cm long. Pups are born black above and paler below. They undergo
their first molt 1–2  months after birth, becoming dark brown. The pelage has a
single hair layer (Jefferson et al. 2015). The ears are small, and their canine teeth are
very large and strong.
14  Marine Mammals: Is the Bahía Blanca Estuary and Its Area of Influence Important… 371

Fig. 14.2  South American sea lion, Otaria flavescens, male assessment in Trinidad Island (Bahía
Blanca Estuary). (Photo by Gisela Giardino)

14.2.1.2  Distribution and Habitat

The South American sea lion is the most abundant marine mammal occurring along
the southern part of South America and it is distributed along the Atlantic and Pacific
coast of South American. On the northern coast of Argentina, there are only four
haulouts (about 2500 individuals) formed only by males, while the Patagonian
region has both reproductive and non-reproductive colonies (about 120,700 indi-
viduals). An additional 7500 animals are found in the Malvinas-Falkland Islands
(Dans et al. 2012). Although Otaria flavescens is considered as not migratory spe-
cies and it remains somewhat concentrated on the coastal zone all around the year,
males are able to travel hundreds of kilometers for breed (Giardino et  al. 2016).
Some sea lions habit of coming up the rivers, as in the case of Rio Negro basins
(Bastida and Rodríguez 2009).
The coastal islands of Uruguay historically represented a major portion of its
population in the Atlantic Ocean, but this site has shown a decline of about 2% per
year during the last few decades, with actual numbers of about 10,000 animals (Páez
1996; Ponce de León 2000). South American sea lion haulout areas include sandy
beaches, flat surfaced slab of rocks, flat bases of cliff, and places with big boulders
(Vaz Ferreira 1981). Some individuals used to rest inside harbors, like in Rio Grande
do Sul (Brazil) (Rosas et  al. 1994; Pavanato et  al. 2013), in Mar del Plata, and
Puerto Quequén harbors in the northern coast of Argentina (Buenos Aires Province).
Sea lions that settled in port areas may lead to negative human interaction if there
are no policies of management (Giardino 2014).
372 G. Giardino et al.

14.2.1.3  Behavior

Terrestrial walking is performed by using all four flippers, while swimming is fun-
damentally powered by the forelimbs. Their breeding colonies are occupied from
middle of December by small number of adult males that take positions and delimit
territories through vocalizing, posturing, and fighting, prior to the arrival of females
several days later. The Southern sea lion is a highly polygynous species. Only one
dominant male can hold 4–10 adult females (harem) although some solitary couples
are found disperse (Campagna 1985; Campagna and Le Boeuf 1988). Once the
breeding season ends, sea lions change their distribution into haulouts of different
age composition, location, and stability throughout the rest of the year (Lewis and
Ximenez 1983). Mother–pup pairs maintain their social bonds ashore until weaning
in late austral spring and remain close to the natal area (Grandi et al. 2008), whereas
males do not provide parental care, have prolonged sexual maturation, and young
males are excluded from breeding opportunities; therefore they tend to disperse
farther away from the breeding grounds. Male sea lions from Buenos Aires Province
travel to Patagonia and Uruguay during the austral summer in order to mate
(Giardino et al. 2016) contributing to around 18% of the gene stock of this species
in the breeding area (Giardino et al. 2017). At sea, South American sea lions fre-
quently raft alone or in small to large groups. They have been also reported in asso-
ciation with feeding cetaceans and seabirds. On the Atlantic coast, most lactating
females have been described as benthic divers and forage in shallow water within
the continental shelf (Campagna et al. 2001; Rodríguez et al. 2013). As generalist
feeders, South American sea lions take a wide variety of prey that varies by location.
Their diet includes many species of benthic and pelagic fishes and invertebrates,
some of them of commercial value (Koen Alonso et al. 2000; Bustos et al. 2012).
Several authors recommended avoiding the physical contact with this species, as
they are, like other wild animals, possible disease vectors (Bernardelli et al. 1996;
Beron Vera et al. 2004; Kiers et al. 2008; Bastida et al. 2011; Arbiza et al. 2012; Bos
et al. 2014; Timi et al. 2014; Dans et al. 2017).

14.2.1.4  Threats and Conservation Status

South American sea lion numbers are increasing in Argentina, and in Buenos Aires
Province, the population grew in recent years (Giardino et al. 2017). The interaction
with different anthropic activities is the main conservation problem such as compe-
tition with artisanal and industrial fisheries (Romero et  al. 2011; Mandiola et  al.
2017). Moreover, they interact negatively with aquaculture in open sea, due to the
entanglement, use of space, chemical and acoustic contamination, industrial waste,
and vessel traffic (Romero et al. 2019).
The Red List IUCN and SADSN-SAREM categorization of Otaria flavescens is
Least Concern (Cardenas-Alayza et al. 2016; Romero et al. 2019).
14  Marine Mammals: Is the Bahía Blanca Estuary and Its Area of Influence Important… 373

 outh American Fur Seal Arctocephalus australis


14.2.2  S
(Zimmerman, 1783)

Common name: Lobo marino de dos pelos Sudamericano, lobo fino austral
(Spanish), South American fur seal (English)

14.2.2.1  Description

Medium size fur seal; male lengths between 1.80 and 2 m. and weight between 150
and 200 kg.; females between 1.20 and 1.40 m and weight between 50 and 60 kg.
The snout has quite a pointed shape, thin and long ears pinnae, and very long and
light-colored vibrissae (Fig. 14.3). Relative long forelimbs with poorly developed
nails. Short and dense fur with a double hair layer, dark grayish brown in the back
of body and a lighter-colored belly. This species shows sexually dimorphic; adult
males are about 1.3 times the length and 3.3 times the weight of females. The canine
teeth of adult males are larger and thicker than females. As their Spanish name says,
they have two types of hair, one very dense inner layer with soft, fine, and short hair
cover by another thicker and longer called guard hair (Bastida and Rodríguez 2009).
This species could be confused, in our area, with Antarctic and Subantarctic fur
seals. Adult male Antarctic fur seals are the same size and almost same color of

Fig. 14.3  South American fur seal, Arctocephalus australis, recorded in Bahía Blanca Estuary.
(Photo by Victoria Massola)
374 G. Giardino et al.

Southern fur seal but have a shorter muzzle and proportionately longer fore- and
hind flippers. Subantarctic fur seals, on the other hand, are unique with a pale blond
face and neck (Jefferson et al. 2015).

14.2.2.2  Distribution and Habitat

South American fur seal has two recognized subspecies, the South American and
Peruvian subspecies. The South American subspecies is distributed along all
Patagonian Sea, from western South Atlantic (southern Brazil) to eastern South
Pacific (southern Chile) coasts. For their distribution at sea, this species inhabit both
the coastal zone and the entire Patagonian platform, reaching the edge of the slope
(Mandiola et al. 2015; Baylis et al. 2018a, b). On the Atlantic side, haulouts can be
found along the coasts of Rio Grande do Sul in Brazil (Muelbert and Oliveira 2006),
although the breeding colonies goes from Islas del Castillo, Uruguay, continue to
Tierra del Fuego-Isla de los Estados (Túnez et al. 2008; Crespo et al. 2015). They
used to travel long distance (Mandiola et al. 2015; Baylis et al. 2018b).

14.2.2.3  Behavior

Males are polygynous and territorial (between 2 and 13 females for each male).
Their fights can result in serious wounds and scars. Individual bulls can occupy ter-
ritories for up to 60  days until most of the females are mated (Cappozzo 1995).
After mating, female begins to make foraging trips punctuated by time attending the
pup ashore (Pavés and Schlatter 2008). Only a few adult males achieve mating
while a large proportion is excluded to peripheral or male exclusive areas. Pupping
peaks take place in middle of November to middle of December, and mating occurs
1–6  days after the female gives birth (Franco-Trecu 2005; Pavés and Schlatter
2008). Colonies are generally found along rocky coasts, on edges above the shore-
line or in boulder-strewn areas. The Southern fur seal food habits vary according to
prey availability; it is an opportunist generalist predator, which feeds mainly on
pelagic and demersal preys, both coastal and continental shelf species. They eat
mainly on prawn, shrimp, squids, and several fish, as croakers, anchovy, and mack-
erel in Buenos Aires waters (Bastida and Rodríguez 2009; Pérez Salles 2015).
Studies on stable isotopes from Patagonia showed that juvenile fur seal feed more
pelagically than subadults and adults (Vales et al. 2015).
Dassis et al. (2012) observed that the most frequent behavior, recorded in Mar
del Plata, was passive flotation, followed by prolonged coastal dives. This behavior
is strongly influenced by the sea state, since when the sea is more “choppy,” fur
seals tend to concentrate while when the sea is calm, they tend to disperse. In addi-
tion, the wind direction and the surface current affect their behavior. This fact could
be a strategy to maintain certain areas with food availability, and for forage energy
optimization.
14  Marine Mammals: Is the Bahía Blanca Estuary and Its Area of Influence Important… 375

14.2.2.4  Threats and Conservation Status

Among its natural predators are the killer whale, Orcinus orca (Bastida et al. 2007),
and the cat shark, Notorynchus cepadianus (Crespi-Abril et al. 2003). There is a
large overlap between the areas used by fur seal and fisheries on the Patagonian
platform (Mandiola et al. 2015; Baylis et al. 2018b), although operational interac-
tions are sporadic (Crespo et al. 1997; Mandiola 2015). Actually, individuals have
been observed feeding during trawling fishing maneuvers in Buenos Aires waters
(Mandiola et al. 2017). The limited number of breeding areas could make this spe-
cies particularly vulnerable to the effects of epidemics and several human activities
that could have bad consequences on the population (Cardenas-Alayza et al. 2016).
The intake of marine litter (mainly plastics derived from fishing activity and remains
of bags) has been recorded in juvenile fur seal stranded on the Buenos Aires coast,
although no lesions were observed in the digestive tract (Denuncio et al. 2017). In
addition, the oil activity carried out on the Patagonian platform (transport route of
oil ships, exploration and exploitation areas) is always a risk.
Arctocephalus australis is categorized as Least Concern by the IUCN Red List
and SADSN-SAREM (Cardenas-Alayza et al. 2016; Vales et al. 2019). Listed in
Appendix II of CITES (Convention on International Trade in Endangered Species
of Wild Fauna and Flora).

 a Plata River Dolphin or Franciscana Pontoporia


14.2.3  L
blainvillei (Gervais and d’ Orbigny, 1844)

Common name: Franciscana, Delfín del Plata (Spanish) Franciscana, La Plata River
dolphin (English)

14.2.3.1  Description

Small-sized dolphin (128–175 cm) that weigh between 35 and 55 kg. Quickly dis-
tinguishes itself from the rest of the dolphins due to their extremely long and narrow
beak (about 12–15% of total length in adults). The long beak is lined with 50–62
fine, pointed teeth per row, more than in nearly any other species of cetacean. The
forehead is steep and rounded with flexible neck. The flippers are broad and spatu-
late, sometimes with an undulating trailing edge; meanwhile dorsal fin is relatively
rounded and small. Newborns have proportionately larger flippers, dorsal fins, and
flukes. The body color is uniform yellowish brown, in some cases lighter on the
belly, which serves to camouflage itself in the murky water where it lives (Bastida
and Rodríguez 2009; Jefferson et al. 2015). The species has a slightly inverted sex-
ual dimorphism, with females being larger than males (Kasuya and Brownell 1979;
Botta et al. 2010; Panebianco et al. 2012) (Fig. 14.4).
376 G. Giardino et al.

Fig. 14.4  La Plata River Dolphin (Franciscana dolphin) Pontoporia blainvillei. (Photo by Ricardo
Bastida)

14.2.3.2  Distribution and Habitat

The Franciscana Dolphin is found only along the east coast of South America
(Brazil, Uruguay, and Argentina), from Golfo San Matías, central Argentina
(42°35′S), to Espirito Santo, southeastern Brazil (18°25′S) (Bastida and Rodríguez
2009; Jefferson et al. 2015). They live mainly in coastal marine waters with a maxi-
mum distribution to the isobath of 50 m, and it is possible to see them swimming
behind the surfing zone (Bastida et al. 2000; Crespo et al. 2010). On the Argentine
coast, abundance was estimated by aerial censuses in the marine area in almost
15,000 animals with a greater density in the northern coastal area of the Province of
Buenos Aires, from Cabo San Antonio to Claromecó (Crespo et al. 2010). In this
region, the highest concentration of dolphins was found between the coastline and
the 30 m deep isobath. Secchi et al. (2003) proposed to divide the distribution area
into four population management units (Franciscana Management Areas, FMAs).
The FMA IV represents the coastal waters of Argentina, including the Buenos Aires
Province, Río Negro, and Chubut. However, the arguments presented for this subdi-
vision are not rigid and are subject to change, as more and better information will
be recorded (Secchi et al. 2003). Subsequent studies, based on new genetic analy-
ses, have suggested at least seven management units should be considered (Mendez
et al. 2008; Cunha et al. 2014; Gariboldi et al. 2016). Studies performed in northern
Buenos Aires Province revealed that in the Samborombón Bay, there would be a
genetically isolated population and suggest that at least there are two populations
only in Buenos Aires coast (Mendez et al. 2008, 2010). Recent studies suggest that
there could be between three and five subpopulations in Argentina (Cunha et  al.
2014; Gariboldi et al. 2016).
14  Marine Mammals: Is the Bahía Blanca Estuary and Its Area of Influence Important… 377

14.2.3.3  Behavior

Franciscanas feed mainly on bottom dwelling fish of the family Sciaenidae


(Rodríguez et al. 2002; Denuncio et al. 2017), but crustacean and mollusks are also
important. They feed mostly near the bottom and appear to be opportunistic, with at
least 58 fish species, six cephalopod species, and six crustacean species known from
the diet. They show cooperative feeding and foraging, and on average, they dedi-
cated three quarters of their time searching for their preys. This dolphin lives in
small groups up to 15 individuals with an average of two to five (Bastida and
Rodríguez 2009). There is no evidence of large seasonal movements and little is
known about daily displacement (Bordino et al. 1999; Bordino 2002). Bordino et al.
(2008), using satellite telemetry, showed that this species has a relatively small
home range (150 km2–345 km2) which barely exceeds 20 linear km to its maximum
extent. In spite of its difficulty to be seen in open sea due to their color, behaviors,
and small size, it is possible to see groups during the summer along the entire coast
of Buenos Aires Province.
Franciscana’s breeding season begin in late spring and summer, and newborn
calves are recorded between the end of October and early April in Bahía
Samborombón and Bahía Anegada with a gestation period between 10 and
11 months (Bordino et al. 1999). Killer whale (Orcinus orca) are known natural
predators and some sharks have attacked franciscanas trapped in fishing nets. They
are seldom observed close to motor boats, suggesting that they are scared and
avoid them (Bastida and Rodríguez 2009). They produce high-frequency narrow-
band echolocation clicks with a maximum frequency recorded 139  kHz and a
bandwidth of 19  kHz (Melcón et  al. 2012) with lower frequencies in calves
(Melcón et al. 2016).

14.2.3.4  Threats and Conservation Status

The main problem this species is facing is incidental mortality in gillnet fisheries.
Every year between 2000 and 3000 dolphins die in coastal fisheries in Brazil,
Uruguay, and Argentina (Bastida and Rodríguez 2009). Only in Buenos Aires coast,
about 500–650 dolphins are entangled every year (Denuncio et  al. 2019). Other
threats include various forms of habitat degradation and pollution. Heavy metals
and several organic chlorine compounds have been detected in individuals of differ-
ent regions (Gerpe et al. 2002; Panebianco et al. 2011, 2012; Romero et al. 2018).
Moreover, more than 30% of the franciscanas which were studied in the Rio de la
Plata estuary and Cabo San Antonio contained plastic debris in their stomach
(Denuncio et al. 2011, 2016).
The Red List of IUCN and SADSN-SAREM categorized this species as
“Vulnerable” (Zerbini et al. 2017; Denuncio et al. 2019). It is the most endangered
cetacean species in the Southwest Atlantic Ocean, listed in Appendix II of CITES.
378 G. Giardino et al.

 ahille’s Bottlenose Dolphin Tursiops truncatus


14.2.4  L
gephyreus (Montagu, 1821)

Common name: Delfin nariz de botella de Lahille, delfin Mular de Lahille (Spanish),
Lahille’s common bottlenose dolphin (English)

14.2.4.1  Description

Robust and large-sized dolphin (max, length 4 m). The adult weight of this South
Western Atlantic subspecies ranges between 200 and 350 kg, but there are maxi-
mum records of 600 kg. At birth calves measure between 85 and 140 cm and weight
between 14 and 30 kg. These general body values are much higher than those of the
other subspecies, Tursiops truncatus truncatus, of the Caribbean region, where they
behave more dynamic. Body color is uniform grey with lighter belly. Large, bul-
bous, and well-differentiated melon. Short and wide beak. Lower jaw slightly
exceeding the upper one. Between 20 and 26 pairs of well-developed teeth in the
upper jaw and between 18 and 24 pairs in the lower jaw. Developed subtriangular
dorsal fin in gephyreus subspecies, and falcate dorsal fin in truncatus subspecies; in
both placed at midback, well-developed typical pectoral fins with a convex anterior
edge and concave posterior edge (Bastida and Rodríguez 2009) (Fig. 14.5).

14.2.4.2  Distribution and Habitat

Lahille’s common bottlenose dolphin is a subspecies distributed along the coastal


waters of Patagonia and Buenos Aires Province; it also inhabits coastal areas of
Uruguay and southern Brazil. In Uruguay and Brazil, the subspecies truncatus is
generally found in offshore waters of the continental shelf. Common bottlenose
dolphins tend to be primarily coastal, but they can also be found in pelagic waters
(Wells and Scott 1999). Individuals that primarily use inshore waters frequents estu-
aries, bays, lagoons, and other shallow coastal regions and occasionally can swim
far up into rivers. Individuals of this ecotype tend to maintain definable, long-term
multi-generational home ranges, but in some locations near the extremes of the spe-
cies range, they show migratory behaviors. On the other hand, the offshore ecotype
is apparently less restricted in range and movement. Some offshore dolphins are
residents around oceanic islands.

14.2.4.3  Behavior

Bottlenose dolphins are commonly associated with many other cetaceans, including
both large whales and other dolphin species (Wells and Scott 1999). This dolphin
species consumes a wide variety of prey, mostly fish and squid (Barros and Odell
14  Marine Mammals: Is the Bahía Blanca Estuary and Its Area of Influence Important… 379

Fig. 14.5  Lahille’s bottlenose dolphin, Tursiops truncatus gephyreus, in Bahía Blanca Estuary.
(Photo by Agustina Mandiola)

1990; Barros and Wells 1998; Blanco et al. 2001; Santos et al. 2001), and some-
times squids, shrimps, and other crustacean. In Buenos Aires Province, Lahille’s
bottlenose dolphins’ diet is mainly based in demersal fish as white mouth croaker,
striped weakfish, and Mugil spp., among other bony fish species (Bastida and
Lichtschein 1986; Mermoz 1977). Lahille’s dolphins use a high variety of prey cap-
ture strategies and techniques, one of these, observed in small streams of Bahía
Samborombón, is to pull mullet fish schools out of water and eat them with the
dolphin body partially stranded (Bastida and Rodríguez 2009). Probably such
behaviors also take place in the shallow areas of Bahía Blanca estuary. This coastal
dolphin can eventually be caught by killer whales; in Argentina, cases of attack have
already been reported in Bahía Samborombón and in Villa Gesell (Buenos Aires
Province) (Bastida and Rodríguez 2009). Probably these attacks may also take place
in the outer zone of the Bahía Blanca estuary. Bottlenose dolphins produce pulsed
signals (echolocation) as well as tonal whistles. These sounds comprise a wide
range of frequencies (40–130 kHz). Moreover, individual sounds known as signa-
ture whistles used in individual recognition and in maintaining group cohesion have
been described. Maritime traffic seems to affect the acoustic behavior of the species
(Buckstaff 2004). This vulnerable dolphin species lives in Argentina in coastal
waters, from Buenos Aires Province to Northern Chubut Province. Although few
records have been made as far south as the provinces of Santa Cruz and Tierra del
Fuego (Goodall et al. 2011). Most behavioral studies conducted on bottlenose dol-
phins in Argentina (e.g., Würsig and Würsig 1979) were discontinued in the 1980s
because of noted population decline and the subsequent lack of sightings (Bastida
and Rodríguez 2009). One of the last remaining resident populations of the country
380 G. Giardino et al.

is suggested to reside in Bahía San Antonio (Río Negro Province) (Vermeulen and
Cammareri 2009). Research conducted in this area described this population as
small, essentially closed, declining (Vermeulen and Bräger 2015), and highly resi-
dent to the study area, indicating this bay as the core region within the larger home
range of this population (Vermeulen and Cammareri 2009).

14.2.4.4  Threats and Conservation Status

The abundance of this species is dramatically decreasing in Argentina, based on


average numbers of sightings since the 1970s, even in areas with increasing or con-
stant survey effort (Bastida and Rodríguez 2009; Vermeulen et al. 2017). For the
Patagonia Sea, it is estimated that there are less than 1000 mature individuals, across
at least two genetically distinct subpopulations, showing genetic evidence of popu-
lation fragmentation (Fruet et al. 2014). Causes of declines in the world – and prob-
ably also in Argentina – are often related to either habitat degradation, prey depletion,
or contamination (Vermeulen and Bräger 2015).
This species is categorized as “Least Concern” by the Red List IUCN (Wells
et al. 2019) and “Vulnerable” by SADSN-SAREM (Vermeulen et al. 2019), listed in
Appendix II of CITES. However, in a preliminary way, the Lahille’s bottlenose
dolphin can be considered the most threatened cetacean currently in Argentina and
probably from Uruguay.

14.2.5  Importance of the Estuary for Marine Mammals

Estuaries have plenty of food and offer coastal protection and habitat for a wide
variety of species, including seabirds, fish, and mammals. Moreover, estuaries are
nursery areas for many fishes (Costanza et al. 1997; Martinho et al. 2007) which are
common prey of marine mammals. In the Bahía Blanca Estuary, 30 fish species
have been reported, where striped weakfishes (Cynoscion guatucupa), whitemouth
croakers (Micropogonias furnieri), and narrownose smooth-hounds (Mustelus
schmitti) are the most important fishing resources (Lopez Cazorla et al. 2004).
Franciscana dolphin, Lahille’s bottlenose dolphin, South American sea lions, and
South American fur seal, the four marine mammals most frequent in Bahía Blanca
Estuary, mainly feed on whitemouth croakers and striped weakfishes of different
sizes. As Failla et al. (2004) mentioned, the distribution and the accessibility of food
may have been a determining factor for the establishment of franciscana dolphins.
Paso-Viola et  al. (2014) found that franciscana dolphins from the Bahía Blanca
Estuary eat striped weakfishes 2.8–28.1 cm long (smaller than commercial size) and
whitemouth croakers 4.1–7.8 cm (far smaller than commercial size). These mam-
mals also eat horse mackerels (Trachurus lathami), Argentine anchovies (Engraulis
anchoita), lantern midshipmans (Porichthys porosissimus), cusk-eels (Raneya
brasiliensis), and Brazilian flatheads (Percophis brasiliensis). Besides fishes, they
14  Marine Mammals: Is the Bahía Blanca Estuary and Its Area of Influence Important… 381

prey on invertebrates as the longfin inshore squids (Loligo sanpaulensis), Tehuelche


octopuses (Octopus tehuelchus), marine shrimps (Artemesia longinaris), and
Argentine red shrimps (Pleoticus muelleri).
Lahille’s bottlenose dolphins from the Buenos Aires Province feed on marine
and estuarine fishes such as the whitemouth croaker, striped weakfish, king weak-
fish (Macrodon ancylodon), drum (Paralonchurus brasiliensis), Argentine menha-
den (Brevoortia pectinate), Brazilian codling (Urophycis brasiliensis), and longfin
inshore squid (Loligo sanpaulensis) (Bastida and Rodríguez 2009). Moreover,
Vermeulen et  al. (2015) had observed bottlenose dolphins feeding on silversides
(Odontesthes sp.). All these species detected as prey for bottlenose dolphins are
frequent fishes of the Bahía Blanca Estuary (see Chap. 11 in this book).
On the other hand, the otariids recorded in Bahía Blanca Estuary are considered
generalist feeders. South American sea lions from Puerto Quequén mostly eat cusk-­
eels, striped weakfishes, horse mackerels, Argentine croakers (Umbrina canosai),
flounders (Paralichthys sp.), and skates (Giardino 2014). South American fur seals
from Buenos Aires Province eat whitemouth croakers, Argentine anchovies, king
weakfishes, striped weakfishes, and squids (Pérez Salles 2015). Except for ancho-
vies, otariids eat fishes smaller than those valuable for commercial interest, indicat-
ing that sea lions and fur seals feed mainly on juveniles (Giardino 2014; Pérez
Salles 2015).
In addition, the Bahía Blanca Estuary offers a sheltered environment. It is known
that relatively small communities of bottlenose dolphins, living in protected coastal
environments with predictable availability of resources, often display a high degree
of residency and long-term site fidelity, as Vermeulen et al. (2017) confirmed for
Argentine groups. Franciscanas have low mobility, with prolonged associations
(Wells et al. 2013) in small groups (Crespo et al. 1998; Failla et al. 2004). This dol-
phin needs turbid waters, depths ranging between 5 and 35 m, favorable conditions
for feeding, and protection against natural predators (Bordino et  al. 1999; Failla
et al. 2004), thus the estuary is an ideal environment. Molecular analyses, carried
out in franciscanas from Buenos Aires Province, allowed to recognize a reduced
mobility of the species and a possibly high level of population isolation (Mendez
et al. 2008, 2010). Sea lions, on the other hand, need a wintering area where they
rest and feed, away from females and their puppies, thus avoiding competition for
food. Thus, the Bahía Blanca Estuary provides a suitable wintering habitat for them.
There are also negative aspects that have to be considered, as the estuary accu-
mulates large amounts of pollutants, becoming a potential threat to resident species.
The Bahía Blanca Estuary has experienced a marked human population increase as
well as industrial development during the past decades (Marcovecchio et al. 2008).
This coastal area also supports an intensive anthropogenic activity, including five
large harbors and one of the biggest industrial parks in South America comprising
refineries, oil terminals, tanks for storing oil products, and multiple docks (Limbozzi
and Leitao 2008; Oliva et al. 2017). Pollution, environmental contaminants, marine
noise, plastic debris, offshore oil and gas activities, shipping, and commercial fish-
eries affect marine mammals. Individuals can die or be negatively impacted by
entanglement or the ingestion of plastic litter (Reijnders et al. 2018). At local level
382 G. Giardino et al.

plastic debris were detected in franciscanas (Denuncio et al. 2011) and in fur seals
(Denuncio et al. 2017). In franciscanas high levels of metals, such as mercury, zinc,
and cooper, were found (Gerpe et al. 2002; Panebianco et al. 2012). In addition,
bottlenose dolphins, South American sea lions, and fur seals from Buenos Aires
Province have mercury, zinc, cadmium, and copper in their tissues (Marcovecchio
et al. 1990; Moreno et al. 1984; Marcovecchio et al. 1994). High concentration of
metals and organic pollutants in the tissues of marine mammals has been associated
with organ anomalies, impaired reproduction, and immune function and, as a con-
sequence of the latter, with the occurrence of large die-offs among seals and ceta-
cean species (Reijnders et al. 2018). However, a clear cause and effect relationship
between residual levels of contaminants and observed effects has been demonstrated
in only a few studies.
No less important is noise pollution. Underwater noise can interfere with key life
functions of marine mammals (e.g., foraging, mating, nursing, resting, migrating)
by impairing hearing sensitivity, masking acoustic signals, eliciting behavioral
responses, or causing physiological stress (Erbe et al. 2018). Although in Argentina
no studies regarding noise pollution in local species have been developed, world-
wide disturbances of shipping traffic, military tests, and oil extraction, among others
were found (Rolland et al. 2012; Tyack and Janik 2013; Fouda et al. 2018; Simonis
et al. 2020). As the Bahía Blanca Estuary has a significant traffic vessel, this type of
pollution also should be considered.

14.2.6  Local Activities for the Conservation


of Marine Mammals

Stranding records provide valuable information about spatial distribution, seasonal


movements, and factors related to mortality (Moore et al. 2018). Observations of
stranded marine fauna on the beach are not new, but are more frequent nowadays.
Many causes, natural or anthropogenic, or both, could lead to the stranding of live,
dying, or dead animals. Natural causes include failure to thrive, predation, disease,
parasites, injuries, and exposure to biotoxins. Other natural threats, such as prey
distribution changes, can be driven by environmental fluctuations or overfishing
(Moore et al. 2018). Directly and indirectly, humans are seriously affecting marine
life. The main anthropic disturbances, among others, that cause pain, suffering, and
often death include incidental entanglement, both in coastal and commercial fisher-
ies; collision with boats, drift nets, and other fishing gear; oil spills; solid waste
floating in the sea; and ocean noise pollution (Moore et al. 2018). In the last three
decades, these events started to be considered by national and provincial govern-
ment agencies, and different public policies were adopted. At the same time, pri-
vate rehabilitation centers have emerged, for rescue, assistance, and research.
These institutions have been coordinating efforts even inter-jurisdictionally, mak-
ing progress in management and conservation along the extensive Atlantic coast of
Argentina.
14  Marine Mammals: Is the Bahía Blanca Estuary and Its Area of Influence Important… 383

Government agencies are responsible for marine fauna. At the national level, the
Ministry of Environment and Sustainable Development (SAyDS) has the compe-
tence to promote regulations and management actions for the conservation of
aquatic resources. In Buenos Aires Province, the Provincial Direction of Natural
Resources – Provincial Agency for Sustainable Development (OPDS) has a network
that oversees the actions of collection centers, share scientific information with
researchers, and promote a better knowledge of ecology, biology, and veterinary
aspects necessary for the conservation of the species.
Locally, in the 1980s, Bahía Blanca Estuary was declared a Protected Provincial
Natural Area; therefore, protection and conservation measures began to be taken
through the legal figure as Bahía Blanca, Bahía Falsa, and Bahía Verde Multiple Use
Nature Reserve. Following work began to be done at ecosystem level. A few years
later, the environmental non-governmental organization Marine Animals Assistance
and Rescue Foundation (FRAAM) was created. Biologists and park rangers started
to work on the environmental awareness with coastal populations adjacent to the
Natural Reserve area, the education continues nowadays. Through environmental
education and interpretation strategies, FRAAM offers training to different social
actors involved in the occasional discovery and assistance of marine fauna, such as
the Argentine Naval Prefecture, park rangers, artisanal fishermen, recreational fish-
ermen, lifeguards, and rescue center volunteers. Primary care assistance includes
direct measures focused on animal health and welfare: systematic data collection,
biometrics, epibionts, necropsies, parasites, and pathological, genetics, and con-
taminants studies.
In the last decade, work on data collection has been strengthened along the
Atlantic coast of Buenos Aires Province, coordinated efforts with national, provin-
cial, multidisciplinary academic groups of national universities, non-governmental
rehabilitation centers, as well as artisanal and recreational fishermen. Altogether,
they use the same protocols for each taxonomic group. All the information recorded
is shared among all the organisms to the National Action Programs for the
Conservation of each one of these species (e.g., PAN-Marine Mammal Conservation).
Accidental observers, in coincidence with the summer touristic season, reported
most of the records. In this sense, FRAAM has an Environmental Education
Program to teach school children and teachers of all levels. In addition, dissemina-
tion and information campaigns are carried out through the media, reinforced with
billboards and brochures. Nowadays, social media contribute and allow direct com-
munication with the informant. However, there is still much to do and direct inter-
vention by the public without knowledge is discouraged. For example, it is not
recommended to feed the animals, force them to return to sea, water, or try to catch
them. Instead, keeping distance and giving immediate notice to the nearest beach
authority helps. In addition, if possible people that discover a stranded animal
should take photographs, keep dogs away, and await the arrival of the authority
without putting themselves at any risk.
Not always animals on the beach are sick, injured, or in trouble. In the case of
pinniped (seals and sea lions), for example, as they alternate periods of their life at
sea and periods in land, they often go out to rest in places far from their settlements,
384 G. Giardino et al.

simply to recover energy. Opposite, cetacean live all their life in the sea, so their
stranding is not a normal condition. Cetaceans can single or mass strand and it is
difficult to know the real reason behind it. However, through examinations, nec-
ropsy, and sampling of deceased animals, it is possible to understand direct and
indirect threats to marine mammal populations (Moore et al. 2018). Each stranding
requires logistics and participatory collaboration of different institutions and profes-
sionals. The faster an animal can be examined, the more accurate the diagno-
sis can be.
In the Bahía Blanca Estuary, when an animal is found alive, a first diagnosis is
made on its body condition, following the corresponding sanitary protocol. Next, all
possible samples and information are collected (biometrics, sex, epibionts, stage of
development, geographic location, pictures, etc.). Meanwhile, authorities delimit an
area to prevent disturbance from people and dogs. If the beach is full of people,
authorities move the animal to a more remote place. If it is necessary due to the
health condition and if the animal size allows it, it is transferred to the rehabilitation
center for veterinary assistance. If the finding is a post-mortem animal, recom-
mended proceeding is the same, followed by a necropsy and carcass recovery.
Species found along the beach may be residents or transients. Transient species pass
through out provincial limits through their lives, or even cross-national and interna-
tional boundaries. For that reason, it is necessary to have coordinated work between
governmental and non-governmental organizations with economic and human
resources optimization.
Understanding how ecosystems function is essential; therefore, it is necessary to
work in coordination, both from official organizations such as universities and non-­
governmental organizations, like FRAAM, for the species conservation and habitat
preservation. The best way to do this is sharing methodologies, protocols, and sci-
entific information.

Box 14.1 Bahía Blanca Estuary: A Point of Connection Between the South
American Breeding Colonies of Sea Lions
South American sea lions (Otaria flavescens) live along the coast of South
America, from Torres, southern Brazil (29° 20′ S; 49° 43′ W) in the Atlantic
Ocean, to Punta Brava, Ecuador (02° 12′ S; 81° 00′ W) in the Pacific Ocean
(Vaz Ferreira 1982; Bastida et al. 2007). Along the Argentine coast, there are
60 settlements, with an estimated total population of more than 200,000 indi-
viduals (Romero et al. 2019); number that is still far from historical popula-
tions (Bastida et al. 2007). The Buenos Aires coast has four rookeries, located
at almost equal distances (ca. 1000 km) from two of the most important focal
breeding areas of this species in the southwest Atlantic (Uruguay and Northern
Patagonia). These four colonies are non-breeding rookeries, two of them
located in Mar del Plata (38° 02′ S, 57° 31′ W) and in Quequén (38° 35′ S; 58°
42′ W) harbors (Fig. 1.2; Chap. 1) and the other two in Isla Trinidad (39° 13′
59″ S; 61° 51′ 14″ W) and Banco Culebra (40° 24′ 30″ S; 61° 58′ 30″ W) (Fig.
2.1; Chap. 2).
14  Marine Mammals: Is the Bahía Blanca Estuary and Its Area of Influence Important… 385

Isla Trinidad is within the limits of the Bahía Blanca, Bahía Falsa, and
Bahía Verde Multiple Use Natural Reserve. The South American sea lions and
South American fur seals are commonly observed within the reserve (Petracci
et al. 2010; Giardino 2014; Martín Sotelo and Victoria Massola pers. obs.).
However, the area of Isla Trinidad has three large bays and an intricate net-
work of channels and streams with extensive sandy beaches in the outermost
islands. Because of the difficult access, surveys must be performed by boat or
plane, and details on the fauna inhabiting Isla Trinidad is largely unknown.
The status of the haulout in Isla Trinidad is almost unknown. Its existence
has been a subject of controversy, because of confusing quotes and erroneous
locations in the literature published over the past few years. By 1935 one his-
torical record mentions the killing of up to 100 sea lions per day in Isla
Trinidad, by the Salvador Di Meglio and Minujin Society (Fidalgo 2004). Dr.
Raúl Arámburu, during a census conducted on December 1973, quantified
900 sea lions in the area (Vaz Ferreira 1982). One of the most recent refer-
ences on the status of the population remarks that this rookery shelters on
average 76 South American sea lions, out of the breeding season with a maxi-
mum of 150 individuals (Petracci et al. 2010; Giardino 2014). Like Mar del
Plata and Quequén rookeries, Isla Trinidad, have, predominantly, juvenile
(3–5  years) and subadult (5–7  years) males (Petracci et  al. 2010; Giardino
2014). Sea lions from Isla Trinidad are also connected with the other winter-
ing grounds in the Buenos Aires coast. In April 2009, one of the sea lions
bleached (Giardino et  al. 2013) in Puerto Quequén, between July and
December 2008, was resighted in Isla Trinidad confirming that connection
(Giardino 2014).
Why estuaries are so important for sea lions? Estuaries play an important
role in coastal regions. Their plant communities provide protection against the
erosion of water and wind. In addition, bottom communities allow for sedi-
ment oxygenation and many species of fish and invertebrates carry out part of
their life cycles in these environments. Several of these fish species are of
commercial interest and are prey for apex predators such as many marine
mammals and seabirds. Among the services provided, estuaries commonly
act as nursery areas for fish (Bortolus 2008). Adults and juveniles of the most
frequent preys of Otaria flavescens as the Brazilian menhaden (Brevoortia
aurea), the whitemouth croaker (Micropogonias furneri), and the striped
weakfish (Cynoscion guatucupa) (Lopez Cazorla et al. 2014; Giardino 2014)
inhabit the Bahía Blanca Estuary.
Female sea lions remain close to the breeding site (Rodríguez et al. 2013;
Grandi et  al. 2008, 2018) nursing their calves, but males do not have any
parental investment. Males tend to disperse away from their mating sites
toward unisexual haulouts. Thus, male sea lions remain in remote places (as
the Bahía Blanca Estuary) during most of the year in order to avoid competing
with females on feeding resources (Giardino et al. 2016).
386 G. Giardino et al.

Like in the other male haulouts from Buenos Aires, sea lions from Isla
Trinidad would be a functional part of the northern Patagonia and Uruguay
breeding aggregations. Every year, during the austral summer, sexually and
socially mature sea lions from Buenos Aires travel to the breeding grounds,
returning once the breeding season is over (Giardino et al. 2016). With these
movements, males maintain genetic flow between the two different genetic
stocks of females (Szapkievich et al. 1999; Túnez et al. 2007, 2010; Feijoo
et al. 2011; Giardino et al. 2016; de Oliveira et al. 2017). This connection is
particularly important for the long-term persistence of this species (Frankham
et al. 2002; Crooks and Sanjayan 2006). Furthermore, these wintering grounds
are also relevant as passage, rest, feeding, training, and maturation areas.
Regarding the South American fur seal (Arctocephalus australis), this spe-
cies is in a process of redistribution and recolonization, and it is observed with
increasing frequency in the Bahía Blanca Estuary, settling on buoys close to
Ingeniero White Port (Mandiola 2015). Even though there are no current stud-
ies on this fur seal in Bahía Blanca, these otariids may also be connected with
Uruguayan colonies, as proven by one animal bleached in Mar del Plata and
sighted afterward in Punta de Lobos, Uruguay.
Connectivity and dispersal are key factors for the long-term persistence of
species, particularly in those with reduced populations or fragmented habitats
(Frankham et al. 2002; Crooks and Sanjayan 2006). Based on the available
information, Isla Trinidad, and in consequence the Bahía Blanca Estuary,
would represent a key point of connection and should be seriously considered
for regional conservation strategies.

Acknowledgments  All this work would be impossible without the help of artisanal fishermen
from Villa del Mar, Puerto Rosales, Ingeniero White, Mariano Natali, Prefectura Naval Argentina
(from Monte Hermoso, Coronel Rosales, Bahía Blanca), park rangers as Martin Sotelo, Ezequiel
Matías Chiatti Heinz, Ariel Tombo, Daniel Jofre and Ariadna Mamani, Dra. Suldrup, Club Nautico,
Oscar Liberman, Ing. Adrían Daño, Verónica Lombad, Sr. Mansi, and all the local people and tour-
ists who help to recover this valuable information.

References

Ainley DG, Henderson RP, Huber HR et al (1985) Dynamics of white shark/pinniped interactions
in the Gulf of the Farallones. South Calif Acad Sci Mem 9:109–122
Amon DJ, Glover AG, Wiklund H et al (2013) The discovery of a natural whale fall in the Antarctic
deep sea. Deep Sea Res Part II 92:87–96. https://doi.org/10.1016/j.dsr2.2013.01.028
Anderson PK (1995) Competition, predation, and the evolution and extinction of Steller’s sea cow.
Mar Mamm Sci 11:391–394. https://doi.org/10.1111/j.1748-­7692.1995.tb00294.x
Andriolo A, Kinas PG, Engel MH et al (2010) Humpback whales within the Brazilian breeding
ground: distribution and population size estimate. Endanger Species Res 11:233–243
14  Marine Mammals: Is the Bahía Blanca Estuary and Its Area of Influence Important… 387

Angeletti S, Cervellini PM, Massola V (2014) Nuevo registro de ballena jorobada (Megaptera
novaeangliae) para el Mar Argentino y notas sobre sus epibiontes. Mastozool Neotrop
21(2):319–324
Arbiza J, Blanc A, Castro-Ramos M et al (2012) Uruguayan pinnipeds (Arctocephalus australis
and Otaria flavescens): evidence of influenza virus and Mycobacterium pinnipedii infections.
In: Romero A, Keith EO (eds) New approaches to the study of marine mammals. Tech Open,
Rijeka, p 248
Barros NB, Odell DK (1990) Food habits of bottlenose dolphins in the southeastern United States.
In: Leatherwood S, Reeves RR (eds) The bottlenose dolphin. Academic, New York, pp 309–328
Barros NB, Wells RS (1998) Prey and feeding patterns of resident bottlenose dolphins (Tursiops trun-
catus) in Sarasota Bay, Florida. J Mammal 79(3):1045–1059. https://doi.org/10.2307/1383114
Bastida R, Lichtschein V (1986) Capturas incidentales de pequeños cetáceos en el área de Mar del
Plata (Provincia de Buenos Aires-Argentina). In: Actas, I Reunión de Trabajo de Expertos en
Mamíferos Acuáticos de América del Sud, 24–29 June, Buenos Aires, pp 14–22
Bastida R, Rodríguez DH (2009) Mamíferos marinos de Patagonia y Antártida. Vazquez Mazzini,
Buenos Aires
Bastida R, Rodriguez D, Moron S (2000) Avistajes costeros y tamaño grupal de Pontoporia bla-
invillei en el sudeste de la Provincia de Buenos Aires (Argentina). In: Crespo EA (ed) Report
of the 3rd workshop for coordinated research and conservation of the Franciscana dolphin
(Pontoporia blainvillei) in the Southwestern Atlantic. Bonn, UNEP/CMS Secretariat, pp 51–54
Bastida R, Rodríguez D, Secchi E et al (2007) Mamíferos Acuáticos de Sudamérica y Antártida.
Vazquez Mazzini Editores, Buenos Aires
Bastida R, Quse V, Guichón R (2011) La tuberculosis en grupos de cazadores recolectores de
Patagonia y Tierra del Fuego: nuevas alternativas de contagio a través de la fauna silvestre. R
Arg Antrop Biol 13:83–95
Baylis AM, Tierney M, Orben RA et al (2018a) Geographic variation in the foraging behaviour of
South American fur seals. Mar Ecol Prog Ser 596:233–245. https://doi.org/10.3354/meps12557
Baylis AM, Tierney M, Staniland IJ et al (2018b) Habitat use of adult male South American fur
seals and a preliminary assessment of spatial overlap with trawl fisheries in the South Atlantic.
Mamm Biol 93(1):76–81. https://doi.org/10.1016/j.mambio.2018.07.007
Bernardelli A, Bastida R, Loureiro J et al (1996) Tuberculosis in sea lions and fur seals from the
south-western Atlantic coast. Rev Sci Tech 15:985–1005
Beron Vera B, Crespo EA, Raga A et al (2004) Uncinaria hamiltoni (Nematoda: Ancylostomatidae)
in South American sea lions, Otaria flavescens, from northern Patagonia, Argentina. J Parasitol
90:860–863. https://doi.org/10.1645/GE-­182R
Berta A, Sumich JL, Kovacs KM (2015) Chapter 1: Introduction. In: Berta A, Sumich JL, Kovacs
KM (eds) Marine mammals, 3rd edn. Academic, San Diego, pp 1–14
Blanco C, Salomon O, Raga JA (2001) Diet of the bottlenose dolphin (Tursiops truncatus) in
the western Mediterranean Sea. J Mar Biol Assoc UK 81:1053–1058. https://doi.org/10.1017/
S0025315401005057
Bordino P (2002) Movement patterns of franciscana dolphins (Pontoporia blainvillei) in Bahia
Anegada, Buenos Aires, Argentina. LAJAM 1(1):71–76. https://doi.org/10.5597/lajam00011
Bordino P, Thompson G, Iñiguez M (1999) Ecology and behaviour of the franciscana (Pontoporia
blainvillei) in Bahía Anegada, Argentina. J Cetacean Res Manag 1:213–222
Bordino P, Wells RS, Stamper MA (2008) Satellite tracking of Franciscana dolphins Pontoporia
blainvillei. In: Argentina: preliminary information on ranging, diving and social patterns.
International whaling commission scientific committee meeting, SC60/SM14, Santiago de
Chile, p 10
Bornatowski H, Wedekin LL, Heithaus MR et al (2012) Shark scavenging and predation on ceta-
ceans at Abrolhos Bank, eastern Brazil. J Mar Biol Assoc UK 92(8):1767–1772. https://doi.
org/10.1017/S0025315412001154
Bortolus A (2008) Influencia de los ambientes costeros patagónicos sobre los ecosistemas
marino-oceánicos: las marismas como caso de estudio. In: Foro para la Conservación del Mar
Patagónico y Áreas de Influencia (eds) Estado de conservación del Mar Patagónico y áreas de
influencia, pp 80–104
388 G. Giardino et al.

Bos KI, Harkins KM, Herbig A et al (2014) Pre-Columbian mycobacterial genomes reveal seals
as a source of New World human tuberculosis. Nature 514:494–497. https://doi.org/10.1038/
nature13591
Botta S, Secchi ER, Muelbert MMC et al (2010) Age and growth of franciscana dolphins, Pontoporia
blainvillei (Cetacea: Pontoporiidae) incidentally caught off southern Brazil and northern
Argentina. J Mar Biol Assoc UK 8:1493–1500. https://doi.org/10.1017/S0025315410001141
Buckstaff KC (2004) Effects of watercraft noise on the acoustic behavior of bottlenose dolphins,
Tursiops truncatus, in Sarasota Bay, Florida. Mar Mamm Sci 20(4):709–725
Budge SM, Springer AM, Iverson SJ et  al (2008) Blubber fatty acid composition of bowhead
whales, Balaena mysticetus: implications for diet assessment and ecosystem monitoring. J Exp
Mar Biol Ecol 359:40–46. https://doi.org/10.1016/j.jembe.2008.02.014
Bustos RL, Daneri GA, Volpedo AV et al (2012) The diet of the South American sea lion (Otaria
flavescens) at Río Negro, Patagonia, Argentina, during the winter-spring period. Iheringia Sér
Zool 102:394–400
Butterworth A (2016) A review of the welfare impact on pinnipeds of plastic marine debris. Front
Mar Sci 3:149. https://doi.org/10.3389/fmars.2016.00149
Campagna C (1985) The breeding cycle of the Southern sea lion, Otaria byronia. Mar Mamm Sci
1:210–218. https://doi.org/10.1111/j.1748-­7692.1985.tb00010.x
Campagna C, Le Boeuf BJ (1988) Reproductive behaviour of southern sea lions. Behaviour
104:233–261. https://doi.org/10.1163/156853988X00539
Campagna C, Werner R, Karesh W et al (2001) Movements and location at sea of South American sea
lions (Otaria flavescens). J Zool 257:205–220. https://doi.org/10.1017/S0952836901001285
Cappozzo HL (1995) Comportamiento Reproductivo en dos Especies de Otáridos de América del
Sur. Tesis de Doctorado, Universidad de Buenos Aires, Buenos Aires, Argentina
Cardenas-Alayza S, Crespo EA, Oliveira L (2016) Otaria byronia, South American sea lion. The
IUCN Red List of Threatened Species 2016:e.T41665A61948292. https://doi.org/10.2305/
IUCN.UK.2016-­1.RLTS.T41665A61948292.en. Accessed 3 Mar 2020
Cockcroft VG, Cliff G, Ross JB (1989) Shark predation on Indian Ocean bottlenose dolphins
Tursiops truncatus off Natal, South Africa. S Afr J Zool:24305–24310. https://doi.org/10.108
0/02541858.1989.11448168
Coscarella MA, Bellazzi G, Gaffet ML et  al (2015) Technique used by killer whales (Orcinus
orca) when hunting for dolphins in Patagonia, Argentina. Aquat Mamm 41:192–197. https://
doi.org/10.1578/am.41.2.2015.192
Costanza R, d’Arge R, De Groot R et al (1997) The value of the world’s ecosystem services and
natural capital. Nature 387(6630):253–260. https://doi.org/10.1038/387253a0
Crespi-Abril AC, Garcia NA, Crespo EA et al (2003) Consumption of marine mammals by broad-
nose sevengill shark Notorynchus cepedianus in the northern and central Patagonian shelf. Lat
Am J Aquat Mamm 2(2):101–107
Crespo EA, Pedraza SN, Dans SL et al (1997) Direct and indirect effects of the highseas fisheries
on the marine mammal populations in the northern and central Patagonian coast. J Northwest
Atl Fish Sci 22:189–207
Crespo EA, Harris G, González R (1998) Group size and distributional range of the francis-
cana, Pontoporia blainvillei. Mar Mamm Sci 14(4):845–849. https://doi.org/10.1111/j.1748-
­7692.1998.tb00768.x
Crespo EA, Pedrasa SN, Grandi MF et  al (2010) Abundance and distribution of endangered
Franciscana dolphins in Argentine waters and conservation implications. Mar Mamm Sci
26:17–35. https://doi.org/10.1111/j.1748-­7692.2009.00313.x
Crespo EA, Schiavini AC, García NA et al (2015) Status, population trend and genetic structure
of South American fur seals, Arctocephalus australis, in southwestern Atlantic waters. Mar
Mamm Sci 31:866–890. https://doi.org/10.1111/mms.12199
Croll DA, Kudela R, Tershy BR (2006) Ecosystem impacts of the decline of large whales in the
North Pacific. In: Estes JA, De Master DP, Doak DF et al (eds) Whales, whaling and ocean
ecosystems, University of California Press, Berkeley Connectivity Conservation. Cambridge
University Press, Cambridge, pp 202–2014
Crooks KR, Sanjayan M (2006) Connectivity conservation, vol 14. Cambridge University Press,
Cambridge
14  Marine Mammals: Is the Bahía Blanca Estuary and Its Area of Influence Important… 389

Cunha HA, Medeiros BV, Barbosa LA et al (2014) Population structure of the endangered francis-
cana dolphin (Pontoporia blainvillei): reassessing management units. PLoS One 9(1). https://
doi.org/10.1371/journal.pone.0085633
Dans SL, Sielfeld W, Aguayo A et al (2012) Estado y tendencia de las poblaciones. In: Crespo EA,
Oliva D, Dans SL et al (eds) Estado desituación del lobo marino común Otaria flavescens en
su área de distribución. Universidad de Valparaiso, Valparaiso, pp 18–35
Dans SL, Crespo EA, Coscarella MA (2017) Wildlife tourism: underwater behavioral responses
of South American sea lions to swimmers. Appl Anim Behav Sci 188:91–96. https://doi.
org/10.1016/j.applanim.2016.12.010
Dassis M, Farenga M, Bastida R (2012) At-sea behavior of South American fur seals: influence
of coastal hydrographic conditions and physiological implication. Mamm Biol 77(1):7–52.
https://doi.org/10.1016/j.mambio.2011.07.003
De Castro RL, Saporiti F, Vales DG et al (2016) What are you eating? A stable isotope insight into
the trophic ecology of short-beaked common dolphins in the Southwestern Atlantic Ocean.
Mamm Biol 81:571–578. https://doi.org/10.1016/j.mambio.2016.07.003
de Oliveira LR, MCM G, Fraga LD et al (2017) Ancient female philopatry, asymmetric male gene
flow, and synchronous population expansion support the influence of climatic oscillations on
the evolution of South American sea lion (Otaria flavescens). PLoS One 12:e0179442. https://
doi.org/10.1371/journal.pone.0179442
Denuncio PE, Bastida R, Dassis M et  al (2011) Plastic debris ingested by franciscana dol-
phins, Pontoporia blainvillei. Mar Pollut Bull 62:1836–1841. https://doi.org/10.1016/j.
marpolbul.2011.05.003
Denuncio PE, Panebianco MV, Del Castillo V et  al (2016) Beak deviations in the skull of
Franciscana dolphins (Pontoporia blainvillei) from Argentina. Dis Aquat Org 120:1–7. https://
doi.org/10.3354/dao03012
Denuncio PM, Mandiola MA, Pérez Salles SB et al (2017) Marine debris ingestion by the South
American Fur Seal from the Southwest Atlantic Ocean. Mar Pollut Bull 122:420–425. https://
doi.org/10.1016/j.marpolbul.2017.07.013
Denuncio PE, Paso Viola N, Cáceres-Saez I et al (2019) Pontoporia blainvillei. Categorización
2019 de los mamíferos de Argentina según su riesgo de extinción. Lista Roja de los mamíferos
de Argentina. Versión digital: http://cma.sarem.org.ar
Dewar WK, Bingham RJ, Iverson RL et al (2006) Does the marine biosphere mix the ocean? J Mar
Res 64:541–561. https://doi.org/10.1357/002224006778715720
Erbe C, Dunlop R, Dolman S (2018) Effects of noise on marine mammals. In: Slabbekoorn H et al
(eds) Effects of anthropogenic noise on animals. Springer handbook of auditory research, vol
66. Springer, New York, pp 277–309
Estes JA, Tinker MT, Williams TM et al (1998) Killer whale predation on sea otters linking oceanic
and nearshore ecosystems. Science 282:473–476. https://doi.org/10.1126/science.282.5388.473
Failla M, Iñiguez MA, Tossenberger V et al (2004) Bioecology of Franciscana (Pontoporia blain-
villei) in Northern Patagonia, Argentina. In: 56th annual meeting of the Scientific Committee
of the International Whaling Commission, Sorrento, p 4
Feijoo M, Lessa EP, Loizaga de Castro R et al (2011) Mitochondrial and microsatellite assessment
of population structure of South American sea lion (Otaria flavescens) in the Southwestern
Atlantic Ocean. Mar Biol 158:1857–1867. https://doi.org/10.1007/s00227-­011-­1697-­4
Fidalgo GL (2004) Mamíferos Marinos. In: Piccolo MC, Hoffmeyer MS (eds) Ecosistema del
Estuario de Bahía Blanca. Instituto Argentino de Oceanografía, Bahía Blanca, pp 221–227
Fouda L, Wingfield JE, Fandel AD et al (2018) Dolphins simplify their vocal calls in response to
increased ambient noise. Biol Lett 14(10):20180484
Franco-Trecu V (2005) Comportamiento maternal y aspectos reproductivos de Arctocephalus aus-
tralis en Isla de Lobos – Uruguay. Tesis de grado, Universidad de la República, Uruguay
Frankham R, Ballou JD, Briscoe DA (2002) Introduction to conservation genetics. Cambridge
University Press, Cambridge
Fruet PF, Secchi ER, Daura-Jorge F et  al (2014) Remarkably low genetic diversity and strong
population structure in common bottlenose dolphins (Tursiops truncatus) from coastal waters
of the Southwestern Atlantic Ocean. Conserv Genet 15(4):879–895. https://doi.org/10.1007/
s10592-­014-­0586-­z
390 G. Giardino et al.

Gariboldi MC, Túnez JI, Failla M et al (2016) Patterns of population structure at microsatellite
and mitochondrial DNA markers in the franciscana dolphin (Pontoporia blainvillei). Ecol Evol
6(24):8764–8776. https://doi.org/10.1002/ece3.2596
Gerpe MS, Rodriguez D, Moreno VJ et al (2002) Accumulation of heavy metals in the Franciscana
(Pontoporia blainvillei) from Buenos Aires province, Argentina. LAJAM 1:95–106. https://doi.
org/10.5597/lajam00013
Giardino G (2014) Estructura y dinámica de las colonias de lobos marinos de un pelo de la
Provincia de Buenos Aires, y su relación con pesquerías de la región. PhD thesis, Universidad
Nacional de Mar del Plata
Giardino G, Mandiola A, Bastida J et al (2013) Técnica de marcado por decoloración de pelo en
el lobo marino Otaria flavescens: descripción y evaluación del método. Mastozool Neotrop
20(2):393–398
Giardino GV, Mandiola MA, Bastida J et al (2016) Travel for sex: long range breeding dispersal and
winter haulout fidelity in southern seal lion males. Mamm Biol Zeitschrift für Säugetierkunde
81:89–95. https://doi.org/10.1016/j.mambio.2014.12.003
Giardino G, Bastida J, Mandiola MA et  al D (2017) Estimated population size of two South
American sea lion male haulouts from the northern coast of Argentina. Mammalia
81(2):197–202 doi:https://doi.org/10.1515/mammalia-­2015-­0147
Gil MN, Torres A, Harvey M et  al (2006) Metales pesados en organismos marinos de la zona
costera de la Patagonia Argentina continental. R Biol Mar Ocean 41:167–176. https://doi.
org/10.4067/S0718-­19572006000200004
Goodall RNP, Marchesi MC, Pimper LE et al (2011) Southernmost records of bottlenose dolphins,
Tursiops truncatus. Polar Biol 34(7):1085–1090. https://doi.org/10.1007/s00300-­010-­0954-­1
Grandi MF, Dans SL, Crespo EA (2008) Social composition and spatial distribution of colonies in
an expanding population of South American sea lions. J Mammal 89:1218–1228. https://doi.
org/10.1644/08-­MAMM-­A-­088.1
Grandi MF, Loizaga de Castro R, Terán E et al (2018) Is recolonization pattern related to female
philopatry? An insight into a colonially breeding mammal. Mamm Biol 89:21–29. https://doi.
org/10.1016/j.mambio.2017.12.002
Hain JHW, Ellis SL, Kenney RD et al (1995) Apparent bottom feeding by humpback whales on
Stellwagen Bank. Mar Mamm Sci 11(4):464–479. https://doi.org/10.1111/j.1748-­7692.1995.
tb00670.x
Harcourt R (1993) Individual variation in predation on fur seals by southern sea lions (Otaria
byronia) in Peru. Can J Zool 71(9):1908–1911. https://doi.org/10.1139/z93-­273
Harwood J (2001) Marine mammals and their environment in the twenty-first century. J Mammal
82:630–640. https://doi.org/10.1644/1545-­1542(2001)082<0630:mmatei>2.0.co;2
Jefferson TA, Webber MA, Pitman RL (2015) Marine mammals of the world – a comprehensive
guide to their identification. Academic, London
Kasuya T, Brownell RL Jr (1979) Age determination, reproduction, and growth of the franciscana
dolphin, Pontoporia blainvillei. Sci Rep Whales Res Inst 31:45–67
Kiers A, Klarenbeek A, Mendelts B et  al (2008) Transmission of Mycobacterium pinnipedii to
humans in a zoo with marine mammals. Int J Tuberc Lung Dis 12:1469–1473
Koen Alonso M, Crespo EA, Pedraza SN et al (2000) Food habits of the South American sea lion,
Otaria flavescens, off Patagonia, Argentina. Fish Bull 98:250–263
Kovacs KM, Aguilar A, Aurioles D et  al (2012) Global threats to pinnipeds. Mar Mamm Sci
28:414–436. https://doi.org/10.1111/j.1748-­7692.2011.00479.x
Lavery TJ, Roudnew B, Gill P et al (2010) Iron defecation by sperm whales stimulates carbon
export in the Southern Ocean. Proc R Soc B Biol Sci 277:3527–3531. https://doi.org/10.1098/
rspb.2010.0863
Leeney RH, Amies R, Broderick AC et al (2008) Spatio-temporal analysis of cetacean strandings
and bycatch in a UK fisheries hotspot. Biodivers Conserv 17:2323. https://doi.org/10.1007/
s10531-­008-­9377-­5
14  Marine Mammals: Is the Bahía Blanca Estuary and Its Area of Influence Important… 391

Lewis M, Ximenez I (1983) Dinámica de la población de Otaria flavescens en el área de


Península Valdés zonas adyacentes (Segunda parte). Contribución No. 79, Centro Nacional
Patagónico, p 21
Limbozzi F, Leitao TE (2008) Characterization of Bahía Blanca main existing pressures and their
effects on the state indicators for surface and groundwater quality. In: Neves R, Baretta J,
Mateus M (eds) Perspectives on integrated coastal zone management in South America. IST
Press, Lisboa, pp 333–349
López Cazorla, A. (2004). Peces. In: Piccolo MC, Hoffmayer MS (eds) El Ecosistema del Estuario
de Bahía Blanca. IADO, Bahía Blanca, Argentina, 191–201
Lopez Cazorla A, Molina JM, Ruarte C (2014) The artisanal fishery of Cynoscion guatucupa in
Argentina: exploring the possible causes of the collapse in Bahía Blanca estuary. J Sea Res
88:29–35. https://doi.org/10.1016/j.seares.2013.12.016
Lundsten L, Schlining KL, Frasier K et  al (2010) Time-series analysis of six whale-fall com-
munities in Monterey Canyon, California, USA.  Deep Sea Res I 57:1573–1584. https://doi.
org/10.1016/j.dsr.2010.09.003
Mandiola MA (2015) Evaluación de la presencia de lobos marinos de dos pelos sudamerica-
nos (Arctocephalus australis) en aguas de la Provincia de Buenos Aires. Mastozool Neotrop
22:419–420
Mandiola MA, Giardino GV, Bastida J et al (2015) Summer records of marine mammal on the
Brazil-Malvinas confluence on Argentine sea shelf break during a seismic survey. Mastozool
Neotrop 22:397–402
Mandiola MA, Blanco G, Rodríguez D (2017) Evaluación de interacciones con Mamíferos
Marinos en la pesquería de anchoíta certificada bajo estándares del Marine Stewardship
Council. Vinculación Tecnológica: de la Universidad Nacional de Mar del Plata al medio socio-­
productivo. vol IV. UNMdP, Mar del Plata, Argentina
Mandiola MA, Giardino G, Bastida J et al (2020) Half a century of sightings data of southern right
whales in Mar del Plata (Buenos Aires, Argentina). J Mar Biol Assoc UK 100:165–171
Marcovecchio JE, Moreno VJ, Bastida RO et  al (1990) Tissue distribution of heavy metals in
small cetaceans from the Southwestern Atlantic Ocean. Mar Pollut Bull 21:299–304. https://
doi.org/10.1016/0025-­326X(90)90595-­Y
Marcovecchio JE, Gerpe MS, Bastida RO et al (1994) Environmental contamination and marine
mammals in coastal waters from Argentina: an overview. Sci Total Environ 154:141–151.
https://doi.org/10.1016/0048-­9697(94)90084-­1
Marcovecchio J, Botté S, Delucchi F et al (2008) Pollution processes in Bahía Blanca estuarine
environment. In: Neves R, Baretta J, Mateus M (eds) Perspectives on integrated coastal zone
management in South America. IST Press, Lisboa, pp 301–314
Martinho F, Leitão R, Neto J et al (2007) The use of nursery areas by juvenile fish in a temperate estu-
ary, Portugal. Hydrobiologia 587:281–290. https://doi.org/10.1016/0025-­326X(90)90595-­Y
Melcón ML, Failla M, Iñíguez MA (2012) Echolocation behavior of franciscana dol-
phins (Pontoporia blainvillei) in the wild. J Acoust Soc Am 131(6):448–453. https://doi.
org/10.1121/1.4710837
Melcón ML, Failla M, Iñíguez MA (2016) Towards understanding the ontogeny of echolocation
in franciscana dolphins (Pontoporia blainvillei). Mar Mamm Sci 32(4):1516–1521. https://doi.
org/10.1111/mms.12336
Mendez M, Rosenbaum HC, Bordino P (2008) Conservation genetics of the franciscana dolphin
in Northern Argentina: population structure, by-catch impacts, and management implications.
Conserv Genet 9(2):419–435. https://doi.org/10.1007/s10592-­007-­9354-­7
Mendez M, Rosenbaum HC, Subramaniam A et al (2010) Isolation by environmental distance in
mobile marine species: molecular ecology of franciscana dolphins at their southern range. Mol
Ecol 19(11):2212–2228. https://doi.org/10.1111/j.1365-­294X.2010.04647.x
Mermoz JF (1977) Sobre el varamiento de un delfín nariz de botella, Tursiops truncatus, en la
desembocadura del Río de la Plata (Buenos Aires, Argentina). Physis Sección C 93:227–235
392 G. Giardino et al.

Moore SE (2008) Marine mammals as ecosystem sentinels. J Mammal 89:534–540. https://doi.


org/10.1644/07-­mamm-­s-­312r1.1
Moore SE (2018) Climate change. In: Würsig B, Thewissen JGM, Kovacs KM (eds) Encyclopedia
of marine mammals, 3rd edn. Academic, New York, pp 194–197
Moore KM, Simeone CA, Brownell RL (2018) Strandings. In: Würsig B, Thewissen JGM, Kovacs
KM (eds) Encyclopedia of marine mammals, 3rd edn. Academic, New York, pp 945–951
Moreno V, Pérez A, Bastida R, Aizpún de Moreno JE et al (1984) Distribución del mercurio total
en los tejidos de un delfín nariz de botella (Tursiops gephyreus Lahille, 1908) de la provincia de
Buenos Aires (Argentina).[Total mercury distribution in tissues of Bottlenose dolphin (Tursiops
gephyreus Lahille, 1908) from Buenos Aires Province (Argentine)]. Revista de Investigación y
Desarrollo Pesquero 4:93–102
Muelbert MMC, Oliveira LD (2006) First records of stranded pregnant female South American fur
seals, Arctocephalus australis, in the southern Brazilian coast. LAJAM 5(1):67–68. https://doi.
org/10.5597/lajam00094
Muñoz-Arnanz J, Chirife AD, Vernazzani BG et al (2019) First assessment of persistent organic
pollutant contamination in blubber of Chilean blue whales from Isla de Chiloé, Southern Chile.
Sci Total Environ 650:1521–1528. https://doi.org/10.1016/j.scitotenv.2018.09.070
Oliva AL, Arias AH, Quintas PY et  al (2017) Polycyclic aromatic hydrocarbons in mussels
from a South American estuary. Arch Environ Contam Toxicol 72(4):540–551. https://doi.
org/10.1007/s00244-­017-­0392-­y
Ott PH, Sucunza F, Wickert J et al (2017) Evidences of attack of a killer whale on a calf southern
right whale in Southern Brazil. Mastozool Neotrop 24:235–240
Páez E (1996) Simulaciones estocásticas en la población de Otaria flavescens en Uruguay. In: 7°
Reunión de Trabajo de Especialistas en Mamíferos Acuáticos de América del Sur, Viña del
Mar, p 116
Panebianco MV, Negri MF, Botte S et al (2011) Metales pesados en el riñón del delfín franciscana,
Pontoporia blainvillei (Cetacea: Pontoporiidae) y su relación con parámetros biológicos. Lat
Am J Aquat Res 39:526–533. https://doi.org/10.3856/vol39-­issue3-­fulltext-­12
Panebianco H, Botteb S, Negria M et al (2012) Heavy metals in liver of the Franciscana dolphin,
Pontoporia blainvillei, from the southern coast of Buenos Aires, Argentina. Ecotoxicol Environ
Contam 7(1). https://doi.org/10.5132/jbse.2012.01.006
Paso-Viola MN, Denuncio P, Negri MF et  al (2014) Diet composition of franciscana dolphin
Pontoporia blainvillei from southern Buenos Aires, Argentina and its interaction with fisheries.
Rev Biol Mar Oceanogr 49(2):393–400
Pavanato H, Silva KG, Estima SC et  al (2013) Occupancy dynamics of South American
sea-lions in Brazilian haul-outs. Braz J Biol 73(4):855–862. https://doi.org/10.1590/
S1519-­69842013000400023
Pavés HJ, Schlatter RP (2008) Temporada reproductiva del lobo fino austral, Arctocephalus austra-
lis (Zimmerman, 1783) en la Isla Guafo, Chiloé, Chile. Revi Chi His Nat 81:137–149. https://
doi.org/10.4067/S0716-­078X2008000100011
Peltier H, Baagøe HJ, Camphuysen KC et al (2013) The stranding anomaly as population indica-
tor: the case of harbour porpoise Phocoena phocoena in North-Western Europe. PLoS One
8(4). https://doi.org/10.1371/journal.pone.0062180
Pérez Salles S (2015). Estudio preliminar de la dieta del lobo marino de dos pelos sudamericano,
Arctocephalus australis, en el sector costero de la provincia de Buenos Aires, Argentina. Tesis
de grado Universidad Nacional de Mar del Plata, Mar del Plata
Perez-Venegas DJ, Seguel M, Pavés H et al (2018) First detection of plastic microfibers in a wild
population of South American fur seals (Arctocephalus australis) in the Chilean Northern
Patagonia. Mar Pollut Bull 136:50–54. https://doi.org/10.1016/j.marpolbul.2018.08.065
Pershing AJ, Christensen LB, Record NR et al (2010) The impact of whaling on the ocean car-
bon cycle: why bigger was better. PLoS One 5:e12444. https://doi.org/10.1371/journal.
pone.0012444
14  Marine Mammals: Is the Bahía Blanca Estuary and Its Area of Influence Important… 393

Petracci PF, Sotelo M, Massola V et  al (2010) Actualización sobre el estado del apostadero de
lobo marino de un pelo sudamericano (Otaria flavescens) en la isla Trinidad, estuario de Bahía
Blanca, Argentina. Mastozool Neotrop 17(1):175–182
Ponce de León A (2000) Taxonomía, sistemática y sinopsis de la biología y ecología de los pinipe-
dios de Uruguay. In: Rey M, Amestoy F (eds) Sinopsis de la biología y ecología de las pobla-
ciones de lobos finos y leones marinos de Uruguay. Pautas para su manejo y Administración.
Parte I. Biología de las especies. Montevideo. p 6–35
Reeves RR, Berger J, Clapham PJ (2006) Killer whales as predators of large baleen whales and
sperm whales. Whales, whaling and ocean ecosystems: 174–187
Reijnders PJH, Borrell A, Van Franeker JA et al (2018) Pollution. In: Würsig B, Thewissen JGM,
Kovacs KM (eds) Encyclopedia of marine mammals, 3rd edn. Academic, London, pp 746–753
Rodríguez D, Rivero L, Bastida R (2002) Feeding ecology of the franciscana (Pontoporia blainvil-
lei) in marine and estuarine waters of Argentina. LAJAM 1(1):77–94. https://doi.org/10.5597/
lajam00012
Rodríguez DH, Dassis M, Ponce de León A et al (2013) Foraging strategies of Southern sea lion
females in the La Plata River Estuary (Argentina–Uruguay). Deep-Sea Res Part II 88:120–130.
https://doi.org/10.1016/j.dsr2.2012.07.012
Rolland RM, Parks SE, Hunt KE et al (2012) Evidence that ship noise increases stress in right
whales. Proc R Soc B Biol Sci 279(1737):2363–2368
Roman J, Estes JA, Morissette L et al (2014) Whales as marine ecosystem engineers. Front Ecol
Environ 12:377–385. https://doi.org/10.1890/130220
Romero MA, Dans S, González R et al (2011) Solapamiento trófico entre el lobo marino de un
pelo Otaria flavescens y la pesquería de arrastre demersal del golfo San Matías, Patagonia,
Argentina. Lat Am J Aquat Res 39:344–358. https://doi.org/10.3856/vol39-­issue2-­fulltext-­16
Romero MB, Polizzi P, Chiodi L et al (2018) Preliminary assessment of legacy and current–use
pesticides in Franciscana dolphins from Argentina. Bull Environ Contam Toxicol 101:14–11.
https://doi.org/10.1007/s00128-­018-­2352-­2
Romero MA, Grandi MF, Túnez JI et  al (2019) Otaria flavescens. Categorización 2019 de
los mamíferos de Argentina según su riesgo de extinción. Lista Roja de los mamíferos de
Argentina. Versión digital: http://cma.sarem.org.ar
Rosas FWC, Pinedo MC, Marmontel M et al (1994) Seasonal movements of the South American
sea lion (Otaria flavescens Shaw) off the Rio Grande do Sul Coast, Brazil. Mammalia 58:51–59.
https://doi.org/10.1515/mamm.1994.58.1.51
Santos MB, Pierce GJ, Reid RJ et al (2001) Stomach contents of bottlenose dolphins (Tursiops
truncatus) in Scottish waters. J Mar Biol Assoc UK 81(5):873–878. https://doi.org/10.1017/
S0025315401004714
Secchi ER, Ott PH, Danilewicz D (2003) Effects of fishing by-catch and conservation status of
the franciscana dolphin, Pontoporia blainvillei. In: Gales N, Hindell M, Kirkwood R (eds)
Marine mammals: fisheries, tourism and management issues. CSIRO Publishing, Collingwood,
pp 174–191
SADSN-SAREM 2019 (Secretaría de Ambiente y Desarrollo Sustentable de la Nación y Sociedad
Argentina para el estudio de los mamíferos) (eds) (2019) Categorización 2019 de los Mamíferos
de Argentina según su riesgo de extinción. Lista Roja de los mamíferos de Argentina. www.
cma.sarem.org.ar
Simonis AE, Brownell RL Jr, Thayre BJ et al (2020) Co-occurrence of beaked whale strandings
and naval sonar in the Mariana Islands, Western Pacific. Proc R Soc B 287(1921):20200070.
https://doi.org/10.1098/rspb.2020.0070
Smith CR (2006) Bigger is better: the role of whales as detritus in marine ecosystems. In: Estes
JA, Demaster DP, Doak DF et al (eds) Whales, whaling and ocean ecosystems. University of
California Press, Berkeley, pp 286–301
Szapkievich VA, Cappozzo HL, Crespo EA et al (1999) Genetics relatedness in two Southern sea
lion (Otaria flavescens) rookeries in the southwestern Atlantic. Mamm Biol 64:1–5. https://doi.
org/10.1111/j.1748-­7692.1999.tb00814.x
394 G. Giardino et al.

Timi JT, Paoletti M, Cimmaruta R et al (2014) Molecular identification, morphological character-
ization and new insights into the ecology of larval Pseudoterranova cattani in fishes from the
Argentine coast with its differentiation from the Antarctic species, P. decipiens sp. E (Nematoda:
Anisakidae). Vet Parasitol 199:59–72. https://doi.org/10.1016/j.vetpar.2013.09.033
Trites AW (1997) The role of pinnipeds in the ecosystem. In: Stone G, Goebel J, Webster S (eds)
Pinniped populations, eastern north Pacific: status trends and issues. A symposium of the 127th
annual meeting of the American Fisheries Society. New England Aquarium, Conservation
Department, Central Wharf, Boston, pp 31–39
Trites AW, Christensen V, Pauly D (1997) Competition between fisheries and marine mammals for
prey and primary production in the Pacific Ocean. J Northwest Atl Fish Sci 22:173–187
Túnez JI, Centrón D, Cappozzo HL et  al (2007) Geographic distribution and diversity of
mitochondrial DNA haplotypes in South American sea lions (Otaria flavescens) and
fur seals (Arctocephalus australis). Mamm Biol 72:193–203. https://doi.org/10.1016/j.
mambio.2006.08.002
Túnez JI, Cappozzo HL, Cassini MH (2008) Natural and anthropogenic factors associated with the
distribution of South American sea lion along the Atlantic coast. Hydrobiology 598:191–202.
https://doi.org/10.1007/s10750-­007-­9150-­x
Túnez J, Cappozzo H, Nardelli M et al (2010) Population genetic structure and historical popula-
tion dynamics of the South American sea lion, Otaria flavescens, in north-central Patagonia.
Genetica 138:831–841. https://doi.org/10.1007/s10709-­010-­9466-­8
Tyack PL, Janik VM (2013) Effects of noise on acoustic signal production in marine mammals. In:
Brumm H (ed) Animal communication and noise. Springer, Berlin, pp 251–271
Vales DG, Cardona L, García NA et al (2015) Ontogenetic dietary changes in male South American
fur seals Arctocephalus australis in Patagonia. Mar Ecol Prog Ser 525:245–260. https://doi.
org/10.3354/meps11214
Vales DG, Mandiola A, Romero MA et al (2019) Arctocephalus australis. Categorización 2019
de los mamíferos de Argentina según su riesgo de extinción. Lista Roja de los mamíferos de
Argentina. Versión digital: http://cma.sarem.org.ar
Vaz Ferreira R (1981) South American sea lion, Otaria flavescens. In: Ridgway SH, Harrison RJ
(eds) Handbook of marine mammals V1. Academic, London, pp 39–65
Vaz Ferreira R (1982) Otaria flavescens (Shaw), South American sea lion. In: FAO-UNEP (ed)
Mammals in the seas IV. FAO, Rome, pp 477–496
Vermeulen E, Bräger S (2015) Demographics of the disappearing bottlenose dolphin in Argentina:
a common species on its way out? PLoS One 10(3). https://doi.org/10.1371/journal.
pone.0119182
Vermeulen E, Cammareri A (2009) Residency, abundance and social composition of bottlenose
dolphins (Tursiops truncatus) in Bahía San Antonio, Patagonia, Argentina. Aquat Mamm
35:379–386. https://doi.org/10.1578/AM.35.3.2009.379
Vermeulen E, Holsbeek L, Das K (2015) Diurnal and seasonal variation in the behaviour of bottle-
nose dolphins (Tursiops truncatus) in Bahía San Antonio, Patagonia Argentina. Aquat Mamm
41(30):272–183. https://doi.org/10.1578/AM.41.3.2015.272
Vermeulen E, Balbiano A, Belenguer F et al (2017) Site-fidelity and movement patterns of bot-
tlenose dolphins (Tursiops truncatus) in central Argentina: essential information for effective
conservation. Aquat Conserv 27(1):282–292. https://doi.org/10.1002/aqc.2618
Vermeulen E, Bastida R, Berninsone LG et al (2018) A review on the distribution, abundance, resi-
dency, survival and population structure of coastal bottlenose dolphins in Argentina. LAJAM
12:2–16. https://doi.org/10.5597/00233
Vermeulen I, Failla M, Loizaga de Castro R et al (2019) Tursiops truncatus. Categorización 2019
de los mamíferos de Argentina según su riesgo de extinción. Lista Roja de los mamíferos de
Argentina. Versión digital: http://cma.sarem.org.ar
Weller DW (2018) Predation on marine mammals. In: Würsig B, Thewissen JGM, Kovacs KM
(eds) Encyclopedia of marine mammals, 3rd edn. Academic, New York, pp 772–780
14  Marine Mammals: Is the Bahía Blanca Estuary and Its Area of Influence Important… 395

Wells RS, Scott MD (1999) Bottlenose dolphin Tursiops truncatus (Montagu, 1821). Handbook of
marine mammals: the second book of dolphins and porpoises 6:137–182
Wells RS, Fauquier DA, Gulland FM et  al (2013) Evaluating postintervention survival of free-­
ranging odontocete cetaceans. Mar Mamm Sci 29(4):463–483. https://doi.org/10.1111/
mms.12007
Wells RS, Natoli A, Braulik G (2019) Tursiops truncatus. The IUCN Red List of Threatened
Species. https://doi.org/10.2305/IUCN.UK.2019-­1.RLTS.T22563A50377908.en
Würsig B, Würsig M (1979) Behavior and ecology of the bottlenose dolphin, Tursiops truncatus,
in the South Atlantic. Fish Bull 77(2):399–412
Zerbini AN, Andriolo A, Da Rocha JM et al (2004) Winter distribution and abundance of humpback
whales (Megaptera novaeangliae) off Northeastern Brazil. J Cetacean Res Manag 6:101–107
Zerbini A, Rosenbaum H, Mendez M, Sucunza F et  al (2016) Tracking southern right whales
through the Southwest Atlantic: an update on movements, migratory routes and feeding
grounds. Scientific Committee annual meeting IWC SC/66a/BRG22, pp 1–10
Zerbini AN, Secchi E, Crespo E et al (2017) Pontoporia blainvillei (errata version published in
2018). The IUCN Red List of Threatened Species. https://doi.org/10.2305/IUCN.UK.2017-­3.
RLTS.T17978A50371075
Zerbini A, Fernandez Ajos A, Andriolo A et al (2018) Satellite tracking of southern right whales
(Eubalaena australis) from Golfo San Matias, Rio Negro Province, Argentina. Scientific
Committee of the International Whaling Commission SC67b, Bled, Slovenia
Chapter 15
Use of Coastal Area Habitats by Land
Mammals

Estela M. Luengos Vidal, Nicolás Caruso, Sabrina Martinez,


Emma Casanave, and Lucherini Mauro

15.1  Introduction

The ocean-land interface forms a major ecosystem, the coastal ecotone, where
marine and terrestrial abiotic and biotic factors intensively interact in a wide array
of forms. At this ecotone, terrestrial mobile species (e.g., mammals, Carlton and
Hodder 2003; reptiles, Lillywhite et al. 2008; ants, Garcia et al. 2011) and freshwa-
ter organisms have the ability to get into nearshore marine and estuarine waters or
fringe intertidal and marginal habitats (i.e., supralittoral zone) to exploit marine-­
derived food resources. These species create connections between marine and ter-
restrial food chains that can have far-reaching implications, including changes in
terrestrial plant communities (Croll et  al. 2005) and effects on the circulation of
pollutants between food webs (Dehn et al. 2006). The inputs from land to the sea
have been extensively studied and proven to promote high primary and secondary
productivity on the coastal zone (Polis and Hurd 1996a). It is less clear how the
commonly large inputs of marine material to adjacent lands affects productivity in
the surrounding area. Although this influence has been traditionally considered
small, this topic has recently increase in importance, and a growing body of work
has focused on the effects of marine inputs to the coastal land (Polis and Hurd
1996a; Spiller et al. 2010; Barreiro et al. 2013).
Besides the growing academic interest related with this topic, assessing the
strength of marine transfers to the coastal land has implications not only for the
ecological theory but also for wildlife management actions, and there is limited and

E. M. Luengos Vidal () · N. Caruso · S. Martinez · E. Casanave · L. Mauro


INBIOSUR (Instituto de Ciencias Biológicas y Biomédicas del Sur), CONICET-UNS,
Bahia Blanca, Argentina
Grupo de Ecología Comportamental de Mamíferos, Laboratorio de Fisiología Animal,
Departamento de Biología, Bioquímica y Farmacia, Universidad Nacional del Sur,
Bahia Blanca, Argentina
e-mail: luengos@criba.edu.ar

© Springer Nature Switzerland AG 2021 397


S. M. Fiori, P. D. Pratolongo (eds.), The Bahía Blanca Estuary,
https://doi.org/10.1007/978-3-030-66486-2_15
398 E. M. Luengos Vidal et al.

disperse information in our region. In the first section of this chapter, we set the
theoretical framework for a greater understanding of the ecological influence and
relevance of marine material inputs to the coastal lands, dismissed until recently. In
the second section, we speculate on the base of the existing information, which ter-
restrial mammal species would use coastal food sources (“maritime mammals”; see
definition later), and provided detailed information on the most relevant species in
nine species data sheet (Sect. 15.4). In the third section, we analyze a study case
focusing on the mammalian carnivores that use a sandy coastal area close to the
Bahía Blanca Estuary. Finally, given the relevance of anthropogenic influences on
this coastal region, we review the effects of human pressures on marine mammal
populations in a text box.

15.2  R
 ole of Coastal Ecosystems in the Ecology
of Land Mammals

The effects of cross-habitat transport of allochthonous marine materials are most


obvious in extreme habitats. In places where the difference between the primary
productivity of the sea and land is ample, such phenomenon is more evident and has
been the focus of a greater number of studies. This is the case, for example, of the
interactions between the Benguela Current and the Namibian Desert in Africa, the
Peru Current and the Atacama Desert in South America, or the effect of the water of
the Gulf of California on the Baja California Peninsula in Mexico (Polis and Hurd
1996a). In the absence or poor production of endogenous energy, the terrestrial
communities found in these regions are largely dependent on marine resources.
However, this process is far more widespread; in fact, it has been found to be nearly
ubiquitous (Polis and Hurd 1996a; Barreiro et al. 2013; Schlacher et al. 2017).
Seaweed, seagrasses, and carcasses of marine vertebrates are commonly washed
ashore and utilized by terrestrial consumers. Strandings of carrion, which are the
evident results of mortality events happening at sea, occur when positively buoyant
carcasses disperse from the original location of death and drift toward the coastline.
The causes of these events are diverse and can be related to biotoxins, bacteria, para-
sites, predators, human interactions, and oceanographic adverse conditions.
Typically, these events are spatially unpredictable phenomena, in which marine
mammal carrions strand in a delimited space and during a limited time and their
occurrence is also dependent on oceanography and coastal topography (Leeney
et al. 2008). The drifting algae depend largely on their residence time in a specific
location, which is mostly conditioned by the flow velocities of currents at that site
and the roughness of the substrate, but a number of physical, chemical, and biologi-
cal factors influence their behavior (Arroyo and Bonsdorff 2016).
In general, marine input supports a great abundance and diversity of terrestrial or
semi-terrestrial arthropods that feed on drift in the interface between land and sea.
These remains are consumed by detritivores and scavengers, which in turn can be
15  Use of Coastal Area Habitats by Land Mammals 399

used by predatory arthropods (secondary consumers) of the supralittoral zone


(Heatwole 1972; Ruiz-Delgado et al. 2016). The work of Colombini et al. (2003)
provides an overview of the importance of the coastal accumulations of macro-
phytes and other organic beach-cast material on the ecology of sandy beach ecosys-
tems. Carrions stranded on the shore represent a different system and constitute a
suitable material for the study of arthropod invasion, utilization, and succession.
The scavenger arthropod community develops primarily as a continuum of gradual
change. Classically, the activity of dipterans and ants has been considered a major
facilitation mechanism on behalf of the early invaders, but the activity of other taxa,
such as vertebrate scavengers (i.e., vultures), can act as facilitation mechanisms.
Shorebirds also prey opportunistically on larvae, flies, and other insects associated
with carrion. Other secondary consumers connected to marine-generated sources of
subsidies are represented by spiders, scorpions, lizards, rodents, and carnivores. In
many cases, allochthonous resources are ephemeral and present seasonal and spatial
fluctuations that probably affect all levels of the food web of a coastal beach. In the
Baja California system of Mexico, Polis and Hurd (1996a) found a complex interac-
tion of marine and terrestrial trophic webs. Terrestrial secondary consumers (spi-
ders, scorpions, and lizards) primarily prey on arthropod species from the marine
food web, which made up 95–99% of their diet. A high marine productivity makes
the abundance of potential preys significantly higher in the supralittoral zone com-
pared to adjacent terrestrial habitats. Consequently, the population of spiders found
along the coast was six times more abundant than that of inland areas (Polis and
Hurd 1995), and lizard populations were estimated to be four times more abundant
in the supralittoral zone than in inland areas (Polis and Hurd 1996b).
For the same geographical area, Rose and Polis (1998) showed that the coyote
(Canis latrans), an opportunist carnivore mammal that lives in coastal habitats, not
only eats a variety of living marine species (arthropods such as crustaceans and
larvae of insects that forage on algal drift, algae, and mollusks) but also consumes
many terrestrial creatures (arthropods, lizards, land birds, and coastal rodents)
exploiting marine resources. This diverse and conspicuous marine input, together
with the in situ terrestrial food supply, has the effect of increasing both the dietary
spectrum and the food intake of coastal coyotes compared with those of inland
populations, favoring coastal population of those secondary consumers.
In the forementioned study, three additional carnivore species were considered:
two canids, the gray fox (Urocyon cinereoargenteus) and the kit fox (Vulpes macro-
tis), and one member of the Procyonidae family, the ringtail (Bassariscus astutus),
which were very rarily recorded (1% of all observations). There were no significant
differences in the abundance of those species in coastal vs. inland areas. Both canids
occurred at almost all sites, but ringtails were spotted only at one coastal site and
one inland site. Nevertheless, Rose and Polis (1998) observed that all these species
foraged on marine resources within the littoral and supralittoral zone (e.g., gray
foxes were recorded eating sea cucumbers, order Holothuroidea).
Bristol Bay, Alaska, is one of the most productive marine ecosystems in the
world, with extensive seabird colonies on land, but ungulate prey limited to low
densities of moose (Alces alces) and migratory caribou (Rangifer tarandus). Watts
400 E. M. Luengos Vidal et al.

et al. (2010) inferred the influence of marine items among wolf preys by combining
direct observation, VHF, and GPS radio-tracking. The strong positive selection of
marine habitats by wolves (Canis lupus) suggests that marine-derived subsidies
may provide an important food source to the coastal populations of this canid (Watts
et al. 2010). Wolves predominantly feed on ungulates, but they are opportunistic-­
generalist carnivores and exhibit considerable dietary plasticity both among and
within populations. Previous studies have documented that wolves can make exten-
sive use of a wide variety of food items, including intertidal organisms (e.g., marine
invertebrates, Klein 1995) and carrion, which may constitute an important food
source in some areas (Lewis and Lafferty 2014). Also, Lewis and Lafferty (2014)
reported evidence of wolves feeding on large marine mammal carcasses (e.g.,
humpback whales) and hunting sea otters (Enhydra lutris). They observed a strong
positive association with the coastline on the base of the intensity of use of this habi-
tat, although it represented 1% of the total area available to pack members. These
subsidies of food are likely to affect the fitness and survival of wolves in these areas
and even increase wolf densities along coastlines, similar to what was observed for
coyote populations by Rose and Polis (1998). This information would be of interest
not only for the ecological theory but also for the planning of management actions
of carnivore populations, especially those related to low-density populations of
ungulates.
Thus, in many cases, marine energy may subsidize land consumers, promoting a
numerical response and leading to population densities that would be impossible
without such subsidy. In other cases, inputs enable some mammal populations to
persist in low productivity areas such as islands.
Marine-based organic materials also enter land food web via colonies of sea
birds nesting on the coast in the form of fish scrape, carcasses of dead chick, bird
feathers, eggs, and guano, in addition to an array of arthropods. This input increases
the abundance of many consumer species (e.g., arthropods) and forms the base of
unique food webs, particularly in small island where land resources are limited and
mammalian and reptilian predators are not present (Polis and Hurd 1996a). In
islands where the native animal communities have not been modified by the anthro-
pogenic introduction of exotic mammals, secondary consumers are typically preda-
tory birds.
Islands may also provide ample opportunity for omnivorous mammalian meso-
predators (i.e., medium-sized predators subordinate to larger predators) to exploit
marine prey (Carlton and Hodder 2003). It has been shown that mesopredators may
benefit substantially from such marine subsidies (Rose and Polis 1998). Consumption
of marine preys has been suggested to exacerbate the impact of mesopredators on
terrestrial communities. Although they are still scarce, some data-based studies
demonstrate that mammalian mesopredators, whose diet is substantially subsidized
by marine items, have a significant impact on terrestrial prey population (Polis and
Strong 1996; Polis and Hurd 1996a; Rose and Polis 1998).
The limited evidence available suggests that such allochthonous input from the
ocean has important effects on the distribution and abundance of omnivorous
15  Use of Coastal Area Habitats by Land Mammals 401

mammals in many adjacent terrestrial systems and that the use of marine resources
by carnivorous mammals is widespread worldwide. The Arctic fox (Vulpes lagopus)
is an opportunistic predator and scavenger distributed in tundra regions and Arctic
islands of Eurasia and North America. Their main preys throughout most of its
range are small mammals such as lemmings (Dicrostonyx spp. and Lemmus spp.;
e.g., Kennedy 1980; Elmhagen et  al. 2000). However, Arctic foxes living in the
islands of the high Arctic Svalbard archipelago, a restricted geographical area,
where no resident rodents are present mostly consume reindeer (Rangifer taran-
dus), rock ptarmigan (Lagopus mutus), geese, seabirds, and eggs, but inland and
coastal foxes have different food habits (Prestrud 1992; Frafjord 1993). In coastal
areas, seabirds and eggs constitute the majority of the diet of foxes (Prestrud 1992;
Hersteinsson and Macdonald 1996). Coastal foxes which depend almost entirely on
ocean cliff birds confined their activity to very small areas, illustrating their strong
affinity to the large concentrated seabird colonies (Jepsen et  al. 2002). However,
once again, the effect of the allochthonous energy subsidy provided by the ocean is
not limited to Artic fox populations but goes beyond the sea-land ecotone. On the
Bylot Island of the Canadian Territory of Nunavut, the top-down control of lem-
ming populations by Arctic foxes is likely strengthened by this allochthonous sub-
sidy (Legagneux et al. 2012).
Studies considering the potential impacts of mammalian mesopredators on inter-
tidal preys are even less numerous. Experimental studies on the importance of ter-
restrial mammals as intertidal predators are scarce, and thus the use of intertidal
resources has been presumed to be insignificant both in terms of energy and flow
pathway and as driving force of intertidal community structures. However, some
studies suggest that this assumption may be incorrect. Suraci et  al. (2014) con-
cluded that terrestrial mammalian mesopredators on islands may directly impact
both the intertidal and shallow subtidal marine communities. They showed that rac-
coon (Procyon lotor) populations in the Gulf Islands of British Columbia, Canada,
with access to marine resources impacted the local abundance of their marine prey,
where top predators are extirpated or are naturally absent. Under these circum-
stances, raccoons have been promoted to the top of both the terrestrial and nearshore
marine food chains and thereby have a direct influence on terrestrial birds, intertidal
vertebrates and invertebrates, and intertidal and shallow subtidal crab populations.
This direct effect by raccoons may in turn initiate trophic cascades within both the
terrestrial and intertidal communities. A similar effect has been reported by Kurle
et al. (2008), who found that predation by introduced brown rats (Rattus norvegicus)
on nesting seabirds on Alaska’s Aleutian Islands indirectly affected intertidal com-
munities through a trophic cascade.
This demonstrates the potential strength of the impact of small mammals on
island ecosystems and, in general, of land mammals on coastal marine ecosystems.
Given these examples and the paucity of information on this topic, it is not surpris-
ing that the effects of terrestrial mammals on the diversity, abundance, and distribu-
tion of marine preys have been identified as a conspicuous gap in the ecological
literature (Carlton and Hodder 2003).
402 E. M. Luengos Vidal et al.

The review of Moore (2002) was a first attempt to join scattered literature about
mammals that use seashores and maritime terrestrial environments. This review
focused on mammalian contribution to ecological processes more than on ethologi-
cal research. Moore (2002) found ecological information on the effect of mam-
mals on marine environments in 11 orders (Marsupialia, Insectivora, Chiroptera,
Lagomorpha, Rodentia, Cetacea, Carnivora, Sirenia, Perissodactyla, Artiodactyla,
and Primates). Moreover, introduced mammals like cats (Felis catus), dogs (Canis
familiaris), rats (Rattus norvegicus and Rattus rattus), and minks (Mustela vison)
showed to be important predators of seabirds. Domestic grazers such as goats and
sheep have been found to feed on seagrasses and seaweed. Salt and minerals could
be a good motivation, but seaweeds and seagrasses are often cited as alternative
protein sources from terrestrial agricultural resources by these domestic animals.
But the grazing pressure in some opportunities causes coastal habitat degradation.
Carlton and Hodder (2003) conducted a more specific review on terrestrial mam-
mals intentionally entering the coastal zone at low tide to prey on living marine
invertebrates, fish, algae, and seagrasses. Those authors introduced the term
“maritime mammals,” to indicate coastal mammalian predators or consumers that
utilize living intertidal resources and transfer their energy to the land. They docu-
mented 135 records of consumption across 45 species of terrestrial mammals in 8
orders (Didelphimorphia, Dasyuromorphia, Insectivora, Primates, Lagomorpha,
Rodentia, Carnivora, Artiodactyla). Maritime mammals were found to occur on all
continental coastlines of the world except Antarctica. Most predation events were
realized by carnivores (59%, mostly by raccoon, mink, black bear Ursus america-
nus, and Arctic fox), followed by rodents (20%) and artiodactyls (14%), but intro-
duced populations of 17 species of mammals were also recorded as maritime
predators.
Carlton and Hodder (2003) identified 228 different prey taxa, representing 12
phyla of marine organisms including 8 animal phyla (mainly bivalve and gastropod
mollusks, crabs, and fish), 3 algae phyla, and 1 plant phylum. Because their review
reports several observations across a wide variety of taxa, they suggest that preda-
tion by maritime mammals is a rarely studied phenomenon rather than a rare phe-
nomenon. They conclude that there is a need for quantitative observations on
predation and experimental studies on mammals as consumers of intertidal energy
resources. Such studies should be facilitated by the use of new technologies, includ-
ing advanced infrared night vision technology, radio telemetry of individuals,
camera traps, and the use of stable isotopes for trophic web analysis.
To our knowledge, since the publication of reviews by Moore (2002) and Carlton
and Hodder (2003), there have been no new global revisions on this subject, but
numerous specific studies have been published that contribute to the ecological
knowledge of these processes (e.g., Watts et al. 2010; Gaydos and Pearson, 2011;
Tarroux et al. 2012; Lafferty et al. 2014; Lewis and Lafferty 2014; Broekhuis et al.
2014; Moss 2017; Lei et al. 2017).
15  Use of Coastal Area Habitats by Land Mammals 403

15.3  T
 he Maritime Mammals Surrounding Bahía
Blanca Estuary

From this section on, we will extend the term “maritime mammals” proposed by
Carlton and Hodder (2003) to include those species that can potentially predate on
seabirds or other vertebrates that use the coastal zone and those that can scavenge
carrion and detritus from marine origin.
To identify the species that may classify as maritime mammals for the region, we
made a literature search on the potential mammals species that, according to their
distribution ranges, could use the coastal and surrounding area of the Bahía Blanca
Estuary, from mouth of the Quequén Salado River in the North-East to the mouth of
Colorado River in the South and from the sea coast up to 1 km inland (Fig. 2.1;
Chap. 2). The actual distribution ranges of many of these species are poorly known
and are based on polygons that connect specific record points (CMA 2019).
Additionally, the information published on their ecology in the study region is
extremely limited. Thus, we collected and summarized in Table  15.1 any further
evidence (including unpublished personal observations and personal communica-
tions from reliable sources such as researchers or park rangers) of the use of coastal
habitats (Table 15.1).
The potential maritime mammal community of the study area comprises 43 spe-
cies (Table 15.1), belonging to 5 out of the 12 orders of wild mammals occurring in
Argentina (Montero and Autino 2018). Rodentia is the most abundant order, with
approximately 20 species. Although there is no literature on the local use of the
intertidal area by any of these native species, we do know that Ctenomys australis
(see species data sheet 15.4.1; Fig. 15.1a) is an endemic species spatially restricted
to the first ridge of dunes in the coastal zone of the Buenos Aires Province, with a
southern limit coincident with the beginning of the estuarine area, in Punta Alta
(Contreras and Reig 1965). The conservation status of this rodent is Endangered
(Austrich et al. 2019), based on evidence of the fragmentation of its habitat due to
the influence of anthropic activity, in addition to the ecological characteristics of the
species (low dispersion rates and relatively low reproduction rates with a high
degree of genetic structuring) (Mora and Mapelli 2010). Ctenomys talarum is
another rodent strictly limited to the coastal zone specifically to the second and third
ridges of dunes. Although this species is listed as Least Concern by the IUCN
(International Union for Conservation of Nature) (Bidau 2016), it was considered as
Vulnerable in Argentina by Fernandez et al. (2019). The capybara (see species data
sheet 15.4.2; Fig. 15.1b) Hydrochoerus hydrochaeris is another rodent that could
use the coastal zone close to Bahía Blanca. It is typical of wetlands and riparian
environments, and its food habit is mainly herbivorous. In recent years, this rodent
has been regularly registered in coastal areas of southern Buenos Aires where there
are freshwater streams or lagoons (e.g., Laguna de Sauce Grande, near the city of
Monte Hermoso) and there is a record of individuals entering the sea (https://www.
404 E. M. Luengos Vidal et al.

Table 15.1  List of the potential species that could use the coastal and surrounding area of the
Bahía Blanca Estuary, from the town of Marisol in the North to the end of Colorado R River in the
South and from the sea coast up to 1 km toward the continent, the scientific name, the local name
in Spanish and the name in English, its conservation category according to CAM (2019) (LC, Least
concerned; NT, near threatened; VU, vulnerable; DD, data deficient; EXO, exotic), and if it has
been cited by reviews of 1 Carlton and Hodder (2003) and 2 Moore (2002)
CC Species Spanish name English name 1 2
LC Lycalopex Zorro gris Pampas fox x
gymnocercus
LC Leopardus geoffroyi Gato montés Geoffroy’s cat
VU Leopardus colocolo Gato de los Pampas cat
pajonales
Carnivora LC Herpailurus Yaguarundi Jaguarundi
yagouaroundi
LC Puma concolor Puma Cougar
LC Galictis cuja Hurón menor Lesser grison x
NT Lyncodon Huroncito Patagonian weasel
patagonicus patagónico
LC Conepatus chinga Zorrino Molina’s hog-nosed
skunk
LC Chaetophractus Peludo Large hairy armadillo
villosus
NT Dasypus hybridus Mulita pampeana Southern long-nosed
armadillo
Cingulata LC Chaetophractus Piche llorón Screaming hairy
vellerosus armadillo
DD Chlamyphorus Pichiciego menor Pink fairy armadillo
truncatus
NT Zaedyus pichiy Piche Pichi
Lagomorpha EXO Lepus europaeus Liebre europea European hare x
LC Lama guanicoe Guanaco Guanaco
Cetartiodactyla EXO Dama dama Ciervo dama Fallow deer x
EXO Sus scrofa Jabalí Wild boar x x
LC Thylamys pallidior Marmosa pálida White-bellied
fat-tailed mouse
opossum
Didelphimorphia LC Monodelphis Colicorto Yellow-sided opossum
dimidiata pampeano
LC Lutreolina Comadreja Red opossum
crassicaudata colorada
LC Didelphis albiventris Comadreja overa White-eared opossum x
LC Oxymycterus rufus Ratón hocicudo Red hocicudo
rojozo
LC Akodon azarae Ratón de pastizal Azara’s grassland x
pampeano mouse
LC Akodon dolores Ratón cordobés Córdoba akodont x
LC Akodon iniscatus Ratón Patagónico Patagonian akodont x
(continued)
15  Use of Coastal Area Habitats by Land Mammals 405

Table 15.1 (continued)
CC Species Spanish name English name 1 2
LC Calomys laucha Laucha chica Little Laucha
LC Calomys musculinus Ratón maicero Corn mouse
LC Eligmodontia typus Laucha colilarga Lowland gerbil mouse
baya
LC Graomys Pericote comun Common pericote
griseoflavus
LC Holochilus vulpinus Rata colorada Crafty marsh rat
LC Oligoryzomys Ratón colilargo Long-tailed colilargo x x
longicaudatus
Rodentia LC Reithrodon auritus Rata conejo Bunny rat
LC Lagostomus Vizcacha Plains viscacha
maximus
Galea musteloides Cuís común Common yellow-­
toothed cavy
LC Microcavia australis Cuis chico Southern mountain
cavy
VU Dolichotis Mara Patagonian hare
patagonum
VU Ctenomys talarum Tuco-tuco de los Tuco-tuco
talares
NT Ctenomys australis Tuco-tuco de los Southern tuco-tuco
medanos
LC Myocastor coypus Coipo Coipo
EXO Rattus norvegicus Rata noruega Brown rat x x
EXO Rattus rattus Rata negra Black rat
LC Hydrochoerus Carpincho Capybara
hydrochaeris
LC Cavia aperea Cuis Pampas cavy

lacapitalmdp.com/sorpresa-­en-­mar-­del-­sud-­aparecio-­un-­carpincho-­nadando-­en-­la-­
playa). It is worth mentioning that this order includes also three exotic cosmopolitan
species (Mus musculus, R. norvegicus, and R. rattus). In particular, the brown rat
has been reported to make an intensive use of intertidal areas in certain sites of
Central Chile (Navarrete and Castilla 1993).
The information about the four species of the order Didelphimorphia potentially
behaving as maritime mammals is too scarce to provide any hint about the possible
use of the coastal zone and its resources, except for the case of Didelphis albiven-
tris. The white-eared opossum is listed as Least Concern by IUCN (Costa et  al.
2015) and under the same category in Argentina (Chimesquy and Martin 2019) is
omnivorous and extremely adaptable, which enables it to take great advantage of
disturbance and changes of land use associated with human presence. Additionally,
D. virginiana, a very similar congeneric species, has been reported to prey on crabs
(Uca pugilator) in the Gulf of Mexico (Rathbun 1918).
406 E. M. Luengos Vidal et al.

Fig. 15.1  Photos of maritime mammals possible to find in areas surrounding the Bahía Blanca
Estuary: a Ctenomys australis, southern tuco-tuco (English), tuco-tuco de las dunas (Spanish), and
b Hydrochoerus hydrochaeris capybara (English), carpincho, capibara (Spanish). (Photos by
Federico Becerra (a) and Irma Gamarra (b))

The only lagomorph (order Lagomorpha), the European hare (Lepus europaeus),
is an exotic species that has dispersed across most of the Argentine territory and
appears to be relatively abundant in the southern Buenos Aires Province, including
the coastal zone. The European hare may graze at on salt marshes and, occasionally,
on mudflats (Moore 2002).
The order Cingulata is represented by five species, but Chaetophractus villosus
(large hairy armadillo; see species data sheet 15.4.3; Fig. 15.2a) is the one that most
likely uses marine food resources due to its omnivorous diet and especially its scav-
enger habits (Arriagada et al. 2017). In accordance to this hypothesis, this armadillo
was relatively frequently recorded in camera trap and sign surveys in a coastal pro-
tected area south of Marisol (unpubl. Data and following section of this chapter).
There are two species of the Cetartiodactyla order that occur in our area of inter-
est. Although there is no site-specific ecological information about the guanacos
(Lama guanicoe), a small population of this native camelid, which is currently very
rare in Buenos Aires Province, persists in islands of the estuary of Bahía Blanca.
15  Use of Coastal Area Habitats by Land Mammals 407

Fig. 15.2  Photos taken by camera trap of maritime mammals possible to find in areas surrounding
the Bahía Blanca Estuary: a Chaetophractus villosus, large hairy armadillo (English), peludo
(Spanish), and b Sus scrofa, wild boar (English), chancho jabali (Spanish). (Photos by Grupo de
Ecologia Comportamental de Mamiferos)

Because the vegetation on the islands is scarce, it is possible that guanacos use
cordgrass (Bortolus et al. 2015) as supplementary source of food. The wild boar
(Sus scrofa) (see species data sheet 15.4.4; Fig. 15.2b) is an exotic species which is
becoming increasingly common in the region. Because of their wide food niche and
opportunistic habits, wild boars could predate on seabird nests and also consume
other biological resources provided by the sea, as observed in the coastal area of
Doñana National Park, Spain (Fernández-Llario et al. 1996).
Carnivores (order Carnivora) that use the ecotone between terrestrial and marine
habitats may act as a strong and dynamic link connecting terrestrial and marine food
webs. There are eight native species from the order Carnivora that reportedly live in
the region of interest. Three of them are rare and/or difficult to observe, including
two small felids, jaguarundi Herpailurus yagouaroundi (categorized as Near
Threatened by IUCN, Caso et  al. 2015, and in Argentina as Least Concern by
408 E. M. Luengos Vidal et al.

Bisceglia et al. 2019) and Pampas cat Leopardus colocolo (categorized as Vulnerable
by IUCN, Lucherini et al. 2016, and the same category in Argentina by Lucherini
et al. 2019), and a mustelid, the Patagonian weasel Lyncodon patagonicus (listed as
Near Threatened by IUCN, Kelt et al. 2016, and the same category in Argentina by
Schiaffini et al. 2019). Although the most recent distribution assessment includes
the southern Buenos Aires Province into the range of the Patagonian ferret, the latest
records of its occurrence are from 1881, and its present-day presence in the area is
exclusively based on a distribution model that predicted a low probability of occur-
rence (Schiaffini et al. 2013a). Although the presence of this carnivore would need
to be confirmed, the coastal habitats of southwestern Buenos Aires appear to repre-
sent a shelter for the populations of carnivores that are rarer elsewhere, most likely
because the comparatively low fertility of the sandy soils along the coast has pre-
vented the development of agriculture and limited livestock farming. This appears
to be especially true for the Pampas cat and the puma (Puma concolor) (see species
data sheet 15.4.5, Fig. 15.3a). Although little data is available (see the next section
of this chapter), local inhabitants of the rural areas report that pumas, which had
almost completely disappeared from the province, have greatly increased their num-
bers in the semi-natural habitats found along the coast in the last 10–15  years
(Canevari and Balboa 2003). The puma is a generalist top predator that could poten-
tially interact with marine and maritime mammals. There are reports of pumas feed-
ing on penguin colonies (Frere et al. 2010; Martínez et al. 2012), and it cannot be
excluded that they prey on other seabirds and that these events have gone undetected.
The Pampas fox (Lycalopex gymnocercus; see species data sheet 15.4.6;
Fig. 15.3b) and Geoffroy’s cat (Leopardus geoffroyi; see species data sheet 15.4.7;
Fig.  15.3c) are two of the most common carnivores in the entire province. The
Pampas fox is a generalist canid that can be found in all the habitats of our region,
being capable of living in areas modified by extensive ranching and agriculture
human activities (Luengos Vidal et al. 2012; Caruso et al. 2016). This carnivore has
a very wide trophic niche (Lucherini and Luengos Vidal 2008). In spite of the fact
that there is no study on Pampas foxes proving that they use the coastal zone or
intertidal habitats, some evidence suggest that they effectively feed in coastal areas.
García and Kittlein (2005) analyzed the summer diet of this fox in Bahıa San Blas
and Isla Gama Provincial Reserve, located in southern Buenos Aires Province, and
found that although fish carrion and crustaceans are not frequent, these items are
present in its diet. They also observed a high opportunistic consumption of fruit
probably obtained in sand dune habitats. Based on their findings, they conclude that
foxes utilize sand dunes and scrublands as feeding sites, and grasslands for other
activities unrelated to food search, such as refuge and breeding (García and Kittlein
2005; Bossi et al. 2019), found that the diet of the Pampas fox from the Brazilian
Pampas included fish, crustaceans of the Decapoda family, and marine crabs
(Brachyura), thus providing evidence that this canid uses marine preys. It is also
worth mentioning that the Darwin fox (Lycalopex fulvipes), a very close relative of
the Pampas fox that inhabits the Chiloe Island (Chile), was observed digging in the
intertidal fringe of sandy beaches where the macrofauna consists of crabs, talidrid
15  Use of Coastal Area Habitats by Land Mammals 409

Fig. 15.3  Photos of maritime mammals possible to find in areas surrounding the Bahía Blanca
Estuary: a Puma concolor, cougar (English), puma (Spanish); b Lycalopex gymnocercus, Pampas
fox (English), zorro pampeano (Spanish); and c Leopardus geoffroyi Geoffroy’s cat (English), gato
montes (Spanish). (Photos by Grupo de Ecologia Comportamental del Mamiferos (a, c) and
Nicolás Mariano Chiaradía (b))
410 E. M. Luengos Vidal et al.

amphipods, and isopods (Elgueta et al. 2007). The above-reported findings suggest
that the Pampas fox is a strong candidate to qualify as maritime mammal.
Geoffroy’s cat, as most felids, has a specialized diet, especially if compared to
that of sympatric canids (Kasper et al. 2016). Nevertheless, fish remains have been
found in fecal samples of this cat from Campos del Tuyú (Manfredi et al. 2004), a
coastal protected area adjacent of northern Buenos Aires Province, and aquatic birds
have been found in Mar Chiquita Biosphere Reserve (Canepuccia et  al. 2007),
another protected area on the Buenos Aires coast.
Molina’s hog-nosed skunk (Conepatus chinga; see species data sheet 15.4.9;
Fig. 15.4a) is the only member of the Mephitidae family found in our region. It is
typically an insect feeder (Castillo et al. 2014) that could potentially benefit from
the trophic chain initiated by seasonal marine items (see Sect. 15.1 in this chapter)
but is also clearly an opportunistic predator, and several other food items may be
included in its diet, such as mammals, reptiles, and amphibians. Interestingly, in
Southern Brazil, Kasper et  al. (2012) recorded several events of Molina’s hog-
nosed skunks predating on turtle and bird nests to eat eggs. The lesser grison
(Galictis cuja) (see species data sheet 15.4.8; Fig. 15.4a) has a diet similar to that
of a small wildcat and is considered strictly carnivorous. The limited data available
on this mustelid indicates that it preys mostly on rodents, but can probably use
different foraging strategies (Kasper et  al. 2016) and frequently consumes birds
and their eggs (Borboroglu and Yorio 2004; Valenzuela et al. 2013; Galende and
Raffaele 2016).

Fig. 15.4  Photos of maritime mammals possible to find in areas surrounding the Bahía Blanca
Estuary: a Galictis cuja, lesser grison (English), hurón menor (Spanish), and b Conepatus chinga
Molina’s hog-nosed skunk (English), zorrino comun (Spanish). (Photos by Luciano Lapolla (a)
and Grupo de Ecologia Comportamental del Mamiferos (b))
15  Use of Coastal Area Habitats by Land Mammals 411

15.4  S
 pecies Data Sheet of Some Relevant
Maritime Mammals

Ctenomys australis, Hydrochoerus hydrochaeris, Sus scrofa, Chaetophractus villo-


sus, Puma concolor, Lycalopex gymnocercus, Leopardus geoffroyi, Conepatus
chinga, Galictis cuja

15.4.1  Ctenomys australis Rusconi, 1943 (Order: Rodentia)

Common name: tuco-tuco de las dunas (Spanish), southern tuco-tuco (English).


Description: It has a chunky and plump body. Total adult length is between 210
and 370 mm and weight is between 248 and 500 g. The head is bulky, short, and
flattened and the neck is short and thick; it has large ears, underdeveloped but not
tiny eyes, and four large incisors left outside the mouth. The short, well-muscled
legs finish in five fingers with nails that are used to dig. The color is brown yellow,
clear than other species of the some genera (Gomez Villafañe et  al. 2005)
(Fig. 15.1a).
Distribution and habitat: C. australis is a territorial rodent that lives exclusively,
in the narrow strip of coastal dunes, characterized by the presence of sandy, poorly
compacted soils and little or no vegetation cover. Its distribution range extends
approximately 300  km between the cities of Necochea (38°37’S, 58°50’O) and
Punta Alta (39°30’S, 61°40’O). This rodent is a strict habitat specialist and shows
low dispersion rates and relatively small population sizes with a high degree of
genetic structuring (Zenuto and Busch 1998; Cutrera et al. 2006).
Behavior: C. australis is solitary, highly territorial, and most likely polygynous
(Zenuto and Busch 1998). It is crepuscular and diurnal. It vocalizes like the other
tuco-tucos, with drums like beats but only in the hottest hours of the day (Gomez
Villafañe et al. 2005). Radio telemetry data suggest that home ranges are greater in
males than females, averaging 1300 m2 for males and 580 m2 for females (Mora
2008). Individuals build large, exclusive burrow systems in coastal sand dunes. The
habit of underground life of these animals contributes to the dynamics of plant com-
munities by altering their composition and maintaining succession over pioneer
states without altering total biomass (Monserrat et al. 2012).
Conservation status: This species was listed in the “Endangered” category under
both the IUCN Red List (Bideau 2016) and the Argentine National Assessment
(Austrich et al. 2019).
Threats: Its peculiar distribution makes this rodent extremely susceptible to the
changes generated in its habitat as a result of tourism, urbanization, forestry devel-
opment, and the progressive advance of artificial pastures. Only 2% of the area of
the sandy dune ecosystem in which the species is found is protected by some type
of legal conservation status (Monserrat et al. 2012).
412 E. M. Luengos Vidal et al.

15.4.2  H
 ydrochoerus hydrochaeris (Linnaeus, 1766)
(Order: Rodentia)

Common name: capybara (English), carpincho, capibara (Spanish).


Description: The capybara is a large, semi-aquatic rodent that is found in regions
of Central and South America. The body is large (approx. 1200  cm) and heavy
(35–70 kg) with a descendent croup; tail is vestigial; limbs are short; feet are peris-
sodactyle, with digits united by moderate webbing; forefeet possess four digits. The
head is broad; ears are short and rounded; snout is large; nostrils are small and
widely separated. Hair is coarse, dark brown to reddish, and light brown to light
yellow (Mones and Ojasti 1986) (Fig. 15.1b).
Distribution and habitat: The species has a wide distribution in Argentina, where
it is found in at least 10 provinces and 25 protected areas, although with varying
degrees of protection. In protected sites, it is usually abundant, but outside of them,
densities drop dramatically. Since 2005, there has been a range expansion in the
Buenos Aires Province where it currently covers the entire eastern and southern
fringe of the province to the proximity of Bahía Blanca, where capybaras were his-
torically found (Doumecq Milieu et al. 2012).
Behavior: These rodents are strictly herbivores. They reproduce throughout the
year, without a specific birth season. They can give birth to between one and eight
offspring, with an average of four offspring, frequently in relation to the environ-
ment and nutritional conditions of the female (Cueto 1999). It is a species active
during the day and night, usually in groups of between 2 and 30 individuals with a
dominant male, and several females, young and male subalterns. The size of their
home ranges between 11 and 27 ha, approximately (Corriale et al. 2013).
Conservation status: Capibara is included in the “Least Concern” category under
both the IUCN Red List (Reid 2016) and the Argentine National Assessment
(Bolkovic et al. 2019).
Threats: The loss of wetlands is one of the most important threatening factors for
this species. The leather of this rodent has great commercial value. Thus, for
decades, direct hunting was the main threat to the species in Argentina because the
extraction was carried out with poor management plans. Although the leather trade
is gradually decreasing, even today, many local communities within their range use
capybaras to meet their protein needs. Population fragmentation is an additional
threat and exotic species can affect capybaras. Finally, capybaras can be affected by
vehicle traffic on roads: on Argentine National Route 174 (connecting the cities of
Rosario and Victoria), 95% of reports of wildlife accidents involving vehicles
corresponded to capybaras (Bolkovic et al. 2019).
15  Use of Coastal Area Habitats by Land Mammals 413

15.4.3  C
 haetophractus villosus (Desmarest, 1804)
(Order: Cingulata)

Common name: peludo (Spanish), large hairy armadillo (English).


Description: It is a dark-colored armadillo, with a body length of 291  mm
approximately without the tail and the tail of 145  mm (Redford and Eisenberg
1992). The weight ranges between 2 and 5 kg (Gallo et al. 2019). The male is gener-
ally larger than the female. Armadillos possess a characteristic protective armor that
consists in a small shield between the ears on the back of the neck and a carapace
that protects the shoulders, back, sides, and rump. This species has hairs projecting
from between the scales of the body armor, and the limbs and belly are covered with
whitish or light brown hairs. Some individuals in this species have 3–4 holes in the
pelvic region of the armor that open to glandular pits (Redford and Eisenberg 1992)
(Fig. 15.2a).
Distribution and habitat: This species has its origin in the Pampas region (Poljak
et al. 2010). It is currently distributed throughout virtually the entire country, and it
is documented that it invaded the Patagonia in historical times. It can be found in
Argentina, Bolivia, Paraguay, and Chile (Abba et al. 2014).
Behavior: C. villosus is an omnivorous species that eats mainly invertebrates and
digs under and into carcasses to eat carrion. It preys potentially on imperial cormo-
rants (Phalacrocorax atriceps), rock shags (Phalacrocorax magellanicus), and kelp
gulls (Larus dominicanus) in Patagonia, Argentina (Punta et al. 2002; Borboroglu
and Yorio 2004). This armadillo is solitary except when rearing young or mating. It
can build and use extensive burrow systems. Gestation lasts 60–75 days and litters
are usually 1–3 offspring. Typically, females produce a litter per year. Birth occurs
inside a burrow, in which the female usually builds a nest with plant material.
Weaning occurs at 55–61 days of age (Gallo et al. 2019).
Conservation status: C. villosus is listed in the “Least Concern” category under
both the IUCN Red List (Abba et al. 2014) and the Argentine National Assessment
(Gallo et al. 2019).
Threats: It is hunted to be used as food by local people and to make musical
instruments and handicrafts. It is persecuted as a pest in agricultural areas for dam-
aging the soil with its burrows and breaking the grain storage bags. It is also accused
of being a nuisance for poultry and for sheep farming during the lambing periods
and is subjected to retaliatory/preventive hunting. Animals may also be accidentally
killed on roads and by dogs.

15.4.4  Sus scrofa Linnaeus, 1758 (Order: Cetartiodactyla)

Common name: chancho jabali (Spanish), wild boar (English).


Description: It is a wild pig whose head and body lengths are between 1100 and
1650 mm, shoulder height is 550–1000 m, and weight is 10–190 kg. Females are
414 E. M. Luengos Vidal et al.

generally smaller in size and weigh less than males. Body color varies from gray to
brown or black or a mixture of colors. The tusks in males may reach 150 mm (Long
2003) (Fig. 15.2b).
Distribution and habitat: Wild boars are native to large parts of Europe, Asia, and
North Africa, and they have become extinct over much of their former range.
Introduced populations, descended from domestic pigs, wild boars, or a combina-
tion of the two, are present in many parts of the world. It is considered by the IUCN
as one of the 100 invasive species most damaging to biodiversity in the world.
S. scrofa was initially introduced in Buenos Aires in 1536 by colonizers as a food
source. Only at the beginning of the twentieth century the wild boar was introduced
from Europe for hunting purposes. Its presence has been confirmed in different
ecoregions of Argentina with preponderance toward environments with high water
availability and vegetation cover (Ballari et al. 2019).
Behavior: The capacity of wild boars to colonize different habitats and become
an invasive species is related not only to their generalistic habitat requirements but
also to its biological characteristics, such as the high reproduction rate, omnivorous
diet, and behavioral plasticity. A research conducted in the southern part of the
Buenos Aires Province Espinal in a non-protected area found that wild boars exhib-
ited a crepuscular and nocturnal activity pattern that peaked around midnight and
that their habitat use was significantly and negatively affected by the proportion of
grass cover but temperature and water availability were two additional and impor-
tant factors affecting distribution and abundance (Caruso et al. 2018). In Argentina,
one or two annual reproductive events have been commonly recorded, although
there are warm climate areas where it is possible to find pregnant females all year
round (Ballari et al. 2019).
Conservation status: Wild board is included in the “Least Concern” category
under the IUCN Red List (Oliver and Leus 2008). The Argentine National
Assessment assigned the same threat category but also listed it as “Invasive species”
(Ballari et al. 2019).
Threats/impacts: In Argentina, numerous negative impacts by wild boars on
native species have been described, and the species is an important vector of para-
sites and diseases that can potentially affect the native mammals which it coexists
with. Additionally, the wild boar causes damage in agricultural plantations and
preys on young livestock, particularly sheep, while its hoofs destroy and deteriorate
grassland and pasture.

15.4.5  Puma concolor (Linnaeus, 1771) (Order: Carnivora)

Common name: puma, leon americano (Spanish), puma, mountain lion, cougar
(English).
Description: Body length is between 85 and 160 cm and weight between 22 and
100 kg. This cat presents a large variation in size across its distribution range. The
coat is fairly uniform in color and unmarked, but color varies from a buff or sandy
15  Use of Coastal Area Habitats by Land Mammals 415

brown to reddish color, through to light silver and slate gray. The head is small with
dark brown to black patches on the muzzle. The ears are short and rounded. The tail
is long gradually darkening toward the tip (Sunquist and Sunquist 2002) (Fig. 15.3a).
Distribution and habitat: Puma concolor has a very extensive geographic distri-
bution in the Americas and is characterized by a wide ecological plasticity. Across
its distribution, the puma is found from high mountains to deserts (Nowell and
Jackson 1995), including landscapes dominated by human activities (Caruso
et al. 2016).
Behavior: It is a very adaptable species, with great leaping and climbing abilities.
The activity pattern is primarily nocturnal and crepuscular. Puma food habits have
been broadly studied across its distribution, showing a generalist foraging behavior
with a diet composition reflecting the specific prey community found in each region.
The preferred size of puma prey ranges from 70 to 165 kg (Carbone et al. 2014).
Large kills are typically covered with scraped-over vegetation and dirt, and pumas
return frequently to feed on the carcass. These felids are solitary carnivores, although
recent studies have found a preciously unreported degree of sociality (Elbroch and
Quigle 2017). Litters are between 1 and 3 kitties and can reach 6; reproductive age
begins at more than 18 months, and the gestation time is approximately 91.5 days
(Logan and Sweanor 2001).
Conservation status: Puma has been listed in the “Least Concern” category under
both the IUCN Red List (Nielsen et al. 2015) and the Argentine National Assessment
(De Angelo et al. 2019). It is listed in Appendix II of CITES.
Threaten: Puma populations are thought to be declining globally primarily due to
anthropogenic pressure (De Angelo et al. 2011). This felid is considered harmful to
livestock, and for that reason, palliative and/or preventive hunting is a common
practice that may lead to population declines (Guerisoli et  al. 2017). In some
Argentine provinces, puma killing is legal, and bounty systems are in place, but this
is not the case of Buenos Aires Province, where hunting of pumas is illegal. Despite
the ability of the puma to adapt to anthropic changes, there are factors that cause the
loss and degradation of its habitat and may lead to a decrease in the abundance of
the species. Car-struck pumas on routes have been recorded in different regions,
leading to considering this source of mortality as a potential conservation threat.
Mascotism and transmission of diseases from domestic animals could also pose a
threat to the species (Foley et al. 2013).

15.4.6  L
 ycalopex gymnocercus Fisher, 1814
(Order: Carnivora)

Common name: zorro pampeano (Spanish), Pampas fox (English).


Description: L. gymnocercus is a medium-sized fox. Body size varies geographi-
cally. Mean measurements of body mass are between 4.21 and 5.95 kg, and total
body length for adults is between 505 and 800 mm (Crespo 1971; Luengos Vidal
et al. 2009). Pelage on the top and sides of the head is reddish and on the dorsal
416 E. M. Luengos Vidal et al.

rostrum is reddish to black. The ventral surface of the head is pale gray to white.
Back, shoulders, and flanks are gray. A blackish line runs along the center of the
back and tail. The tail is relatively long (50% of the length of head and body), bushy,
and gray with a black tip (Lucherini and Luengos Vidal 2008) (Fig. 15.3b).
Distribution and habitat: The Pampas fox inhabits the Southern Cone of South
America and is one of the most common and widespread carnivores within its geo-
graphic range. This canid prefers open habitats but also occurs in areas of Pampas
grassland modified by extensive ranching and agricultural activities (Lucherini and
Luengos Vida 2008; Caruso et al. 2016).
Behavior: Pampas foxes are considered either abundant or common in most
areas. It is a generalist and adaptable predator. Diet varies geographically and
includes both domestic and wild vertebrates, as well as fruit, insects, carrion, and
garbage. Gestation lasts 55–60 days. It is a monoestric species with a well-defined
period of reproduction (Crespo 1971). Observations of free-ranging foxes in Buenos
Aires Province indicate that litter sizes vary between two and four individuals and
that both parents participate in pup care (Sassola 2006). In the center of Argentina,
home range sizes average 213.3 ± 136.8 ha, without significant variations between
females and males or variations related to natural or modified environments
(Luengos Vidal 2009).
Conservation status: The Pampas fox is included in the “Least Concern” cate-
gory under both the IUCN Red List (Lucherini 2016) and the Argentine National
Assessment (Luengos Vidal et al. 2019). It is listed in Appendix II of CITES.
Threats: This canid is apparently capable to adapt to strong alterations of natural
habitats, and the main threat to its conservation is direct illegal hunting related to the
high level of conflict with livestock, mainly sheep. Although this practice is illegal,
a large part of the ranchers in the center of Argentina uses poison to control the
population of this species, deploying poisoned baits before the sheep shedding
stage, with reportedly very efficient results (Luengos Vidal et al. 2019). There are
reports of predation and harassment of dogs on gray foxes.

15.4.7  Leopardus geoffroyi (d’Orbigny and Gervais, 1844)

Common name: gato montes (Spanish), Geoffroy’s cat (English).


Description: This felid is a medium-sized (430 and 880 cm in length) spotted cat
that weights between 2.2 and 7.8 kg. Males are larger than females (Lucherini et al.
2016). Thus its size is similar to a domestic cat. Its coat color and size vary. The fur
is from gray to ochraceous with numerous black spots about 15 to 20 mm in diam-
eter which tend to group in the central parts of the dorsum two by two so as to
enclose lighter areas. The tail is shorter than that of other Neotropical small cat spe-
cies (Sunquist and Sunquist 2002) (Fig. 15.3c).
Distribution and habitat: L. geoffroyi is widespread in the southern part of South
America. It is also found in a variety of habitats, open and closed and wet and
dry. It is tolerant to a certain degree of habitat alteration and degradation and is
15  Use of Coastal Area Habitats by Land Mammals 417

present in well-preserved areas as well as in sites highly modified by anthropic


activity, either by agricultural activity or by urbanization (Caruso et al. 2016, but see
Castillo et al. 2008).
Behavior: It feeds mainly on small rodents and sometimes birds and lagomorphs.
It has been shown that the relative importance of waterbirds in the diet changed with
seasonal fluctuations in their availability (Manfredi et al. 2004, Canepuccia et al.
2007). This wildcat is mainly, but not exclusively, nocturnal; human disturbance did
not affect the patterns of activity, but cats appeared to be forced to spend more time
active and move over greater distances in more disturbed areas (Manfredi et  al.
2011). Geoffroy’s cats show a very similar use of the space despite the largely
diverse habitats that they inhabit (Tirelli et  al. 2018). In Buenos Aires Province,
home range size varied from 248 ha to 342 ha for a wetland area (Manfredi et al.
2006) and from 220 to 280 ha for the Espinal (Castillo et al. 2019), with male home
ranges approximately 2.5–4.1 larger than those of females.
Conservation status: This cat is listed in the “Least Concern” category under the
IUCN Red List (Pereira et  al. 2015) as well as under the Argentine National
Assessment (Pereira et al. 2019). It is also listed in Appendix II of CITES.
Threats: This cat is primarily threatened by habitat loss and fragmentation, but
also by the hunting by farmers because of its attacks to domestic animals. Predation
by domestic dogs and vehicle collisions are additional sources of mortality (Pereira
et  al. 2010; Caruso et  al. 2017). Although international trade has dramatically
decreased after it was declared illegal, the commercialization of Geoffroy’s cat pelts
can be locally important. Individuals of the species exposed to various pathogens
and parasites shared with domestic dogs and cats have been reported, which consti-
tutes a health risk for the affected populations (Vega et al. 2018).

15.4.8  Galictis cuja (Molina, 1782) (Order: Carnivora)

Common name: hurón menor (Spanish), lesser grison (English).


Description: The lesser grison is a small mustelid that weighs between 800 and
2400 g, with a thin, elongate body with a long neck, narrow chest, short legs, and a
short and bushy tail. Head is small and flat with short, broad, rounded ears. Body
length is between 425 and 667 mm, and there is a clear pattern of sexual size dimor-
phism, with females smaller than males. The coat is gray-yellowish in the upper
parts and black on the face, with a diagonal narrow strip that runs from the forehead
to the shoulder (Redford and Eisenberg 1992; Yensen and Tarifa 2003) (Fig. 15.4a).
Distribution and habitat: Its distribution range includes Argentina, Peru, Bolivia,
Paraguay, and Chile. Although no estimates of population density are available, it is
considered common in grassland, savanna, and steppe habitats, while it appears to
be naturally rare in some high altitude, arid, very dense vegetation habitats.
Behavior: The lesser grison is active mostly during daylight (Tellaeche et  al.
2014; Luengos Vidal et al. 2016). It is not uncommon to observe two or three indi-
viduals feeding or moving together, probably from the same family groups. The
418 E. M. Luengos Vidal et al.

litter size is 2–5 offspring. This mustelid is an active predator. Its diet resembles that
of small felid and includes reptiles, small birds, lagomorphs, and rodents (Redford
and Eisenberg 1992; Punta et al. 2002; Galende and Raffaele 2016).
Conservation status: The lesser grison is listed in the “Least Concern” category
under both the IUCN Red List (Helgen and Schiaffini 2016) and the Argentine
National Assessment (Aprile et al. 2019).
Threats: Habitat degradation due to agricultural activities, excessive overgrazing,
and soil erosion by livestock are the most important threat they face. Harassment
and predation by dogs (feral or domestic) were also documented. Likewise, the
transmission of pathogens by dogs, such as distemper, could also affect this
mustelid. Lesser grisons are occasionally hunted for considering it a good luck
charm. They are kept as pets or trained to eliminate rodents in sheds and hatcheries
(Aprile et al. 2019).

15.4.9  Conepatus chinga (Molina, 1782) (Order: Carnivora)

Common name: zorrino comun (Spanish), Molina’s hog-nosed skunk (English).


Description: C. chinga is a small-sized mephitid, weighing approximately
between 1 and 3 kg and with a body length from 410 to 685 mm from nose to tail.
It has the characteristic skunk coloring with generally black fur and two white
stripes running from the top of the head down the sides of the body to a mostly white
tail, although in some cases the two stripes are absent (Redfor and Eisenberg 1992)
(Fig. 15.4b).
Behavior: It is an omnivore species whose diet includes mainly insects but also
reptile and bird and plats bulbs (Castillo et al. 2014; Kasper et al. 2016). It is mostly
active at night and twilight. Although Molina’s hog-nosed skunks are solitary ani-
mals, data from Buenos Aires Province indicate that males and females may be able
to share dens, presumably in the breeding season (Castillo et al. 2013). The repro-
ductive period of C. chinga seems related to climatic seasons (Kasper et al. 2009).
The gestation period is close to 40 days. Litters that vary from two to three offspring
stay between 4 and 5 months with the mother (Kasper et al. 2009). The home ranges
vary significantly in size between males (243.7  ±  76.5  ha) and females
(120.4 ± 77.6 ha) in Buenos Aires Province and averaged 163 ± 1.17 ha in Southern
Brazil (Kasper et al. 2012). Overlap between home range and core area was exten-
sive between and within sexes (Castillo et al. 2011, Kasper et al. 2004). It digs its
own borrow but may also use borrows dug by other animals (Redford and
Eisenberg 1992).
Distribution and habitat: It is distributed in Argentina, Bolivia, Brazil, Chile,
Paraguay, Peru, and Uruguay. In Argentina, it is present in all provinces, except
Tierra del Fuego. Currently, only one species is recognized throughout Argentina
(Schiaffini et al. 2013b).
15  Use of Coastal Area Habitats by Land Mammals 419

Conservation status: C. chinga is included in the “Least Concern” category under


both the IUCN Red List (Emmons et  al. 2016) and the Argentine National
Assessment (Castillo and Schiaffini 2019). It is listed in Appendix II of CITES.
Threats: Although the abundance of its population is poorly known (Castillo
et  al. 2011; Kasper et  al. 2012), this carnivore appears to be relatively common
throughout its range. Molina’s hog-nosed skunks were heavily hunted for their fur
in Argentina during the 1970s and early 1980s. Additionally, extensive areas of
skunk habitat have been severely degraded through overgrazing and soil erosion by
livestock (primarily sheep) and feral, exotic species. It is often used as a pet and
persecuted by rural inhabitants for the consumption of eggs and poultry (Castillo
and Schiaffini 2019).

15.5  A
 Study Case: Maritime Carnivores of a Coastal Area
of the Buenos Aires Province

Buenos Aires is the second largest and one of the most developed provinces in
Argentina. It concentrates most livestock and cropland activities. This province
once hosted the natural grasslands typical of the Pampas ecoregion, but since the
arrival of the European colonizers, it has been going through a human-driven pro-
cess of fragmentation, which caused the loss of more than 60% of its original habi-
tats (Brown et  al. 2006). This situation has had a great impact on the ecology,
behavior, and conservation of several vertebrate species and particularly of mam-
malian carnivores, along with the role that they play in the ecosystems (see previous
section).
Although the basic ecological information available for carnivores in the prov-
ince has increased during the last decades, most of the projects studying these spe-
cies focused on inland areas; almost no data is available for coastal areas. One of
those projects was carried out on Campos del Tuyú Wildlife Reserve (36°20′S,
56°50′W) and mainly focused on Geoffroy’s cat. Data on its food habits were
reported (Manfredi et  al. 2004), along with characterization of defecation sites
(Soler et al. 2009), and spatial ecology (home range size and habitat use) (Manfredi
et al. 2006), and the general activity pattern (Manfredi et al. 2011) of radio-tracked
Geoffroy’s cat individuals. Mar Chiquita Biosphere Reserve (37°46′S, 57°27′W) is
another coastal area represented in the literature. Manfredi et  al. (2004) and
Canepuccia et al. (2007) studied Geoffroy’s cat food habits in general and in rela-
tion to waterbird abundance, size, and distance before attack; and Farias and Kittlein
(2008) reported data on Pampas fox food habits. Finally, Caruso et al. (2012) pub-
lished the first record of occurrence of Geoffroy’s cat in the Bahía Blanca, Bahía
Falsa, and Bahía Verde Natural Reserve (38°56’S; 62°14’W).
Given this clear paucity of information, we present the ecological data on the
carnivore community obtained in a coastal area that represents a relic of grasslands
in the Buenos Aires Province. Because the natural grasslands have almost
420 E. M. Luengos Vidal et al.

disappeared due to their transformation into croplands, these relicts have a high
conservation value (Miñarro 2004).
Fieldwork was carried out in the “Arroyo Los Gauchos Nature Reserve,” located
in the southwestern Buenos Aires Province (Coronel Dorrego county, 38°56′1.93”S,
60°45′9.37”O; Fig. 2.1; Chap. 2 of this book). The “Arroyo Los Gauchos Nature
Reserve” is a coastal protected area of 7.07 km2 that belongs to the Pampas ecore-
gion and is characterized by the presence of dunes, both non-vegetated and covered
by psammophyte vegetation, and a general low topography with interdune depres-
sions occupied by relatively small wetlands and halophyte grasses. The climate is
temperate, the annual mean temperature is 14.1 °C, and the annual mean precipita-
tion is approximately 850 mm (Monserrat et al. 2012; Celsi and Giussani 2019).
Although the study area is formally protected by the provincial government since
2011, it lacks implementation, and its habitats are preserved primarily because of
the difficulty of access. Until now, human disturbance has had a low impact in com-
parison to the other coastal areas of the province. Activities such as cattle raising,
forestry, and tourism are still limited, and urban development is scarce (Monserrat
and Celsi 2009; Monserrat and Codignotto 2013). The area was proposed as a valu-
able grassland area (Miñarro 2004) well before its establishment as a protected area
due to its high biodiversity (more than 72 vascular plant species) and the presence
of endemic rare or endangered vertebrate species, among which are Ctenomys aus-
tralis and a lizard, Liolaemus multimaculatus (Monserrat and Celsi 2009).
To characterize the carnivore community of the study area, we conducted a sur-
vey using sign counts and camera trapping from January 2009 to March 2009.
To study the habitat use by carnivores, we conducted 1-km-long linear transects,
looking for indirect evidences of the presence of mostly feces and tracks. Both the
starting point and the bearing of transects were placed randomly over the study area.
We sampled each transect walking at a constant speed along 1 km and recorded the
GPS position of each evidence found. We characterized the habitat surrounding
each evidence within a radius of 20 m by estimating visually the proportion of each
habitat type. Seven habitat types were identified as the most characteristics in the
study area: bare ground, mostly non-vegetated dunes or any other portion of land
without vegetation cover; low grassland, grasses (Poaceae) less than 30  cm tall;
medium-tall grassland, areas covered by grassy vegetation (Poaceae) taller than
30 cm; areas covered by Juncus spp.; areas covered by Cortaderia spp.; areas cov-
ered by Hyalis argentea; and shrubland, mostly dominated by Lycium chilense,
Baccharis divaricata, and Discaria americana. Additionally, we used camera traps
to increase the chances of recording rarer species that are more difficult to detect by
sign counts and thus complete the information on the composition of the carnivore
community.
In 14 line transects, we recorded 67 mammal evidences from 5 different taxa:
Pampas fox, Geoffroy’s cat, skunk, armadillo (probably large hairy armadillo), and
rodents. Feces represented 62.8% of the evidences, followed by tracks (34.3%). The
Pampas fox was the most frequent species, with 33 records, followed by Geoffroy’s
cat, with 15 records. Given the lack of reliable information for the rest of species on
the base of their scarce evidence, we focused our analysis on the two most common
carnivore species: the Pampas fox and Geoffroy’s cat. We used logistic regressions
15  Use of Coastal Area Habitats by Land Mammals 421

to study if any of the variables registered affected the probability of finding evi-
dences of these species. Evidences were treated as a dummy variable (i.e., evidence
corresponding to L. gymnocercus or evidence corresponding to L. geoffroyi), and
the habitat variables were used as predictors. Using the function “dredge” of the
package MuMIn in R (R Core Team 2013), we created a list of all possible models
given the list of predictors and assuming only additive effects. We ranked all models
using Akaike’s information criterion adjusted for small samples (AICc) and used
the ΔAICc and the AICc weights to evaluate the relative importance of each model
within the final set of models (Burnham and Anderson 2002). We used a multimodel
inference approach (Burnham and Anderson 2002) and averaged the set of models
with ΔAICc <2. The regression coefficients (β) were considered to have a signifi-
cant effect when the corresponding 95% confidence intervals (CI) excluded zero
(Zeller et  al. 2011). With respect to the habitat surrounding each evidence, five
models formed the set with ΔAICc <2 (Table 15.2). Proportion of Cortaderia spp.
and distance to shoreline were the only variables included in the five models. The
averaged model and the confidence intervals suggested that a greater proportion of
Cortaderia ssp. reduced the chance of detecting the presence of L. geoffroyi in rela-
tion to L. gymnocercus, which would indicate that Pampas foxes use more inten-
sively the sites with more Cortaderia spp. than Geoffroy’s cats. The distance to the
shoreline increased the chance of presence of L. gymnocercus signs, thus indicating
that Geoffroy’s cat tended to use site closer to the shoreline.
Camera trapping produced records for five carnivore species: Pampas fox,
Geoffroy’s cat, Molina’s hog-nosed skunk Conepatus chinga, puma Puma con-
color, and lesser grison Galictis cuja. These are the most common carnivores in the
Pampas agroecosystems (Luengos Vidal et al. 2005) and also in a sandy dune area
in the Espinal ecoregion located not very far from our study area (Caruso et  al.
2012). Herpailurus yagouaroundi and Leopardus colocolo were not detected during
the study and are presumably absent from our study area because the local people

Table 15.2  Parameters of the generalized linear models (GLMs) contrasting the habitat use of
L. gymnocercus and L. geoffroyi. Only parameters for the set of models with ΔAICc <2 are reported
1 Int. S BG C D J df logLik AICc delta weight
1 0.76 −0.68 −0.77 3 −27.85 62.23 0.00 0.29
2 0.47 −2.46 −0.70 −0.76 4 −26.80 62.51 0.28 0.25
3 0.48 −2.55 −0.56 −0.98 −1.02 5 −25.90 63.20 0.98 0.18
4 0.76 −0.41 −0.88 −0.97 4 −27.32 63.55 1.32 0.15
5 0.76 −0.78 −0.95 0.35 4 −27.35 63.62 1.39 0.14
Avr. 0.64 −2.50 −0.49 −0.78 −0.87 0.35
CI −63.59 −693.10 −1.35 −1.55 −1.71 −0.37
(2.5%)
CI 64.87 688.10 0.36 −0.01 −0.02 −1.07
(97.5%)
Int intercept, S proportion of shrubland, BG proportion of bare ground, C proportion of Cortaderia
spp., D distance to shoreline, J proportion of Juncus spp., d.f. degrees of freedom, logLik = log
likelihood function, ΔAICc  =  difference in value of Akaike’s information criterion between the
focal model and the top-ranked model, Avr. average model coefficients, CI confidence interval
422 E. M. Luengos Vidal et al.

did not recognize those (Caruso et al. 2020). This is not surprising given the rarity
of both species in more typical cropland areas far from the coast (Benzaquín et al.
2009; Luengos Vidal et al. 2017). The following non-carnivore species were also
found in the study area: the red deer Cervus elaphus, European hare Lepus euro-
paeus, wild boar Sus scrofa, greater rhea Rhea americana, and several rodents and
armadillos. L. gymnocercus and L. geoffroyi were the most frequently recorded spe-
cies, followed by C. chinga, P. concolor, and G. cuja.
With a relatively limited surveying effort, camera traps enabled us to detect five
of the seven species of carnivores that may potentially occur in this coastal area.
They also confirmed that the Pampas fox and Geoffroy’s cat were the most common
species in the carnivore community of southern Buenos Aires coastal habitats.
There are many factors that can affect the presence and use of habitat of a spe-
cies, and it is very difficult to draw conclusions with a relatively brief survey in a
heterogeneous habitat. For example, the stranding of marine vertebrate carcasses on
the coast, which did not occur in the period during which the survey was carried out,
could have substantially changed the results, given the importance of carcasses on
the ecology of coastal scavengers such as canids or vultures (see the previous sec-
tions of this chapter). However, our results suggest a prevalence of foxes in environ-
ments with Cortaderia. This plant grows in coarse widespread patches, is associated
with lightly humid soils, and is used as shelter and food resources by different spe-
cies of rodents (Bonaventura et al. 2003) and birds (Pretelli et al. 2013). Thus, the
use by a trophic generalist species like the Pampas fox (García and Kittlein 2005;
Lucherini and Luengos Vidal 2008; Bossi et al. 2019) may be associated with the
presence of these potential sources of food. Geoffroy’s cat is more specialized in its
food habits and may find the areas close to the shoreline more suitable for hunting
birds, as reported in another coastal area (Canepuccia et al. 2007). However, because
intraguild competition is a structuring factor in carnivores (e.g., May et al. 2008; Di
Bitetti et al. 2010), based on our data, we cannot exclude the possibility that inter-
specific competition plays an important role in explaining carnivore habitat use in
addition to the distribution of resources.
Although further studies would be needed to confirm and complete the informa-
tion reported here, our findings suggest that the coastal dune areas to the east of
Bahía Blanca are inhabited by a relatively diverse vertebrate community and shelter
species that are considered rare in the human-dominated habitats that characterize
today’s Pampas landscapes, such as P. concolor, L. geoffroyi, and R. americana,
thus supporting the conservation.

15.6  Final Remarks

Understanding the relevance of the role played by the contribution of the sea toward
the terrestrial environment, as we have tried to show in this chapter, is of great utility
not only in providing information that allows changing the ecological theoretical
framework on the influence of the ocean on terrestrial trophic networks but also for
15  Use of Coastal Area Habitats by Land Mammals 423

the management and conservation of coastal biota. Our understanding of the inter-
action between marine and terrestrial ecosystems is very scarce for the area of the
estuary of Bahía Blanca, and we have only been able to lay down some baseline
information here to guide future investigations that are urgently needed to contrib-
ute to the conservation of the coasts of the southern part of Buenos Aires Province
and all the biodiversity that they host.

Box 15.1 The Impact of Coastal Urbanization on Land Mammals


Urbanization (defined as the natural growth of urban centers associated with
an increased in their population or with a high density of inhabitants and the
need to cover their requirement) is particularly concentrated along the world’s
coasts and is growing rapidly, which makes of these areas represent very
important hotspots of habitat modification (Martínez et al. 2007). Globally,
one of the main factors affecting coastal biodiversity is urbanization, because
its main consequences are habitat loss and fragmentation, but also because it
causes overexploitation of fish stocks, pollution, and depletion of populations
of coastal wildlife (McKinney 2002; McDonald et al. 2008).
The coastal infrastructure associated with recreational activities may be
vast, including roads and paths to access beaches, waterfront housing devel-
opment, car parking, picnic spots, swimming pools, service areas, beach facil-
ities (toilets, lifeguard facilities), navigation canals, drainage, and stream
canalization to reclaim coastal wetlands. These infrastructures lead to varied
conflicts since they may alter the quality of the landscape and disturb ecosys-
tem functioning, as they modify the coastline; cause erosion; interrupt sedi-
ment transport; alter dunes, reefs, and wetlands; increase traffic intensity,
noise, and the contamination of air and water; and alter the patches of dune
vegetation for the construction of houses which leads to the loss of native
plant species and favors their replacement by exotic (Jędrzejczak 2004).
Although the information about the impacts of this infrastructure on the
coastal mammals is still scarce, some effects have been demonstrated and oth-
ers can be predicted.
Barriers to Movement of Animals and Wrack
Along with habitat loss, the alteration of the physical processes that affect the
deposition and retention of sediments on modificated coasts may also affect
the deposition and retention of buoyant material, including macrophyte
wrack, driftwood, and other natural allochthonous debris, which can be
important to biota as food or habitat. Predators, such as shorebirds or mari-
time mammals, which usually move between the dunes and the beach, mostly
to feed on stranded material (see Sect. 15.1 in this chapter), can respond to a
combination of habitat loss, decreased accessibility at particular sites, and
reduced food availability. For instance, structures such as seawalls that are
424 E. M. Luengos Vidal et al.

placed on the coastline negatively affected some animals such as small rodents
and other land mammals (Carlton and Hodder 2003; Bird et al. 2004).
Changes in Species Distribution or Abundance
Negative effects of coastal urbanization such as the decrease in the abundance
of species that are threatened of extinction due to loss of habitats have been
reported in different vertebrate taxa. Schmidt et al. (2012) found evidence of
the synergistic impacts from rising sea level and coastal development in
Lower Florida Keys (USA) on population decline in an endangered mammal,
the Lower Keys marsh rabbit (Sylvilagus palustris hefneri), and that several
bird species modified the selection of nesting sites, due to anthropic distur-
bances on the beaches, such as vehicle access. Meager et al. (2012) observed
that human activities occurred mostly in the mid to lower intertidal zone of the
beach in Moreton and North Stradbroke Island, eastern Australia, overlapping
closely with the preferred habitats of silver gulls (Larus novaehollandiae),
pied oystercatchers (Haematopus longirostris), red-capped plovers
(Charadrius ruficapillus), and endangered little terns (Sternula albifrons).
Wrack and Carrion Removal
Animal carcasses are fundamental resources in many food webs, supporting a
diverse and abundant suite of scavengers and structuring ecosystems (e.g.,
Wilson and Wolkovich 2011; Moleón et al. 2014). In fact, carcasses of marine
animals washing up on sandy beaches are a critical input for scavengers feed-
ing at the sea-­land interface (see Sect. 15.1 in this chapter). The consumption
of beach-cast marine carrion on sandy shorelines is rapid and often complete,
suggesting evolutionary pathways resulting in a pivotal role of carrion in food
webs of these ecosystems. Moreover, carrion is utilized by a variety of spe-
cies, both native and introduced, that are highly adapted to changes in resource
supply and capable of reacting rapidly to these changes. A comparison
between scavenging dynamics on sandy beaches found at two extreme levels
of urbanization in Australia showed that invasive mammals replaced native
raptors as scavengers on urban beaches, resulting in a significant decline in
scavenging efficiency (Huijbers et al. 2013; Schlacher et al. 2015). The coastal
areas of Buenos Aires Province are probably no exception to this kind of con-
flict. The removal of carrion on the coasts by dogs affects, through competi-
tion for resources, different local scavenger species such as the Pampas fox
(obs. Pers.).
Introduction of Exotic Species
In general, the presence of dogs (Canis familiaris) and domestic cats (Felis
catus) in urban areas represents one of the greatest negative threats to local
biodiversity (Hughes and Macdonald 2013; Weston and Stankowich 2014;
Loss et al. 2013). They can be predators or efficient competitors of native spe-
cies and carriers of diseases that can affect wildlife. Hughes and Macdonald
15  Use of Coastal Area Habitats by Land Mammals 425

(2013) found that mammals were the most studied taxa regarding the interac-
tion between dogs and wildlife. Predation was one of the greatest impacts
caused by dogs, followed by the transmission of diseases, the disturbance on
wild species (which carries energetic costs for them), hybridization, and com-
petition for food. The transmission of dog diseases to wildlife may be caused
by direct interactions between species, but the fecal contamination of dogs in
waterways has indirect negative effects also on marine mammal health
(Holderness-Roddam 2011). Holderness-Roddam’s (2011) review deter-
mined that feral and domestic cat predation on small vertebrates is significant,
but both in poorly modified urban areas and on beaches and their surround-
ings, domestic dogs prey even more significantly on some species. Some
damage may be caused by people walking through these areas with their pets,
but the impact is even greater if dogs are allowed to go unleashed.
Rats are other highly invasive species whose impact over the coastal fauna
has been unambiguously documented. Several authors have suggested that in
Madagascar, Rattus rattus competes for resources with rodents of the endemic
subfamily Nesomyinae and subsequently replaces them (Goodman 1995).
Jones et  al. (2008) examined 94 manuscripts that demonstrated effects of
exotic rats on seabirds and found that they affect 75 species of seabirds and
that the consequences of rat predation on seabirds are independent of time
since rat introduction. Martin et al. (2000) studied islands in the Mediterranean
and discovered that storm petrels (Hydrobates pelagicus) are limited to rat-
free islands. Finally, rats may cause extinctions of endemic mammals on
Australian islands through introduction of novel diseases and competition
(Smith and Banks 2014).
Although there are essentially no studies on the impact of coastal urbaniza-
tion on terrestrial mammals in the Buenos Aires Province, based on our review
of the literature on this topic, we conclude that there is a great need to assess
not only the changes in habitat but also the effects of urban development pro-
cesses on native species and use this information to design plans for reducing
this type of environmental impact.
As a recommendation, strategies that preserve the size and connectivity of
patches of native vegetation and limit the introduction of non-native predators
such as rats and domestic dogs and cats, in coastal urban areas where man and
terrestrial mammals cohabit, should be considered priorities for the expected
future advance of cities on the coasts.

References

Abba AM, Poljak S, Superina M (2014) Chaetophractus villosus. The IUCN Red List of
Threatened Species: e.T4369A47438745. https://doi.org/10.2305/IUCN.UK.2014-­1.RLTS.
T4369A47438745.en. Accessed 13 Feb 2020
Aprile G, Cirignoli S, Varela D et al (2019) Galictis cuja. In: SAyDS–SAREM (eds) Categorización
2019 de los mamíferos de Argentina según su riesgo de extinción. Lista Roja de los mamíferos
de Argentina. http://cma.sarem.org.ar. Accessed 13 Feb 2020
426 E. M. Luengos Vidal et al.

Arriagada A, Baessolo L, Saucedo C et al (2017) Hábitos alimenticios de poblaciones periféricas


de Zaedyus pichiy y Chaetophractus villosus (Cingulata, Chlamyphoridae) en la Patagonia
chilena. Iheringia Sér Zool 107:1–8. https://doi.org/10.1590/1678-­4766e2017013
Arroyo NL, Bonsdorff E (2016) The role of drifting algae for marine biodiversity In: E Olafson
(ed) Mar Macrophytes as Found Species CRC Press Boca Raton p 100
Austrich A, Mapelli FJ, Mora M et  al (2019) Ctenomys australis. In: SAyDS–SAREM (eds)
Categorización 2019 de los mamíferos de Argentina según su riesgo de extinción. Lista Roja
de los mamíferos de Argentina. http://cma.sarem.org.ar. Accessed 13 Feb 2020
Ballari SA, Cirignoli S, Winter M et al (2019) Sus scrofa. In: SAyDS–SAREM (eds) Categorización
2019 de los mamíferos de Argentina según su riesgo de extinción. Lista Roja de los mamíferos
de Argentina. http://cma.sarem.org.ar. Accessed 13 Feb 2020
Barreiro F, Gómez M, López J et  al (2013) Coupling between macroalgal inputs and nutrients
outcrop in exposed sandy beaches. Hydrobiologia 700:73–84. https://doi.org/10.1007/
s10750-­012-­1220-­z
Benzaquín M, Manfredi C, Luengos Vidal EM et al (2009) Estudio de la comunidad de felinos
en una región del espinal del sudoeste de la Provincia de Buenos Aires.In: Cazzaniga NJ,
Arelovich H (eds) Ambientes y recursos naturales del sudoeste bonaerense: Producción, con-
taminación y conservación. (Actas de las V Jornadas Interdisciplinarias del SO Bonaerense).
EdiUNS, Bahía Blanca
Bidau C (2016) Ctenomys talarum. The IUCN Red List of Threatened Species: e.T5828A22195175.
https://doi.org/10.2305/IUCN.UK.2016-­2.RLTS.T5828A22195175.en. Accessed on 14
Feb 2020
Bird BL, Branch LC, Miller DL (2004) Effects of coastal lighting on foraging behavior of beach
mice. Conserv Biol 18:1435–1439. https://doi.org/10.1111/j.1523-­1739.2004.00349.x
Bisceglia S, Palacios R, Quiroga VA et al (2019) Herpailurus yagouaroundi. In: SAyDS–SAREM
(eds) Categorización 2019 de los mamíferos de Argentina según su riesgo de extinción. Lista
Roja de los mamíferos de Argentina. Versión digital: http://cma.sarem.org.ar. Accessed 13
Feb 2020
Bolkovic ML, Quintana R, Cirignoli S et  al (2019) Hydrochoerus hydrochaeris. In: SAyDS–
SAREM (eds) Categorización 2019 de los mamíferos de Argentina según su riesgo de extin-
ción. Lista Roja de los mamíferos de Argentina. http://cma.sarem.org.ar
Bonaventura SM, Pancotto V, Madanes N et al (2003) Microhabitat use and density of sigmodon-
tine rodents in Spartina densiflora freshwater marshes, Argentina. Mammalia 67(3):367–378.
https://doi.org/10.1515/mamm.2003.67.3.367
Borboroglu PG, Yorio P (2004) Habitat requirements and selection by kelp gulls (Larus dominica-
nus) in central and northern Patagonia, Argentina. Auk 121:243–252. https://doi.org/10.1093/
auk/121.1.243
Bortolus A, Carlton JT, Schwind E (2015) Reimagining South America coasts: unveiling the hid-
den invasion history of an iconic ecological engineer. Divers Distrib 21(11):1267–1283. https://
doi.org/10.1111/ddi.12377
Bossi MAS, Migliorini RP, Santos TG et al (2019) Comparative trophic ecology of two sympatric
canids in the Brazilian Pampa. J Zool 307:215–222. https://doi.org/10.1111/jzo.12636
Broekhuis F, Grünewälder S, McNutt JW (2014) Optimal hunting conditions drive circalu-
nar behavior of a diurnal carnivore. Behav Ecol 25:1268–1275. https://doi.org/10.1093/
beheco/aru122
Brown A, Martinez Ortiz U, Acerbo M et  al (2006) La situación ambiental argentina 2005.
Fundación Vida Silvestre, Buenos Aires
Burnham K, Anderson D (2002) Model selection and multi-model inference, 2nd edn. Springer,
New York
Canepuccia AD, Martinez MM, Vassallo AI (2007) Selection of waterbirds by Geoffroy’s
cat: effects of prey abundance, size, and distance. Mamm Biol 72:163–173. https://doi.
org/10.1016/j.mambio.2006.07.003
Canevari M, Balboa F (2003) 100 Mamíferos argentinos. Editorial Albatros, Buenos Aires
15  Use of Coastal Area Habitats by Land Mammals 427

Carbone C, Codron D, Scofield C et  al (2014) Geometric factors influencing the diet of verte-
brate predators in marine and terrestrial environments. Ecol Lett 17:1553–1559. https://doi.
org/10.1111/ele.12375
Carlton JT, Hodder J (2003) Maritime mammals: terrestrial mammals as consumers in
marine intertidal communities. Mar Ecol Prog Ser 256:271–286. https://doi.org/10.3354/
meps256271
Caruso NC, Sotelo M, Luengos Vidal EM (2012) Primer registro de gato montés, Leopardus
geoffroyi, en la reserva Bahía Blanca, Bahía Falsa, Bahía Verde (RNBBBFBV), provincia de
Buenos Aires. BioScriba 5:54–59
Caruso N, Lucherini M, Fortin D et al (2016) Species-specific responses of carnivores to human-­
induced landscape changes in Central Argentina. PLoS One 11(3). https://doi.org/10.1371/
journal.pone.0150488
Caruso N, Luengos Vidal EM, Lucherini M et al (2017) Carnivores in the southwest of the prov-
ince of Buenos Aires: ecology and conflicts with farmers. RIA, Rev Inves Agropec 43:165–174
Caruso N, Valenzuela AEJ, Burdett CL et al (2018) Summer habitat use and activity patterns of
wild boar Sus scrofa in rangelands of Central Argentina. PLoS One 13:e0206513. https://doi.
org/10.1371/journal.pone.0206513
Caso A, de Oliveira T, Carvajal SV (2015) Herpailurus yagouaroundi. The IUCN Red List of
Threatened Species: e.T9948A50653167. https://doi.org/10.2305/IUCN.UK.2015-­2.RLTS.
T9948A50653167.en. Accessed on 15 Feb 2020
Castillo DF, Schiaffini M (2019) Conepatus chinga. In: SAyDS–SAREM (eds) Categorización
2019 de los mamíferos de Argentina según su riesgo de extinción. Lista Roja de los mamíferos
de Argentina. http://cma.sarem.org.ar. Accessed 13 Feb 2020
Castillo D, Luengos Vidal EM, Lucherini M et al (2008) First report on the geoffroyi cat in a highly
modificated area of the argentine areas. Cat News 49:27–28
Castillo DF, Lucherini M, Luengos Vidal EM et al (2011) Spatial organization of Molina’s hog-­
nosed skunk (Conepatus chinga) in two landscapes of the pampas grassland of Argentina. Can
J Zool 89:229–238. https://doi.org/10.1139/Z10-­110
Castillo DF, Luengos Vidal EM, Caruso N et  al (2013) Denning ecology of Conepatus chinga
(Carnivora: Mephitidae) in a grassland relict of Central Argentina. Mastozool Neotrop
20:373–379
Castillo DF, Luengos Vidal EM, Casanave EB (2014) Feeding habits of Molina’s hog-nosed
skunk in the pampas grassland of Argentina. Mammalia 78:473–447. https://doi.org/10.1515/
mammalia-­2013-­0020
Castillo DF, Luengos Vidal EM, Caruso NC et al (2019) Spatial organization and habitat selection
of Geoffroy’s cat in the Espinal of Central Argentina. Mamm Biol 94(1):30–37. https://doi.
org/10.1016/j.mambio.2018.12.003
Caruso NC, Sotelo M, Luengos Vidal EM (2012) Primer registro de gato montés, Leopardus
geoffroyi, en la reserva Bahía Blanca, Bahía Falsa, Bahía Verde (RNBBBFBV), provincia de
Buenos Aires. BioScriba 5:54–59
Caruso NC, Luengos Vidal EM, Manfredi MC, Araujo MS, Lucherini M, Casanave EB (2020).
Spatio-temporal interactions of carnivore species in a coastal ecosystem in Argentina. Ocean
Coast Manag 198:105311. https://doi.org/10.1016/j.ocecoaman.2020.105311
Celsi CE, Giussani LM (2019) Geographical distribution and habitat characterization of Poa
schizantha (Poaceae), a narrow endemic of the coastal sand dunes of the southern pampas,
Argentina. Bot J Linn Soc 192:296–313. https://doi.org/10.1093/botlinnean/boz069
Chemisquy MA, Martin GM (2019) Didelphis albiventris. In: SAyDS–SAREM (eds)
Categorización 2019 de los mamíferos de Argentina según su riesgo de extinción. Lista Roja
de los mamíferos de Argentina. Versión digital: http://cma.sarem.org.ar. Accessed 13 Feb 2020
CMA-Categorización 2019 de los Mamíferos de Argentina según su riesgo de extinción. Lista
Roja de los mamíferos de Argentina (2019) SAREM http://cma.sarem.org.ar
Colombini I, Chelazzi L, Gibson RN et al (2003) Influence of marine allochthonous input on sandy
beach communities. Oceanogr Mar Biol An Annu Rev 41:115–159
428 E. M. Luengos Vidal et al.

Contreras JR, Reig OA (1965) Datos sobre la distribución del género Ctenomys (Rodentia,
Octodontidae) en la zona costera de la Provincia de Buenos Aires comprendida entre Necochea
y Bahía Blanca. Physis 25:169–186
Corriale MJ, Muschetto E, Herrera EA (2013) Influence of group sizes and food resources
in home-range sizes of capybaras from Argentina. J Mammal 94:19–28. https://doi.
org/10.1644/12-­MAMM-­A-­030.1
Costa LP, Astua de Moraes D, Brito D et al (2015) Didelphis albiventris. The IUCN Red List of
Threatened Species: e.T40489A22176404. https://doi.org/10.2305/IUCN.UK.2015-­4.RLTS.
T40489A22176404.en. Accessed on 15 Feb 2020
Crespo JA (1971) Ecología del zorro gris Dusicyon gymnocercus antiquus (Ameghino) en la pro-
vincia de La Pampa. Rev Mus Arg Cs Nat 1:147–205
Croll DA, Maron JL, Estes JA et al (2005) Introduced predators transform subarctic islands from
grassland to tundra. Sci 307:1959–1961. https://doi.org/10.1126/science.1108485
Cueto GR (1999) Biología reproductiva y crecimiento del carpincho (Hydrochoerus hydrochaeris)
en cautiverio: Una interpretación de las estrategias poblacionales. PhD Thesis. Universidad de
Buenos Aires, Buenos Aires
Cutrera AP, Antinuchi CD, Mora MS et al (2006) Home-range and activity patterns of the south
American subterranean rodent Ctenomys talarum. J Mammal 87:1183–1191. https://doi.
org/10.1644/05-­MAMM-­A-­386R1.1
De Angelo C, Paviolo A, Rode D et al (2011) Participatory networks for large-scale monitoring
of large carnivores: pumas and jaguars of the upper Paraná Atlantic Forest. Oryx 45:534–545.
https://doi.org/10.1017/S0030605310000840
De Angelo C, Llanos R, Guerisoli MM et al. (2019) Puma concolor. In: SAyDS–SAREM (eds)
Categorización 2019 de los mamíferos de Argentina según su riesgo de extinción. Lista Roja
de los mamíferos de Argentina. http://cma.sarem.org.ar
Dehn LA, Follmann EH, Thomas DL et al (2006) Trophic relationships in an Arctic food web and
implications for trace metal transfer. Sci Total Environ 362:103–123. https://doi.org/10.1016/j.
scitotenv.2005.11.012
Di Bitetti MS, De Angelo CD, Di Blanco YE et al (2010) Niche partitioning and species coexis-
tence in a Neotropical felid assemblage. Acta Oecol 36(4):403–412. https://doi.org/10.1016/j.
actao.2010.04.001
Doumecq Milieu RE, Morici A, Nigro NA (2012) Ampliación de la distribución austral del car-
pincho (Hydrochoerus hydrochaeris) en la provincia de Buenos Aires, Argentina. Nótulas
Faunísticas 92:1–10
Elbroch LM, Quigle H (2017) Social interactions in a solitary carnivore. Current zoology
63(4):357–362. https://doi.org/10.1093/cz/zow080
Elgueta EI, Valenzuela J, Rau JR (2007) New insights into the prey spectrum of Darwin′ s fox
(Pseudalopex fulvipes Martin, 1837) on Chiloé Island, Chile. Mamm Biol 72:179–185. https://
doi.org/10.1016/j.mambio.2006.07.004
Elmhagen B, Tannerfeldt M, Verucci P et al (2000) The arctic fox (Alopex lagopus): an opportunis-
tic specialist. J Zool 251:139–149. https://doi.org/10.1111/j.1469-­7998.2000.tb00599.x
Emmons L, Schiaffini M, Schipper J (2016) Conepatus chinga. The IUCN Red List of
Threatened Species: e.T41630A45210528. https://doi.org/10.2305/IUCN.UK.2016-­1.RLTS.
T41630A45210528.en. Accessed on 13 Feb 2020
Farias AA, Kittlein MJ (2008) Small-scale spatial variability in the diet of pampas foxes
(Pseudalopex gymnocercus) and human-induced changes in prey base. Ecol Res 23:543–550.
https://doi.org/10.1007/s11284-­007-­0407-­7
Fernández GP, Carnovale CS, Mora MS (2019) Ctenomys talarum. En: SAyDS–SAREM (eds)
Categorización 2019 de los mamíferos de Argentina según su riesgo de extinción. Lista Roja
de los mamíferos de Argentina. http://cma.sarem.org.ar
Fernández-Llario P, Carranza J, De Trucios SJH (1996) Social organization of the wild boar (Sus
scrofa) in Doñana National Park. Misc Zoo 19:9–18
Foley JE, Swift P, Fleer KA et al (2013) Risk factors for exposure to feline pathogens in California
mountain lions (Puma concolor). J Wildl Dis 49:279–293. https://doi.org/10.7589/2012-
­08-­206
15  Use of Coastal Area Habitats by Land Mammals 429

Frafjord K (1993) Food habits of arctic foxes (Alopex lagopus) on the western coast of Svalbard.
Arctic 49–54
Frere E, Millones A, Morgenthaler A et al (2010) High predation rates by pumas on Magellanic
Penguin adults: new conflicts in coastal protected areas in Argentina. 1st World Seabird
Conference Poster Session
Galende GI, Raffaele E (2016) Predator feeding ecology on Patagonian rocky outcrops: implica-
tions for colonies of mountain vizcacha (Lagidium viscacia). Stud Neotrop Fauna Environ
51:104–111. https://doi.org/10.1080/01650521.2016.1185270
Gallo JA, Poljak S, Abba, AM et al (2019) Chaetophractus villosus. In: SAyDS–SAREM (eds)
Categorización 2019 de los mamíferos de Argentina según su riesgo de extinción. Lista Roja
de los mamíferos de Argentina. http://cma.sarem.org.ar
García VB, Kittlein MJ (2005) Diet, habitat use, and relative abundance of pampas fox
(Pseudalopex gymnocercus) in northern Patagonia, Argentina. Mamm Biol 70:218–226.
https://doi.org/10.1016/j.mambio.2004.11.019
Garcia EA, Bertness MD, Alberti J et  al (2011) Crab regulation of cross-ecosystem resource
transfer by marine foraging fire ants. Oecologia 166:1111–1119. https://doi.org/10.1007/
s00442-­011-­1952-­x
Gaydos JK, Pearson SF (2011) Birds and mammals that depend on the Salish Sea: a compilation.
Northwest Nat 92:79–94. https://doi.org/10.1898/10-­04.1
Gomez Villafañe IE, Miño M, Cavia R et al (2005) Guía de Roedores de la provincia de Buenos
Aires. Editor LOLA (Literature Lat Am Argentina) Buenos Aires
Goodman SM (1995) Rattus on Madagascar and the dilemma of protecting the endemic rodent
fauna. Conserv Biol 9:450–453. https://doi.org/10.1046/j.1523-­1739.1995.9020450.x
Guerisoli M d l M, Luengos Vidal EM, Franchini M et al (2017) Characterization of puma–live-
stock conflicts in rangelands of Central Argentina. R Soc Open Sci 4:170852. https://doi.
org/10.1098/rsos.170852
Heatwole H (1972) Marine-dependent terrestrial biotic communities on some cays in the Coral
Sea. Ecology 52:363–366. https://doi.org/10.2307/1934598
Helgen K, Schiaffini M (2016) Galictis cuja. The IUCN Red List of Threatened Species:
e.T41639A45211832. https://doi.org/10.2305/IUCN.UK.2016-­1.RLTS.T41639A45211832.
en. Accessed on 13 Feb 2020
Hersteinsson P, Macdonald DW (1996) Diet of arctic foxes (Alopex lagopus) in Iceland. J Zool
240:457–474. https://doi.org/10.1111/j.1469-­7998.1996.tb05298.x
Holderness-Roddam B (2011) The effects of domestic dogs (Canis familiaris) as a disturbance
agent on the natural environment. PhD Thesis. University of Tasmania
Hughes J, Macdonald DW (2013) A review of the interactions between free-roaming domestic
dogs and wildlife. Biol Conserv 157:341–351. https://doi.org/10.1016/j.biocon.2012.07.005
Huijbers CM, Schlacher TA, Schoeman DS et  al (2013) Urbanisation alters processing of
marine carrion on sandy beaches. Landsc Urban Plan 119:1–8. https://doi.org/10.1016/j.
landurbplan.2013.06.004
Jędrzejczak MF (2004) The modern tourist’s perception of the beach: is the sandy beach a place of
conflict between tourism and biodiversity. J Coast Res 2:109–119
Jepsen JU, Eide NE, Prestrud P et al (2002) The importance of prey distribution in habitat use by
arctic foxes (Alopex lagopus). Can J Zool 80:418–429. https://doi.org/10.1139/z02-­023
Jones HP, Tershy BR, Zavaleta ES et al (2008) Severity of the effects of invasive rats on seabirds: a
global review. Conserv Biol 22:16–26. https://doi.org/10.1111/j.1523-­1739.2007.00859.x
Kasper CB, Fontoura-Rodrigues ML, Cavalcanti GN et al (2004) Recent advances in the knowl-
edge of Molina’s hog-nosed skunk Conepatus chinga and striped hog-nosed skunk C. semis-
triatus in South America. Small Carniv Conserv 41:25–28
Kasper CB, Soares JBG, Freitas TRO (2012) Differential patterns of home-range, net displacement
and resting sites use of Conepatus chinga in southern Brazil. Mamm Biol 77:358–362. https://
doi.org/10.1016/j.mambio.2012.03.006
Kasper CB, Peters FB, Christoff AU et al (2016) Trophic relationships of sympatric small carni-
vores in fragmented landscapes of southern Brazil: niche overlap and potential for competition.
Mammalia 80:143–152. https://doi.org/10.1515/mammalia-­2014-­0126
430 E. M. Luengos Vidal et al.

Kasper CB, Fontoura-Rodrigues ML, Cavalcanti GN, Freitas TRO, Rodrigues, FHG, Oliveira TG,
Eizirik E (2009). Recent advances in the knowledge of Molina’s hog-nosed skunk Conepatus
chinga and striped hog-nosed skunk C. semistriatus in South America. Small Carnivore
Conservation 41:25–28
Kelt D, Pardiñas U, Schiaffini M et  al (2016) Lyncodon patagonicus. The IUCN Red List of
Threatened Species: e.T41647A45212747. https://doi.org/10.2305/IUCN.UK.2016-­1.RLTS.
T41647A45212747.en. Accessed on 15 Feb 2020
Kennedy AJ (1980) Site variation in summer foods of arctic fox, Prince of Wales Island. Northwest
Territories Arctic:366–368
Klein DR (1995) The introduction, increase, and demise of wolves on Coronation Island, Alaska.
Ecol Conserv wolves a Chang world 275:280
Kurle CM, Croll DA, Tershy BR (2008) Introduced rats indirectly change marine rocky intertidal
communities from algae-to invertebrate-dominated. Proc Natl Acad Sci 05:3800–3804. https://
doi.org/10.1073/pnas.0800570105
Lafferty DJR, Belant JL, White KS et al (2014) Linking wolf diet to changes in marine and
terrestrial prey abundance. Arctic:143–148
Leeney RH, Amies R, Broderick AC et al (2008) Spatio-temporal analysis of cetacean strandings
and bycatch in a UK fisheries hotspot. Biodivers Conserv 17:2323
Legagneux P, Gauthier G, Berteaux D et al (2012) Disentangling trophic relationships in a high
Arctic tundra ecosystem through food web modeling. Ecology 93:1707–1716. https://doi.
org/10.1890/11-­1973.1
Lei J, Booth DT, Dwyer RG (2017) Spatial ecology of yellow-spotted goannas adjacent to a sea
turtle nesting beach. Aust J Zool 65:77–86. https://doi.org/10.1071/ZO17006
Lewis TM, Lafferty DJR (2014) Brown bears and wolves scavenge humpback whale carcass in
Alaska. Ursus 25:8–13. https://doi.org/10.2192/URSUS-­D-­14-­00004.1
Lillywhite HB, Sheehy CM III, Zaidan III (2008) Pitviper scavenging at the intertidal zone:
an evolutionary scenario for invasion of the sea. Bioscience 58:947–955. https://doi.
org/10.1641/B581008
Logan KA, Sweanor L (2001) Desert puma: evolutionary ecology and conservation of an enduring
carnivore. Island press, Washington
Long JL (2003) Introduced mammals of the world: their history, distribution and influence. CABI
Publishing, xxi, Wallingford
Loss SR, Will T, Marra PP (2013) The impact of free-ranging domestic cats on wildlife of the
United States. Nat Commun 4(1):1–8. https://doi.org/10.1038/ncomms2380
Lucherini M (2016) Lycalopex gymnocercus. The IUCN Red List of Threatened Species:
e.T6928A85371194. https://doi.org/10.2305/IUCN.UK.2016-­1.RLTS.T6928A85371194.en.
Accessed on 19 Feb 2020
Lucherini M, Luengos Vidal EM (2008) Lycalopex gymnocercus (Carnivora: Canidae). Mamm
Species 2008:1–9. https://doi.org/10.1644/820.1
Lucherini M, Eizirik E, de Oliveira T et  al (2016) Leopardus colocolo. The IUCN Red List of
Threatened Species: e.T15309A97204446. https://doi.org/10.2305/IUCN.UK.2016-­1.RLTS.
T15309A97204446.en. Accessed on 15 Feb 2020
Lucherini M, Cuyckens GAE, Morales MM et al (2019) Leopardus colocolo. In: SAyDS–SAREM
(eds.) Categorización 2019 de los mamíferos de Argentina según su riesgo de extinción. Lista
Roja de los mamíferos de Argentina. Versión digital: http://cma.sarem.org.ar. Accessed 13
Feb 2020
Luengos Vidal EM (2009) Ecología espacial de Pseudalopex gymnocercus en el pastizal pam-
peano. PhD Thesis. Universidad Nacional del Sur, Bahia Blanca
Luengos Vidal EM, Manfredi C, Castillo D et al (2005) Variaciones en la composición del gremio
de carnívoros en la región pampeana. In: Vaquero M, Cernadas de Bulnes M (eds) Producción,
recursos y medioambiente en el SO Bonaerense. Actas III Jornadas Interdisciplinarias del
Sudoeste Bonaerense. EdiUNS, Bahía Blanca, pp 97–106
Luengos Vidal EM, Lucherini M, Casanave E et  al (2009) Morphometrics of pampas foxes
(Pseudalopex gymnocercus) in the argentine pampas. Mammalia 73(1):63–67. https://doi.
org/10.1515/MAMM.2009.013
15  Use of Coastal Area Habitats by Land Mammals 431

Luengos Vidal EM, Sillero-Zubiri C, Marino J (2012) Spatial organization of the pampas fox
in a grassland relict of central argentina: a flexible system. J Zool 287:133–141. https://doi.
org/10.1111/j.1469-­7998.2011.00896.x
Luengos Vidal EM, Castillo DF, Caruso NC et al (2016) Field capture, chemical immobilization,
and morphometrics of a little-studied south American carnivore, the lesser grison. Wildl Soc
Bull 40:400–405. https://doi.org/10.1002/wsb.654
Luengos Vidal EM, Guerisoli M d l M, Caruso N et al (2017) Updating the distribution and popu-
lation status of jaguarundi Puma yagouaroundi (É. Geoffroy saint-Hilaire, 1803) Mammalia,
Carnivora, Felidae, in the southernmost part of its distribution range. Check List: J Biodivers
13(4):75–79
Luengos Vidal EM, Farías A, Valenzuela AEJ et al (2019) Lycalopex gymnocercus. En: SAyDS–
SAREM (eds) Categorización 2019 de los mamíferos de Argentina según su riesgo de extinción.
Lista Roja de los mamíferos de Argentina. http://cma.sarem.org.ar
Manfredi C, Lucherini M, Canepuccia AD et  al (2004) Geographical variation in the diet of
Geoffroy’s cat (Oncifelis geoffroyi) in pampas grassland of Argentina. J Mammal 85:1111–1115.
https://doi.org/10.1644/BWG-­133.1
Manfredi C, Soler L, Lucherini M et  al (2006) Home range and habitat use by Geoffroy’s cat
(Oncifelis geoffroyi) in a wet grassland in Argentina. J Zool 268:381–387. https://doi.
org/10.1111/j.1469-­7998.2005.00033.x
Manfredi C, Lucherini M, Soler L (2011) Activity and movement patterns of Geoffroy’s
cat in the grasslands of Argentina. Mamm Biol 76:313–319. https://doi.org/10.1016/j.
mambio.2011.01.009
Martin J, Thibault J, Bretagnolle V (2000) Black rats, island characteristics, and colonial nest-
ing birds in the Mediterranean: consequences of an ancient introduction. Conserv Biol
14:1452–1466. https://doi.org/10.1046/j.1523-­1739.2000.99190.x
Martínez Zanon JI, Travaini A, Zapata S et al (2012) The ecological role of native and introduced
species in the diet of the puma Puma concolor in southern Patagonia. Oryx 46:106–111. https://
doi.org/10.1017/S0030605310001821
Martínez ML, Intralawan A, Vázquez G et al (2007) The coasts of our world: ecological, economic
and social importance. Ecol Econ 63:254–272. https://doi.org/10.1016/j.ecolecon.2006.10.022
May R, Van Dijk J, Wabakken P et  al (2008) Habitat differentiation within the large-carnivore
community of Norway’s multiple-use landscapes. J Appl Ecol 45(5):1382–1391. https://doi.
org/10.1111/1365-­2664.12581
McDonald RI, Kareiva P, Forman RT et al (2008) The implications of current and future urbaniza-
tion for global protected areas and biodiversity conservation. Biol Conserv 141:1695–1703.
https://doi.org/10.1016/j.biocon.2008.04.025
McKinney ML (2002) Urbanization, biodiversity, and conservation. Bio Sci 52:883–890. https://
doi.org/10.1641/0006-­3568(2002)052[0883:UBAC]2.0.CO;2
Meager JJ, Schlacher TA, Nielsen T (2012) Humans alter habitat selection of birds on ocean-
exposed sandy beaches. Divers Distrib 18:294–306. https://doi.org/10.1111/j.1472-­4642.2011.
00873.x
Miñarro F (2004) Identificación de áreas valiosas de pastizal en las pampas y campos de Argentina,
Uruguay y sur de Brasil. Fundación Vida Silvestre Argentina, Buenos Aires
Moleón M, Sànchez-Zapata JA, Selva N et  al (2014) Inter-specific interactions linking preda-
tion and scavenging in terrestrial vertebrate assemblages. Biol Rev 89:1042–1054. https://doi.
org/10.1111/brv.12097
Mones A, Ojasti J (1986) Hydrochoerus hydrochaeris. Mamm Species 1986:1–7. https://doi.
org/10.2307/3503784
Monserrat AL, Celsi CE (2009) Análisis regional de la costa pampeana austral en el marco del
sistema de áreas protegidas y caracterización de un área clave como reserva, en el partido de
coronel dorrego. Bioscriba 2:1–23
Monserrat AL, Codignotto J (2013) Geodiversidad pampeana: Geomorfología y conservación de
los paisajes de dunas costeras. Comunicacoes Geologicas 100:21–32
Montero R, Autino A (2018) Sistemática y filogenia de los Vertebrados. Con énfasis en la fauna
argentina. Editorial Independiente. San Miguel de Tucuman.
432 E. M. Luengos Vidal et al.

Monserrat AL, Celsi CE, Fontana SL (2012) Coastal dune vegetation of the southern pampas
(Buenos Aires, Argentina) and its value for conservation. J Coast Res 279:23–35. https://doi.
org/10.2112/JCOASTRES-­D-­10-­00061.1
Moore PG (2002) Mammals in intertidal and maritime ecosystems: interactions, impacts and
implications. Oceanogr Mar Biol Annu Rev 40:491–608
Mora MS (2008) Metapopulation biology of the sand dune tuco–tuco (Ctenomys australis): effects
of habitat spatial structure on the ecology and population genetics. PhD Thesis. Universidad
Nacional de Mar del Plata. Argentina
Mora MS, Mapelli FJ (2010) Conservación en médanos: Fragmentación del hábitat y dinámica
poblacional del tuco–tuco de las dunas. In: Isla F, Lasta C (eds) Manejo barreras medanosas
para la Prov Buenos Aires, Editor la Univ Nac Mar del Plata, Mar del Plata, p 161
Moss B (2017) Marine reptiles, birds and mammals and nutrient transfers among the seas and
the land: an appraisal of current knowledge. J Exp Mar Bio Ecol 492:63–80. https://doi.
org/10.1016/j.jembe.2017.01.018
Navarrete SA, Castilla JC (1993) Predation by Norway rats in the intertidal zone of Central Chile.
Mar Ecol Ser 92:187
Nielsen C, Thompson D, Kelly M et al (2015) Puma concolor (errata version published in 2016).
The IUCN Red List of Threatened Species: e.T18868A97216466. https://doi.org/10.2305/
IUCN.UK.2015-­4.RLTS.T18868A50663436.en. Accessed on 05 Feb 2020
Nowell K, Jackson P (1995) Wild cats: status survey and conservation action plan. IUCN, Gland
Oliver W, Leus K (2008) Sus scrofa. The IUCN Red List of Threatened Species:
e.T41775A10559847. https://doi.org/10.2305/IUCN.UK.2008.RLTS.T41775A10559847.en.
Accessed on 12 Feb 2020
Pereira JA, Fracassi NG, Rago V et al (2010) Causes of mortality in a Geoffroy’s cat population
a long-term survey using diverse recording methods. Eur J Wildl Res 56:939–942. https://doi.
org/10.1007/s10344-­010-­0423-­8
Pereira J, Lucherini M, Trigo T (2015) Leopardus geoffroyi. The IUCN Red List of Threatened
Species: e.T15310A50657011. https://doi.org/10.2305/IUCN.UK.2015-­2.RLTS.T15310A
50657011.en. Accessed on 05 Feb 2020
Pereira JA, Lucherini M, Cuyckens GAE et al (2019) Leopardus geoffroyi. Categorización 2019
de los mamíferos de Argentina según su riesgo de extinción. Lista Roja de los mamíferos de
Argentina. http://cma.sarem.org.ar
Polis GA, Hurd SD (1995) Extraordinarily high spider densities on islands: flow of energy
from the marine to terrestrial food webs and the absence of predation. Proc Natl Acad Sci
92:4382–4386. https://doi.org/10.1073/pnas.92.10.4382
Polis GA, Hurd SD (1996a) Allochthonous input across habitats, subsidized consumers, and
apparent trophic cascades: examples from the ocean-land interface. In: Polis GA, Winemiller
KO (eds) Food webs. Springer, Boston, p 275
Polis GA, Hurd SD (1996b) Linking marine and terrestrial food webs: allochthonous input from
the ocean supports high secondary productivity on small islands and coastal land communities.
Am Nat 147:396–423
Polis GA, Strong DR (1996) Food web complexity and community dynamics. Am Nat 147:813–846
Poljak S, Confalonieri V, Fasanella M et al (2010) Phylogeography of the armadillo Chaetophractus
villosus (Dasypodidae Xenarthra): post-glacial range expansion from pampas to Patagonia
(Argentina). Mol Phylogenet Evol 55:38–46. https://doi.org/10.1016/j.ympev.2009.12.021
Prestrud P (1992) Food habits and observations of the hunting behaviour of Arctic foxes, Alopex
lagopus, in Svalbard. Can Field-Naturalist Ottawa 106:225–236
Pretelli MG, Isacch JP, Cardoni DA (2013) Year-round abundance, richness and nesting of
the bird assemblage of tall grasslands in the south-east pampas region, Argentina. Ardeola
60(2):327–343. https://doi.org/10.13157/arla.60.2.2013.327
Punta G, Yorio P, Saravia J, Borboroglu PG (2002) Breeding habitat requirements of the Imperial
cormorant and rock shag in Central Patagonia, Argentina. Waterbirds 26:176–183. https://doi.
org/10.1675/1524-­4695(2003)026[0176:BHROTI]2.0.CO;2
15  Use of Coastal Area Habitats by Land Mammals 433

R Development Core Team (2013) R: A Language and Environment for Statistical Computing.
Vienna, Austria: R Foundation for Statistical Computing. www.R-­project.org
Rathbun MJ (1918) The grapsoid crabs of America. US Nat Mus Bull 97:461
Redford KH, Eisenberg JF (1992) Mammals of the Neotropics. The southern cone. Chile,
Argentina, Uruguay, Paraguay. University of Chicago Press, Chicago
Reid F (2016) Hydrochoerus hydrochaeris. The IUCN Red List of Threatened Species:
e.T10300A22190005.https://doi.org/10.2305/IUCN.UK.2016-­2.RLTS.T10300A22190005.en.
Accessed on 12 Feb 2020
Rose MD, Polis GA (1998) The distribution and abundance of coyotes: the effects of allochtho-
nous food subsidies from the sea. Ecology 79:998–1007. https://doi.org/10.1890/0012-­965
8(1998)079[0998:TDAAOC]2.0.CO;2
Ruiz-Delgado MC, Vieira JV, Reyes-Martínez MJ et al (2016) Colonisation patterns of supralit-
toral arthropods in naturally stranded wrack debris on Atlantic sandy beaches of Brazil and
Spain. Mar Freshw Res 67:1634–1643. https://doi.org/10.1071/MF14407
Sassola N (2006) Caracterización del comportamiento del zorro pampeano (Pseudalopex gym-
nocercus) en estado silvestre en el período de crías postemergentes, en la región de la pampa
Argentina. Licenciatura Thesis Univesidad Nacional del Sur, Bahía Blanca
Schiaffini MI, Gabrielli M, Prevosti FJ et al (2013a) Taxonomic status of southern south American
Conepatus (Carnivora: Mephitidae). Zool J Linnean Soc 167:327–344. https://doi.org/10.1111/
zoj.12006
Schiaffini MI, Martin GM, Giménez AL (2013b) Distribution of Lyncodon patagonicus (Carnivora,
Mustelidae): changes from the last glacial maximum to the present. J Mammal 94:339–350.
https://doi.org/10.1644/12-­MAMM-­A-­155.1
Schiaffini MI, Ercoli MD, Díaz, G (2019) Lyncodon patagonicus. In: SAyDS–SAREM (eds)
Categorización 2019 de los mamíferos de Argentina según su riesgo de extinción. Lista Roja
de los mamíferos de Argentina. Versión digital: http://cma.sarem.org.ar. Accessed 13 Feb 2020
Schlacher TA, Weston MA, Lynn D et al (2015) Conservation gone to the dogs: when canids rule
the beach in small coastal reserves. Biodivers Conserv 24(3):493–509. https://doi.org/10.1007/
s10531-­014-­0830-­3
Schlacher TA, Hutton BM, Gilby BL et al (2017) Algal subsidies enhance invertebrate prey for
threatened shorebirds: a novel conservation tool on ocean beaches? Estuar Coast Shelf Sci
191:28–38. https://doi.org/10.1016/j.ecss.2017.04.004
Schmidt JA, McCleery R, Seavey JR et  al (2012) Impacts of a half century of sea-level rise
and development on an endangered mammal. Glob Chang Biol 18:3536–3542. https://doi.
org/10.1111/gcb.12024
Smith HM, Banks PB (2014) Disease and competition, not just predation, as drivers of impacts of
the black rat (Rattus rattus) on island mammals. Glob Ecol Biogeogr 23:1485–1488. https://
doi.org/10.1111/geb.12220
Soler L, Lucherini M, Manfredi C et al (2009) Characteristics of defecation sites of the Geoffroy’s
cat Leopardus geoffroyi. Mastozoología Neotropical 16:485–489. Sociedad Argentina para el
Estudio de los Mamíferos
Spiller DA, Piovia-Scott J, Wright AN et  al (2010) Marine subsidies have multiple effects on
coastal food webs. Ecology 91:1424–1434. https://doi.org/10.1890/09-­0715.1
Sunquist M, Sunquist F (2002) Wild cats of the world. University of Chicago Press, Chicago
Suraci JP, Clinchy M, Zanette LY et al (2014) Mammalian mesopredators on islands directly impact
both terrestrial and marine communities. Oecologia 176:1087–1100. https://doi.org/10.1007/
s00442-­014-­3085-­5
Tarroux A, Bety J, Gauthier G et  al (2012) The marine side of a terrestrial carnivore: intra-­
population variation in use of allochthonous resources by arctic foxes. PLoS One 7(8)
Tellaeche CG, Reppucci JI, Luengos Vidal EM et  al (2014) New data on the distribution and
natural history of the lesser grison (Galictis cuja), hog-nosed skunk (Conepatus chinga), and
culpeo (Lycalopex culpaeus) in northwestern Argentina. Mammalia 78(2):261–266. https://doi.
org/10.1515/mammalia-­2013-­0005
434 E. M. Luengos Vidal et al.

Tirelli FP, Trigo TC, Trinca CS et al (2018) Spatial organization and social dynamics of Geoffroy’s
cat in the Brazilian pampas. J Mammal 99:859–873. https://doi.org/10.1093/jmammal/gyy064
Valenzuela AEJ, Rey AR, Fasola L et al (2013) Trophic ecology of a top predator colonizing the
southern extreme of South America: feeding habits of invasive American mink (Neovison vison)
in Tierra del Fuego. Mamm Biol 78:104–110. https://doi.org/10.1016/j.mambio.2012.11.007
Vega RM, Prous CG, Krivokapich S et  al (2018) Toxocariasis in Carnivora from Argentinean
Patagonia: species molecular identification, hosts, and geographical distribution. Int J Parasitol
Parasites Wildl 7:106–110. https://doi.org/10.1016/j.ijppaw.2018.02.004
Watts DE, Butler LG, Dale BW et al (2010) The Ilnik wolf Canis lupus pack: use of marine mam-
mals and offshore sea ice. Wildlife Biol 16:144–149
Weston MA, Stankowich T (2014) Dogs as agents of disturbance. In: Gompper ME (ed) Free-­
ranging dogs and wildlife conservation. Oxford University Press, Oxford, p  94. https://doi.
org/10.2981/09-­040
Wilson EE, Wolkovich EM (2011) Scavenging: how carnivores and carrion structure communities.
Trends Ecol Evol 26:129–135. https://doi.org/10.1016/j.tree.2010.12.011
Yensen E, Tarifa T (2003) Galictis cuja. Mamm Species 2003:1–8. https://doi.org/10.1644/728
Zeller KA, Nijhawan S, Salom-Pérez R et al (2011) Integrating occupancy modeling and inter-
view data for corridor identification: a case study for jaguars in Nicaragua. Biol Conserv
144(2):892–901. https://doi.org/10.1016/j.biocon.2010.12.003
Zenuto R, Busch C (1998) Population biology of the subterranean rodent Ctenomys australis
(tuco-tuco) in a coastal dunefield in Argentina. Zeitschrift fur Saugetierkd 63:357–367
Chapter 16
Coastal Wetlands of the Bahía Blanca
Estuary: Landscape Structure and Plant
Associations

Paula Daniela Pratolongo, Flavia Funk, María Julia Piovan, Carla Celleri,


and Vanesa L. Negrín

16.1  Introduction

Coastal wetlands include a wide variety of environments located within coastal


watersheds. As coastal watersheds can extend many kilometers inland from the
coast, the associated wetlands can be tidal or non-tidal and freshwater or saltwater.
These types of ecosystems support several natural functions providing essential ser-
vices to human societies. For instance, they can protect upland areas from flooding
due to sea level rise and storms. Vegetation can also prevent coastline erosion due to
their ability to dissipate the energy of waves and ocean currents, which would oth-
erwise damage human infrastructure. Coastal wetlands provide habitat to waterfowl
and shorebirds, especially many globally threatened and endangered species. Many
commercial fish and shellfish species also rely on coastal wetlands. From a biogeo-
chemical point of view, coastal wetlands filter pollutants and excess of nutrients out

P. D. Pratolongo () · F. Funk


Centro de Recursos Naturales Renovables de la Zona Semiárida, CERZOS (Universidad
Nacional del Sur-CONICET), Bahía Blanca, Argentina
Departamento de Biología, Bioquímica y Farmacia, Universidad Nacional del Sur,
Bahía Blanca, Argentina
M. J. Piovan
Instituto Argentino de Oceanografía, IADO (Universidad Nacional del Sur-CONICET),
Bahía Blanca, Argentina
C. Celleri
Centro de Recursos Naturales Renovables de la Zona Semiárida, CERZOS (Universidad
Nacional del Sur-CONICET), Bahía Blanca, Argentina
V. L. Negrin
Departamento de Biología, Bioquímica y Farmacia, Universidad Nacional del Sur,
Bahía Blanca, Argentina
Instituto Argentino de Oceanografía, IADO (Universidad Nacional del Sur-CONICET),
Bahía Blanca, Argentina

© Springer Nature Switzerland AG 2021 435


S. M. Fiori, P. D. Pratolongo (eds.), The Bahía Blanca Estuary,
https://doi.org/10.1007/978-3-030-66486-2_16
436 P. D. Pratolongo et al.

of the water before it reaches the ocean. They also sequester and store large amounts
of carbon due to their rapid growth, high sedimentation rates, and slow decomposi-
tion (Barbier 2019).
In temperate regions, tidal salt marshes are the dominant vegetation type in the
intertidal zone of protected coasts. Instead, extensive mangroves characterize the
tidal fringe in tropical latitudes. Salt marsh plants are mostly herbaceous and salt-­
tolerant macrophytes. They are well adapted to face regular inundation by marine
water, especially those species occupying the lowest portion of a marsh, flooded
twice a day by tides. At higher elevations, as tidal influence becomes less frequent,
the terrestrial hydrology dominates, and a different type of vegetation may appear
in response to varying salt stress (Cronk and Fennessy 2001). In arid and semiarid
coastal locations, high evaporation rates may produce muddy salt flats instead of a
typical high marsh. At the other end of the scale, large volumes of freshwater dis-
charging into the upper salt marsh may support less salt-tolerant or even freshwater
wetland plants (Pratolongo et al. 2019).
The coastal wetland continuum encompasses a broad spectrum of ecosystems,
from salt marshes and tidal flats to non-tidal wetlands landward. The unifying attri-
bute along all these contrasting ecosystems is that sea level acts as a significant
driver of their hydrology. While sea level distinctly sets the limits of the tidal fringe,
its influence is less obvious landward. According to Hageman (1969), the perima-
rine zone is the area where non-tidal wetlands persist under control of the relative
sea level. In humid climates, freshwater seepage and high groundwater levels keep
the saturated conditions necessary for the persistence and growth of perimarine
freshwater swamps, non-tidal marshes, and fens (Waller et al. 1999; Gardner et al.
2000). On the contrary, a combination of high evaporation and low freshwater
inputs characterizes the perimarine zone of arid climates. In these areas, soils
develop extremely high salinities, and salt flats commonly replace freshwater wet-
lands (Pratolongo et al. 2009).

16.1.1  Habitat Zonation and Response to Sea Level Changes

Within the intertidal zone, salt marsh plants commonly display distinct patterns of
zonation, driven by the individual species tolerance to physical stress and biological
interactions acting across abiotic gradients (Pennings and Bertness 2001). The ele-
vation gradient within the marsh establishes flood frequency, depth, and duration
(the hydroperiod). Thus, abiotic factors linked to hydroperiod, like soil moisture
content, redox state, nutrient limitation, and salt concentration, commonly correlate
with elevation (Silvestri et  al. 2005; Moffett et  al. 2012). Significant biological
interactions, like competition and facilitation, also perform in close association with
abiotic stress (Bertness 1991; Pennings et al. 2005). The interplay between stress
gradients and biological responses results in the typical shore-parallel zonation of
plants, which may be more complex and spatially variable depending on the micro-
topography of the marsh surface.
16  Coastal Wetlands of the Bahía Blanca Estuary: Landscape Structure and Plant… 437

At the seaward margin of the elevation gradient, saltwater regularly inundates the
low marsh. A few resistant plant genera such as Sarcocornia, Suaeda, Aster, and
Spartina can tolerate these stressful abiotic conditions imposed by tidal influence
(Doody 1992). In low marsh areas, frequent tidal flooding prevents the accumula-
tion of salts, and soil salinity is comparatively low. Instead, substrate stability, oxy-
genation, and sulfide toxicity may control plant establishment and survival at the
seaward margin (Adam 1990; Mendelssohn and Morris 2000). Landward, soil salin-
ity levels vary across marsh elevations, and the shape of the salt gradient depends on
climate. In warm, humid regions, freshwater input from rain and upland sources
moderates soil salinity in the high marsh. Thus, salts concentrate by evaporation at
intermediate marsh elevations, where tidal flushing is inefficient. Mid-marshes
characterized by higher salinities support distinctive salt-tolerant vegetation, and
salt accumulations may also lead to the development of bare areas known as salt
pans (Davy and Costa 1992). In colder climates, on the other hand, salinity decreases
with increasing elevation (Pennings and Bertness 1999), and the opposite pattern
prevails in arid climates, that is, increasing salt concentration along with elevation
(Zedler 1982; Callaway et al. 1990).
Besides the salt marsh zonation of vascular plants, another biota also occupies
distinct zones across the elevation gradient. For sessile or low mobility benthic spe-
cies, their position within the tidal range reflects the relative ecological tolerance to
a combination of abiotic and biotic stressors (Zapperi et al. 2016). Flooding dura-
tion, which depends on elevation, also drives the distribution of microorganisms on
salt marsh sediments (Gehrels 2000), allowing for paleoecological studies and
reconstructions of past relative sea levels (e.g., Horton and Edwards 2006). The
elevation gradient across the marsh also creates different stress gradients to mobile
aquatic and terrestrial animal species. Different species must either tolerate or avoid
submergence and emergence, constraining their distributions (Pennings and
Bertness 2001).
The transition from the coastal wetland continuum to the upland is also depen-
dent on the regional patterns of rainfall or freshwater discharge. In perimarine envi-
ronments with enough freshwater inputs, waterlogging is caused by freshwater
seepage, rainfall, and poor drainage due to the low elevation respect to the mean sea
level. Here, there is a progressive transition from the upper salt marsh into freshwa-
ter wetland communities. Depending on the nutrient status, an array of floristically
different wetland communities may develop, including freshwater marshes and for-
ested swamps. In arid climates with low freshwater inputs, evaporation increases
with elevation, and soils develop extremely high salinities in the perimarine zones,
precluding vegetation development. In arid tropical and subtropical regions, where
evaporation by far exceeds precipitation, extensive, barren, salt-encrusted flats char-
acterize coastal lowlands (Barth and Böer 2002). Barren salt flats, covered by evap-
orite accumulations, also appear in arid coastal settings within temperate climates
(e.g., the Bahía Blanca Estuary and the Shark Bay, in Western Australia). Coastal
salt flats in the perimarine zone of arid environments are above the limits of the tidal
inundation, but sea level keeps shallow groundwater near the surface. From a
438 P. D. Pratolongo et al.

hydrogeomorphic point of view, these salt flats are comparable to the perimarine
freshwater fens and swamps typical of humid regions (Pratolongo et al. 2016).
A central process in coastal wetland development is ecosystem response to
changes in the relative sea level. A changing sea level produces a modification in the
hydroperiod and, thus, in the ecological state of wetlands. In response to the differ-
ent hydrologic conditions, the different plant zones within the coastal wetland con-
tinuum are expected to migrate. In coastal settings along the mid-Atlantic and
Pacific coasts of North America, for instance, the relative sea level was continually
rising from the last glacial maximum to the present due to the combination of iso-
static and eustatic effects. In these regions, Holocene estuaries drowned, and new
wetlands formed landward. The process of wetlands migration upslope, in response
to a rising relative sea level during the Holocene, has been widely studied along the
eastern coast of North America. Coastal barrier ecosystems along the seaward mar-
gin of the Delmarva Peninsula, on the Atlantic coast of North America, underwent
a sustained sea level rise during the Holocene (Oertel et al. 1989). Changes in sea
level have been driving a succession of state changes in wetlands along the main-
land edge of the Peninsula in which high marshes replace former upland forests.
Following this sequence, the low marsh replaces a high marsh, and bare mudflats
replace the former low marshes (Brinson et al. 1995; Christian et al. 2002).
An opposite pattern was described in high latitude coastal regions (e.g.,
Fennoscandia, Finland, Labrador), where the relative sea level was continually fall-
ing from the last glacial maximum to the present. In these regions, the different
plant zones within the coastal wetland continuum have been typically migrating
seaward. Along with the falling sea level, new land emerged, and pioneer low marsh
plants colonized mudflats. The coastal area of western Finland rose during the
Holocene as a result of rapid land uplift that followed the last glacial maximum. In
the northern Gulf of Bothnia, the relative sea level fell and the shoreline displaced
seaward. Shoreline retreat still occurs on this relatively flat coast, at a rate exceeding
8 mm a year, along with downward migration of plant zones (Vartiainen 1988; Ecke
and Rydin 2000) and seaward expansion of pioneer plant communities (Zobel and
Kont 1992). A more complex dynamics characterizes wetland environments in the
temperate Atlantic coast of South America, where the relative sea level reached a
transgressive maximum during the Holocene. In these systems, the late Holocene
marine regression resulted in significant low-lying coastal landforms inherited from
the former estuarine dynamics, which are presently occupied by extensive perima-
rine wetlands.
16  Coastal Wetlands of the Bahía Blanca Estuary: Landscape Structure and Plant… 439

16.2  The Bahía Blanca Estuary

16.2.1  Coastal Landscape Evolution

The Holocene marine transgression has deeply modeled the coastal landscape
around the Bahía Blanca Estuary. On the northern coasts of Argentina, the relative
sea level reached around 6  m above present, about 6000  years ago (Isla 1989;
Violante and Parker 2000; Cavallotto et  al. 2004). The falling sea level after the
highstand (i.e., the regressive phase) resulted in large low-lying areas of former
estuarine environments. Typical regressive forms, like extensive plains composed of
beach ridge and lagoonal deposits, characterize many coastal settings in Argentina,
from Río de la Plata Coastal Plain to San Sebastián Bay, in Tierra del Fuego.
Raised Holocene deposits in the Bahía Blanca Estuary (Fig. 16.1) formed after
the mid-Holocene highstand (ca. 6000 BP) during the regressive phase. According
to González and Weiler (1983), these deposits correspond to relatively short periods
of high energy identified as Transgressive Stages (e.g., pulses of rising sea level
within the general falling trend). On the northern shore of the Principal Channel,
there is a succession of Holocene beach ridges and tidal flat deposits, corresponding
to high- and low-energy periods (González 1989). The oldest and highest deposits,
located in the inner section of the Principal Channel, form a spit composed of sev-
eral sand-shell ridges up to 10 m above the present sea level (González et al. 1983;
Aliotta and Farinati 1990). Based on 14C dates, González (1989) described at least
five beach ridges representing significant episodes of high wave energy within the
regressive phase, after the highstand. These episodes were named Transgressive
Stages I to V and dated between 5990 ± 115 and 3560 ± 100 years BP. The inland,
oldest beach ridge in this sequence represents the maximum transgressive episode.
Holocene deposits also appear further East, close to Punta Alta City. These
deposits form a relatively continuous sand ridge parallel to the coast located about
6 to 7 m above the mean sea level (14C age of 4615 ± 110 years BP). This ridge
indicates the end of the maximum transgressive event before a regressive pulse
(Aliotta and Farinati 1990). In this area, there is a second (seaward) ridge, at a lower
elevation, assigned to a younger transgressive pulse (Aliotta and Farinati 1990).
Extensive coastal flats prograded seaward from the second ridge during the late
Holocene, under a falling sea level. Their radiocarbon ages range between 3300 and
3900 14C years BP, determined from fossils in life position (Farinati et al. 1992).
The raised Holocene deposits in the northern margin of the Principal Channel
define the base of a scarp that divides two different domains. La Vidriera Salt Flat
is a narrow, depressed landform that extends inland, aligned with the Principal
Channel (Fig. 16.1). The axis along Principal Channel-La Vidriera Salt Flat sepa-
rates the elevated Positive of Ventania to the North and the Colorado River Basin to
the South (González-Uriarte 1984). South of the alignment Principal Channel-La
Vidriera Salt Flat, the northwestern margin of the Bahía Blanca Estuary belongs to
440 P. D. Pratolongo et al.

Fig. 16.1  Major landscape units in the coastal zone of Bahía Blanca. (Modified from Pratolongo
et al. 2016, 2017)

the Colorado River Basin and has a gentler slope. In Verde Erin, a cattle ranch
located 20 km southwest from Bahía Blanca (Fig. 16.1), Farinati (1983) described a
shelly ridge at the base of a paleo-cliff, which would correspond to storm deposits
indicating the inland limits of the marine transgression, at a 14C age of
5406 ± 227 years BP. In the coastal plain extending from the ridge toward Principal
Channel, two successive terrace levels formed at different ages during the regressive
phase (González-Uriarte 1984). The Old Marine Plain (OMP), at an average eleva-
tion of 5 m above the present sea level, is a nearly continuous flat surface covered
by different types of halophytic shrub communities. The Recent Marine Plain
(RMP) is 2–3 m above the mean sea level. The unit is a mosaic of topographic highs
and elongated depressions, corresponding to former tidal channels. In the RMP, the
gentle slope creates a gradual transition to the Present Marine Plain (PMP), com-
posed of active tidal channels, mudflats, and salt marshes, currently affected by the
estuarine dynamics.
16  Coastal Wetlands of the Bahía Blanca Estuary: Landscape Structure and Plant… 441

In a more recent paleoenvironmental reconstruction, Pratolongo et  al. (2017)


found well-preserved mollusks at the base of a core obtained in the OMP (530 cm
deep), with ages between 5660 ± 30 and 5470 ± 30 years BP. It suggests that, by the
time of the Holocene transgressive maximum, the presently OMP was a sandy bot-
tom at a subtidal elevation. Thinner laminated sediments lay on top of the sand
deposits, containing abundant shells of a brackish gastropod species. It suggests
that, after the transgressive peak, the estuarine system progressively infilled during
a low-energy period. A thick deposit of massive, grayish, organic-rich muds overlies
the laminated sediments (Pratolongo et al. 2017). This sequence would correspond
to rapid sedimentation during a high sea level stillstand or slow regression. A series
of creeks, west of the Bahía Blanca Estuary, showed a similar evolution, with infill
estuarine sequences composed of grayish muds forming flat plains in estuaries rap-
idly filled during the Holocene transgression (Isla et al. 1996; Marquez et al. 2016).
In the Bahía Blanca Estuary, González-Uriarte (1984) described a paleodrainage
that occupied the aligned depressions along La Vidriera Salt Flat-Principal Channel
(Fig. 16.1). It indicates the former presence of a large river, which brought sedi-
ments to the region within a deltaic environment. Enhanced sedimentation under
low-energy conditions may have filled the estuarine area and deposited the top mas-
sive mud layer. This model agrees with fluvial paleochannel structures described
along the Principal Channel (Spagnuolo 2005; Giagante et al. 2008, 2011) and par-
ticularly with the fluvial-deltaic environment described by Aliotta et al. (2014).
In contrast with the uninterrupted, elevated OMP, there is fragmented RMP at a
lower elevation, with old tidal channels dissecting topographic highs. The RMP
formed at a later stage during the regressive phase, after sea level dropped to about
3 m above the present mean tidal level (the approximate highest elevation of the
RMP). Then, the flat and continuous surface corresponding to the OMP emerged,
and a more recent coastal dynamics shaped the fragmented landscape pattern of the
RMP. The old tidal channels formed during this later stage, after tidal working dis-
sected the older surface of mud deposits. Conversely, there is no well-defined
boundary between the RMP and the PMP, but a gradual transition. The differentia-
tion between the RMP and PMP would be then a response to the increased tidal
influence on the RMP, because of the recent rising trends in the relative sea level
(Pratolongo et al. 2013).

16.2.2  Coastal Wetlands and Landscape Patterns

Plant associations in the study area are reliable indicators of the dominant hydrogeo-
morphic conditions imposed by landscape position (Pratolongo et al. 2013, 2016).
From a hydrogeomorphic perspective, coastal wetlands in Bahía Blanca classify
into intertidal and inland (i.e., perimarine) wetlands. Within the intertidal zone (i.e.,
the area below the elevation of mean high tides), extensive barren mudflats are the
dominant land cover type. Pure stands of Spartina alterniflora are commonly
restricted to low marshes in the middle reach of the estuary but are extremely rare
442 P. D. Pratolongo et al.

in the inner zone. Spartina densiflora marshes occasionally appear as pure stands
that form a transition zone between S. alterniflora and Sarcocornia ambigua in
places influenced by freshwater discharges.
Through the shallow inner section of the Principal Channel, vegetated marshes
constrain to elevations close to the mean high tide level, with S. ambigua as the
dominant species. Seasonally hypersaline conditions, because of the higher evapo-
ration rates, would impose physiological limitations to vegetation development,
precluding the establishment of less tolerant species (Pratolongo et  al. 2010). A
supralittoral zone develops above the elevation of the mean high tide and below the
limits of the highest tides (spring high tides and storm surges). Seawater inundates
this elevational fringe irregularly, which enhances evaporation and soil salinization.
Vegetation in the supralittoral zone is sparse, forming a mosaic of salt flats, halo-
phytic steppes, and shrubs. Plants and bare soil commonly assemble in mound-­
intermound complexes (Piovan et al. 2014).
Soils in these intertidal and supralittoral wetlands classify as Aquisalids (for-
merly Salorthids). They usually have an A horizon, enriched in organic matter com-
pared to the underlying strata. The organic-rich layer may extend up to 25–50 cm
depth in highly productive intertidal marshes. In supratidal halophytic steppes, A
horizons rarely exceed 8–10 cm depth. In barren areas, the A horizon may not be
present at all, and the soil surface exposes parental material (Kruger 1989). More
elevated landscape positions in the supralittoral zone are less affected by tidal inun-
dation, but soil properties are influenced by a highly saline groundwater table. In
situ determinations of groundwater salinity ranged from 41 to 45 (seawater salinity
is around 35). During dry years, groundwater level fluctuates from 3 to 0.3 m below
the soil surface, but it rises to the soil surface during wet years, leaving soils ponded
for several months. Soluble salt crystals form surface salt crusts when the soil is dry.
Salt crusts disappear after tidal flooding and rain events but reappear after a few
days of high evaporation (Piovan 2016).
Intertidal and supralittoral wetlands form a gently sloped continuum that extends
from PMP through the RMP. The narrow depressions that typically dissect the land-
scape pattern in the RMP correspond to former tidal channels that presently drain
rainfall excess. Within channels, soils are permanently flooded or saturated, even in
dry years. The water table fluctuates around the soil surface (from below 0.4 to
above 0.5 m), rapidly responding to rainfall events in the catchment area. Continuous
records of groundwater levels in wells showed increases of more than 50 cm within
hours after a single precipitation event (Piovan et  al. 2014; Piovan 2016).
Groundwater salinity in these channels is significantly lower (23–26) than in more
elevated environments within the unit. The lower salinity allows for the develop-
ment of highly productive S. densiflora marshes and soils with a deeper layer
enriched with organic matter.
The OMP is not affected by tidal flooding. Beyond tidal influence, an irregularly
humid zone develops close to the inland limit of the OMP. In this landscape posi-
tion, a saline shallow water table fluctuates close to the soil surface, fed by seepage
from the surrounding uplands. The permanent field indicator of a dryland saline
seepage in the area is salt accumulation in a fringe at the base of the scarp in the
16  Coastal Wetlands of the Bahía Blanca Estuary: Landscape Structure and Plant… 443

inland limits of the marine transgression. During humid periods with anomalously
high rainfall, a saturated zone develops in this area due to the water table rising to
the soil surface. Under dry conditions, especially in hot summer months, evapora-
tion intensifies the capillary rise of saline groundwater, leading to salt accumula-
tions in the soil surface (Celleri et al. 2019). Soils in the seepage zone classify as
Petrogypsids (formerly Petrogypsic Gypsiorthids) because of the presence of a sub-
surface soil horizon cemented by gypsum (Kruger 1989). Gypsum is an evaporite
mineral that frequently results from concentration and crystallization by evapora-
tion of a fluctuating water table enriched in calcium and sulfate (Gómez-Miguel
et al. 1984). Soils in this saline seepage area develop a loose cover of salt crystals
that form occasionally when the capillary fringe rises to the soil surface. In the
peripheral zone downslope from the discharge area, soils classify as Aquic
Ustifluvents and show typical characteristics indicating saturation.
Wetlands in the OMP are dry-end wetlands, that is, wetlands that occur in land-
scape positions where flooding is irregular and soil saturation to the surface occurs
occasionally (Whigham 1999). These wetlands are subjected to episodic disappear-
ance and may get dry for several years, relying on the occurrence of rainy years
during humid periods to resurge. Plant communities in these environments are very
reactive to small changes in climate and quickly respond to exceptionally humid or
dry periods. Because of their highly dynamic nature, and the climatic settings of the
area, these wetlands are sensitive indicators of climate change and variations in
large-scale circulation patterns. Out of the seepage area, soils in the OMP are well-­
drained and have lower contents of soluble salts.

16.2.3  Wetlands Indicators in Arid Environments

In arid and semiarid environments, coastal wetlands beyond the limits of the tidal
inundation, like those in the OMP of the Bahía Blanca Estuary, may have a pulsed
ephemeral nature that makes them difficult to identify (Cintron-Molero and
Schaeffer Novelli 2002). Thus, soils are useful instruments for wetland identifica-
tion and delineation (Tiner 1999). Extended flooding has a significant effect on
soil-forming processes, which produce a set of unique and noticeable soil proper-
ties. When porewater replaces the air in the soil pore spaces, the supply of oxygen
depletes, and chemical processes change (Henderson and Patrick 1982). Dissolved
oxygen is virtually absent in wetland soils that are continuously saturated with
water. These soils chemically reduce and develop an aquic moisture regime (Soil-­
Survey-­Staff 1975). Aquic soils (Vepraskas 1996) have morphological indicators of
saturation and reduced conditions (i.e., redoximorphic features). Common redoxi-
morphic indicators of wetland soil include gray colors, iron, and manganese concre-
tions near the surface (Tiner 1999).
Wetland soils in arid climates concentrate carbonates, gypsum, and even more
soluble salts on their surface. The combination of high salinities and low organic
matter content hampers typical processes expected in humid climates, such as
444 P. D. Pratolongo et al.

microbial activity and chemical reduction of iron. If redoximorphic features and


hydric soil indicators do not form, seasonally dry or intermittent wetlands would be
challenging to identify. Boettinger (1997) suggests that salts more soluble than gyp-
sum in the upper 30 cm indicate a hydric soil. Thus, the presence of a salt crust
should be enough as a field indicator of wetland soil, subject to periodic saturation
from a saline water table.

16.2.4  Plant Communities

From the physiognomic-floristic point of view, the spontaneous vegetation of the


Bahía Blanca region lies within three Phytogeographic Provinces: Monte, Pampeana,
and Espinal (Canepuccia et al. 2013; Oyarzabal et al. 2018). This ecotone is charac-
terized by the dominance of shrubs and grasses, mostly perennial. Due to climatic
constraints, this type of vegetation is typically composed by a high number of C4
grasses and xerophytic species (León et al. 1998; Paruelo et al. 1998). Plant leaves
are frequently small, succulent, reduced to thorns, or absent. Photosynthetic stems
are another common adaptation to arid conditions. Vegetation in the area has a
patchy structure, with vegetation islands that contain different species and even dif-
ferent growth forms, but with low species richness.
Several authors described plant associations in the coastal zone of Bahía Blanca
(e.g., Verettoni 1974; Kruger and Peinemann 1996; Nebbia and Zalba 2007). Based
on this previous information and extensive field work carried out in the area, Piovan
et al. (2014) defined nine major land cover classes, which were used to derive a land
cover map at the landscape level. The methodological approach involved the pro-
cessing of a 30-year-long time series of Landsat satellite images to obtain surface
reflectance data, and the implementation of numerical classification algorithms,
masking, and reclassification, to produce the final land cover map presented in
Fig. 16.2. Examples of the more representative land cover classes are presented in
Fig. 16.3.
Southwestern Atlantic salt marshes are characterized by the presence of the halo-
phytic genera Spartina and Sarcocornia (Isacch et  al. 2006). In Argentina,
Sarcocornia appears all along the coast, from Buenos Aires to Tierra del Fuego,
commonly forming marshes below the elevation of the highest astronomical tides
(González et al. 2019). Within the intertidal fringe of the Bahía Blanca Estuary, bare
mudflats are the dominant land cover type (836  km2). Salt marshes of Spartina
alterniflora (196  km2) appear as monospecific associations with variable plant
heights and densities, from sparse plants, up to 20 cm tall (González Trilla et al.
2013), to dense stands taller than 1 m in places subject to high sedimentation rates
(Pratolongo et al. 2010). Intertidal marshes of Sarcocornia ambigua are less repre-
sented (72 km2), and the dominant species commonly forms circular mounds, some-
times in association with Heterostachys ritteriana or S. densiflora at higher
elevations (Pratolongo et  al. 2013). Vegetation cover of this type of marshes is
16  Coastal Wetlands of the Bahía Blanca Estuary: Landscape Structure and Plant… 445

Fig. 16.2  Land cover map showing major plant associations in the coastal zone of Bahía Blanca.
Land cover classes were previously defined, based on field surveys, and the map was further
obtained through digital classification of satellite images from the Landsat series. (Modified from
Pratolongo et al. 2016, 2017)
446 P. D. Pratolongo et al.

Fig. 16.3  Most representative plant communities. (a) Marshes of Sarcocornia ambigua and plant
detail. (b) Marshes of Spartina alterniflora and plant detail. (c) Halophytic steppe with sparse
patches of vegetation. (d) Bare mudflats with sparse plants of S. alterniflora. (e)  Thickets
of Allenrolfea patagonica (f) Cyclolepis genistoides

highly uneven, from a few isolated patches in a matrix of bare soil to an almost
continuous carpet at some locations.
Intertidal marshes dominated by Spartina densiflora are rarely observed and
typically associate to freshwater discharge of the few permanent small rivers in the
estuary (e.g., Napostá, Sauce Chico, and Maldonado Channel). In the supralittoral
zone, however, S. densiflora marshes occupy the old tidal channels of the RMP,
covering 1.8 km2. S. densiflora is the clear dominant species in these later marshes,
16  Coastal Wetlands of the Bahía Blanca Estuary: Landscape Structure and Plant… 447

forming dense stands about 50 cm tall, and S. ambigua may be also present with
relatively high percent cover (Piovan 2016).
Tidal inundation becomes less frequent upslope from the mean high tide level,
and there is a spatial transition from S. ambigua marshes to halophytic steppes.
Vegetation in these steppes keeps a similar structure, with plant patches forming
vegetation islands. Barren areas increase their cover close to the elevation of the
highest tides, and soils develop more permanent salt crusts. At some locations, veg-
etation patches are sparse enough to allow for the distinction of salt flats as a pure
land cover type. S. ambigua is also a dominant species in halophytic steppes, but, as
tidal inundation decreases, the number of species within vegetation patches
increases. Typical accompanying species are H. ritteriana and S. densiflora, with
higher abundances than those observed in marshes downslope. H. ritteriana is a
salt-tolerant species whose presence is considered as an indicator of soils with high
concentrations of chlorides and sodium (Cantero et al. 1998). At higher elevations,
vegetation is more diverse, and species like Atriplex undulata, Limonium brasilien-
sis, and Frankenia juniperoides may also appear in vegetation islands.
Three different types of woody associations can be distinguished in the coastal
zone of Bahía Blanca. Dense thickets dominated by Allenrolfea patagonica form a
continuous fringe in the inland limits of the OMP. This association covers 45 km2
within the OMP. Vegetation height is about 1 m, and A. patagonica has an average
60% cover, sometimes in association with Cyclolepis genistoides. The plant asso-
ciation of A. patagonica and C. genistoides commonly denotes soils with high con-
centrations of sulfates and calcium (Cantero et  al. 1998). Most species in the
undergrowth appear in response to episodic rainfall, but S. ambigua and Lycium
chilense are always present, even in dry periods. Grahamia bracteata and cactus
from the genus Trichocereus are also commonly observed.
Downslope from thickets of A. patagonica, vegetation is slightly taller, and the
amount of bare soil increases. This landscape position is typically covered by bushes
dominated by C. genistoides (261  km2). In this plant association, A. undulata
appears as a common species in the underbrush, replacing A. patagonica. Other
common species in this lower stratum are S. ambigua, F. juniperoides, Cressa truxil-
lensis, and L. brasiliensis. Bushes dominated by C. genistoides also occur in inland
saline environments of central Argentina (Cantero et al. 1998; Karlin et al. 2012;
Cabido et al. 2018). In the RMP, these two plant associations (thickets of A. pata-
gonica and bushes of C. genistoides) appear as a discontinuous mosaic jointly
mapped as halophytic shrubs in Fig.  16.2. In this landscape unit, A. patagonica,
C. genistoides, A. undulata, and S. ambigua are all common species, whose relative
abundance is determined by specific site conditions.
The last woody association described for the Bahía Blanca Estuary is the brush-
wood dominated by Geoffroea decorticans (150 km2), which is indicative of well-­
drained soils and lower salinities. The dominant species, G. decorticans, is a tall
shrub, exceeding 2 m height, which forms dense thorny bushes. This plant commu-
nity is characterized by the predominance of medium to tall shrub associations with
several strata and relatively high vegetation cover, typical of the Monte
Phytogeographic Province (Gaitan et al. 2019). The most frequent shrub species are
448 P. D. Pratolongo et al.

Prosopidastrum globosum, Monttea aphylla, and Condalia microphylla. The diverse


underbrush is composed of L. chilense, A. undulata, Atriplex heterosperma,
Grindelia brachystephana, Ephedra triandra, and several grasses (Poa lanuginosa,
Trichloris crinita, Jarava spp., and Nassella spp., among others).
Suaeda divaricata is an emblematic shrub species in the area, which usually
appears within woody stands in highly saline soils. Locally called Vidriera, the
extensive salt flat La Vidriera was named after this species. Endemic species of the
Cactaceae family also occur in vegetated patches within the salt flat, like Echinopsis
leucantha, Opuntia sulphurea, Opuntia penicilligera, and Trichocereus candicans.
Other frequent species are Sesuvium portulacastrum, A. patagonica, Heterostachys
olivascens, H. ritteriana, Nitrophila australis, Suaeda argentinensis, and G. brac-
teata (Pérez Cuadra 2008).
Along the salinity and flooding gradient, it is possible to identify changes in
vegetation structure from isolated and cushion plants to steppes of greater height
and higher species diversity. Salinity and flooding act as ecological filters that
reduce plant diversity in the salt marsh (Rogel et al. 2000). Therefore, the environ-
mental gradient and its unique physicochemical characteristics set an ecotone in the
transition between intertidal marshes and upland plant communities. Much of the
variation in vegetation structure between intertidal marshes, perimarine wetlands,
and upland communities can be related to a gradient in elevation, salinity, soil water
content, and severity of hypoxia (Traut 2005).

16.2.5  Plant Adaptations to Life in Salt Marshes

Salt marshes are exposed to extreme environmental conditions as high salinity, reg-
ular tidal flooding, and the mechanical impact of waves and storms (González et al.
2019). Salt marsh environments are stressful, and plant species must survive inter-
vals of complete inundation, as well as changes in water and soil salinity. The spa-
tial distribution of vegetation within salt marshes is not random or spatially
uncorrelated, and the role of adaptations to soil salinity and tidal inundation in
determining this distribution has been widely studied (Traut 2005; Bao-Shan et al.
2011). Several adaptations allow salt marsh plants to thrive in high salinity and low
oxygen environments. For some species, morphological adaptations or anatomical
structures provide a strategy to tolerate environmental stress. Morphological adap-
tations include smaller leaves, fewer stomata, increased succulence, thick cuticle,
and deposition of wax. Anatomical changes include salt secretory trichomes and
glands, located on leaves, and well-developed water storing tissues (Wahid 2003).
Oxygen availability is essential for plant growth, and it is governed by soil type,
topographic position, and the frequency and the duration of salt marsh flooding
(Silvestri et al. 2005). Various adaptations allow salt marsh plants to survive under
low oxygen levels imposed by tidal flooding. The most common adaptation is the
formation of aerenchyma in leaves, stems, and roots. This tissue enables the vertical
gas transport within plants, allowing oxygen to reach the flooded roots (Cronk and
16  Coastal Wetlands of the Bahía Blanca Estuary: Landscape Structure and Plant… 449

Box 16.1 Most Emblematic Native Plant Species in the Bahía


Blanca Estuary
Family Amaranthaceae
At present, the Amaranthaceae and Chenopodiaceae form a monophyletic
group (Kadereit et  al. 2003) and are considered as a single family
Amaranthaceae (Stevens 2001). This family comprises about 180 genera and
around 2500 species widely distributed. Preferred habitats are warm temper-
ate and tropical grasslands, savannas, sand dunes, salt marshes, semideserts,
and deserts (Kadereit et al. 2003). Most species in this family have xeromor-
phic characteristics. In Argentina, 14 genera are represented with nearly 121
species (Múlgura and Marticorena 2008).
Sarcocornia ambigua
Subfamily Salicornioideae
Genus Sarcocornia
Species ambigua
Sigla sp. (Michx.) M.A. Alonso & M.B. Crespo
Common name Jume
Plants of the Sarcocornia genus grow in saline areas, usually near the coast
where they form salt marshes. Sarcocornia genus comprises about 28 succu-
lent perennial species (Steffen et al. 2010). Five South American species of
Sarcocornia have been identified, and four different morphological types are
found in South America and the Mediterranean countries. All Sarcocornia
species from South America were first described as Salicornia. Recent studies
indicate that Sarcocornia and Salicornia are paraphyletic groups that show
strong morphological and ecological similarity (Alonso and Crespo 2008).
Sarcocornia and Salicornia can be distinguished only by inflorescence char-
acters and life form (perennial and annual, respectively) (Steffen et al. 2015).
The taxonomy and nomenclature of South American species have recently
been updated, and they are now referred as Sarcocornia spp. (Alonso and
Crespo 2008). Sarcocornia ambigua is the species currently accepted for the
area of Bahía Blanca. However, most of the existing bibliography regarding
marshes in Bahía Blanca refers to this species as Sarcocornia perennis.
Sarcocornia ambigua is a perennial erect shrub. The stems are succulent
and articulated, leaves opposite, and flowers hidden in cavities of the inflores-
cence axis. This is a widespread species and grows in areas with highly vari-
able salinity, which makes it a potential candidate for the development of a
novel crop (Freitas and Costa 2014). The aerial parts of Salicornia and
Sarcocornia species have been introduced into the European market for
human consumption, due to their high nutritional value in terms of minerals
and antioxidant vitamins (Ventura et al. 2011; Bertin et al. 2016).
450 P. D. Pratolongo et al.

Allenrolfea patagonica
Subfamily Salicornioideae
Genus Allenrolfea
Species patagonica
Sigla sp. (Moq.) Kuntze
Common name Black jume
The genus Allenrolfea has three species: Allenrolfea occidentalis is found
in North America, while Allenrolfea patagonica and Allenrolfea vaginata are
exclusive from Argentina. A. patagonica grows in a wide range of salinities,
in soils not affected by tidal flooding, from Salta to southern Buenos Aires
(Kruger and Peinemann 1996). A. patagonica is a 1-m-tall perennial shrub,
with pyramidal leaves and flowers grouped in terminal inflorescences. This
species is used for the preparation of bleach (del Vitto et al. 1997).
Atriplex undulata
Subfamily Chenopodioideae
Genus Atriplex
Species undulata
Sigla sp. (Moq.) D. Dietr.
Common name Wavy-leaf saltbush
The genus Atriplex is represented in Argentina by 34 species. This is a
widespread genus that colonizes many arid and semiarid regions (Giusti
1997). Atriplex species can be used to increase productivity in arid or semi-
arid lands because of their salt tolerance and high productivity. Several spe-
cies have been planted as a foraging shrub in marginal agricultural lands in
many countries (Salem et al. 2010). Atriplex undulata is a perennial shrub,
native to arid and semiarid rangelands of central Argentina (Piovan et  al.
2014). A. undulata is a 1-m-high shrub with ramified erect stems. Leaves are
alternate, obovate-oblong, grayish-white on both sides, and with undulate
edges. Male and female flowers are on separate plants (dioecious). Each
round and soft fruit contains one single seed that matures in autumn.
Suaeda divaricata
Subfamily Chenopodioideae
Genus Suaeda
Species divaricata
Sigla sp. Moq.
Common name Vidriera
The genus Suaeda is globally distributed in saline or alkaline habitats. In
Argentina, there are three native species: Suaeda argentinensis, Suaeda divar-
icata, and Suaeda neuquenensis (Alonso et  al. 2004). S. divaricata is a
16  Coastal Wetlands of the Bahía Blanca Estuary: Landscape Structure and Plant… 451

perennial shrub 1–3  m tall that inhabits saline soils from Jujuy to Chubut.
Leaves are succulent, semi-cylindrical, and positioned perpendicular to stems.
They are dioecious plants with sessile, axillary flowers that bloom in spring
and fructify in summer (Gates et al. 2018). S. divaricata has vacuoles contain-
ing a high salinity solution, which facilitates survival in arid environments.
These salts are partially eliminated by crystal formation in stomata, what
makes their leaves reflect sunlight, and the name Vidriera (crystal window)
alludes to this optical property. Its ashes are rich in sodium carbonate and can
be employed for traditional saponification.
Heterostachys ritteriana
Subfamily Salicornioideae
Genus Heterostachys
Species ritteriana
Sigla sp. (Moq.)Ung.-Sternb.
Common name Jumecillo
The genus Heterostachys comprises two species, Heterostachys olivas-
cens, found from southern Buenos Aires to Río Negro, and Heterostachys
ritteriana, from Central America to South America (Soriano 1947). H. ritteri-
ana is a halophyte shrub 0.3–0.5 m tall that grows in arid environments. This
is the most frequent species in inland saline steppes of Central and South
America and tolerates saline soils and periodic flooding events (Karlin et al.
2012). The small leaves are succulent, alternate, and caduceus. Young stems
are succulent and turn woody afterward (Pérez Cuadra and Hermann 2014).
This species has solitary flowers located in the axil of succulent bracts.
Family Asteraceae
The largest family of flowering plants, Asteraceae, has around 1535 genera
and 23,000 species (Bremer 1994). This family includes many economically
important species (sunflower, lettuce), as well as many ornamentals. This
group has a cosmopolitan distribution and is highly diversified in its habitat
preference and life forms. It includes aquatics, herbs, and shrubby trees in
temperate, tropical, and arid environments, as well as trees in tropical rainfor-
ests (Jansen and Palmer 1987).
Cyclolepis genistoides
Subfamily Gochnatioideae
Genus Cyclolepis
Species genistoides
Sigla sp. Gillies ex D. Don
Common name Palo Azul
452 P. D. Pratolongo et al.

Cyclolepis genistoides is a native shrub distributed from Chaco to Patagonia


regions, in arid environments and saline soils (Cabrera 1978). C. genistoides
is 1–2.5 m tall, with grayish-green striated stems and branches perpendicular
to the stem. Leaves are alternate, entire, linear-lanceolate, and pubescent on
both sides. The capitulum is yellowish-white, with 10–15 flowers. It blooms
in September–October. It is widely used in folk medicine as a diuretic, anti-
rheumatic, and antispasmodic agent (Sosa et al. 2011).
Family Fabaceae
The Fabaceae is a large and economically important family of angiosperms,
with 730 genera and over 19,400 species (Schrire et  al. 2005). It includes
trees, shrubs, and herbaceous plants, perennials or annuals, which are easily
recognized by their fruits (legume). The group is widely distributed and
becomes the most common family in tropical rainforests and dry forests of the
Americas and Africa (Lavin et al. 2005).
Geoffroea decorticans
Subfamily Faboideae
Genus Geoffroea
Species decorticans
Sigla sp. (Gillies ex Hook. &Arn.) Burkart
Common name Chañar
Geoffroea decorticans is a native South American tree, distributed from
Jujuy to northern Patagonia (Silva et al. 2004). It occurs in dry forests to regu-
larly flooded areas, as well as saline soils. G. decorticans trees can rise up to
12 m tall and form extensive uniform colonies due to their reproduction by
gemmiferous roots (Costamagna et al. 2013). The bark detaches itself in strips
uncovering lighter layers of the same hue. Their leaves are compound, impari-
pinnate, and alternate or arranged in fascicles. The species blooms in
September–October, with shiny yellow flowers arranged in dense 3–8 flower
racemes. Fruits mature in January (Eynard and Galetto 2002). In folk medi-
cine, flowers, fruit, bark, and leaves are used for their emollient, balsamic,
antitussive, and expectorant properties (Silva et al. 2004).

Fennessy 2001). Several salt marsh plants can have as much as 60% of air space in
their root, shoot, and leaf cortex, and species such as S. alterniflora have a continu-
ous air space that extends from leaves to roots (Maricle and Lee 2002; Callaway
et al. 2007). The ability of some salt marsh plants to allow gas transport might have
implications not only for the root itself but also to the surrounding sediments.
Oxygen can be released out of the roots in a process called “radial oxygen loss”
16  Coastal Wetlands of the Bahía Blanca Estuary: Landscape Structure and Plant… 453

which oxidizes the rhizosphere (Armstrong 1979; Colmer 2003) and changes the
redox conditions and chemical forms of several elements (Sundy et  al. 2003).
Another adaptation of many salt marsh species is the development of extensive root
systems to increase contact with oxygen-rich water o with the air. For example, root
and rhizome growth would contribute 50–90% of S. alterniflora productivity
(Redelstein et al. 2018), and S. ambigua has stems with adventitious roots at some
of their nodes (Alonso and Crespo 2008).
Excess in soil salinity has an immediate effect on plant growth and development
(Munns and Tester 2008). For instance, a decline in biomass and plant height with
increasing salt stress has been observed in S. alterniflora (Wang et  al. 2006;
Canepuccia et  al. 2013). Halophytes are plants that have specific adaptations to
survive and grow under high soil salinities. However, different adaptations lead to
large differences in salt tolerance among species (Kruger and Peinemann 1996).
Halophytes can avoid or tolerate salts through exclusion, secretion, shedding, and
succulence (Cronk and Fennessy 2001). Salt exclusion is the most important mech-
anism to cope with high salt concentrations. S. alterniflora excludes at the root level
91–97% of the ions present in saltwater (Bradley and Morris 1991). However, for
most species, salt exclusion at the roots is not enough, and other mechanisms arise.
For example, F. juniperoides and L. brasiliensis have salt glands on their leaves for
salt secretion, and A. undulata secretes salt through salt secretory trichomes located
on leaves and stems (Pérez Cuadra and Cambi 2016). These secreted or excreted
salts are released back to the environment by leaf shedding or by wind and rain
(Wahid 2003).
A great challenge for plants in saline soils is water acquisition. In addition, high
concentrations of salts within the plant can be toxic. Some plants can concentrate
salts in their cell solutions, which results in a high osmotic pressure. Through accu-
mulation of Na+ and Cl− inside the cell, these plants maintain an osmotic gradient
that allows water uptake from saline sediments (Bradley and Morris 1991), but this
increased internal solute concentration can damage the plant. Thus, to avoid toxic
concentrations within the cytoplasm, there is a compartmentalization of the cell,
and Na+ and Cl− get sequestered in vacuoles (Munns and Tester 2008). Succulence
is an increase in cell size induced by salts, due to the large vacuole volume. Well-­
developed water storing tissues in the cortex and pith characterize succulent stems
(Dickison 2000). Succulence is widespread in halophytes and occurs in
Heterostachys, Allenrolfea, and Sarcocornia species in the Bahía Blanca Estuary.
Succulence provides these species with the advantage of diluting their internal
media and decreasing the negative effects of salts (Box 16.1).
454 P. D. Pratolongo et al.

16.3  L
 andscape Dynamics and Changes
in Ecosystem Functions

Despite the limited extent of the area covered by marine vascular plants (<2% of the
ocean surface), salt marshes, mangroves, and seagrass beds play a significant role in
sequestering carbon dioxide. The carbon sequestered in these vegetated coastal eco-
systems has been termed blue carbon. Tidal saline wetlands account for an impor-
tant fraction of the carbon storage in the ocean (Chmura et al. 2003; Mcleod et al.
2011). Although their global area is one to two orders of magnitude smaller than
that of terrestrial forests, the contribution of tidal saline wetlands per unit area to
long-term carbon sequestration is much greater. It is in part due to their generally
high primary productivity and their efficiency in trapping suspended matter and
associated organic carbon from outside their ecosystem boundaries during tidal
inundation. Despite their importance, these blue carbon sinks are being lost at criti-
cal rates in response to multiple stressors (Theuerkauf et al. 2015). Global estimates
of carbon released by salt marsh land-use change are large, ranging from 0.02 to
0.24 Pg CO2 year−1 (Pendleton et al. 2012), and shoreline erosion is a significant
mechanism for current global salt marsh loss, enhanced by relative sea level rise and
human activities (Mariotti et al. 2010).
Based on historical aerial photographs and high-resolution satellite images,
Pratolongo et  al. (2013) quantified changes in land cover in four sites along
Principal Channel, for the years 1967, 1996, and 2005. This work was a first attempt
to analyze changes in size and position of the plant communities within the coastal
zone of Bahía Blanca, as well as primary human modifications like dredge spoil
deposition and landfilling. Results showed that the influence of anthropic transfor-
mations decreases to the west of the port area, that is, through the head of the
Principal Channel. The loss of S. ambigua marshes was a recurring pattern through
the four sites, and, to a lesser extent, the area covered by halophytic steppes also
reduced. In the case of S. ambigua marshes, most of the area changed to mudflats.
This type of replacement involves lowering surface elevation and losing vegetation
cover. The loss of halophytic steppes was mainly due to the replacement of the
original land cover by human land uses (e.g., landfilling and dredge spoil deposits).
Through the four sites, most of the original halophytic steppes area was replaced by
landfills and dredge spoil deposits, and a smaller fraction of these natural environ-
ments eroded to mudflats and channels. Besides these general patterns, the erosion
of S. ambigua marshes dominated landscape dynamics through the head of Principal
Channel and direct human transformations in the port area. While S. ambigua
marshes and halophytic steppes reduced, S. alterniflora marshes expanded their
cover. The expansion of S. alterniflora marshes was widespread west of the port
area, in a process entailing sediment accretion, surface elevation, and colonization
by plants.
More recent work based on these previously identified trends, aimed at quantify-
ing the impacts of land cover replacements on ecosystem function. Maps based on
aerial photographs (year 1967) and high-resolution satellite images (years 2005 and
16  Coastal Wetlands of the Bahía Blanca Estuary: Landscape Structure and Plant… 455

2017) served to quantify changes in area between years, and the updated land cover
maps, over a larger area, confirmed most of the observed tendencies (Fig. 16.4). In
the North section of the Bahía Blanca Estuary (i.e., the area represented in Fig. 16.4),
erosion rates of S. ambigua marshes (% loss per year) accelerated from 0.8% year
−1
(from 1967 to 2005) to 2.0% year −1 (from 2005 to 2017). For this later period,
S. ambigua marshes lost to mudflats at an average rate of 93 ha year −1. Halophytic
shrubs and steppes also reduced their cover, but erosion was less pronounced and
steady through time. For the entire period (1964–2017), 1750  ha of halophytic
steppes, equivalent to 14% of the original area, eroded at an average rate of
40 ha year −1. In turn, halophytic shrubs eroded at an average rate of 11 ha year −1
(545 ha for the entire period).
Elevation profiles obtained in the field were used to estimate how deep the marsh
soil was excavated during shoreline transgression. To assess the volume of material
eroded from S. ambigua marshes, the total area of salt marshes replaced by mudflats
for a given pair of dates was considered, along with the depth of erosion at the
shoreline. Considering the average erosion depths and soil organic carbon densities,
the northern section of the Bahía Blanca Estuary exported soil organic carbon at a
rate of 2893 t C year −1, because of marsh erosion between 2005 and 2017. The soil
organic material that is being eroded presents a very low C:N ratio (8–10), and its
bulk isotopic composition (δ13C −24.7 to −16.8‰; δ15N + 8.7 to +11.3‰) reflects
a dominance of organic matter of marine origin (unpublished data). Along with the
loss of salt marsh soils, erosion involves a large amount of above- and belowground
biomass of S. ambigua, which may represent an additional export of between ~450
and 1400  t C year−1 (based on biomass data and carbon contents in Negrin
et al. 2016).
According to the landscape evolution described for the area, the presently inter-
tidal platform occupied by S. ambigua marshes is composed of marine deposits that
formed under a higher relative sea level, after the transgressive maximum in the
Holocene. The present relative sea level trend in the Bahía Blanca Estuary has not
been evaluated, but the closest estimation is 0.85 mm year−1, with a 95% confidence
interval of ±0.31 mm year−1, based on tide gauge station measurements obtained
from 1918 to 1982 in Puerto Quequén (290 km north-east from Bahía Blanca). In
Puerto Madryn (about 450 km south-west from Bahía Blanca), the mean relative sea
level trend is 1.5 mm year−1 with a 95% confidence interval of ±0.79 mm year−1
based on data from 1944 to 2000 (Permanent Service for Mean Sea Level, available
through NOAA’s Sea Levels Online website at http://tidesandcurrents.noaa.gov/
sltrends/sltrends.html). Under the current rates of relative sea level rise, there is
accelerated erosion, and salt marsh soils act as a significant net source of organic
carbon to estuarine waters.
These estimations do not discern the proportion of labile or refractory carbon in
the eroded material. Moreover, the fate of eroded carbon is complex and can follow
multiple pathways that are highly dependent on the individual characteristics of the
estuary. Some eroded material would be transported in suspension and deposited
elsewhere in the estuary. Relatively young and bioavailable carbon would be
respired within the estuary and metabolized by microbes (Canuel et al. 2012). On
456 P. D. Pratolongo et al.

Fig. 16.4  Land cover changes between years 1967 and 2017
16  Coastal Wetlands of the Bahía Blanca Estuary: Landscape Structure and Plant… 457

the contrary, older and more refractory soil carbon that is not processed within the
estuary would be exported to open waters, where it may enter detrital ocean webs or
the microbial loop or it may sink to the ocean floor (Cai et al. 2003). Thus, the rela-
tive importance of carbon export from marshes for the global carbon budget depends
on the fate and biochemical characteristics of the eroded carbon.
While S. ambigua marshes are eroding, S. alterniflora marshes are expanding
their distribution, and part of the eroded material may be redeposited in these
expanding marshes. In 1967, for the same region in the northern section of the estu-
ary, S. alterniflora marshes covered 4454 ha. The area increased at accelerated rate
from 1967 to 2005 (62 ha year−1), but the expansion slowed down more recently.
S. alterniflora marshes covered 7090 ha by 2017, and their expansion rate for this
later period was 22.5 ha year−1. Islote del Puerto is an island in the harbor area that
hosts a large breeding colony of the endangered species Olrog’s gull (Larus atlanti-
cus). In this island, S. alterniflora marshes covered 178  ha by 1967. Vegetation
cover also encompassed S. ambigua marshes and halophytic shrubs (208.2 and
70.8  ha, respectively), but these areas were completely covered with a massive
dredge spoil deposit in 1989. In 2005, S. alterniflora covered 566  ha occupying
most of the intertidal fringe, and marshes did not expand from 2005 to 2017.
Spartina alterniflora is presently increasing in abundance and expanding its
regional extent in many South American countries, including several coastal loca-
tions in Argentina. This species is globally regarded as a powerful ecosystem engi-
neer, based on the ability to cause significant modifications to the abiotic
environment. Worldwide, the establishment and expansion of S. alterniflora marshes
fostered large-scale alterations in ecosystem processes (see Box 16.2), including the
extinction of native species, loss of functional native diversity, changes in nutrient
cycling and organic matter storage, and loss of habitat (Ayres et al. 2004). The con-
sequences of S. alterniflora expansion in South America are largely unexplored.
Besides carbon sequestration, there are essential ecosystem services provided by
salt marshes that relate to elemental cycling (Barbier et al. 2011). For instance, for
biomass production and growth, plants take up nutrients from the sediments. Both
essential and toxic metals can also be absorbed during plant growth, providing tools
for mitigation of eutrophication and metal pollution in coastal waters (Tangahu
et al. 2011). The uptake of nutrients and metals depends mainly on the elemental
concentrations in the porewater and sediment characteristics like grain size, organic
matter content, pH, and redox potential (Wang et al. 2013). Although mineral uptake
occurs at the root level, elements can be further translocated to leaves and stems,
setting a distinctive distribution within aerial and belowground tissues and linking
their cycling to the fate of detritus. After senescence, plant tissues release elements
back to the environment through decomposition, whose rate depends on environ-
mental conditions like climate, but also on detritus quality. Thus, decomposition
rates usually differ between species and between organs within a species (Simões
et al., 2011; Tong et al. 2011).
Negrin et al. (2016) reviewed the available information on ecological and bio-
geochemical processes in salt marshes of the Bahía Blanca Estuary. Evaluations of
the belowground and aboveground biomass dynamics, decomposition, and nutrient
458 P. D. Pratolongo et al.

Box 16.2 Spartina alterniflora: An Invasive Exotic or a Native Increasing


Its Range?
Invasive species are those introduced to an ecosystem that they did not occupy
previously and establish a population that spreads autonomously (Simberloff
2010). Once established in the non-native environment, invasive species can
affect the ecosystem in many ways, including modifications that affect most
of the originally resident species. Occasionally, native plant species spread
into formerly unoccupied habitats and become invasive (Valéry et al. 2009).
An example is the large expansion of Phragmites australis in North American
coastal marshes, which spread rapidly into new habitats after the mid-­
nineteenth century. The introduction of Old-World genotypes in North
America has triggered the invasive behavior of this native species (Saltonstall
2002). Native species may also become invasive after human modifications of
the environment. For instance, Elymus athericus was commonly present in
low abundances in high marshes of western Europe. Since the mid-1980, this
native grass has been aggressively invading the middle and low marshes,
where it often forms dense monospecific stands replacing natural Atriplex
portulacoides low marshes. This seaward expansion is facilitated by increased
anthropogenic nitrogen in aerial depositions and runoff (Valéry et al. 2016).
The grass group Spartina encompasses several successfully invasive spe-
cies of intertidal mudflats and salt marshes worldwide (Ainouche and Gray
2016). The critical ecological role of several Spartina species as “ecosystem
engineers” on coastal salt marshes and their remarkable history punctuated by
natural or human-mediated introductions outside their native range, rapid
expansion, and propensity to interspecific hybridization and polyploid specia-
tion have long captured the attention of researchers and institutions involved
in land management. In 2014, a DNA-based phylogenetic study confirmed the
paraphyly of the grass subtribe Sporobolinae, and the creation of a large
monophyletic genus Sporobolus was proposed, which includes species previ-
ously included in the genus Spartina (Peterson et al. 2014). The name Spartina
designates a morphologically well-circumscribed group of grasses, represent-
ing a monophyletic clade. Spartina species have a tremendous impact in many
scientific and non-scientific fields, provided their global ecological impor-
tance (Bortolus et al. 2019). Therefore, for the sake of simplicity, the newly
designated subgenus Spartina will be treated here as the genus Spartina.
Spartina anglica appeared by the end of the 1880s in Southampton Water,
UK, by chromosomal doubling from Spartina x townsendii, a hybrid between
the introduced Spartina alterniflora and the native Spartina maritima (Ayres
and Strong, 2001). S. anglica invaded large areas during the first 30  years
(Raybould, 1997), and the species continues to expand along the north-east
and north-west coasts of England, despite the aggressive control methods
employed (Lacambra et al. 2004). S. anglica is also increasing in abundance
and spreading in marshes of the Wadden Sea, possibly due to warmer spring
16  Coastal Wetlands of the Bahía Blanca Estuary: Landscape Structure and Plant… 459

temperatures (Nehring and Hesse 2008). Spartina anglica was later intro-
duced to Australia in the 1920s. The species was planted across many states,
but populations established in Port Gawler, South Australia, Bass River and
Western Port Bay, and Victoria, as well as northern and eastern Tasmania.
After management for control and eradication of the species, the largest infes-
tations persist in the Tamar River and the Rubicon Estuary, in Tasmania, as
well as the Anderson Inlet, in Victoria (Beasy and Ellison 2013).
In Willapa Bay, in the Pacific coast of North America, S. alterniflora
invaded nearly one third of the original mudflat area. This species was intro-
duced during the late 1800s but was not reported until the 1940s. During the
first 50 years, salt marshes gradually expanded throughout the bay, resulting
in radical changes to the ecosystem functions (Simenstad and Thom 1995).
During the past decades, after an aggressive eradication campaign that
included widespread application of herbicides, S. alterniflora has virtually
disappeared from Willapa Bay (Strong and Ayres 2016).
Spartina alterniflora was first transplanted in coastal China in 1979, to
stabilize tidal flats. Since the mid-1990s, this species began to spread in the
Yangtze Estuary (Ouyang et  al. 2013) gradually invading native marshes
(Cheng et al. 2006). Before the introduction of S. alterniflora, coastal wet-
lands in the Jiangsu Province typically contained P. australis and Suaeda
salsa high marshes, with extensive bare mudflats at lower elevations. At pres-
ent, S. alterniflora characterizes plant zonation in the Jiangsu Province. It is
the dominant species forming marshes in the upper intertidal zone, while
marshes of S. salsa and P. australis persist landward. In estuarine areas of the
Yangtze Delta, S. alterniflora occupies an elevation range that overlaps with
the native salt marsh species, often leading to its disappearance (Gao
et al. 2014).
In Atlantic South America, Spartina alterniflora increased its abundance
and expanded its regional extent over the twentieth century. Early records and
indirect observations indicate that S. alterniflora may have expanded from
Rio de Janeiro, where it was first collected in 1817. To the North, the species
was not registered on the coasts of Suriname and Guyana until the 1830–1840s.
To the South, S. alterniflora was first detected in 1880, in Uruguay. In
Argentina, this species was first collected in the Bahía Blanca Estuary in
1902, near Punta Alta. Noteworthy, S. alterniflora was never reported in the
many significant botanical accounts made across the Argentine coast during
the 1800s (Bortolus et al. 2015). More recently, this species has increased in
abundance in many South American countries. In Argentina, salt marshes of
S. alterniflora are aggressively expanding along mudflats of the Bahía Blanca
Estuary. In San Antonio Bay (Río Negro province), extensive marshes
appeared after 1914 (Willis 1914), and marsh expansions were also docu-
mented in Península Valdés (42°24’ S), at the southernmost limit of its distri-
bution (Bortolus et al. 2015).
460 P. D. Pratolongo et al.

Although S. alterniflora has been historically described as a species native


to the Atlantic coasts of South America, recent work proposed that this spe-
cies would be either an exotic invader in this region or an infrequent native
plant, whose modern abundance and geographic expansion are due to habitat
changes (Bortolus et al. 2015; Schwindt et al. 2018). In Argentina, S. alterni-
flora was traditionally regarded as a native species (e.g., Zuloaga et al. 2019).
Accordingly, S. alterniflora does not appear in the national database of exotic
species (http://www.inbiar.uns.edu.ar/). Yet, the native status of many species
may have been wrongly assigned, leading to substantial misinterpretations of
ecosystem change (Carlton 2009). Based on several historical, environmental,
biological, ecological, geographic, vector, physiological, and morphological
criteria, Bortolus et al. (2015) proposed that S. alterniflora would be a non-
native species, which was accidentally introduced from either North America
or Europe, sometime prior to 1817. In case that additional evidence corrobo-
rates that S. alterniflora is introduced to South America, this would be one of
the largest biological invasions involving this species, with these alien marshes
entirely reshaping coastal systems (Bortolus et al. 2015).

contents allowed for the identification of major pathways in the biogeochemical


cycling of carbon, nitrogen, and phosphorus. Comparing result obtained in
S. ambigua and S. alterniflora marshes, located close to the elevation of the mean
high tide, both types of marshes would be similar in terms of their net aerial primary
production. Thus, they would play similar roles in the sequestration of carbon and
nitrogen. For both types of marshes, published values on sequestration rates are in
the range 181–247 g C m−2 year−1 and 4.9–6.7 g N m−2 year−1. Regarding phospho-
rus sequestration, however, marshes of S. alterniflora would be more efficient
because of the higher phosphorus concentration in aboveground tissues. While the
average phosphorus sequestration in S. ambigua is 0.33  g P m−2  year−1, values
reported for S. alterniflora marshes range 0.42–0.57 g P m−2 year−1.
Both types of marshes also differ in their decomposition rates. Although aboveg-
round tissues decompose at relatively similar rates, belowground tissues of
S. ambigua decompose faster than S. alterniflora biomass (Negrin et al. 2016). The
average difference of 70 versus 29% during the first year for belowground material
suggests that both types of marshes have a differential impact on elemental cycling.
Accordingly, the loss of elements bound to fine particulate and dissolved organic
and inorganic compounds would be higher in S. ambigua marshes, but marshes of
S. alterniflora would be more efficient for the accumulation in situ of elements
bound to undecomposed detritus. The nature of the exported material (particulate or
dissolved, organic or inorganic) defines the ultimate fate of elements and their abil-
ity to re-enter grazing and detrital food webs, but S. ambigua marshes would be
more efficient in the recycling of elements, while S. alterniflora marshes would play
16  Coastal Wetlands of the Bahía Blanca Estuary: Landscape Structure and Plant… 461

a more substantial role in the long-term storage of elements, and the building up of
elevation, through the accumulation of undecomposed plant material.
Considering the different biogeochemical roles of S. alterniflora and S. ambigua
marshes, the observed land cover changes may have a large impact on the overall
elemental cycling within the Bahía Blanca Estuary, and the actual rates of change
may accelerate in response to global warming and sea level rise (Kirwan and Murray
2008). Direct human modifications derived from dredging of navigation channels,
urban expansion, and industrial development in the coastal zone might also modify
salt marsh cover. Moreover, net primary productivity and decomposition rates are
also sensitive to climate change (Kirwan and Blum 2011). Under this changing
scenario, further research is needed to completely understand the complex dynam-
ics of biogeochemical cycling in these salt marshes.

16.4  Conclusions

This chapter offers an insight on the factors shaping landscape structure and wet-
land dynamics in the Bahía Blanca Estuary. The Holocene marine transgression and
subsequent regressive stages had a remarkable influence on the coastal landscape
structure and define hydrogeomorphic wetland presence and characteristics.
Vegetation is dominated by halophytic species, like S. alterniflora, S. ambigua,
A. patagonica, C. genistoides, and A. undulata. These species associate to form salt
marshes, steppes, and shrubs. Barren surfaces are also extensive, with tidal mudflats
covering most of the intertidal fringe. Salt flats and bare soil in sparsely vegetated
steppes thoroughly spread beyond the limits of tidal influence. S. alterniflora has
been increasing its abundance over the last decades, and new marshes appeared in
the area. As a counterpart, S. ambigua marshes eroded at accelerated rate, with a
large exportation of particulate organic carbon from the marsh soil pool to estuarine
waters. S. alterniflora and S. ambigua marshes play different roles in sequestration
and remineralization of C, N, and P. Thus, the observed land cover changes may
have a large impact on the overall elemental cycling within the Bahía Blanca
Estuary.

References

Adam P (1990) Saltmarsh ecology. Cambridge University Press, New York


Ainouche M, Gray A (2016) Invasive Spartina: lessons and challenges. Biol Inv18:2119–2122
Aliotta S, Farinati EA (1990) Stratigraphy of Holocene sand-shell ridges in the Bahía Blanca
Estuary, Argentina. Mar Geol 34:353–360
Aliotta S, Ginsberg SS, Giagante D et al (2014) Seismic stratigraphy of Pleistocene deltaic depos-
its in Bahía Blanca Estuary, Argentina. An Acad Brase Cienc 86:649–662
Alonso MÁ, Crespo MB (2008) Taxonomic and nomenclatural notes on South American taxa of
Sarcocornia (Chenopodiaceae). Ann Bot Fenn 5:241–255
462 P. D. Pratolongo et al.

Alonso MA, Conticello L, Cerazo MB (2004) Suaeda neuquenensis (Chenopodiaceae), a new spe-
cies from Argentina. Novon 14:1–5
Armstrong W (1979) Aeration in higher plants. Advances in Botanical Research 7:225–332
Ayres DR, Strong DR (2001) Origin and genetic diversity of Spartina anglica (Poaceae) using
nuclear DNA markers. Am J Bot 88:1863–1867
Ayres DR, Zaremba K, Strong DR (2004) Extinction of a Common Native Species by Hybridization
with an Invasive Congener. Weed Technol 18:1288–1292
Bao-Shan CUI, Qiang HE, Yuan AN (2011) Community structure and abiotic determinants of salt
marsh plant zonation vary across topographic gradients. Estuar Coast 34:459–469
Barbier EB (2019) The Value of Coastal Wetland Ecosystem Services. In: Perillo GME, Wolanski
E, Cahoon DR, Hopkinson CS (eds) Coastal Wetlands, 2nd edn. Elsevier, pp 947–964
Barbier EB, Hacker SD, Kennedy C et al (2011) The value of estuarine and coastal ecosystem
services. Ecol Monogr 81:169–193
Barth HJ, Böer B (2002) Introduction. In: Barth HJ, Böer B (eds) Sabkha Ecosyst. Vol. I Arab.
Penins. Adjac. Ctries. Kluwer Academic Publishers, Dordrecht, Netherlands, pp 1–5
Beasy KM, Ellison JC (2013) Comparison of three methods for the quantification of sediment
organic carbon in salt marshes of the Rubicon Estuary, Tasmania. Australia. Int J Biol 5:1–13
Bertin RL, Maltez HF, de Gois JS et  al (2016) Mineral composition and bioaccessibility in
Sarcocornia ambigua using ICP-MS. J Food Compos Anal 47:45–51
Bertness MD (1991) Zonation of Spartina patens and Spartina alterniflora in New England salt
marsh. Ecology 72:138–148
Boettinger JL (1997) Aquisalids (Salorthids) and other wet saline and alkaline soils: Problems
identifying aquic conditions and hydric soils. In: Vepraskas MJ, Sprecher SW (eds) Aquic
Cond. hydric soils Probl. soils. SSSA Spec. Publ. No. 50. Soil Science Society of America,
Madison, WI, pp p79–p97
Bortolus A, Carlton JT, Schwindt E (2015) Reimagining South American coasts: unveiling the hid-
den invasion history of an iconic ecological engineer. Divers Distrib 21:1267–1283
Bortolus A, Adam P, Adams JB et  al (2019) Supporting Spartina: Interdisciplinary perspective
shows Spartina as a distinct solid genus. Ecology 100:e02863
Bradley PM, Morris JT (1991) Relative importance of ion exclusion, secretion and accumulation
in Spartina alterniflora Loisel. J Exp Bot 42:1525–1532
Bremer K (1994) Asteraceae: cladistics and classification. Timber Press, Portland, Ore
Brinson MM, Christian RR, Blum LK (1995) Multiple states in the sea-level induced transition
from terrestrial forest to estuary. Estuaries 18:648–659
Cabido M, Zeballos SR, Zak M et al (2018) Native woody vegetation in central Argentina: clas-
sification of Chaco and Espinal forests. Appl Veg Sci 21:298–311
Cabrera AL (1978) Flora de la provincia de Jujuy, República Argentina: Compositae. INTA,
Buenos Aires
Cai WJ, Wang Y, Krest J et al (2003) The geochemistry of dissolved inorganic carbon in a surficial
groundwater aquifer in North Inlet, South Carolina, and the carbon fluxes to the coastal ocean.
Geochim Cosmochim Ac 67:631–639
Callaway RM, Jones S, Ferren WR et  al (1990) Ecology of a mediterranean-climate estuarine
wetland at Carpinteria, California: plant distributions and soil salinity in the upper marsh. Can
J Botany 68:1139–1146
Callaway JC, Parker VT, Vasey MC et al (2007) Emerging issues for the restoration of tidal marsh
ecosystems in the context of predicted climate change. Madrono 54:234–249
Canepuccia AD, Pérez CF, Farina JL et al (2013) Dissimilarity in plant species diversity between
salt marsh and neighboring environments decreases as environmental harshness increases. Mar
Ecol Prog Ser 494:135–148
Cantero JJ, Cisneros JM, Zobel M et al (1998) Environmental relationships of vegetation patterns
in saltmarshes of central Argentina. Folia Geobot 33:133–145
Canuel EA, Cammer SS, McIntosh HA et al (2012) Climate change impacts on the organic carbon
cycle at the land-ocean interface. Ann Rev Earth Pl Sc 40:685–711
16  Coastal Wetlands of the Bahía Blanca Estuary: Landscape Structure and Plant… 463

Carlton JT (2009) Deep Invasion Ecology and the Assembly of Communities in Historical Time.
In: Rilov G, Crooks JA (eds) Biological Invasions in Marine Ecosystems. Springer, Berlinp,
pp 13–56
Cavallotto JL, Violante RA, Parker G (2004) Sea-level fluctuations during the last 8600 years in the
de la Plata river (Argentina). Quat Int 114:155–165
Celleri C, Zapperi G, González Trilla G et al (2019) Assessing the capability of broadband indices
derived from Landsat 8 Operational Land Imager to monitor above ground biomass and salin-
ity in semiarid saline environments of the Bahía Blanca Estuary, Argentina. Int J Remote Sens
40:4817–4838
Cheng X, Luo Y, Chen J et al (2006) Short-term C4 plant Spartina alterniflora invasions change
the soil carbon in C3 plant-dominated tidal wetlands on a growing estuarine Island. Soil Biol
Biogeochemestry 38:3380–3386
Chmura GL, Anisfeld SC, Cahoon DR et al (2003) Global carbon sequestration in tidal, saline
wetland soils. Global Biogeochem Cy 17(4). https://doi.org/10.1029/2002GB001917
Christian RR, Stasavich LE, Thomas CR et al (2002) References is a Moving Target in Sea-Level
Controlled Wetlands. In: Weinstein MP, Kreeger DA (eds) Concepts and Controversies in
Tidal Marsh Ecology. Springer, Dordrecht, pp 805–825
Cintron-Molero G, Schaeffer Novelli Y (2002) A guide to wetlands on arid and semiarid zones.
USFWS División of International Conservation, USAColmer TD (2003) Long-distance
­transport of gases in plants: a perspective on internal aeration and radial oxygen loss from
roots. Plant. Cell and Environ 26:17–36
Colmer TD (2003) Long‐distance transport of gases in plants: a perspective on internal aeration
and radial oxygen loss from roots. Plant Cell Environ 26:17–36
Costamagna MS, Ordóñez RM, Zampini IC et al (2013) Nutritional and antioxidant properties of
Geoffroea decorticans, an Argentinean fruit, and derived products (flour, arrope, decoction and
hydroalcoholic beverage). Food Res Int 54:160–168
Cronk JK, Fennessy MS (2001) Wetland plants: biology and ecology. Lewis Publishers, CRC
Press, Boca Raton, USA
Davy AJ, Costa CSB (1992) Development and Organization of Saltmarsh Communities. In:
Seeliger U (ed) Coastal plant communities of Latin America. Academic Press, pp 157–178
del Vitto LA, Petenatti EM, Petenatti ME (1997) Recursos herbolarios de San Luis (República
Argentina) primera parte: plantas nativas. Multequina 6:49–66
Dickison WC (2000) Integrative Plant Anatomy. Harcourt Academic Press, San Diego
Doody JP (1992) The conservation of British saltmarshes. In: Allen JR, Pye K (eds) Saltmarshes:
Morphodynamics, conservation and engineering significance. Cambridge University Press,
New York, pp 80–114
Ecke F, Rydin H (2000) Succession on a land uplift coast in relation to plant strategy theory. Ann
Bot Fennici 37:163–171
Eynard C, Galetto L (2002) Pollination ecology of Geoffroea decorticans (Fabaceae) in central
Argentine dry forest. J Arid Environ 51:79–88
Farinati EA (1983) Paleontología, Paleoecología y Paleogeografía de los sedimentos marinos de
los alrededores de Bahía Blanca. PhD Thesis. Universidad Nacional del Sur, Bahía Blanca,
Argentina
Farinati EA, Aliotta S, Ginsberg SS (1992) Mass mortality of a Holocene Tagelus plebeius
(Mollusca, Bivalvia) population in the Bahía Blanca Estuary, Argentina. Mar Geol 106:301–308
Freitas RF, Costa CS (2014) Germination responses to salt stress of two intertidal populations of
the perennial glasswort Sarcocornia ambigua. Aquat bot 117:12–17
Gaitan JJ, Bran DE, Oliva GE et  al (2019) Patagonian desert. In: Reference module in Earth
systems and environmental sciences. Elsevier, Amsterdam. https://doi.org/10.1016/
B978-0-12-409548-9.11929-3
Gao S, Du Y, Xie W et al (2014) Environment-ecosystem dynamic processes of Spartina alterni-
flora salt-marshes along the eastern China coastlines. Sci China Earth Sci 57:2567–2586
464 P. D. Pratolongo et al.

Gardner LR, Reeves HW, Thibodeau PM (2000) Groundwater dynamics along forest-marsh tran-
sects in a southeastern salt marsh, USA: description, interpretation, and challenges for numeri-
cal modeling. Wetl Ecol Manag 10:145–159
Gates MW, Torrens J, Fidalgo P et al (2018) The gall associates of Asphondylia poss. swaedicola
Kieffer & Jörgensen (Diptera: Cecidomyiidae) on Suaeda divaricata Moq.(Amaranthaceae) in
the semiarid Argentina and summary of parasitic Hymenoptera associated with Suaeda world-
wide. Neotrop entomol 47:598–609
Gehrels WR (2000) Using foraminiferal transfer functions to produce high-resolution sea-level
records from salt-marsh deposits, Maine, USA. The Holocene 10:367–376
Giagante DA, Aliotta S, Ginsberg SS (2008) Análisis sismoestratigráfico de paleocanales en el
subsuelo marino del estuario de Bahía Blanca. Rev de la Asociación Geológica Argentina
63:65–75
Giagante DA, Aliotta S, Ginsberg SS et  al (2011) Evolution of a coastal alluvial deposit in
response to the last Quaternary marine transgression, Bahía Blanca estuary, Argentina. Quat
Res 75:614–623
Giusti L (1997) Chenopodiaceae. In: Hunziker AT (ed) Flora Fanerogámica Argentina. Proflora
Conicet, Buenos Aires, pp 1–53
Gómez-Miguel V, Peres Arias J, Guerrero F, et al (1984) The soils and water table properties of the
Polder area “Castillo de Dona Blanca”, Puerto de Santa Maria, Cadiz Spain. In: Polders of the
world. International symposium, Netherlands, pp 374–383
González MA (1989) Holocene levels in the Bahía Blanca Estuary, Argentina Republic. J Coast
Res 5:65–77
González Trilla G, Pratolongo P, Beget ME et  al (2013) Relating Biophysical Parameters of
Coastal Marshes to Hyperspectral Reflectance Data in the Bahía Blanca Estuary, Argentina. J
Coast Res 286:231–238
González MA, Weiler NE (1983) Ciclicidad de niveles marinos holocénicos en Bahía Blanca y
en el Delta del Río Colorado (Provincia de Buenos Aires), en base a edades de Carbono – 14.
Oscil. del Niv. del mar durante el último hemiciclo deglacial en la Argentina. Mar del Plata,
Bahía Blanca, pp 69–90
González MA, Panarello HO, Marino H et al (1983) Niveles marinos del Holoceno en el estu-
ario de Bahía Blanca (Argentina). Isótopos estables y microfósiles calcáreos como indicadores
paleoambientales. Oscil. del Niv. del mar durante el último hemiciclo deglacial en la Argentina.
Mar del Plata, Bahía Blanca, pp 48–68
González E, Gonzalez Trilla G, San Martin L et al (2019) Vegetation patterns in a South American
coastal wetland using high-resolution imagery. J Maps 15:642–650
González-Uriarte M (1984) Características geomorfológicas de la porción continental que rodea
la Bahía Blanca, Provincia de Buenos Aires. IX Congr. Argentino Geol. Bariloche, Argentina,
pp 556–576
Hageman BP (1969) Development of the western part of the Netherlands during the Holocene.
Geol en Mijnb 48:373–386
Henderson RE, Patrick JR (1982) Soil aeration and plant productivity. In: Rechcigl M Jr (ed)
Handb. Agric. Product. CRC Press, Boca Raton, USA, pp 51–69
Horton BP, Edwards RJ (2006) Quantifying Holocene sea level change using intertidal foramin-
ifera: lessons from the British Isles. Departmental Papers http://repository.upenn.edu/ees_
papers/50. Accessed 21 Dec 2019
Isacch JP, Costa CSB, Rodríguez-Gallego L et  al (2006) Distribution of saltmarsh plant com-
munities associated with environmental factors along a latitudinal gradient on the south-west
Atlantic coast. J Biogeogr 33:888–900
Isla FI (1989) Holocene sea-level fluctuation in the southern hemisphere. Quat Sci Rev 8:359–368
Isla FI, Cortizo LC, Schnack EJ (1996) Pleistocene and Holocene beaches and estuaries along the
Southern Barrier of Buenos Aires, Argentina. Quat Sci Rev 15:833–841
Jansen RK, Palmer JD (1987) A chloroplast DNA inversion marks an ancient evolutionary split in
the sunflower family (Asteraceae). P Natl Acad Sci USA 84:5818–5822
16  Coastal Wetlands of the Bahía Blanca Estuary: Landscape Structure and Plant… 465

Kadereit G, Borsch T, Weising K et al (2003) Phylogeny of Amaranthaceae and Chenopodiaceae


and the evolution of C4 photosynthesis. Int J Plant Sci 164:959–986
Karlin MS, Buffa EV, Karlin UO et al (2012) Relations among soil properties, plant communities
and cattle receptivity in saline environments (Salinas Grandes, Catamarca, Argentina). Rev
Latinoamericana de Recursos Naturales 8:30–45
Kirwan ML, Blum LK (2011) Enhanced decomposition offsets enhanced productivity and soil car-
bon accumulation in coastal wetlands responding to climate change. Biogeosciences 8:987–993
Kirwan ML, Murray AB (2008) Ecological and morphological response of brackish tidal marsh-
land to the next century of sea level rise: Westham Island, British Columbia. Global Planet
Change 60:471–486
Kruger HR (1989) Suelos halomórficos de la Bahía Blanca. Dinámica de sales y relación con la
vegetación. PhD Thesis. Universidad Nacional del Sur, Bahía Blanca, Argentina
Kruger HR, Peinemann N (1996) Coastal plain halophytes and their relation to soil ionic composi-
tion. Vegetatio 122:143–150
Lacambra C, Cutts N, Allen J, et al. (2004) Spartina anglica: a review of its status, dynamics and
management. English Nature Research Reports No. 527, English Nature, Peterborough, UK
Lavin M, Herendeen PS, Wojciechowski MF (2005) Evolutionary rates analysis of Leguminosae
implicates a rapid diversification of lineages during the tertiary. Syst Biol 54:575–594
León RJC, Bran D, Collantes M et al (1998) Grandes unidades de vegetación de la Patagonia extra
andina. Ecología Austral 8:123–141
Maricle BR, Lee RW (2002) Aerenchyma development and oxygen transport in the estuarine cord-
grasses Spartina alterniflora and S-anglica. Aquat Bot 74:109–120
Mariotti G, Fagherazzi S, Wiberg PL et al (2010) Influence of storm surges and sea level on shallow
tidal basin erosive processes. J Geophys Res. 115(C11). https://doi.org/10.1029/2009JC005892
Marquez M, Ferrero L, Cusminsky GC (2016) Holocene palaeoenvironmental evolution of the
Pampa coastal plain (Argentina) based on calcareous microfossils. Rev Bras Paleontol 19:25–40
Mcleod E, Chmura GL, Bouillon S et al (2011) A blueprint for blue carbon: toward an improved
understanding of the role of vegetated coastal habitats in sequestering CO2. Front Ecol Environ
9:552–560
Mendelssohn IA, Morris JT (2000) Eco-physiological controls on the primary productivity of
Spartina alterniflora. Concepts and Controversies in Tidal Marsh. Ecology:59–80
Moffett KB, Gorelick SM, McLaren RG et  al (2012) Salt marsh ecohydrological zonation due
to heterogeneous vegetation–groundwater–surface water interactions. Water Resour Res
48(2):W02516. https://doi.org/10.1029/2011WR010874
Múlgura ME, Marticorena A (2008) Chenopodiaceae. In: Zuloaga FO, Morrone O, Belgrano MJ
(eds) Catálogo de las Plantas Vasculares del Cono Sur, vol II. Missouri Botanical Garden Press,
Missouri, pp 1909–1929
Munns R, Tester M (2008) Mechanisms of salinity tolerance. Annu Rev Plant Biol 59:651–681
Nebbia AJ, Zalba S (2007) Comunidades Halófilas de la costa de la Bahía Blanca (Argentina):
Caracterización, mapeo y cambios durante los últimos cincuenta años. Boletín la Soc Argentina
Botánica 42:261–271
Negrin VL, Botté SE, Pratolongo PD et al (2016) Ecological processes and biogeochemical cycling
in salt marshes: synthesis of studies in the Bahía Blanca estuary (Argentina). Hydrobiologia
774:217–235
Nehring S, Hesse KJ (2008) Invasive alien plants in marine protected areas: the Spartina anglica
affair in the European Wadden Sea. Biol. Invasions 10:937–950
Oertel GF, Kearney MS, Leatherman SP et al (1989) Anatomy of a barrier platform: outer barrier
lagoon, southern Delmarva Peninsula, Virginia. Mar Geol 88:303–318
Ouyang ZT, Gao Y, Xie X et al (2013) Spectral discrimination of the invasive plant Spartina alter-
niflora at multiple phenological stages in a saltmarsh wetland. PLoS One 8:e67315
Oyarzabal M, Clavijo JR, Oakley LJ et al (2018) Unidades de vegetación de la Argentina. Ecología
Austral 28:40–63
466 P. D. Pratolongo et al.

Paruelo JM, Beltrán B, Jobbagy EG et al (1998) The climate of Patagonia: general patterns and
controls on biotic processes. Ecología Austral 8:85–101
Pendleton L, Donato DC, Murray BC et  al (2012) Estimating global “blue carbon” emissions
from conversion and degradation of vegetated coastal ecosystems. PloS One 7(9). https://doi.
org/10.1371/journal.pone.0043542
Pennings SC, Bertness MD (1999) Using latitudinal variation to examine effects of climate on
coastal salt marsh pattern and process. Current Topics in Wetland Biogeochemistry 3:100–111
Pennings SC, Bertness MD (2001) Salt marsh communities. In: Bertness MD, Gaines SD, Hay ME
(eds) Marine community ecology. Sinauer Associates, Sunderland, pp 289–316
Pennings SC, Grant MB, Bertness MD (2005) Plant zonation in low-latitude salt marshes: disen-
tangling the roles of flooding, salinity and competition. J Ecol 93:159–167
Pérez Cuadra V (2008) Salitral de la Vidriera, un refugio para especies vegetales con característi-
cas particulares. Boletín de la Sociedad Latinoamericana y del Caribe de Cactáceas y otras
Suculentas 5:6–8
Pérez Cuadra V, Cambi VN (2016) Caracteres epidérmicos de 30 especies xero-halófilas:¿ es el
ambiente el principal factor determinante? Lilloa 53:282–298
Pérez Cuadra V, Hermann R (2014) Anatomía foliar y caulinar de tres Salicornieae (Chenopodiaceae)
halófilas argentinas. Phyton-Int J Exp Bot 83:369–377
Peterson PM, Romaschenko K, Herrera Arrieta Y (2014) A molecular phylogeny and classification
of the Cteniinae, Farragininae, Gouiniinae, Gymnopogoninae, Perotidinae, and Trichoneurinae
(Poaceae: Chloridoideae: Cynodonteae). Taxon 63:275–286
Piovan MJ (2016) Controles geomorfológicos sobre la presencia y estructura de humedales coste-
ros en el estuario de Bahía Blanca. PhD Thesis. Universidad Nacional del Sur, Bahía Blanca,
Argentina
Piovan MJ, Zapperi GM, Pratolongo PD (2014) Seed germination of Atriplex undulata under
saline and alkaline conditions. Seed Sci Technol 42:286–292
Pratolongo P, Kirby JR, Plater A et al (2009) Temperate coastal wetlands: morphology, sediment
processes, and plant communities. In: Perillo GME, Wolanski E, Cahoon DR, Brinson MM
(eds) Coastal Wetlands 1 st edn. Elsevier, pp 89–118
Pratolongo P, Perillo GME, Píccolo MC (2010) Combined effects of waves and plants on a mud
deposition event at a mudflat-saltmarsh edge in the Bahía Blanca estuary. Estuar Coast Shelf
Sci 87:207–212
Pratolongo P, Mazzon C, Zapperi G et al (2013) Land cover changes in tidal salt marshes of the
Bahía Blanca estuary (Argentina) during the past 40 years. Estuar Coast Shelf Sci 133:23–31
Pratolongo PD, Piovan MJ, Cuadrado DG (2016) Coastal Environments in the Bahía Blanca
Estuary, Argentina. In: Khan M, Boër B, Ȫzturk M et al (eds) Sabkha Ecosystems. Tasks for
Vegetation Science, vol 48. Springer, Cham, pp 205–224
Pratolongo P, Piovan MJ, Cuadrado DG et al (2017) Coastal landscape evolution on the western
margin of the Bahía Blanca Estuary (Argentina) mirrors a non-uniform sea-level fall after the
mid-Holocene highstand. Geo-Mar Lett 37:373–384
Pratolongo P, Leonardi N, Kirby JR et al (2019) Temperate coastal wetlands: morphology, sedi-
ment processes, and plant communities. In: Perillo GME, Wolanski E, Cahoon DR, Hopkinson
CS (eds) Coastal Wetlands, 2nd edn. Elsevier, pp 105–152
Raybould AF (1997) The history and ecology of Spartina anglica in Poole Harbour. Proc. Dorset
Nat. Hist. Archaeol. Soc. 119:147–158
Redelstein R, Dinter T, Hertel D et al (2018) Effects of inundation, nutrient availability and plant
species diversity on fine root mass and morphology across a saltmarsh flooding gradient. Front
Plant Sci 9. https://doi.org/10.3389/fpls.2018.00098
Rogel JA, Ariza FA, Silla RO (2000) Soil salinity and moisture gradients and plant zonation in
Mediterranean salt marshes of Southeast Spain. Wetlands 20:357–372
Salem HB, Norman HC, Nefzaoui A et al (2010) Potential use of oldman saltbush (Atriplex num-
mularia Lindl.) in sheep and goat feeding. Small Ruminant Res 91:13–28
16  Coastal Wetlands of the Bahía Blanca Estuary: Landscape Structure and Plant… 467

Saltonstall K (2002) Cryptic invasion by a non-native genotype of the common reed, Phragmites
australis, into North America. Proc. Natl. Acad. Sci. 99:2445–2449
Schrire BD, Lavin M, Lewis GP (2005) Global distribution patterns of the Leguminosae: insights
from recent phylogenies. Biol Skrif 55:375–422
Schwindt E, Battini N, Giachetti C et al (2018) Especies exóticas marino-costeras (Argentina).
Vázquez-Mazzini, Buenos Aires
Silva RA, López de Ruiz RE, Ruiz SO (2004) Estudio fitoquímico de flores de Geoffroea decorti-
cans (Gill. ex Hook. et Arm.) Burk, Leguminoseae (Fabaceae). Acta Farmacéutica Bonaerense
23:524–526
Silvestri S, Defina A, Marani M (2005) Tidal regime, salinity and salt marsh plant zonation. Estuar
Coast Shelf S 62:119–130
Simberloff D (2010) Invasions of Plant Communities. More of the Same, Something Very
Different, or Both? The Am Mid Nat 163:220–233
Simenstad C, Thom R (1995) Spartina alterniflora (smooth cordgrass) as an invasive halophyte in
Pacific northwest estuaries. Hortus Northwest 6:9–13
Simões MP, Calado ML, Madeira M et al (2011) Decomposition and nutrient release in halophytes
of a Mediterranean salt marsh. Aquat. Bot. 94:119–126
Soil-Survey-Staff (1975) Soil taxonomy: a basic system of soil classification for making
and interpreting soil surveys, Agricultural handbook no. 436. Soil Conservation Service,
U.S. Department of Agriculture, U.S. Government Printing Office, Washington, DC
Soriano A (1947) Las Quenopodiáceas de la Tribu “Salicornieae” en la República Argentina. Rev
Argentina de Agronomía 14:148–172
Sosa A, Fusco MR, Rossomando P et al (2011) Anti-inflammatory properties from isolated com-
pounds of Cyclolepis genistoides. Pharm biol 49:675–678
Spagnuolo JO (2005) Evolución geológica de la región costera-marina de Punta Alta, provincia de
Buenos Aires. PhD Thesis. Universidad Nacional del Sur, Bahía Blanca, Argentina
Steffen S, Mucina L, Kadereit G (2010) A revision of Sarcocornia in South Africa, Namibia and
Mozambique. Syst Bot 35:390–408
Steffen S, Ball P, Mucina L et al (2015) Phylogeny, biogeography and ecological diversification of
Sarcocornia (Salicornioideae, Amaranthaceae). Ann Bot-London 15:353–368
Stevens PF (2001 onwards). Angiosperm Phylogeny Website. Version 14, July 2017. http://www.
mobot.org/MOBOT/research/APweb/. Acceded Diciembre 2019
Strong DR, Ayres DA (2016) Control and consequences of Spartina spp. invasions with focus upon
San Francisco Bay. Biol Invasions 18:2237–2246
Tangahu BV, Abdullah S, Rozaimah S et al (2011) A review on heavy metals (As, Pb, and Hg)
uptake by plants through phytoremediation. International Journal of Chemical Engineering.
https://doi.org/10.1155/2011/939161
Theuerkauf EJ, Stephens JD, Ridge JT et al (2015) Carbon export from fringing saltmarsh shore-
line erosion overwhelms carbon storage across a critical width threshold. Estuar Coast Shelf
S 164:367–378
Tiner R (1999) Wetland Indicators; A Guide to Wetland Identification, Delineation, Classification,
and Mapping. Lewis Publishers/CRC Press, Boca Raton, USA
Tong C, Zhang L, Wang W et al (2011) Contrasting nutrient stocks and litter decomposition in stands
of native and invasive species in a sub-tropical estuarine marsh. Environ. Res. 111:909–916
Traut BH (2005) The role of coastal ecotones: a case study of the salt marsh/upland transition zone
in California. J Ecol 93:279–290
Valéry L, Fritz H, Lefeuvre JC, Simberloff D (2009) Invasive species can also be native…. Trends
Ecol Evol 24:585
Valéry L, Radureau A, Lefeuvre JC (2016) Spread of the native grass Elymus athericus in salt
marshes of Mont-Saint-Michel bay as an unusual case of coastal eutrophication. J Coast
Conserv 21:421–433
Vartiainen T (1988) Vegetation development on the outer islands of the Bothnian Bay. Vegetatio
77:149–158
468 P. D. Pratolongo et al.

Ventura Y, Wuddineh WA, Myrzabayeva M et al (2011) Effect of seawater concentration on the
productivity and nutritional value of annual Salicornia and perennial Sarcocornia halophytes
as leafy vegetable crops. Sci Hortic 128:189–196
Vepraskas MJ (1996) Redoximorphic features for identifying aquic conditions. North Carolina
Agr. Res. Serv. Tech. Bull. 301. North Carolina State University, Raleigh, NC
Verettoni HN (1974) Las comunidades vegetales de la región de Bahía Blanca. UNS, Bahía Blanca
Violante RA, Parker G (2000) El Holoceno en las regiones marinas y costeras del nordeste de
Buenos Aires. Rev la Asoc Geológica Argentina 55:337–351
Wahid A (2003) Physiological significance of morpho-anatomical features of halophytes with par-
ticular reference to Cholistan flora. Int J Agric Biol 5:207–212
Waller MP, Long AJ, Long D et al (1999) Patterns and processes in the development of coastal
more vegetation: multi-site investigations from Walland Marsh, Southeast England. Quat Sci
Rev 18:1419–1444
Wang Q, Wang CH, Zhao B et  al (2006) Effects of growing conditions on the growth of and
interactions between salt marsh plants: implications for invisibility of habitats. Biol Invasions
8:1547–1560
Wang Y, Zhou L, Zheng X et al (2013) Influence of Spartina alterniflora on the mobility of heavy
metals in salt marsh sediments of the Yangtze River Estuary, China. Environ Sci Pollut R
20:1675–1685
Whigham DF (1999) Ecological issues related to wetland preservation, restoration, creation and
assessment. J Total Environ 240:31–40
Willis B (1914) Northern Patagonia, character and resources. Scribner Press, New York, USA
Zapperi G, Pratolongo P, Piovan MJ et al (2016) Benthic-Pelagic Coupling in an Intertidal Mudflat
in the Bahía Blanca Estuary (SW Atlantic). J Coast Res 32:629–637
Zedler JB (1982) The ecology of southern California coastal salt marshes: a community profile.
Fish and Wildlife Service, Washington
Zobel M, Kont A (1992) Formation and succession of alvar communities in the Baltic land uplift
area. Nord J Bot 12:249–256
Zuloaga FO, Belgrano MJ, Zanotti CA (2019) Actualización del Catálogo de las Plantas Vasculares
del Cono Sur. Darwiniana, nueva serie 7:208–278
Chapter 17
Environmental Diagnosis of the Protected
Coastal Areas of the Bahía Blanca Estuary

M. Elizabeth Carbone, María Ángeles Speake, and Walter Daniel Melo

17.1  Introduction

The interrelation between natural environments and the various human activities in
coastal areas generates conflicts between conservation and development. For this
reason, the compatibility of both concepts is very important in areas with high ter-
ritorial occupation pressure and presence of vulnerable natural resources.
Recreational and productive activities are currently being developed in coastal
areas, which directly or indirectly affect natural resources. Human pressure on natu-
ral environments is imposed by space demand for new settlements, installation of
infrastructure and equipment, sand extraction, and commercial fishing among oth-
ers. These human activities are the main stressors on coastal ecosystems (Charlier
and Bologa 2003). Sea level variations, erosion, and sedimentation are major natu-
ral processes also involved in the present evolution of coastal areas. All these trans-
formations can damage coastal environments when they are not accompanied by an
adequate system of orderly and sustainable management guidelines. The formula-
tion and implementation of appropriate coastal planning and management measures
arise from a concrete current diagnosis that can project future changes and modifi-
cations in this kind of environments (Maelfait et al. 2006). Coastal protected areas
contribute to the conservation of the natural and cultural heritage of a region and
reduce pressures caused by human activities on these environments. The proposal
and implementation of programs and policies for a sustainable coastal management
of these areas are strengthened with an adequate monitoring of the environmental

M. E. Carbone () · M. Á. Speake · W. D. Melo


Instituto Argentino de Oceanografía IADO (Universidad Nacional del Sur-CONICET),
Bahía Blanca, Argentina
Departamento de Geografía y Turismo, Universidad Nacional del Sur,
Bahía Blanca, Argentina
e-mail: ecarbone@criba.edu.ar

© Springer Nature Switzerland AG 2021 469


S. M. Fiori, P. D. Pratolongo (eds.), The Bahía Blanca Estuary,
https://doi.org/10.1007/978-3-030-66486-2_17
470 M. E. Carbone et al.

quality of the site and the integrated knowledge of the current processes that occur
in these environments.
The coastal areas of Latin America encompass a large variety of geo-­
environmental units that include sandy beaches, estuaries, mangroves, coral reefs,
marshes, coastal lagoons, islands, and fjords. Likewise, the sea that bathes its coasts
is very important in terms of biodiversity and productivity of the regional ecosys-
tems. Coastal waters also hold a high productivity, which supports one of the five
major commercial fisheries worldwide (Food and Agricultural Organization 2012).
The coastal and marine areas of Latin America showed for more than 20 years the
effects caused by human settlements. Urban expansion implies a significant change
in land uses, especially on coastal ecosystems (Barragán Muñoz 2014). The 60% of
the population is concentrated less than 100 km from the coast, and this behavior
induces overexploitation of marine resources (Food and Agricultural
Organization 2014).
Recreational activities and tourism also exacerbate impacts on coastal areas,
along with uncontrolled waste discharge into the oceans, and the expansion of aqua-
culture. Non-compliance with the objectives of creating coastal protected areas is
another negative impact on these coastal areas (Aldana Mazorra and Hernández
Zanuy 2018). The most degraded coastal ecosystems in Latin America are coastal
wetlands and coral reefs, which implies the loss of valuable ecosystem services
essential to many economies in the region. These coastal habitats also play an
important role of coastal protection and shoreline stabilization, in the face of
extreme meteorological phenomena, occurring at increasing frequency and with
greater intensity (Food and Agricultural Organization 2012). The decreasing abun-
dance of fishing resources is noticeable at the local scale, and also regional extinc-
tions of some commercial fish species were reported, along with profound changes
in the structure and functioning of ecosystems. The current policies of coastal ter-
ritorial planning through the creation of protected areas, as an instrument of control
and regulation of the use of the coastline, are no longer sufficient to address the
complexity and magnitude of the problems that arise in the coastal ecosystems of
different countries of Latin America (Day et al. 2012; Barragán Muñoz 2014).
Marine protected areas in Latin American countries are the product of a rela-
tively recent conservation management. Although there are experiences since the
1970s of the twentieth century. Until 2011, more than 700 marine and coastal pro-
tected areas have been established in Latin America covering 250,000 km2 or 1.5%
of its coastal waters. These numbers reflect the fact that the region has a significant
delay in meeting the goal of the working program on protected areas of the
Convention on Biological Diversity that recommends conserving at least 10% of the
ocean surface (Aldana Mazorra and Hernández Zanuy 2018). Latin America has
made considerable progress in aspects related to the protection of its marine and
coastal biodiversity. It is possible to verify that all the countries of the region have
protected coastal areas, although not all of them are structured in the same way. The
problems that are noticed in these geographical spaces are the low level of imple-
mentation of protected areas in the marine environment in relation to the continental
sector, less knowledge of the adjacent marine environment, high costs for the
17  Environmental Diagnosis of the Protected Coastal Areas of the Bahía Blanca Estuary 471

management of coastal areas, insufficient and not specialized staff, and budget
restrictions (Day et al. 2012; Barragán Muñoz 2014; Aldana Mazorra and Hernández
Zanuy 2018).
Latin American countries such as Argentina and Uruguay host some of the most
biodiverse areas in the world and intensify the protection of their coastal seas and
the surrounding oceans. It is done by creating coastal protected areas and by extend-
ing the conservation objectives to adjacent waters. The initiative aligned with the
overall objective of safeguarding at least 10% of the marine and coastal areas world-
wide by 2020. Marine coastal protected areas (MPAs), extensions of water managed
for conservation, recently expanded and now cover 8.4% of the oceanic territory of
Latin America (Food and Agricultural Organization 2014).
The jurisdiction of the Bahía Blanca Estuary is complex (Fig. 2.1; Chap. 2),
covering areas of national, provincial, and municipal level of public and private
domain. Six sites are analyzed in this chapter: three of these geographical areas have
management categories: (1) Bahía Blanca, Bahía Falsa, and Bahía Verde Nature
Reserve; (2) Islote del Puerto Nature Reserve (both natural reserves are provin-
cials); and (3) Bahía Blanca Coastal Reserve (Fig. 17.1). The following sites are of
preservation and conservation interest due to their predominant ecological charac-
teristics: (4) Arroyo Parejas-Isla Cantarelli Natural Area, (5) Villa del Mar Wetlands,
and (6) the Puerto Cuatreros Wetland. Due to the ornithological importance of these
coastal areas in 2008, the Bahía Blanca, Bahía Falsa, and Bahía Verde Multiple
Use Nature Reserve was designated as a Bird Conservation Area (IBA) by BirdLife
International, and in 2016 the six sub-sites were designated Site of Regional
Importance by the Western Hemisphere Shorebird Reserve Network (Fig. 17.2).

17.2  G
 eo-environmental Characteristics of Coastal
Protected Areas

The Bahía Blanca Estuary is an extensive and complex system of natural marshes in
the Buenos Aires Province, Argentina (Fig. 2.1; Chap. 1). It is a coastal wetland
system with unique geographical and biological characteristics. It is a diverse and
very dynamic geographical space that has an important natural capital. The inter-
tidal ecotone is the spine of coastal zones and represents the area of greatest interest
in coastal area management (Barragán Muñoz 2014). The legalization of these geo-
graphical spaces through the creation of coastal protected areas is one of the most
versatile tools for the preservation and conservation of the natural physical system
of a coastal region.
The intertidal zone that characterizes the Bahía Blanca Estuary is the main object
of interest in terms of integrated management of natural coastal areas, besides the
islands and tidal channels. The jurisdiction of the ecosystems of the Bahía Blanca
Estuary is complicated by the different levels of administration that involve both the
local, provincial, and national level. It should be noted that in the coastal study area,
472 M. E. Carbone et al.

Fig. 17.1  Location of the natural reserves of the Bahía Blanca Estuary. In detail, the limits of the
Bahía Blanca Coastal Reserve and the Islote de la Gaviota Cangrejera Nature Reserve are observed.
The total area of the Bahía Blanca, Bahía Falsa, and Bahía Verde Multiple Use Nature Reserve
provincial is shown in Fig. 2.1; Chap. 2. (Modified from Speake et al. 2018)

some sites have been declared of international interest due to their importance for
migratory birds.
The geomorphological units that present the Bahía Blanca, Falsa, and Verde
Multiple Use Nature Reserve are made up of tidal channels, intertidal zones, and
islands. Elongated tidal channels have different dimensions, the smaller ones being
those at the west end of the Bahía Blanca Estuary head, and the predominant orien-
tation of them is northwest-southeast and represents 39% of the study area. The
drainage design is dendritic and subparallel and drains into the Principal Channel
(CP) which is 97 km long in total. In the southern sector of this area, the design is
17  Environmental Diagnosis of the Protected Coastal Areas of the Bahía Blanca Estuary 473

Fig. 17.2  Areas of ecological interest in the Bahía Blanca Estuary, location of the Western
Hemisphere Shorebird Reserve Network sub-sites and nature reserves

anastomosed presenting various meanders, which separate plains and islands


(Carbone et al. 2016).
The Bahía Blanca, Bahía Falsa, and Bahía Verde Nature Reserve has islands,
tidal channels, and intertidal zones represented by silt-clay tidal flats and vegetated
marshes (Fig. 2.1; Chap. 2). The intertidal environments occupy 67.3% of the area,
with tidal plains being the predominant ones with 38.1%, and the marshes of
Spartina alterniflora and Sarcocornia perennis occupy 10.2% and 19%, respec-
tively, while the islands and tidal channels of various dimensions occupy 32.7% of
the reserve.
The islands are made up of two different highly fragmented morphological
groups, and these occupy the outermost sector of the Bahía Blanca Estuary. In the
north, the Zuraita, Bermejo, and Trinidad islands are the largest, while in the south,
the Monte and Ariadna islands are smaller, characterized by being more compact
than those located to the north of the reserve (Melo 2007). These geo-environmental
units have very complex dynamics. Inside these there are abandoned tidal channels,
while in the southeast sector of Trinidad Island, a tidal plain develops with more
than 10 km of development with a predominance of silty clay silt.
The southern islands have tidal flats with a high percentage of fine sand (Melo
2007). The current coastal processes generated the growth of geoforms, which give
protection to these environments with intense dynamics.
The exposed surfaces are covered with scrubland typical of the shrub steppe and
halophilic scrubs (Angeles 2001). The phytogeographic provinces the Espinal and
the Monte predominate. The vegetation is the xerophilous and halophilic shrub
steppe and grasslands. The vegetation that can be found on the islands is dominated
by argentine Spartina densiflora, Atriplex undulata, and perennial glasswort
474 M. E. Carbone et al.

Sarcocornia perennis, which do not usually exceed the meter. This environment
represents the breeding and feeding habitat of the crab gull and the breeding of
Larus atlanticus, Larus dominicanus, Egretta ardesiaca, and Ardea alba (Petracci
and Sotelo 2013).
The nature reserve was created on March 21, 1991, and then regulated as a
Multiple Uses Natural Reserve by Law N° 11.074, ratified latter in 1998 by
Provincial Law N°. 12.101. The reserve covers an area of 30.000 ha of land and
180.000 ha of water. The reserve has a basic implementation of regularization and
control of permitted activities compatible with the conservation of the area and its
integration into the management plan in force since 2007. As part of the implemen-
tation, the reserve has a park ranger and infrastructure service, with a service office,
checkpoints and surveillance in island sectors, a motor vehicle, semi-rigid boat for
nautical control, and radio communication means (Massola and Cinti 2012).
The Islote de  la Gaviota Cangrejera Nature Reserve (115  ha) protects a very
large breeding colony of Larus atlanticus, with 3800 nests censored in 2005 and
with similar population values recorded for 2006 (Petracci et al. 2008). This species
is in risk of extinction, and it is only found on the Southwestern Atlantic coast
(BirdLife International 2019). For this reason, the purpose of this reserve is the
conservation and protection of this colony. It was declared of provincial interest by
Decree 469/11. This species in Argentina is considered “Threatened” by Resolution
N° 348/10 of the Environment and Sustainable Development Ministry of the Nation
(Petracci and Sotelo 2013). In Bahía Blanca, it was also declared as “Emblematic
Species” by Ordinance N°. 12671/04. The greatest impacts on this species are
caused by the reduction of the nesting and breeding area, the dumping of hydrocar-
bons in the Principal Channel, illegal subtraction of eggs and pigeons, and alteration
of the surrounding island environment through urban and sewage wastes thrown
into the sea (Sotelo and Mazzola 2008; Petracci and Sotelo 2013; Carbone et al.
2016). The species is also currently protected by the International Convention on
Migratory Species (CMS) through Law N° 23918/91.
It should be noted that due to the dredging activities to deepen the Principal
Channel of the Bahía Blanca Estuary, these gulls stopped nesting in 1989 and
returned in 1999, in the sectors formed by the dredged material. Due to its impor-
tance for the conservation of the species, the Government of the Buenos Aires
Province declared of “Provincial Interest” the conservation and protection activities
carried out in the Islote del Puerto Nature Reserve y Resolution N°. 04/05 (Mazzola
and Cinti 2012; Petracci and Sotelo 2013).
The Bahía Blanca Coastal Reserve (Ordinance No 13.892/2006) has an area of
approximately 320 ha. The main objective is the conservation and protection of the
coastal environment, and environmental research and education are allowed. The
typical geomorphological units of this reserve belong to a marine-continental transi-
tion environment, with plains and tidal channels, some of the plains covered by
halophilic vegetation (Angeles 2001; Nebbia and Zalba 2007). The fauna stands out
for the presence of crabs in the tidal plains and the prairies of Spartina densiflora.
Petracci and Sotelo (2013) in the nearby marine sector recorded the presence of
adult and juvenile individuals with Thalassarche melanophris. In rural areas
17  Environmental Diagnosis of the Protected Coastal Areas of the Bahía Blanca Estuary 475

peripheral to the reserve, they observed flocks of Sturnella defilippii. Also in the
vicinity of the area, the Porzana spiloptera has been registered.

17.3  H
 uman Occupation of the Coastal Areas of the Bahía
Blanca Estuary Before Being Reserves

The islands that make up the Bahía Blanca, Bahía Falsa, and Bahía Verde Multiple
Use Natural Reserve were used by man since 1890 through productive and extrac-
tive activities. The occupation of the islands was generated by taking advantage of
the natural resources that they generated, such as the exploitation of Geoffroea
decorticans that were on the islands and constituted a source of cheap and abundant
energy for those boats that roamed the area, activity regulated by the State, impos-
ing the payment of 10% of the profits obtained (Amarfil 2006).
From the beginning of the twentieth century on the islands, facilities were built,
which changed their original ecosystem of herbaceous vegetation with the introduc-
tion of sheep, goats, cows, and horses.  Isla Verde was olso exploited was also
exploited with sheep farming reaching a production of up to 5000 sheep, by the year
1852 also in Bermejo Island and Trinidad (Amarfil 2006), and there are currently
remains of these past activities. From 1944, the Trinidad Island was in private
domain, and the same families administered the Bermejo Island, dedicating mainly
to the fishing of Galeorhinus galeus. Later they installed gouges, mills, and pens, to
raise sheep and grow barley. In 1961, the concession to the first settlers ceased; on
that occasion, the new tenants were engaged in the breeding of goats and sheep
(Cinti 2017). The administration of these islands depended on the agency of Fiscal
Lands of La Plata (Buenos Aires Province) and then passed to the Ministry of
Agrarian Affairs, which declared the protected area that included these islands, in
1998. Livestock activities continued in those environments through the lease of land
to those who inhabited the islands. During 2007, the Provincial Organization for
Sustainable Development is created, which currently manages this coastal nature
reserve.
The Ariadna, Monte, and Verde islands have sandy beaches, low tidal plains, and
cliffs of reddish-colored siltstones with herbaceous steppe and halophilic scrub-
lands, subsequently replaced by agricultural crops (Melo 2007). Ariadna Island has
an inn where overnight is possible. Some of the islets near this area have been colo-
nized by Larus atlanticus and Larus dominicanus, registering up to 300 nests of the
former (Petracci and Sotelo 2013). The permitted uses of the reserve include scien-
tific research activities, inter-island walks, guided tours of islands, and interpretive
trails where sightings of marine and terrestrial fauna are made. In addition to activi-
ties related to tourism and recreation, artisanal and sport fishing is carried out, which
must be controlled.
The Bahía Blanca Coastal Reserve is located on municipal land, which from
1950 to 2006 was used as a solid waste deposit in the open. The usual practice
476 M. E. Carbone et al.

consisted of the massive deposit of waste, both organic and inorganic, and their
burning. Despite this situation and facing the need for more recreational spaces on
the city’s waterfront, in 1978, the location of a new urban park in this sector called
Almirante Brown Maritime Park was decided (Speake and Carbone 2017).

17.4  Environmental Quality Index (EQI)

The results presented in this section are part of an investigation work where the
protected areas of the southwest of Buenos Aires Province and the north of the
Patagonian region are evaluated to compare the naturalness and state of the different
geo-environments of the different reserves present throughout these coastal zones.
The methodological procedure used to determine the EQI for the natural reserves of
the Bahía Blanca Estuary was carried out through 36 indicators for the year 2018,
obtained from:
(a) Topographic charts scale 1:100,000, satellite image charts 1:250,000, aerial
photographs scale 1:20,000, and Landsat satellite images 5 and 7 processed
(b) Institutional environmental monitoring data
(c) Specific bibliographic background
(d) Statistical information from the 2010 Census of Home and Housing, Agricultural
Census 2004, Meteorological Statistics 2010–2015, Information from the
Municipality of Bahía Blanca, and the Provincial Organization for Sustainable
Development
The DPSIR model (Organiztion for Economic Cooperation and Development
1993; EEA 1998) is applied, which provides a framework for organizing informa-
tion to structure the indicators. It implies elaborating in a general way a causal
progression of human actions that cause pressure on the environment and natural
resources, which lead to a change in the state of the environment and to which soci-
ety responds with measures or actions to reduce or prevent the impact. The degree
of naturalness (NI) of a region, the absence of modifications introduced by man, the
role of the physical environment can be conceived as a source of various resources
(RI), as a sink for different types of waste (WI) generated for human activities (SSI)
and as support for these activities (Organization for Economic  Cooperation and
Development 1993). During the second stage, the data of the selected environmen-
tal indicators were normalized, and a nominal scale of 0 to 1 was assigned, where 0
corresponds to the worst possible situation with respect to the indicator in question
and 1 to the best possible situation with respect to the indicator analyzed according
to the following expression:

V   Im  Imin  /  Imax  Imin  (17.1)

where V is the normalized value; Im, value of the indicator; Imax, maximum
value in the study area; and Imin, minimum value in the zone. The third stage
17  Environmental Diagnosis of the Protected Coastal Areas of the Bahía Blanca Estuary 477

consists in the calculation of the aggregate indices of pressure, state, and response
(in turn, indicators of motive forces and impacts are included within them) for each
of the four functions mentioned according to:

NIp   Vi·Wi (17.2)

where NIp is the pressure index on the natural condition of the environment, Vi
is the standardized value of the indicator, and Wi is the weight of the indicator; i
= (1 − n), where n is the total number of pressure indicators. Similarly, the calcula-
tion of status and response indices is carried out, both for naturalness and for the
source, sink, and activities-services functions. The pressure, status, and response
indices are integrated into indices of each of the functions:

NI   NIp  NIs  NIr  / 3 (17.3)

where NI is the index of the naturalness function, NIp the index of pressure on
naturalness, NIs the index of state of naturalness, and NIr the index of response on
naturalness.
Proceeding also for each of the functions addressed, RI is the index of the
resource source function; WI, index of the waste sink function; and SSI, index of the
activities-services function. Finally, the environmental quality index (EQI) obtained
from the following expression is calculated:

EQI   NI  RI  WI  SSI  / 4 (17.4)

where IN is the naturalness index, RI the resource source index, WI the waste
sink index, and SSI the index of activities/services. The interpretation of the final
value of the environmental quality index is expressed on a scale of five classes
whose maximum value is one and the minimum zero, the highest values corre-
sponding to the most optimal environmental situation.
The different indices obtained from the indicators in the complete list present
significant differences between the different coastal protected areas. It should be
noted that each of them has an administrative legal jurisdiction of different levels,
namely, the Bahía Blanca, Bahía Falsa, and Bahía Verde Multiple Use Natural
Reserve and the Islote de la Gaviota Cangrejera Nature Reserve belong to the pro-
vincial scope, while the Bahía Blanca Coastal Nature Reserve is under the munici-
pal domain. This characteristic is stated as the response indicators correspond to
government actions with different criteria for the coastal protected areas (CPA)
mentioned.
The values correspond to the interaction of human activities with the physical
environment in each of the geographical spaces analyzed in the Bahía Blanca
Estuary area and that interrelation in each of the geo-environmental units of each of
the reserves also refers to the function and execution of the management guidelines
through the management plans that regulate each of the reserves. IN values refer to
the natural and pristine state of the ecosystems involved, being 0.27 for the Bahía
Blanca Coastal Reserve, while the value of the Islote del Puerto Nature Reserve is
478 M. E. Carbone et al.

0.20, and the minimum corresponds to the first reserve with a value of 0.18. The
presence of the large intertidal extension in the first mentioned reserve and the dif-
ficulty in accessing and connecting to it are reflected in the maximum obtained with
respect to the NI. The values of the NIs and NIr for the same reached values of 0.33
and 0.39 representing the maximums for this index in the analyzed areas. The pres-
sure and response dimensions obtained their maximum and minimum values in the
Bahía Blanca Coastal Reserve and Islote de la Gaviota Cangrejera Nature Reserve,
respectively.
The minimums with respect to the response obtained for the coastal reserve can
be analyzed from the complex situation that has characterized this sector that appro-
priate sanitation measures have not yet been taken since it is still used as a solid
waste deposit (Speake and Carbone 2017). The natural environments of this area are
vulnerable to this type of anthropic intervention, which is reflected in the decrease
in the ecological condition of these environments.
In the Islote de la Gaviota Cangrejera Nature Reserve, the naturalness index has
the intermediate value with 0.19 composed of the NIp and NIr values of 0.17 and
0.23, respectively. It is worth mentioning that this sector was used as a deposit of
dredging material after the deepening of the Principal Channel between 1989 and
1992. On that occasion, 50  million cubic meters of material were extracted that
were deposited on the island that today in this reserve.
The minimum values of the IN for the Bahía Blanca Coastal Reserve ranged
between 0.13 and 0.15 for the status and response indicators, respectively. In the
three coastal areas, the highest value of the index corresponds to the state of the
ecosystems involved in two environments, being 0.39 for the Bahía Blanca, Bahía
Falsa, and Bahía Verde Multiple Use Nature Reserve, 0.25 for the Islote del Puerto
Nature Reserve, and 0.15 for the Bahía Blanca Coastal Reserve. The lower values
of the naturalness index in terms of the state dimension reflect the loss of the natural
condition of these environments, mainly associated with land uses in areas adjacent
to the reserves. The productive and industrial economic activities are located in the
north sector of the Principal Channel. The response index values are closely related
to the control and control measures of productive and industrial activities near natu-
ral reserves. These measures are particularly difficult in the Bahía Blanca, Bahía
Falsa, and Bahía Verde Multiple Use Nature Reserve given the extension and variety
of geo-environmental units that integrate it as detailed in the previous section.
The NI values indicate the intermediate nature condition considering that the
natural coastal protected areas in this sector do not have a marked demographic
pressure and extractive activities are prohibited, as are the old productive activities.
It should be noted that in the places where these practices were currently allowed,
the cattle that live on the Bahía Blanca, Bahía Falsa, and Bahía Verde Multiple Use
Nature Reserve islands are beginning to be extracted.
The index that involves the environment, as a source of IR resources, presents the
highest value for the Bahía Blanca Coastal Reserve, where the sector with the great-
est presence of pre-existing human activities with a value of 0.36 for the pressure
variable is concentrated. The lowest data corresponds to the state indicators with
0.12 referring to the diversity of land uses of this reserve and adjacent area
17  Environmental Diagnosis of the Protected Coastal Areas of the Bahía Blanca Estuary 479

(agricultural activities, fishing). One of the most important objectives in the man-
agement plan is to make the concept of extracting the mentioned resources compat-
ible with conservation and sustainability. The highest values of the RI are evidenced
in the Bahía Blanca, Bahía Falsa, and Bahía Verde Nature Reserve in the response
category, the fishing activities are regulated and controlled, and the species of com-
mercial interest are the Micropogonias furnieri and Artemesia longinaris.
Currently, the fishing activity has suffered a significant decrease; the main causes
are associated with the increase in port activity, maintenance and dredging of the
ports, sewage discharges, and the overexploitation of the resource (Conde et  al.
2009; López Cazorla et al. 2014). The status and response indicators ranged from
0.38 to 0.09 due to the increase in the importance of conservation management in
this area. The highest values were observed in the Bahía Blanca, Bahía Falsa, and
Bahía Verde Multiple Use Nature Reserve, while the minimums are observed in the
Bahía Blanca Coastal Reserve; IR values range between 0.35 and 0.19, respectively.
The dimensions that reached intermediate values correspond to the pressure and
response indicators with values ranging between 0.46 and 0.36.
The analysis of this index for the sectors adjacent to the urbanizations also
revealed lower numbers with respect to the state of the same, throwing values of
0.16 and 0.22; it should be noted that the reserves of the Islote de la Gaviota
Cangrejera Nature Reserve and the Bahía Blanca Coastal Reserve were located near
the industrial zone and urban waste discharges, respectively. The indices with the
highest value correspond to those of pressure with a maximum of up to 0.38 the
reserve of multiple uses. It should be noted that the islands that make up the Bahía
Blanca, Bahía Falsa, and Bahía Verde Multiple Use Nature Reserve before being
declared protected areas were provided for agricultural practices by provincial orga-
nizations; today, there are vestiges of these activities as part of the infrastructure and
equipment (Angeles 2001; Massola and Cinti 2012).
The waste index that considers the environment as a waste disposal IW, the maxi-
mum value corresponds to the Bahía Blanca Coastal Reserve; it should be noted that
in this environment it has not yet been completely remedied and household waste
continues to be deposited in the study area. The dimensions that formed this index
yielded higher values corresponding to pressure and response reaching values of for
the second variable mentioned.
The variables that stand out mainly in this index correspond to the pressure and
response, with values ranging from for the first reserve and from 0.35 to 0.20 for
Islote de la Gaviota Cangrejera Nature Reserve and Bahía Blanca Coastal Reserve.
The state variable ranged between 0.22 and 0.18, respectively (Fig. 17.3), maintain-
ing a lower range of values which represents approaching an optimal situation
according to the scale considered. With respect to places adjacent to the analyzed
reserves, although there are no human settlements inside the perimeter of the
reserves, it is worth mentioning that the proximity to the city of Bahía Blanca with
more than 300,000 inhabitants (Indec 2010), the industrial port area, the urban and
industrial effluent discharge sites, and waste treatment plants act as direct driving
forces and impact their natural environment (Spetter et al. 2019).
480 M. E. Carbone et al.

a IP NR BB CR BBFV NR b IP NR BB CR BBFV NR

NIr RIr

NIs RIs

NIp RIp

0 0.2 0.4 0.6 0 0,1 0,2 0,3 0,4 0,5


c d
WIr SSIr

WIs SSIs

WIp SSIp

0 0,1 0,2 0,3 0,4 0 0,2 0,4 0,6

Fig. 17.3 (a) Naturalness indices, (b) response, (c) waste sink and activities, and (d) services for
the Bahía Blanca Estuary of Islote de la Gaviota Cangrejera Nature Reserve (IP NR), Bahía Blanca
Coastal Reserve (BB CR), and Bahía Blanca, Bahía Falsa, and Bahía Verde Multiple Use Nature
Reserve (BBFV NR)

The capacity of the medium as a support for activities and services through the
ISS presented a significant variability of values mainly among coastal reserves. In
this case, activities and services derived from tourism were considered considering
the restrictions on its compatibility with the legislation of the areas and management
plans. The main tourist activities that are carried out are excursions embarked to
Bermejo Island, to Puerto Cuatreros and tour along the seafront under the ecotour-
ism modality, with specialized guides.
In order to analyze this index, the relation of the tourist activities with the state
of the environments that make up the areas of the coastal reserves is also considered,
and the degree of deterioration of the environments is an indicator of their status.
The Bahía Blanca Coastal Reserve currently presents a significant deterioration of
the ecological system whose main causes are clandestine final disposal of urban
solid waste generated by the inhabitants of the city of Bahía Blanca, open burning
of waste, land and water pollution by leaching of the same, and presence of rodents
and predators of conservation species (Speake and Carbone 2017). Another problem
is the expansion of invasive alien species, such as Magallana gigas, which repre-
sents a threat to the conservation of the biodiversity of nature reserves exacerbated
by the absence of a management plan.
For the state variables, the better indices for the Bahía Blanca, Bahía Falsa, and
Bahía Verde Multiple Use Nature Reserve are observed with values ranging between
0.34 and 0.46, while for the municipal natural reserve, the index has much lower
17  Environmental Diagnosis of the Protected Coastal Areas of the Bahía Blanca Estuary 481

values ranging between 0.12 and 0.09. It should be noted that this reserve response
rate is 0.05 and it is related to what has been detailed above about the current envi-
ronmental deterioration processes and lack of effective planning. The maximum
value of the SSI is presented in the multipurpose reserve with a value of 0.38, and
the minimum value of 0.07 is of the municipal natural reserve, while the Islote de la
Gaviota Cangrejera Nature Reserve has a value of 0.15 (Fig. 17.3).
The aggregate indices for these areas were also obtained from simplified lists of
indicators. The results of the application of 12 previously selected indicators pre-
sented higher values with respect to those previously analyzed. The naturalness
index for the Bahía Blanca, Bahía Falsa, and Bahía Verde Multiple Use Nature
Reserve was 0.52. State  and response  are the predominant dimentions. For the
Islote de la Gaviota Cangrejera Nature reserve, it was 0.25 and 0.14 for the munici-
pal reserve (Fig. 17.3). Indices that involve the environment as a source of resources,
as waste sinks, and as support for different activities and services varied among the
three coastal reserves.
The values of the dimensions that constitute these indices in the three reserves
varied significantly, with pressure and status prevailing in the majority of the
response involved. This condition arises because the weight given to the response
indicators as to the other dimensions is 1 and considering that these sites depend on
different administrative units, where public investments are different for the three
reserves.
Obtaining disaggregated indices allowed zoning the study area according to the
EQI values in three classes. Each of these belongs and represents Class 1 to Class 3
(low, moderately low, and medium) of the study area. The highest value corresponds
to the Bahía Blanca, Bahía Falsa, and Bahía Verde Multiple Use Nature Reserve
with 0.39, followed by the Islote del Puerto reserve with a value of 0.21 and the
municipal reserve with a value of 0.09 which has the Class 1 that represents a low
environmental quality. These results arise from the interrelation of all the indices
analyzed above using the complete list of indicators where all the dimensions of the
DPSIR model were calculated.
The values obtained allowed the area to be zoned in the following way: the
reserve with the largest areal dimension presents values corresponding to Class 3
that represents an average environmental quality, which reflects a higher NI related
to the state of the geo-environmental units that comprise it. The pressure exerted on
the physical environment from the extraction of fishing resources, the discharge of
sewage effluents without treatment in the surrounding area, and the location of areas
of urban household waste deposits in the municipal coastal area alters and deterio-
rates the natural condition of the environments. The use of the response indicators
showed lower quality values that highlight the current concern to begin to counter-
act the pressures on the environment.
The different jurisdictional levels of administration of these reserves make it dif-
ficult to apply appropriate remediation and mitigation measures in this geographical
area. This is reflected in the state indices that marked a better condition of natural-
ness among the most external estuary reserves.
482 M. E. Carbone et al.

The coastal reserve has lower values of environmental quality, and they fit it into
the lower category. The indices obtained in this ecosystem show an environmental
problem accentuated by poor planning, the lack of a diagnosis of its current situa-
tion, and delay in the implementation of appropriate management guidelines. The
response indicators do not show the formulation and application of public policies
to mitigate impacts in the coastal zone.
The various geomorphological units that make up the reserves (tidal plains,
marshes, secondary tidal channels, salt, islands, etc.) have difficult accessibility
which restricts access to them. This condition also hinders the tasks of surveillance
and control of the protected areas analyzed; another problem with the reserves is the
limited staff and means available for patrolling them.
The continental sector adjacent to the multi-use reserve is characterized by the
concentration of livestock farming activities. Likewise, in each of the islands that
make it up, it is possible to differentiate conflict points where direct actions should
be implemented to restore, mitigate, conserve, and rehabilitate sectors vulnerable to
the pressures exerted regarding the presence of cattle and rabbits, mainly on the
Trinidad Island and Isla de los Conejos and Ariadna.
The integrated management guidelines that should be applied in this coastal zone
are related to ecosystem conservation, rehabilitation of areas used, and uses and
appropriation of renewable natural resources with low-impact techniques with effi-
cient and sustainable management of them.
The indices obtained through the indicators made possible the organization, sys-
tematization, quantification, and communication of the relative information of dif-
ferent aspects of this coastal zone, which is essential for making decisions about
environmental policies that will be applied in the short term. The variability of the
site, the infrastructure, the equipment, and the natural and anthropic processes that
act in each of the units considered allowed zoning the area according to the values
of the environmental index obtained.
The application of the environmental quality index determined that the Bahía
Blanca, Bahía Falsa, and Bahía Verde Multiple Use Nature Reserve has an average
environmental quality according to the results obtained from the aggregated indices
(NI, RI, WI, SSI). The Islote del Puerto reserve has the intermediate value (0.21),
while the municipal reserve has the lowest values, between 0.12 and 0.09. The most
optimal environmental quality was presented by the two reservations mentioned
above. Likewise, the three natural reserves acquire maximum value from the point
of view of ecological and geomorphological importance when evaluating the geo-­
environmental units that characterize them.
The interrelation of the DPSIR indicators and the various functions of the
medium allowed the zoning obtained from the EQI where it was determined which
are the sectors in this area that need the rapid implementation of strategic manage-
ment guidelines. The values indicate a high antthropic pressure on the natural qual-
ity of these ecosystems, generating an overexploitation of the physical environment,
pollution of the same and in turn the low application of environmental and urban
policies in the near past could be inferred. According to the response rate, some
indicators mark a change in the formulation and application of sustainable measures
17  Environmental Diagnosis of the Protected Coastal Areas of the Bahía Blanca Estuary 483

to be able to modify the present and past situations through different levels of
administration.
Continuous monitoring of the environmental condition of coastal areas is essen-
tial to guide the actions and policies to be implemented. In the Bahía Blanca Estuary
reserves, the application of the EQI allowed us to observe a vulnerable and environ-
mental risk area, but it also arises from this study that there is an incipient change in
conservation policy and adequate management of resources when the conservation
of these is included in the government agenda spaces. These types of studies must
be carried out continuously due to the dynamics of this coastal environment. The
application of appropriate environmental policies will perhaps be reflected in the
following assessment of the environmental quality of this coastal zone.

17.5  A
 reas of Ecological Interest in the Northern Sector
of the Bahía Blanca Estuary

The estuary was formally included in Western Hemisphere Shorebird Reserve


Network (WHSRN) during 2016 due to its importance for the conservation of
migratory shorebirds and its unique wetland character due to the ecosystem services
they provide to human beings (WHSRN 2016). The areas described below do not
have an official management category but are sites in the WHSRN and represent
areas of ecological interest and are being requested to be included in the protected
area system.
Puerto Cuatreros is located in both coast of the main estuary channel between
Puerto Galvan and the La Vidriera Salt Flat, has areas of tidal, island, and shrubland
plains (Fig. 2.1, Chap. 2). The objectives set are intended to preserve the structure,
composition, and functioning of the estuarial ecosystem and promote the imple-
mentation of sustainable productive activities that allow a socially fair and environ-
mentally sustainable local and regional development (Langhoff 2015).
The sector was called a site of interest by the Western Hemisphere Shorebird
Reserve Network (WHSRN). The presence of certain species indicates the good
state of health of ecosystems. Birds migrate from Canada and the Arctic area and
can fly up to 11,000 km without descent. The birds feed and rest in the Bahia Blanca
wetlands, in Tierra del Fuego, and in different parts of Brazil (Petracci et al. 2008).
Different social actors of the Municipality of Bahía Blanca and Coronel Rosales,
of the Management Consortium of Puerto Rosales, of the OPDS, of the Yacht and
Fishing Club of Gral. Daniel Cerri, and of the Management Consortium of the Port
of Bahía Blanca participate in the management of the integral ecosystem to promote
not only the conservation of these birds but the ecosystem as a whole.
Villa del Mar the continental sector is dominated by the salt fats, and in the tidal
plains, the channels have a parallel design to the Principal Channel. In the internal
sectors of the district in general, the modelers of the coastal processes are the reflux
currents, while in the ambit the active processes are waves and coastal drift. The
484 M. E. Carbone et al.

presence of old bars and spikes was observed that reveal the previous marine activ-
ity in the sector.
Wetlands the predominant environments of this area are the tidal plains and the
marshes. Both cover approximately 47% in this sector. The tidal plains are areas that
have little slope and lack of vegetation. In turn, they are crossed by numerous tidal
channels that interconnect with each other and lead to larger channels. At low tide,
they have a greater air exposure than the plains, allowing the formation of plant
populations (mainly by the Spartina sp. and Sarcocornia sp.) tolerant to partial
immersion, anoxia, and edaphic hypersalinization (Cabrera 1971). Below the lower
limit of the vegetation, there are extensive tidal plains covering approximately
4.5 km2, and in some sectors, they are interrupted by rocky outcrops, as in the east-
ern sector of the wetland. Another geomorphological feature to highlight is the
underwater dunes in the bottoms in front of Villa del Mar, geoforms of mesomareal
estuaries (Melo 2007; Spagnuolo 2005).
In the Arroyo Parejas Wetland located south of the town of Punta Alta with an
area of 753 ha, the internal tidal plains belong to an old marine inlet, while the exter-
nal plains correspond to current sedimentation processes. The dominant water sup-
ply is the sewer discharge channel of the mentioned locality, which in its final route
is incorporated into the old tidal channel (Melo 2007). The Arroyo Parejas Wetland-­
Cantarelli Island has a marine coastal environment populated by Sarcocornia peren-
nis and Spartina alterniflora forming large marshes that are exposed at low tide,
where crab caves are found (Petracci and Sotelo 2013); in addition to these species,
this site is characterized by the presence of great biodiversity, with birds being noted
for their great abundance and diversity (Petracci et al. 2008). Every year there are
birds that arrive from the north of the American continent to rest and feed; some of
these species are threatened, and such is the case of the Calidris canutus that flies
from the Canadian Tundra to Tierra del Fuego, across the entire American continent
(Petracci et  al.  2008; Petracci  and Sotelo 2013). Hirundo rustica is another bird
commonly observed using tidal flats and the semi-industrial zone located in
Cantarelli Island to feed and nest (Larracoechea et  al. 2012). A project has been
presented to request to incorporate this natural environment into the provincial pro-
tected areas.

17.6  E
 valuation of the Management Effectiveness
of the Protected Coastal Areas of the Bahía
Blanca Estuary

The declaration of the three coastal protected areas in the Bahía Blanca Estuary has
been of utmost importance to conserve and protect the environments involved. In
order to make a comprehensive diagnosis of the management of the mentioned
coastal areas, the evaluation of their effectiveness was carried out, and in this sec-
tion, the partial results of this work are presented. The evaluation was carried out
17  Environmental Diagnosis of the Protected Coastal Areas of the Bahía Blanca Estuary 485

using the framework proposed by the International Union for Conservation of


Nature (IUCN) and the World Commission on Protected Areas (WCPA), adapted to
the particular conditions of the study area. From it, the five fundamental elements of
the management cycle were considered: context, planning, inputs, processes, and
results.
The Rapid Assessment Prioritization Protected Area Management evaluation
questionnaire was applied for each of the reservations. The objective is to evaluate
the management effectiveness of the coastal protected areas of the Bahía Blanca
Estuary region through variables and indicators grouped into five areas: human
resources infrastructure and equipment, financial resources: planning and political-­
Institutional referred to the legislation in the coastals protected areas.
The criteria and The indicators evaluated in these natural reserves were biologi-
cal importance, socioeconomic importance and vulnerability for the context vari-
able, for the planning variable the objectives, legal security and site design are
considered, for the inputs variable the resources are considered human and financial
assigned to each reserve, for the process variables are involved the management
plans, the operational plans and the monitoring of the results are evaluated through
the restoration, planning and management of the natural resources of the reserves.
With respect to the context indicators, the highest value is held by the multipur-
pose reserve with 42% (Fig. 17.4), and the one that acquires the lowest value is the
municipal reserve (24%), where the vulnerability of its ecosystem acquires greater
prominence (becomes more important since out door landfill has not been reused
(Speake et al. 2018).
All reserves have a high biological importance due to the presence of emblematic
species that represent part of the ecological criteria used for the foundation of their
creation. In the natural reserves of the estuary, sites of high value for the conserva-
tion of certain species have been identified; hence, there is an importance of the
indicators that make up the context variable.
The extension of the protected areas of the estuary is not adequate to maintain
natural processes at the landscape level; this variable is mainly intensified in the

Fig. 17.4 Management Context Planning Supplies Processes Results


effectiveness of Bahía 100
Blanca Estuary natural
reserves [Islote de la
80
Gaviota Cangrejera Nature
Reserve (IP NR), Bahía
Blanca Coastal Reserve 60
(BB CR), and Bahía
Blanca, Bahía Falsa, and
40
Bahía Verde Multiple
Use Nature Reserve
(BBFV NR)] (Modified 20
from Speake et al. 2018)
0
BBFVNR IPNR BBCR
486 M. E. Carbone et al.

Bahía Blanca, Bahía Falsa, and Bahía Verde Multiple Use Nature Reserve. The area
bordering the north of the Principal Channel must be preserved, and controls must
be carried out in port and industrial uses incompatible with the conservation objec-
tives of these areas. The main threats to these areas are urban effluent drains, indus-
trial discharges, oil spills, and untreated waste (Spetter et  al. 2015; Carbone
et al. 2016).
It should be noted that the Bahía Blanca, Bahía Falsa, and Bahía Verde Multiple
Use Nature Reserve created in 1991 has a management plan in force since 2007. It
was developed from inter- and multidisciplinary work with the participation of pub-
lic officials, representatives of private institutions, civil associations, and NGO
responding to a strategic planning process with public participation mechanisms,
has not been updated, and has partial approval. The smaller reserves, the Islote de la
Gaviota Cangrejera Nature Reserve and the Bahía Blanca Coastal Reserve, do not
have a management plan. In 2008, the Bahía Blanca, Bahía Falsa, and Bahía Verde
Multiple Use Nature Reserve was designated as an Important Area for Bird
Conservation (AICA) by BirdLife International, and in 2016, most of the Bahía
Blanca Estuary was listed as a “Regional Importance Site” by the Western
Hemisphere Shorebird Reserve Network (WHSRN), becoming the eighth site in
Argentina to obtain such recognition as mentioned above (Speake et al. 2018). In
the same vein, at the local level, the Municipality of Bahía Blanca declared a list of
“emblematic” species through ordinance 12.671/2004, including the Larus atlanti-
cus, Pontoporia blainvillei, Carcharhinus brachyurus, and Charadriidae, while the
Municipality of Coronel Rosales declared a list of endemic species of the region by
Decree 3408/2013, becoming one of the first municipalities in the country to protect
sea turtles (Petracci et al. 2008; Matamala 2013).
The indicators where the variable inputs of the protected areas referred to the
human and economic resources assigned to each of the Bahía Blanca Estuary natu-
ral reserves were evaluated varied between 7.3 and 4.1% for the multiple use nature
reserve and Islote de la Gaviota Cangrejera Nature Reserve. The number of assigned
park rangers is scarce for the two areas mentioned and null in the case of the munici-
pal reserve. Administrative and dissemination tasks are carried out in collaboration
with non-governmental organizations. The areal dimension of the reserves makes it
difficult to monitor and control these environments by assigned personnel. The
media and infrastructure are insufficient. Likewise, the financial resources allocated
are also insufficient.
Regarding the planning variables, these include objectives and design of the area
and cover the legal aspects, and the values obtained varied between 24 and 7%. The
indicators of the policies applied show little follow-up of the objectives and opera-
tional plans in the estuary reserves, although they are present in the Bahía Blanca
Coastal Reserve, given that it has more staff and resources allocated. The municipal
reserve currently has no resources or monitoring of operational plans since it does
not have a management plan either. Nor is a complete inventory of the biological
biodiversity of the entire region. There is only one list of species that were surveyed
by international organizations that evaluated Bahía Blanca Estuary to be included as
a site of regional interest in the WHSRN and a list of interest from the municipality.
17  Environmental Diagnosis of the Protected Coastal Areas of the Bahía Blanca Estuary 487

The surveys of flora and fauna species were carried out when the management plan
of the Bahía Blanca, Bahía Falsa, and Bahía Verde Multiple Use Nature Reserve,
currently outdated, was prepared. The management plan of these area proposes
operational objectives to carry out an environmental biological monitoring; regulate
the activities of artisanal, sport fishing, and recreational tourism; monitor exotic
species such as the Magallana gigas, in addition to implementing an education and
dissemination program; and also intensify marine turtle conservation studies as a
species of accidental capture by fishermen, in addition to other objectives related to
emblematic species of the aforementioned natural reserves. This management plan
should be updated continuously due to the complexity and dynamism of the ecosys-
tems that are conserved and protected in these coastal areas.
There is a slight direct and continuous communication between government
agencies that manage natural reserves in the estuary. Likewise, there is no efficient
application of the laws and ordinances related to the protected area, given that in the
islands that are part of the Bahía Blanca, Bahía Falsa, and Bahía Verde Multiple Use
Nature Reserve, there are irregularities with respect to the livestock use observed in
them. With respect to the Islote de la Gaviota Cangrejera Nature Reserve, its limits
are not well defined, and the legislation in force has a lack of representation. The
effective application of current legislation would allow the detection of illegal activ-
ities to identify and penalize those responsible.
The environmental problems that directly affect the integrity of the estuarine
protected area system also arise from the insufficient application of the laws that
regulate the different activities allowed and not allowed in natural reserves and adja-
cent areas. This condition is observed in the indicators of the process variables that
refer to management practices and its planning. The values varied between 14 and
2%, with the lowest value for the municipal reserve.
The laws related to protect area must be complementary to their objectives and
promote management effectiveness. Laws related to protected areas must include
land use planning and laws to regulate agriculture and regulations and policies at
national, regional, and local levels. With respect to legal security, the reservations
are analyzed by having different jurisdictional levels present delays in the applica-
tion of the legislation. Environmental protection goals should be incorporated into
all aspects of the legislation of protected coastal areas. Environmental impact
assessments must be carried out for the development of pre-existing infrastructure;
to the expansion of the economic activities incompatible with the conservation mea-
sures of the reserves, controlled by land use planning guidelines that in these envi-
ronments make implementation difficult, although the assigned human resources
are highly trained, the allocated resources are not sufficient.
The comparative analysis allowed to identify systemic strengths and weaknesses
of the natural protected area system of the estuary. The main problems are linked to
inputs (insufficient infrastructure and financing) and processes (lack of adequate
monitoring programs). The Bahía Blanca Coastal Reserve operates in the absence
of equipment and infrastructure, while the Bahía Blanca, Bahía Falsa, and Bahía
Verde Multiple Use Nature Reserve and Islote de la Gaviota Cangrejera  Nature
Reserve present significant deficiencies.
488 M. E. Carbone et al.

On the other hand, although the staff is highly trained to carry out critical man-
agement activities, the number is insufficient. As for planning, all reservations have
legal protection. However, only one of them has a partially approved management
plan. Finally, the indicators that detected the highest score are those related to the
context (biological and socioeconomic importance).
The study area stands out for its great biodiversity and the provision of important
ecosystem services. The reserves protect the habitat of numerous animal and plant
species of high social, cultural, and economic importance, as well as geoforms and
natural processes of interest, holding great tourist, educational, and scientific value.
The use of this methodology is presented as a tool of great relevance for the deter-
mination of aspects to strengthen and prioritization of the actions, in order to
improve the effective management of protected areas (Speake et al. 2018; Speake
and Carbone 2017).
The results highlight an index with values of 10.2 and 7.9% for provincial
reserves, for the education and management of visitors, which could indicate that
within the management of Bahía Blanca Estuary protected areas, greater emphasis
is given to tourist use and the management of the resources to be protected. This
situation is also evident in the environmental education programs that are developed
for the dissemination and knowledge of the estuary reserves.
The multipurpose nature reserve achieved the best results, with the highest over-
all driving effectiveness index. The inputs are more scarce than in the other reserved
considered. However, in terms of defining objectives and management planning,
they outperform the other areas. This area has the limits and management objectives
better defined than the areas declared later.
The strengths identified in the estuary coastal protected areas refer to the natural
physical system where biological and socioeconomic importance are highlighted.
Most of these units of analysis cover a size that is not sufficient to ensure effective
long-term conservation. The system of the analyzed reserves also has conservation
objectives, both at the level of provincial policy and at the level of each conservation
area, although at the municipal level there is a lack of setting specific application
objectives. The majority of the units that make up the system have adequate legal
support and the necessary basic infrastructure, in accordance with the established
objectives, although it is the municipal reserve that does not have adequate legal
support. Regarding the processes, the lack of complete and updated inventory of
natural resources and the collaboration between protected area personnel, commu-
nities, and other organizations stand out.
The weaknesses observed for planning are inadequate limits, including inappro-
priate size and land uses in adjacent areas of protected areas, which are the greatest
threats to Bahía Blanca Estuary reserves. For effective management, inputs consti-
tute the aspect that needs the greatest improvement, highlighting the lack of funds
and the low number of personnel assigned to conservation units.
The scarcity of sustainable financial resources for the management of protected
areas and the administration of the coastal protected area system complicates their
effectiveness. The results also show the limited availability of direct personnel
assigned to protected areas, inadequacies in threat control, and the results of recent
17  Environmental Diagnosis of the Protected Coastal Areas of the Bahía Blanca Estuary 489

investigations carried out by various government agencies such as the Instituto


Argentino de Oceanografía and the Universidad Nacional del Sur that should be
incorporated into the plans of handling these reserves. The pressures that most
affect the coastal protected areas of the Bahía Blanca Estuary are garbage, liquid
waste, invasive species, and the development of port and industrial activities.
The values of the indicators of the different areas analyzed allowed the classifica-
tion of reserves into two categories. They represent the lowest ranges with values
below 25% (Islote de la Gaviota Cangrejera Nature Reserve 13% and Bahía Blanca
Coastal Reserve 7%). These areas lack the minimum resources necessary for basic
management, and only one of the reserves reaches the value of 29% (Bahía Blanca,
Bahía Falsa, and Bahía Verde Multiple Use Nature Reserve) reaching the second
category that represents an unsatisfactory effectiveness; the area has resources and
means that are indispensable for its management, but it lacks elements to reach an
acceptable minimum level.

17.7  I mportance of the Wetlands of the Bahía Blanca


Estuary in Compliance with the Sustainable
Development Goals (SDG)

Given the advance of threats to wetlands, the international approach is to implement


the SDGs proposed in the 2030 Agenda to achieve harmony between the natural
physical subsystem, the socioeconomic subsystem, and the administrative political
subsystem. To address the study of coastal areas in this case of Bahía Blanca Estuary
wetlands in an integrated manner, the subsystems mentioned must be taken into
account.
The objective of the Convention on Wetlands is “the conservation and wise use
of all wetlands through local, regional and national action and international coop-
eration as a means of achieving sustainable development throughout the world”.
The term “rational use” refers to the sustainable use of wetlands to keep their eco-
logical functions as pristine as possible for future generations.
The advance of populations over coastal areas with 600 million people lives in
coastal areas that are less than 10 m above sea level shows the importance of con-
servation and protection of coastal environments (Barragán Muñoz 2014). Economic
and extractive activities should be managed in such a way that spawning areas that
are often found in coastal marine areas are protected to ensure that populations are
restored (Barragán Muñoz 2014). Considering the effects that climate change is
likely to have on coastal ecosystems in the medium and short term, the governing
bodies of the different administrative legal levels should generate integrated man-
agement legislation to mitigate those effects.
“Wetlands play an important role in disaster risk reduction if they are effectively
managed and restored if necessary.” Coastal ecosystems and, in particular, saltwater
marshes called “blue carbon ecosystems” play a role important in mitigating
490 M. E. Carbone et al.

climate change. This carbon is stored in the long term in wetland soils (Barragán
Muñoz 2014).
The International Alliance on Blue Carbon  – announced during the Climate
Change Conference held in Paris in 2015 – aims to bring together different partners,
from government organizations to non-governmental and research organizations, to
conserve coastal ecosystems. Particularly the estuary wetlands are of great biologi-
cal importance and acquire a fundamental function as regulators of natural and eco-
logical processes.
The planning of coastal protected areas should include the sustainable develop-
ment objectives that consider not only the environmental quality of the natural capi-
tal of coastal ecosystems but human well-being; not only the objective of the quality
of life of marine ecosystems should be considered but also the objectives for the
increase of the ecosystem services of the wetlands.

17.8  Conclusion

The creation of three marine coastal protected areas in the estuary sets a precedent
for the importance attached to the value of this environment and the need to preserve
it. The creation of coastal protected areas (CPA) constitutes a very versatile strategy
for the conservation of natural and cultural heritage. Although there are problems
that put the long term viability of the ecological systems they preserve at risk. The
diagnosis updated in adaptive management processes and continuous monitoring
can achieve the objectives of its creation. However, there are problems of execution
where the legal creation rule is the only measure promoted by the administrator in
favor of the area. In order to achieve effective management, it is imperative that the
respective management plans be developed and updated, given that it is not only a
territorial planning instrument that clearly indicates the permitted and prohibited
activities (FAO 2014). Likewise, it should be noted that the distinctions granted by
international organizations to these wetland environments (WHSRN and AICA) do
not have a legislated management category, which makes it difficult to lodge protec-
tion against land uses incompatible with the protection of these areas. The lack of
adequate financing and equipment is unleashed (Speake and Carbone 2017) which
represents a direct threat against biodiversity conservation objectives. The same
situation occurs with assets of ecological interest that do not have laws or manage-
ment strategies for their value and conservation.
The main pressures received by the natural reservoirs of the estuary and that
favor the degradation of the estuary are the concentration of infrastructure on the
northern coastal edge (port and industrial), environmental pollution, dredging with
the subsequent artificial filling of wetlands, and the presence of invasive species.
The direct consequences on human well-being are derived from the decrease in
food from fishing, the loss of first-class tourist resources, degradation of the land-
scape, loss of biodiversity and health problems due to air quality, and conflicts
between human activities and the physical environment.
17  Environmental Diagnosis of the Protected Coastal Areas of the Bahía Blanca Estuary 491

References

Aldana Mazorra O, Hernández Zanuy A (2018) La Planificación Espacial Marina: marco opera-
tivo para conservar la diversidad biológica marina y promover el uso sostenible del potencial
económico de los recursos marinos en el Caribe. En: Hernández Zanuy A. C. (Ed.). Adaptación
basada en Ecosistemas: alternativa para la gestión sostenible de los recursos marinos y coste-
ros del Caribe. Red CYTED 410RT0396. (E.  Book). Editorial Instituto de Oceanología, La
Habana. p171
Amarfil R (2006) Actividades económicas puntaltenses en la Ría de la Bahía Blanca, en revista El
Archivo. Número 16:6–9
Angeles G (2001) Estudio Integrado del Estuario de Bahía Blanca. Tesis Doctoral. Departamento
de Geografía. Universidad Nacional del Sur, p 166
Barragán Muñoz JM (2014) Política, gestión y litoral. Una nueva visión de la Gestión Integrada de
Áreas Litorales. Madrid: Ed. Tébar Flores p 456
BirdLife International (2019) Important Bird Areas (IBA) factsheet: Reserva de Uso Múltiple de
Bahía Blanca, Bahía Falsa y Bahía Verde. http://www.birdlife.org Consultado el 28/11/2019
Cabrera AL (1971) Fitogeografía de la República Argentina. Sociedad Argentina Botánica
14(1):1–42
Carbone ME, Spetter CV, Marcovecchio JE (2016) Seasonal and spatial variability of macronutri-
ents and Chlorophyll a based on GIS in the South American estuary Bahía Blanca, Argentina.
Env Earth Sci 75:1–15
Charlier RH, Bologa AS (2003) Coastal zone under siege  – is there realistic relief available? J
Coast Res 19(4):884–889
Cinti S (2017) Las islas de la Bahía Blanca y Los forjadores de su historia, Ed Vacasagrada p 302
Conde AA, Piccolo MC, Pizarro N (2009) Análisis histórico de las capturas de la flota costera
en el puerto de Bahía Blanca. Período 1983–2007. Actas del 12° Encuentro de Geógrafos de
América Latina, p 123
Day J, Dudley N, Hockings M, Holmes G, Laffoley D, Stolton S, Wells S (2012) Guidelines for
applying the IUCN protected area management categories to marine protected areas. IUCN,
Gland, p 36
European Environmental Agency EEA (1998). Europe’s environment: the second assessment.
Copenhagen 132 p
Food and Agricultural Organization (2012) Estado de las áreas marinas y costeras protegidas en
América Latina. Elaborado por Aylem Hernández Avila. REDPARQUES Cuba. Santiago de
Chile, p 620
Food and Agricultural Organization (2014) Manejo de las Áreas Marino Costeras Protegidas para
garantizar medios de vida sustentables y seguridad alimentaria. Elaborado por A. Rocío Motrán
Ferrándiz y Vanessa Dávila. REDPARQUES Santiago de Chile
INDEC Instituto Nacional de Estadísticas y Censos (2010) Censo Nacional de Población, Hogares
y Viviendas 2010. Buenos Aires p 345
Langhoff ML (2015) La reacción de la comunidad local ante el proyecto de dragado en el área de
Puerto Cuatreros. ¿Es posible recuperar la relación comunidad naturaleza en Bahía Blanca?
En: V Jornadas de investigación en Humanidades, 76–89
Larracoechea G, Durán H, D’Acunto C (2012) Nidificación de la Golondrina Tijerita (Hirundo
Rustica) en el Balneario Arroyo Pareja, Buenos Aires, Argentina. Nuestras Aves 57:18–19
López Cazorla A, Molina JM, Ruarte C (2014) The artisanal fishery of Cynoscion guatucupa in
Argentina: exploring the possible causes of the collapse in Bahía Blanca estuary. J Sea Res
88:29–35
Maelfait H, Belpaeme K, Lescrauwaet AK, Mees J (2006) Indicators as reliable guides for
Integrated Coastal Zone Management. In: Forkiewicz, M. 2006. Integrated Coastal Zone
Management: theory and practice. EuroCoast – Littoral 180–186
Massola MV, Cinti S (2012) Valoración del humedal costero del sudoeste bonaerense desde el
abordaje holístico IX Jornadas Nacionales de Geografía Física Bahía Blanca 20–26
492 M. E. Carbone et al.

Matamala R (2013). Ecoturismo accesible como alternativa turístico-recreativa para personas con
discapacidad visual en la localidad de Villa del Mar, partido de Coronel Rosales. Departamento
de Geografía y Turismo, Universidad Nacional del Sur p 145
Melo, W. D. (2007). Orígenes morfológicos. En: M. C. Piccolo, M. C. y M. Hoffmeyer (Eds.).
Ecosistema del estuario de Bahía Blanca 21–27
Nebbia AJ, Zalba SM (2007) Comunidades halófitas de la costa de la Bahía Blanca (Argentina):
caracterización, mapeo y cambios durante los últimos cincuenta años. Boletín de la Sociedad
Argentina de Botánica 42:161–171
Organization for Economic Cooperation and Development (1993) Core set of indicators for
environmental performance reviews. A synthesis report by the Group on the State of the
Environment. Environment Monographs, p 83
Petracci P, Sotelo M (2013) Aves del estuario de Bahía Blanca: Una herramienta Para su cono-
cimiento y conservación. Grupo Editorial Muelle Sur, Bahía Blanca, p 192
Petracci PF, Sotelo MR, Díaz LI (2008) Nuevo registro de nidificación de la Gaviota Cangrejera
(Larus atlanticus) en la Reserva Natural Bahía Blanca, Bahía Falsa y Bahía Verde, Buenos
Aires, Argentina. Hornero 023:037–040
Sotelo M, Massola V (Eds.) (2008) Propuesta Plan de Manejo Reserva Natural Provincial de Uso
Múltiple Bahía Blanca, Bahía Falsa, Bahía Verde. Bahía Blanca, p 167
Spagnuolo JO (2005) Evolución geológica de la región costera-marina de Punta Alta, provincia de
Buenos Aires. Bahía Blanca Tesis de Doctorado Departamento de Geología UNS p 269
Speake MA, Carbone ME (2017) Evaluación y priorización rápida del manejo de las áreas pro-
tegidas costeras en el estuario de la Bahía Blanca. En: Cenizo M, Celsi C. (Eds.). Libro de
resúmenes de Segundas Jornadas Bonaerenses sobre Conservación de Ambientes y Patrimonio
Costero. Villa Gesell, Buenos Aires, p 453
Speake MA, Carbone ME, Spetter CV (2018) Ocurrencia de eventos de emergencia ambiental en
el área costera de Bahía Blanca, provincia de Buenos Aires. En: Libro de resúmenes de las XII
Jornadas Nacionales de Geografía Física 91–95
Spetter CV, Buzzi NS, Fernández EM, Cuadrado DG, Marcovecchio JE (2015) Assessment of the
physicochemical conditions sediments in a polluted tidal flat colonized by microbial mats in
Bahía Blanca Estuary (Argentina). Mar Poll Bull 9:491–505
Spetter CV, Fernández EM, Carbone ME, Negrin V et al (2019) Estudio de línea de base sobre
la dinámica de nutrientes en una planicie de marea del estuario de Bahía Blanca previo a
la instalación de una planta de tratamiento de desechos cloacales. V Reunión Argentina de
Geoquímica de la Superficie La Plata, Asociación Argentina de Sedimentología, pp 330–333
Chapter 18
Small-Scale Artisanal Fishers and Socio-­
environmental Conflicts in Estuarine
and Coastal Wetlands

Daniela M. Truchet and M. Belén Noceti

18.1  Introduction

From the paradigmatic constitution of what we call modernity, the relationship


nature-culture has been introduced as a dichotomic form in the scientific area, prob-
ably as a strategy of legitimation of the extractive practices on nature (Svampa
2012; Giarraca and Teubal 2013). In academic environments, such dichotomy has
been overcome with the contributions to the social sciences of authors like Descola
and Pálsson (2003), Ingold (2000, 2003), Ellen (2003), and Latour (2004), among
others. For these authors, the nature/environment is not only a simple scenario
where the social actors and actresses unfold their survival activities, but it is more
like a sociocultural resource. In this sense, nature has a plus value in the policy
design and execution oriented to the habitat conformation and organization (Carman
2011). These different forms of inhabiting nature are defined as continuous disputes
and may fall under the category of socio-environmental conflicts (Wagner 2016):
conflicts where different actors and actresses struggle for the access and control of
territories and natural resources, which are referred as divergent interests and values
in scenarios with strong politic asymmetries. In these conflicts, there are also estab-
lished meanings and feelings that acquire the notions of progress and development

D. M. Truchet (*)
Área de Oceanografía Química. Instituto Argentino de Oceanografía (IADO), Universidad
Nacional del Sur (UNS), CCT-CONICET, Buenos Aires, Argentina
Departamento de Biología, Bioquímica y Farmacia, Universidad Nacional del Sur (UNS),
Buenos Aires, Argentina
e-mail: dmtruchet@iado-conicet.gob.ar
M. B. Noceti
Instituto de Investigaciones Económicas y Sociales del Sur (IIESS), Universidad Nacional del
Sur (UNS) – CONICET Bahía Blanca, Buenos Aires, Argentina
Departamento de Economía, Universidad Nacional del Sur (UNS), Buenos Aires, Argentina

© Springer Nature Switzerland AG 2021 493


S. M. Fiori, P. D. Pratolongo (eds.), The Bahía Blanca Estuary,
https://doi.org/10.1007/978-3-030-66486-2_18
494 D. M. Truchet and M. B. Noceti

and, in a certain way, of democracy. Occasionally, the ways in which the state
decides about the organization and exploitation of spaces are imperative, that is,
imposed without the population consensus, assigning sacrificial territories or
socially empty spaces (Sack 1986; Quist 2019), a situation that enhances the process
of social fragmentation. This fragmentation produces, in one hand, the restriction
and gradual exclusion of certain social sectors from the access to and the availability
of natural resources (Latour 2004), and on the other hand, it produces inequality in
the distribution of the benefits and externalities of certain productive activities
(Girado 2012; Merlinsky 2013).
This context of social, ecological, and cultural crisis finds no answers on the
mechanist paradigms. Complexity thus becomes the new emergent boarding per-
spective because environmental risks do not know about social classes. However,
risk distribution is unequal and follows the logics of a class inequality, typical of the
neoliberalism, known as “environmental racism.” By denying identities and cul-
tures, the environmental racism deepens the fragmentation and widens the gap
between different social sectors (Machado Aráoz 2010). The so-called “neo-­
extractivism” in South American countries raised in dependence on the global econ-
omies. Commonly, underdeveloped South American countries that produce primary
goods (commodities) have a peripheral position in the world economy and eco-
nomic dependence on the rich developed countries. In this scenario, Latin America
becomes a territory of widespread loss of natural resources, along with the displace-
ment of several communities and cultures. Social resistance to displacement and
inequality tends to create a spiral of violence, criminalization, and repression
(Svampa 2012), with large power asymmetry: large empowered companies and
political lobbies against indigenous communities and unpowered populations.
In South America, ecological damages raise social and territorial demands, and
the ecological legislation often tends to consider these conflicts as “creative trans-
formations” based on the tensions and antagonisms of the different actors and
actresses (Merlinsky 2013). In other words, the struggles that arise within the kalei-
doscope of actors and actresses create opportunities for making the popular demands
more visible.
In several regions of this continent, small-scale artisanal fishers are actors often
related to socio-environmental conflicts, as they are affected by the consequences of
climate change, industrial development and large fisheries, and, sometimes, tourism
and large aquaculture projects (Charles 1992; Coulthard et al. 2011; DuBois and
Zografos 2012; Noceti 2017; McGregor 2018). For example, augmented emissions
of CO2 and climate change are expected to diminish fishing stocks in emerging
countries, coinciding with the present socio-economic vulnerable areas (Cheung
et  al. 2010). In these changing scenarios, artisanal fisheries are at serious risk
because entire families could be driven into poverty and social destitution (Allison
et al. 2009). Thus, poverty and vulnerability are, and will continue to be, endemic
conditions in small-scale fishing communities.
Thus, artisanal fishing keeps entire families away from marginal conditions and
stands for a set of values over the profit-oriented and capitalist enterprises typical of
large and industrial fisheries (Johnson 2018). These values are context-specific,
18  Small-Scale Artisanal Fishers and Socio-environmental Conflicts in Estuarine… 495

committed to a community and a place and their socio-economic and historical rela-
tions. They are also according and respectful to what nowadays is called the “rights
of nature” (Truchet 2018): small-scale artisanal fishers have an intrinsic relation
with nature and are invaluable actors in contributing to alternative forms of develop-
ment in harmony with human and nonhuman beings.1 Although, in most cases, they
lack formal education and specialized training, fishers exhibit a great empirical
knowledge about biology and ecology, which has proven to be as valuable as scien-
tific information and provides significant tools for ecological academic research and
animal conservation (García-Quijano and Valdés-Pizzini 2015). Furthermore, the
arts of artisanal fishing are based on the interaction with what we call nonhuman
beings, and their function is to protect and respect the different species that develop
their life cycle in the seas and coastal wetlands. This form of knowledge is not only
based on biological academic knowledge but also in learnings acquired through
ancestral practices and traditional ontologies.
In addition, several studies have demonstrated that fishers’ knowledge about ecol-
ogy (hereafter, FEK) is priceless for ecological studies, since fishers have been navi-
gating and observing aquatic ecosystems for decades, acquiring a unique comprehension
of ecological interactions (Johannes et al. 2000). Although the empirical knowledge of
fisher communities should be well appreciated, Johannes et  al. (2000) found that
marine ecological researches often neglect local knowledge and dismiss fishers’ cul-
ture. In our experience, the dismissal of FEK and their cosmology is the main reason
for which biologists and ecologists often arrive at uncertain or wrong conclusions
about some marine species traits and the status of coastal ecosystems. In this sense,
Quist (2019) also argues that scientific indetermination about the causes of depleting
fish populations and the weakness of environmental legislation result from the exclu-
sion of FEK from politics. Industrial knowledge, such as that coming from the petro-
leum industry, is often included in policy and lawmaking and is, therefore, considered
valid. Thus, considering FEK perspectives into research, conservation, and manage-
ment strategies represents a new approach that includes local knowledge.
From what we exposed so far, the main objectives of this chapter are a) to under-
stand which socio-ecological conflicts take place in the Bahía Blanca Estuary; b) to
comprehend the different conservation paradigms coexisting in the Bahía Blanca
Estuary through the voice of different actors; c) to characterize small-scale artisanal
fisheries in the Bahía Blanca Estuary from their past, present, and future status; and
d) to assess fishers’ ecological knowledge (FEK) in the Bahía Blanca Estuary and
how it can be useful for future monitoring programs in estuaries that are presently
seen as sacrificial territories.

1
 Descola (2005) studied indigenous societies (jíbaros) in the Amazonas describing their relation-
ships with nature and was able to identify different forms of inhabitants in the world: human and
nonhuman beings. In a reductionist and vague explanation, nonhuman beings are rocks, plants,
animals, and spirits, and humans are also able to establish a relationship or a way to socialize with
them. Thus, the relations are not only conceived as human-human like the occidental world per-
ceives but also as humans-nonhumans. This was also observed in other local communities, such as
fishers and hunters by Pálsson (2003).
496 D. M. Truchet and M. B. Noceti

18.2  The Bahía Blanca Estuary

18.2.1  S
 ea-Land Dispute and Maritime Possession:
An Ontological and Territorial Perspective

The Bahía Blanca Estuary (southwest of the Buenos Aires Province) is an area in
dispute regarding the use of its maritime-coastal resources since it was occupied for
the government of the United Provinces of the Río de la Plata in the nineteenth
century. This area was known as the “Liverpool of the South” or the “City of the
Seven Ports” because Bahía Blanca was the only city in Argentina with seven docks
by the first half of the twentieth century (Lugones 1883). Until the 1980s, the inhab-
itants of this region had free access to the coasts of the estuary, building their identi-
ties attached to the maritime-coastal territory. Following Peron and Rieucau (1996),
this singular type of identity is called “maritimity,” and it constitutes lifestyles in
reference to the link human being-sea. The upcoming installation of the petrochemi-
cal complex of Bahía Blanca in the 1980s triggered a gradual privatization of the
access to the estuary. The state policies were afterward oriented to alienate the estu-
ary from the public sphere, which ended up with the sociability contexts that were
carried for almost a century. This constitution of the territory in an economically
productive and socially empty scenario and where the monetary income justifies the
negative environmental externalities will be called from now on “maritimization”
(Peron and Rieucau 1996; Noceti 2017). Therefore, in this section, we will use an
anthropological and ethno-ecological perspective to explain the dispute for inhabit-
ing the space of the Bahía Blanca Estuary, a conflict named as “fishers’ conflict”
(see Text Box 18.1: Fishers’ conflict) in juridical and social media.
In the present point, we will try to give an account of the identity process built in
the estuarine space, from an ontological perspective that denies the existence of a
unique world and a single representation of itself (cosmovisions). Thus, we consider
the existence of multiple forms of nature, and the alterity is in function of the exis-
tence of these compositions (Tola 2011). The ontological turn is built on the multi-
naturalism (Viveiros de Castro 1996) and might be helpful when it comes to explain
how and why the socio-environmental conflicts persist in the Bahía Blanca Estuary
(Noceti 2018). We will assume that the conflict is materialized in socio-­environmental
coordinates, but it hides a political-ontological discussion. Therefore, it is possible
to explain why these conflicts persist in time and why it was not possible to find so
far any solutions that allow the coexistence between actors. We hereby assume that
it does not exist in one single and universal form of nature over which different
actors develop their representations. Instead, the space is perceived as a territory to
be inhabited (Yory 2013, 2015), and it emerges as an ontological composition raised
toward different coordinates: ethics, cognitive, perceptual, social, cultural, politics,
and affective, including topophilia developments (Tuan 1974). The debate is not
about domain and control of a common good such as the sea and its coasts. More
likely, it happens because of the identification of the existence of multiple territories
(De la Cadena and Blaser 2018).
18  Small-Scale Artisanal Fishers and Socio-environmental Conflicts in Estuarine… 497

In the Bahía Blanca Estuary, we identified three territories disputing their exis-
tence in the same space and at the same time. Each one refers to different collectives
who are in a conflictive interrelation: the fishers of the Bahía Blanca Estuary; seniors
and state officials, along with businessmen of private companies of the Bahía Blanca
Port Management Consortium (Consorcio General de Puertos de Bahía Blanca,
CGPBB); and the conservationists (most of them environmental professionals of
nongovernmental organizations  – NGOs). These territories are, respectively, the
estuary-ria, the estuary-port, and the estuary-protected areas (Fig. 18.1) and coex-
ist in permanent conflict, trying to impose themselves in a scenario of strong p­ olitical
asymmetry in favor of the CGPBB and in detriment of the popular sectors, like the
artisanal fishers. It reinforces the existence of a fragmentation process, a socio-­
spatial segmentation, and the alienation of the productive spaces, with a strong
meaning for fishers. Through the ethnographic exploration, we were able to identify
the following characteristics of each territorial moment that could help us to explain

Fig. 18.1  The Bahía Blanca Estuary and the three different forms of territory that coexist in per-
manent conflict: the estuary-ria, the estuary-port, and the estuary-protected areas
498 D. M. Truchet and M. B. Noceti

the areas where the socio-environmental conflicts in the Bahía Blanca Estuary arose
and why:
–– The estuary-ria: this territorial form has its foundation in artisanal fishers, mostly
from Italian immigrant families,2 which named the estuary as a “ría,” though, in
terms of geomorphology, it is not correct. This territory denotes the forms in
which the alterity is constructed between humans and between human and non-
human beings (i.e., physical, spirits, animals, vegetation, other forms) (Descola
2005). Artisanal fishers have built their identities differentiating from “people
that only fish” due to the forms in which they configure their lives in space and
time. The space and time are in function of their fishing activity, which is sea-
sonal. In terms of space, it is observed that the artisanal fishers have their own
maps besides the official naval chart. These maps are built in function of the
places that are identified as “fishes’ hot spots” and are classified and named
based on the experiences lived by their forefathers. As an example, the “pajar-
era,” “palito,” and “hundido” are places that are invisible for the eyes of people
that do not inhabit the estuary nor belong to the history of fisheries in the region.
Instead, they received their names from stories that every fisher knows but remain
unknown for those that do not share their livelihood as an indentitary form.
–– The estuary-port: This territory integrates the petrochemical complex of Bahía
Blanca and the Port Management Consortium (CGPBB), because both have a
similar speech and petrochemical activities are linked to the port development
through the exportation of industrial products. In the estuary-port, the relation-
ship human-nonhuman is virtually null or inexistent from the perspective of
people that belong to the petrochemical complex and the CGPBB. Nature appears
as a mere “object,” where human activities (those producing high economic
income) take place. The rest of the activities (sports, tourism, fisheries) that
develop in the same area are ignored, and they are not even conceived as existent,
even when they are easily observable, and therefore they do not appear in the
estuary-port cognitive map. In fact, they do not even recognize the existence of
beaches in the area.
The protected areas (especially Bahía Blanca, Bahía Falsa, and Bahía Verde
Natural Reserve) are considered out of the territory and contact restricted to specific
situations. Companies/enterprises of the petrochemical complex relate to the pro-
tected areas through money disbursement by their corporate social responsibility
programs, pursuing tax reduction. Companies like Profertil, Dow Chemical,

2
 Small-scale artisanal fishers of the Bahía Blanca Estuary are fully described in the following point
18.2.2.
18  Small-Scale Artisanal Fishers and Socio-environmental Conflicts in Estuarine… 499

Petrobras, and YPF maintain economic cooperation agreements with city govern-
ments of the area and even with the local universities (Heredia Chaz 2014).
Moreover, for these actors, the existence of the “other” is a stumbling block to be
overcome, where the relationship mode, called rapacity by Descola (2003), is use-
ful to describe the relationship not only between nonhuman and humans but also
between humans belonging to the port system and fishers. Their discourse based on
development makes evident the strong political asymmetry and deepens the differ-
ences between fishers and the port system.
–– The estuary-protected areas: this scenario is represented by a series of actors that
are dissimilar between each other but with a strong speech based on the defense
of the nonhuman environment, above the existence of human productive activi-
ties, such as fisheries and the industrial context. In the view of conservationists,
nature is an entity that must be protected from human actions that cause damage.
Nature exists if it can be protected by these actors, and therefore we can verify a
passive role of the nonhuman beings that differs from the human nature, making
clear its non-belonging to nature.
This cognitive formation presumes in the social imaginary a great diversity of
nonhumans that depend on humans for their well-being. This dependence is utilitar-
ian: the protection of nonhumans ensures the benefits of humans that found their
actions on the speeches of defending and protecting the environment. Under this
protectionist and paternalist scenario, there is a place for the workers of the pro-
tected areas and the conservationists from the local NGOs. The ways in which these
actors catalogue the rest of humans and nonhumans are in function of what they
consider that must be controlled and what it is expected to be protected. This group
of actors is not homogeneous when it comes to claim for the protection of nonhu-
man life and holding other actors responsible for the damage to the environment.
While the workers of the protected areas develop their programs of reeducation
oriented to modify the arts of fishing  – considering this collective as the main
responsible of the decrease of fishing stocks in the estuary – the conservationists of
the environmental NGOs have accompanied artisanal fishers in their strikes and
legal claims against the CGPBB and even against the Natural Reserve, because they
adduce that the Argentine government is the main responsible for the loss of fishing
stocks, since they do not control the human activities that affect the ecosystem.
500 D. M. Truchet and M. B. Noceti

Box 18.1 Fishers’ conflict


The emergent marine pollution that was registered in the Bahía Blanca Estuary
drove a constant and fierce socio-ecological conflict known as the “fishers’
conflict.” The main actors were represented by the artisanal fishers’ families,
which initiated a lawsuit against the industrial complex and a sewage water
treatment plant, which, according to them, were the major causes of the
decline in fishing. Noceti (2017) and, recently, explained that, during the
2009–2012 period, the fishers’ families suffered from a drastic reduction in
their incomes due to the collapse of the fisheries. The community claimed this
was due to overfishing in El Rincón area and the pollution of the estuary,
which resulted in a rapid impoverishment of the families. Fishers’ communi-
ties started a series of multiple claims to the municipal government, demand-
ing economic help to support basic survival conditions to sustain their
livelihoods. Given this situation, both the municipality and the provincial gov-
ernment designed the fishery “reconversion program,” which involved the
purchase of fishing permits by the provincial government from those fishers
who chose to stop fishing or the purchase of new, more powerful boats with
outboard motors for fishing offshore (mostly in the external part of the estu-
ary), where there was a greater biomass of commercial species. Most of the
families chose to sell their permits to obtain money that they spent quickly
and failed to reinvest in another type of productive activity given the recession
periods that hit Argentina over the last few years (especially during
2015–2019).
These families had to start fishing again in situations of greater vulnerabil-
ity, because most of them did (and still do it) under conditions of illegality,
since they sold their permissions and use alternative places as “ports” to avoid
being legally enrolled and to pay extensive taxes. On the other hand, fishers
who chose larger boats failed to sustain the activity due to the amount of fuel
needed to operate these boats and the continual increase in fuel prices, which
led them to sell some of these boats and return to fishing in the inner part of
the estuary. However, artisanal fishing in the Bahía Blanca Estuary is now an
activity that is at risk of being lost, along with the fishers’ culture, due to the
decrease in the number of fleets. Nevertheless, the “illegality” in fishers’
activities is a constant discussion. For example, Nahuelhual et  al. (2018,
2019) explained that small-scale artisanal fishers that do their activities in an
illegal form cause less damage to the ecosystems that “legal” activities such
as the extractives large fisheries companies, that are sometimes  – if not
always – extent of high taxes.
In summary, fishers’ conflict exists since 2009, and legal proceedings are
still underway. At present, fishers’ lawsuit against companies, the port, and
the sewage plant was not fully resolved, and contrasting reports about pollut-
ant levels in the estuary raise questions and concern in the fishers’ community.
Therefore, scientific commitment and responsibility are required to provide
this vulnerable collective with a sound evaluation on the health of the environ-
ment that they are part of and where they develop their livelihoods.
18  Small-Scale Artisanal Fishers and Socio-environmental Conflicts in Estuarine… 501

18.2.2  S
 mall-Scale Artisanal Fisheries in the Estuary:
The Past, Present, and Future Status

For the Argentine law, small-scale artisanal fishers are defined as people who extract
fish for their commercialization, using artisanal beach nets, such as trammel nets,
gillnets, fixed nets, traps, hook and line, or any other fishing gear that is not forbid-
den, by manual extraction, diving, from the coast or boats, and/or manual collection
from the intertidal zone (Noceti 2017). In this context, the artisanal fishers from the
Bahía Blanca Estuary are linked to Italian immigrants, especially from the Island of
Ponza, who arrived to the area around 1890–1920 and whose forefathers handed
down sustainable fishing methods from “parent to child” (London et  al. 2017;
Noceti 2018). Some of the Italian traditions are still observed like their religious
belief to San Silverio, who is considered in the Italian tradition “the keeper and
guardian of the fishers.” Also, some of the foods, fishing arts, vocabulary, and words
are Italian and are present in their livelihoods. The fishers’ circuits find their extreme
situations in the child labor of male children and teenagers in families with extreme
socioeconomic vulnerability and low levels of schooling, where kids work in order
to earn money to sustain their families (Truchet et al. 2020a, b): “When we came
from Ponza (Nápoles, Italy), I was ten years… At the school in Italy, I was in fifth
grade, but when I came to Ingeniero White, they put me in third grade! The verbs,
the language… they were very difficult for me. When I got to the fifth grade here, I
missed my exams, and I had to take the year again. And, honestly, I didn’t want to
do it, and my mom sent me to fish. I was the only support to my family; my father
earned too bad; he was a hairdresser. At that time, (Ingeniero) White was a small
village and fifty percent Italians, we knew each other, so my mom immediately
looked for a shipment for me and at the age of thirteen, I was already fishing. I
remember we were seven boys at ‘La Envidio’, Antonio was the owner, and at that
time it was different… Not like now, where the kids can earn their own money when
they work. I worked all day and the money was earned by my mom and we built our
home in that way. Then at the age of sixteen I started fishing with my uncle…” (SG,
ex-fisher).
Despite traditions, artisanal fishing has also been an economic refuge for fami-
lies during the economic crises that have hit Argentina during the last decades, espe-
cially in 2001 and from 2014 to the current days. In this way, unemployed and/or
less experienced men learned the job from other fishers, in order to provide eco-
nomical support for their families. As some of our interviewees claimed: “I started
fishing in 2001 (Argentine economic crisis) at twenty-one… why? Because I lost
my job at a supermarket. And mi brother in law brought me here, to the fishing. He
had a boat and he used to fish. And he said to me ‘let us try, you might like it’. And
honestly, I did. I did like it. And slowly I started to buy the materials; it was a great
effort… But I like it if you don’t like this, then it’s useless. And slowly I learned to
fish by watching and paying attention to the environment…” (PB, fisher).
The most commercialized species in the estuary have always been the “len-
guado” (Paralichthys orbignyanus), the silverside (Odontesthes argentinensis), the
502 D. M. Truchet and M. B. Noceti

striped weakfish (Cynoscion guatucupa), the whitemouth croaker (Micropogonias


furnieri), and the narrownose smooth-hound (Mustelus schmitti), which constitute
the varied fishery, and also shellfishes, like the Argentine shrimp and prawn
(Pleoticus muelleri and Artemesia longinaris). The arts of fishing have changed
from those described by Lopez Cazorla (2004) and have been simplified. Nowadays
fishers only use gillnets and trammel nets for the different species and a larger net
for shellfishes. Also, few fishers use a modify version of the shellfishes’ net for
C. guatucupa in order to avoid the possible damage caused to the nets by the South
American sea lion (Otaria flavescens). In the Bahía Blanca Estuary, fishing activi-
ties are done in groups of four in good seasons for shellfishes or in groups of two,
when the economic conditions are not suitable to afford more sailors. In the past, the
job was done by a single person, the owner of the boat. Overall, there are 12 work-
ing hours: 6 h during flood tide and 6 h during ebb, especially on shellfish seasons.
In the past, fishers could afford to work less hours, but nowadays, they spent almost
2 days in the sea because of the high prices of the fuel.
Presently, there are 72 families of artisanal fishers of the Bahía Blanca Estuary
distributed in the coastal towns and villages of the estuary, especially in Villa del
Mar (alternative port), Punta Alta (Puerto Rosales), Ingeniero White (Port of
Ingeniero White or Puerto Piojo, as it is named by the fishers’ collective), and
General Daniel Cerri (Puerto Cuatreros) (Noceti 2018) (Fig.  18.2). One hundred
fifty families were recorded in 2013, which are less than a half of the 420 that were
recorded in the 40s of the twentieth century (Noceti 2013). After the fishers’ conflict
(Box 18.1) between 2011 and 2012, the number has decreased, and some of them
are fishing illegally; therefore, the information for the main artisanal fishers’ ports
(Port of Ingeniero White, Puerto Rosales, Villa del Mar, and Puerto Cuatreros) is
quite outdated or obsolete. To our knowledge, almost 50% of the fishers retired after
the reconversion plan. The unknown sociodemographic profile has been recon-
structed from the several interviews that we carried out and public information on
the social media (Truchet et al. under review).
In accordance to our data, the gross profit per fishing boat is from 500 to 700
dollars per week, under good climatic conditions in the varied fishing. From 1000 to
1700 dollars is the estimated profit for a good week of shellfishes. The owner of the
boat gets the bigger share (almost 40%), and the rest is equally distributed between
sailors. However, most of the profit has to be used to cover the expensive costs of
fuel, repairing nets, and maintaining the equipment. Under bad weather or fishing
conditions, weekly profit may be as low as 14 dollars, and it does not cover the fuel
costs. At this point, it is important to point out that good weather conditions in the
Bahía Blanca Estuary are not common, due to cold temperatures in winter and
strong wind storms during summers. Therefore, fishers have a quite low average
income, considering the amount of days with no activity.
Active fishers’ ages go from 25 to 50 years, while more than 50-year-old fishers
are mostly retired because they sell their permissions. It is a risky job, and fishers
retire younger than the usual age of 65  years for the Argentine men (Fig.  18.3).
Gendered roles and activities are described in Box 18.2, but in general, women are
secluded at home, raising their children and doing unpaid domestic labors. Some of
18  Small-Scale Artisanal Fishers and Socio-environmental Conflicts in Estuarine… 503

Fig. 18.2  Small-scale artisanal fishers’ ports: (a) Port of Ingeniero White-Puerto Piojo, (b) Villa
del Mar alternative port, and (c) Puerto Cuatreros (General Daniel Cerri). (Photos by Lic. Rocío
M. Truchet)

them are involved only in postharvest activities, and a few of them sail but only in
bad economic periods, where men cannot afford for external help. The formal edu-
cation of active and retired fishers includes only primary to high school, and none of
them have achieved superior education. However, some of their children, especially
the women, did: “For me, sailing wasn’t dangerous… I sailed with storms and
winds. I had four daughters and I had to give feed them; if I didn’t go fishing, it was
a lost day. I fished to eat and to feed my family. With that, I could educate and raise
my daughters… the older one studied psychology, another one is a teacher… and
people don’t like us or blame us for overfishing, when we use nets that do no harm
for nature” (AB, ex-fisher).
The high taxes, fuel costs, the overfishing of large ships at El Rincón area, the
pollution of the estuary, and the policies that have been taken in the last years, favor-
ing big companies and the port consortium, make small-scale artisanal fisheries a
collective in risk of disappearing. Our recent research (Truchet et al. under review)
showed that artisanal fishers reached a catch of 500 tons in 2008, while in 2016
(after the fishers’ conflict) reported catches were less than 100 tons. Actual data on
the present catches are not available, but we suggest that it could reach less than 50
504 D. M. Truchet and M. B. Noceti

Fig. 18.3 (a) and (b) encounters with artisanal fishers of the Bahía Blanca Estuary, (c) and (d)
artisanal fishers arriving to the Port of Ingeniero White (Puerto Piojo) after a working day where
they fished Micropogonias furnieri, and (e) recreational fishers in Puerto Cuatreros. (Photos by
Lic. Rocío M. Truchet)

tons. On the other hand, large fishing ships continue operating in the outer area of
the estuary with much higher landings per year. In this scenario, small-scale arti-
sanal fishers are in risk to be lost in a nearby future, along with their ancestral cul-
ture and knowledge.
18  Small-Scale Artisanal Fishers and Socio-environmental Conflicts in Estuarine… 505

Box 18.2 Gender roles


Globally, women and men in fisheries have different but complementary roles
and activities, but at the same time, they are influenced by the socio-economic
and cultural contexts of their region and country. According to FAO (2017 a,
b), in most regions worldwide, fishing is carried out by men, especially off-
shore the coastal zones, while women, in a smaller proportion, are secluded to
coastal areas in small precarious boats or canoes with rows. The international
literature places women in artisanal fisheries as key actresses in the local
economies and in the food security and sovereignty. According to Harper
et al. (2013), the roles of women are variable according to the regions: in most
coastal countries of the Central and North Pacific Ocean, women contribute
with 56% of the total captures in small-scale artisanal fisheries, with a positive
impact in the small economies, and help to reverse poverty situations.
However, in other continents like Africa and Latin America, only 47% of the
jobs promoted by artisanal fisheries are occupied by women, being the post-
harvest jobs (cleaning, cutting, packing, selling), the jobs that are allowed for
women, while sailing is usually forbidden, because they are symbol of bad
luck in the maritime tradition (Rodríguez 2009; Luna López 2011;
Vasconcellos et al. 2011; Oliveira and Silva 2012; Álvarez et al. 2017; FAO
2017b; Rojas Herrera 2018).
Concerning gendered fishing activities in the Bahía Blanca Estuary and
other nearby coastal areas (i.e., Pehuen-có and Monte Hermoso), London
et al. (2017) described fishing activities as being related to men, whereas the
women are limited to domestic work and to repairing nets. A recent study by
Truchet et al. (2020a, b) disagrees, pointing out that some women assist men
in the postharvest and help with fishing when families cannot afford to pay for
external help of another male sailor. However, women are not considered
“fishers” in the Bahía Blanca Estuary (nor according to themselves or males),
only fishers’ “assistants” or “women of fishermen”; they are still restricted to
their homes for processing the final product and not paid equally for their
work: “You won’t find any women that knows how to knit her own nets, it’s
weird, that’s a men thing. Except for my niece, I taught her so she could help
me sometimes, and she knows something. Pablo, her husband, is a fisher and
also knows how to knit nets, but no so much because this’s a family business,
and that’s my job. I sold Pablo my permission and the boat he has used to be
mine. Now I’m helping to repair some nets. Anyway, regarding women, I’ve
never seen a women fishing, except for a woman I’ve met one day, she was a
widow and used to fish here… Otherwise, it’s uncommon” (MF, ex-fisher).
Moreover, in the maritime tradition of the southwestern Atlantic, a strong
patriarchal society, women are also a symbol of bad luck for the number of
catches. As it was explained by one of our informants: “It’s uncommon to see
women fishing… we have that expression ‘don’t board a woman nor kill a
seagull!’, it’s bad luck for the job, you know…’” (AB, ex-fisher). On the other

(continued)
506 D. M. Truchet and M. B. Noceti

Box 18.2 (continued)
hand, fishermen claim that the job is only for those that “have more balls” and
“face the sea,” being men the only ones considered as “fishers”: “The only
ones that fishes are the one that has balls, the ones that has guts; it’s not for
everyone” (MF, ex-fisher).
In this context, we corroborated that the gender condition of women is a
sub-alterity to the point in which they are not considered fishers, even when
they are fully involved in the activity. Fishermen assume that the sea is his
territory because they develop their ontologies, focused in values such as the
prestige and the virility.

18.2.3  A Clash of Cosmologies: The Conservation Paradigms

In order to study the different forms in which conservation was considered in many
communities, often with different cosmovisions, Pálsson (2003) established three
anthropological categories based on the contributions of human ecology and social
sciences: orientalism, paternalism, and communalism (Fig. 18.4). According to this
author, each paradigm represents a particular stance with respect to human-­
environment or a particular way in which humans relate with nonhuman beings. The
communalism differs from others because it rejects the modern imposed radical
separation of nature from society, arguing that a constant dialogue has always
flowed between them (Pálsson 2003).
The orientalism and paternalism take for granted the division between nature and
society and emphasize the contrast between domination, exploitation, and protec-
tion to the environment. This is to say, that one tends to see nature as “something”
that is an “object” of human domination and exploitation, while the other one
believes that nature should always be protected by humans. The orientalism sug-
gests a negative relation, reciprocity, and break in human-nature relation, with a
misbalance between them and a clear colonial regime. However, this paradigm not
only belongs to the view of industrial and extractive companies and policy makers
but also belongs to a large group of scientists that are unaffected by ethical consid-
erations and are only present as mere analysts of the material world, without consid-
ering the human compound (Pálsson 2003; Carman and Gónzalez Carman 2016).
The paternalism assumes that the conservation of nature is a human responsibil-
ity, and it is characterized by protectionism instead of exploitation (Pálsson 2003).
However, though they propose radical thoughts, in both paradigms, humans are the
“masters of nature” just for believing they can exploit it or by thinking that they
should always protect it. Thus, there are clear evidences of modern traces in the
speeches of both paradigms.
Finally, the communalism suggests a reciprocity in human-nonhuman/environ-
mental relations, with notions of participation and dialogue and where humans are
not considered the masters of nature. This paradigm rejects any division between
humans and the environment in ecological studies, an assumption that nowadays is
widely recognized but still poorly studied and therefore poorly understood. Pálsson
18  Small-Scale Artisanal Fishers and Socio-environmental Conflicts in Estuarine… 507

a
Continuity Communalism
Generalised reciprocity

Discontinuity Orientalism Paternalism


Negative reciprocity Balanced reciprocity

Domination Protection

Orientalism

Paternalism Communalism

Fig. 18.4 (a) Types of human-environment relations (conservation paradigms) adapted from


Pálsson (2003): orientalism, paternalism, and communalism. (b) A co-existence of paradigms in
the speeches of small-scale artisanal fishers. Adapted from González de Carman and Carman
(2018) for the Bahía Blanca Estuary

(2003) attributes this paradigm to indigenous and some traditional fishers and hunt-
ers’ societies that have no fundamental distinction between nature and society.
At this point, our main question is if these paradigms are present in the arenas
where the socio-environmental conflicts take place in the Bahía Blanca Estuary. For
instance, we interviewed different actors like fishers, employees of the CGPBB, the
petrochemical complex, and the protected areas of the Bahía Blanca Estuary. We
found several dissimilarities between the speeches of the CGPBB, the petrochemi-
cal complex people, and some dialogues we carried out with workers of the nongov-
ernmental organizations (NGOs) and protected areas (Noceti 2017; Truchet et al. in
review): “The development of this region (the southwest of Buenos Aires province)
depends on the port and that the society understands the importance of the founda-
tion of the city-port of Bahía Blanca, which we used to call the ‘third foundation of
508 D. M. Truchet and M. B. Noceti

Bahía Blanca’… fisheries here are marginal; it is expected that they disappear” (VC,
CGPBB worker).
“They (fishers) insist in fishing here, even when the consortium bought their sail-
ing permissions and give them larger boats with better motors, so they can fish
outside of the estuary… they are a bunch of thieves” (EE, worker of a petrochemical
company).
In words of Pálsson (2003), it is possible to verify an orientalist speech regarding
the relations with the environment, where society and nature are extremely dissoci-
ated and where nature is also a colonized space. The words that overflow speeches
of these interviews are development, production, conquer of new spaces, and
increased profitability. In these speeches, the history of the port-petrochemical
complex is founded in ontologies based in capitalists’ values and ethics where
nature is a simple scenario, where some actors are erased to achieve, first, the objec-
tives and ideals of progressivism and then the development proposed by an elite
congregated in the CGPBB. The strikes and protests are criminalized, and an enemy
is built in the construction of a new port complex. This enemy is local, and it is a
threat to the local development: we are referring to the collective of the small-scale
artisanal fishers of the Bahía Blanca Estuary (Noceti 2018).
The government and the CGPBB created different devices for social control and
fragmentation of social movements. In the context of these interviewees, it can be
inferred that the functionaries of the CGPBB and the petrochemical complex tend
to homogenize the fishers’ sector, even when they are already fragmented by previ-
ous policies created during the fishers’ conflict. Overall, fishers are considered abu-
sive, violent, unpredictable, ignorant, and “living from state social plans,” instead of
a hard-working collective (Noceti 2017). It is clear that between the petrochemical
complex, the CGPBB, NGOs, and the protected areas, there is a close relationship,
where artisanal fishers are forgotten and misplaced in the arenas where decisions are
taken regarding conservation strategies, as it can be inferred by their speeches: “Our
company delivers a thousand dollars in cultural projects and also in conservation
projects…several environmental NGOs asked us for financing for some promo-
tional activity, and we’re ok with it; we offer money to them, to FRAAM (a NGO)
to be more specific. These people are really wise about the ecology of this area”
(MC, worker of a petrochemical company).
“Fisheries here are a joke, there’s no fishing here, fisheries are the ones from Mar
del Plata, for example. But, here? They’ve never existed… We can’t compare the
magnitude of the money incomes by artisanal fisheries with what the port and the
petrochemical activities delivers” (EE, worker of a petrochemical company).
On the other hand, in the speeches of the NGOs and the workers of the protected
areas of the Bahía Blanca Estuary, there is a clear paternalist paradigm: humans are
not part of the nature, and the role of humans is to protect a form of nature where
there is no place for other human beings and cultures: “We (FRAAM, the NGO)
constantly do actions for nature, which is constantly attacked by men and their dif-
ferent actions, like pollution due to untreated sewage waters and the actions of the
dredge. During all these years, we’ve done several campaigns to raise awareness of
the habitat and nesting of the Olrog’s gull (Larus atlanticus), the American oyster-
18  Small-Scale Artisanal Fishers and Socio-environmental Conflicts in Estuarine… 509

catcher (Haematopus palliatus). Our last campaign was with the help of the Tamar
Project for the sea turtles” (VM, NGO FRAAM).
Besides having a paternalist-environmental view, the NGOs and the workers of
the protected areas of the Bahía Blanca Estuary want to reconvert the arts of fishing,
and they do not recognize the intrinsic relation that fishers have with the environ-
ment, denying their culture and education, which might not be formal but come
from several generations and more than 100 years of history. There is also a misin-
formation regarding the arts of fishing, because fishers elaborate manually their nets
to avoid catching juveniles. It has also been proven that recreational and industrial
fishing has a larger impact on the bycatch of juveniles of certain species like the
narrownose smooth-hound (Mustelus schmitti) (Colautti et al. 2010; Llompart et al.
2017): “Our main goal is to transform some of the arts of fishing… I mean, fishers
don’t get how dangerous they’re… they tend to catch juveniles with consequences
of the reproduction of the species. Sometimes they understand it; sometimes they
don’t. It’s complicated to work with them. On the other side, it’s easier to work with
the recreate and sport fishers because they have another education…” (MS, worker
of a protected area).
Several conversations and dialogues carried out with fishers lead us to infer a
coexistence of paradigms in the human-environmental relations. The occidentalism
appears in the case of the profits they can obtain from shellfishes (Pleoticus muelleri
and Artemesia longinaris) considering that they are “fool animals” and “easy to
fish.” There is a clear separation between these species and the artisanal fishers, who
have an extractive behavior toward these animals: “The shrimp and prawn are silly
animals, easy to catch, there isn’t much to say about them. They came and you took
them, and if you catch them, they’re good profit. I remember that in the Easter of
2007; I could buy my last boat…you see? That’s the profit we can obtain” (DG,
ex-fisher).
Regarding the South American sea lion (Otaria flavescens), there is a coexis-
tence of two paradigms: the orientalism and communalism. The orientalism because
some fishers see these animals as a threat that they have the right to kill, in order to
avoid losing their fishing and profits: “The worst enemy of the fishers is the South
American sea lion (O. flavescens); they destroy our nets. That’s a day of lost work.
They ruin the fishing and you lose the fish…If we can kill them, we do it. But the
thing is that we can’t; I mean, they’re really smart” (DG, fisher).
But also, the communalism is present, since they believe they share the same
environment with sea lions and have to respect them: “Those who shot the South
American sea lion (Otaria flavescens) or kill them with guns, they aren’t real fisher;
let’s say it’s a person that only fishes” (MD, ex-fisher).
“Being a fisher implies to respect the animals. If I’m already in the environment
of the animal (the South American sea lion), why would I kill it? I’m already doing
damage by taking its food” (MF, ex-fisher).
The communalism is also preset in speeches where fishers claim that the best
fishing is the sole, because it is risky and they have to prove themselves as males.
Thus fishers get into the mud, though profits are lower than those obtained with
shellfishes. The virility of fishers of the Bahía Blanca Estuary is present along all of
510 D. M. Truchet and M. B. Noceti

Fig. 18.5  An artisanal fisher from Villa del Mar repairing a shellfish net that was manually knitted.
(Photos by Lic. Rocío M. Truchet)

their speeches about loneliness. Virility is proved by getting into the sea in order to
face any kind of risks including the situation of no fishing. Fishing means uncertain-
ties, abilities, wisdom, courage, and boldness to cope with the inclemency of
weather. It also means “freedom and no owner,” living by themselves and for them-
selves. In such a way of life, women just wait on the coast, and they are not included
in the fishing mud.
The paternalism is present along with the communalism in the dialogues about
the net building (Fig. 18.5). Fishers build their net based on their wide knowledge
about the geomorphology of the estuary, the species, and their life cycles: “I have
nets for every species… we try to take care of the animals. We know their reproduc-
tive cycles, where we can find them, which sediment, which channel, we have refer-
ences and take notes. We know where to fish. I do admit that we, the fishers,
sometimes overexploit our resources. But nothing in comparison with those big
boats out there (in reference to El Rincón area). We try to avoid overfishing by using
these nets that are made and designed only for this estuary and the species we fish”
(DG, fisher).
Finally, based on their speeches, it was possible to identify that they feel excluded
by the conservationist, after the actions are taken in order protect the estuary, and on
the other side, they consider that the petrochemical complex is the main cause of the
18  Small-Scale Artisanal Fishers and Socio-environmental Conflicts in Estuarine… 511

ecosystem decline, in complicity with the CGPBB. Fishers do not see these actors
as part of the estuary, and a great distance can be inferred between these actors in
one side and, on the other side, fishers and the estuary: “Everything it’s ok with the
workers of the protected areas, but I want them far… When an endangered animal
appears, we always call the people from FRAAM to take care of them, but they
come and give us an endless speech of conservation. Anyway, they live a different
reality in comparison to us. Unlike us, they have the food guaranteed every day”
(DG, fisher).
“Well… they’re good people, but they know everything, you know? Regarding
the sea, turtles and the shellfishes they’re kind of annoying, but we have to let them
speak, they like to talk a lot, like a burrowing parakeet (laughs, in reference to the
noise emitted by Cyanoliseus patagonus). Anyway, they’re good people, but too
idealistic… I don’t know, they live in their own world, they’re from a different
social class and they understand too little about fishing” (LF, fisher).
“They’re a bunch of vultures, where they can get benefits, they do it. They don’t
care about anything in order to fill their pockets with money. They don’t care if the
ria dies, if we sail in rotten mud. Nothing… If you don’t speak Chinese or English,
these guys won’t sit and dialogue with you” (MF, ex-fisher in reference to
the CGPBB).

18.3  U
 se of Fishers’ Ecological Knowledge (FEK)
in the Conservation of Coastal Systems: The Bahía
Blanca Estuary Case

Over the past decades, many ecological studies were built using fishers’ knowledge,
but only at the end of the 1990s that fishers’ knowledge was taking seriously in
scientific studies. A study by Johannes et al. (2000) reveals several cases in which
fishers’ knowledge reflected doubt and was discarded by the scientific community,
what led them to arrive to some uncertain and wrong conclusions about fishes and
marine mammals’ biology, ethology, and ecology. Therefore, the use of fishers’
ecological knowledge (hereafter, FEK) in science is recent and still discussed
among the most rational and modern sectors of science.
Leaving those sectors aside, FEK is now widely used in ecological studies in the
North Hemisphere (Johannes 1993; Neis 1998; Neis et al. 1999; Zukowski et al.
2011; Ambrose et al. 2014; Anbleyth-Evans 2018; Anbleyth-Evans and Lacy 2019,
among others) and has proven to be a priceless management tool in the South,
­especially in Brazil (Leite and Gasalla 2013; Silvano and Begossi 2016; Messias
et al. 2019; Thykjaer et al. 2019, others).
In the last decades, marine environments started to show “symptoms” of the
anthropic activities (i.e., dredging, oil spills, industrial development, tourism, big-
ger urban settlements with untreated sewage discharges, agricultural and cattle
activities, among others) (Marcovecchio et  al. 2008); eutrophication process; the
512 D. M. Truchet and M. B. Noceti

inputs of toxics substances like PCB, PAHs, OC trace metals, and other emerging
concern pollutants (Arias et al. 2010; Spetter et al. 2015; Buzzi and Marcovecchio
2018; Fernández Severini et  al. 2019); and the acidification of the water and the
increasing CO2 emissions. Degradation affected trophic food webs and worldwide
fisheries (Allison et al. 2009; López Abbate et al. 2017), and the complexity of these
environmental problems stresses more the need to develop new ways of manage-
ment and conservation of the marine ecosystems that combine scientific and local
knowledge, specifically FEK.
Unlike the Brazilian coastal areas, studies exploring the possible application of
FEK in conservation are few in Argentina, and they are still in development, espe-
cially in the Samborombón Bay (Carman and Gónzalez Carman 2016; González
Carman and Carman 2018) and more recently in the Bahía Blanca Estuary (Truchet
et al. in review). In the Samborombón Bay, FEK was used along with conservation-
ists and NOGs to help in reducing the bycatch of endangered animals (Caretta
caretta, Chelonia mydas, Dermochelys coriacea, Pontoporia blainvillei).
Reconversion involved the implementation of a fishing device using hooks. It was
not fully successful because fish biomass was lower with hooks than with nets, and
hence, the profits were lower. The importance of these studies was that fishers dem-
onstrated to have a wide ecological knowledge about the species and how to help
endangered animals in case of possible bycatch.
On the other hand, as we stated before, in the Bahía Blanca Estuary, the studies
are incipient. The characteristics of fisheries have been barely described (Noceti
2017; Truchet et  al. 2020a, b), though fishers’ ecological knowledge has been
widely used in biological samplings, scientific studies, and doctoral thesis (Lopez
Cazorla 1987; Lopez Cazorla 2004). Recently, demonstrated that fishers’ percep-
tions and knowledge toward the Bahía Blanca Estuary as a human-impacted sacrifi-
cial zone are in accordance to several scientific studies (Marcovecchio et al. 2008;
Arias et al. 2010), but there are some gaps in knowledge that could be fulfilled with
fishers’ experiences. There are some examples that reenforce what we stated before,
like the case of the flat fish, the “lenguado” (Paralichthys orbygnatus), the mussel
(Brachidontes rodriguezii), and the Pacific oyster (Magallana gigas). In the case of
P. orbygnatus, the ethnobiological and anthropological research with fishers led us
to think that they present malformations due to pollution: in this case, fishers recog-
nize the healthy anatomical and morphometric body conditions of fishes: “I’ve
observed soles with spots and protuberances as if they were tumors… I don’t know
if it’s because of petroleum as many fishers believe. I think it’s due to all the pollut-
ants and the sewage water without treatment from Rosales that are thrown into the
estuary” (DG, fisher).
Flat fishes’ malformations have been ignored in the scientific literature until a
study by Castellini et al. (2018); however these authors indicate that the abnormali-
ties are due to natural constrains during the ontogeny and do not mention the pos-
sible effects of pollution on a highly impacted estuary. Nevertheless, fishers declare
that they have not seen them in the past, before the industrial settlements and the
untreated sewage plants. The same thing happened with the decrease in the popula-
tions of B. rodriguezii; though there are no studies regarding this problem, fishers
18  Small-Scale Artisanal Fishers and Socio-environmental Conflicts in Estuarine… 513

strongly believe that constant dredging of the principal channel is a possible expla-
nation to the decline of this species: “There was a mortality mass of the scorched
mussel. I’m sure that was due to the dredging because the mud buried them and they
couldn’t breathe through their siphons” (DG, fisher). Recent studies on this species
by Buzzi and Marcovecchio (2018) recorded medium levels of some toxic trace
metals and PAHs that could be harmful for human consumption, but these studies
do not give a perspective on why this species tended to decrease its populations in
the estuary as fishers stated. In all these cases, where the scientific and fishers’
observations might not agree, and when observations diverge, they should both be
re-examined so that we can understand our ecosystem better (Johannes et al. 2000).
FEK and local knowledge have proven to be also a valuable tool to understand
long-term modifications in coastal areas and the distribution and changes of the
­species pools (Sievanen 2014; Gonçalves et al. 2019; Thykjaer et al. 2019), because
fishers have a sense of observation of the entire ecosystem since they navigate more
than 12 h daily over most of their lives. This information is also useful to detect
alien species and their possible ecological damage: “Decades ago, that grass (in
reference to Spartina alterniflora) wasn’t observed here in Villa del Mar, it’s not
native, people from FRAAM (the NGO) say it is, but I think it’s not. I believe it
came from the great cargo ships and the sediments removed from dredging the prin-
cipal channel that was thrown here. Now you can see it everywhere, and we don’t
know what to do with it” (MD, ex fisher). A parallel was detected here because a
study by Bortolus et al. (2015) demonstrated that S. alterniflora is a species original
from North America, while the older fishers always knew it was an alien species that
they did not observe in the estuary. It is also remarkable that through observation,
fishers know that S. alterniflora is an alien species, while conservationists still deny
it – probably through misinformation – and want to protect it, instead of trying to
find different managing strategies to mitigate the possible effects on the ecosystem
of invasive species.
Our recent unpublished data (Truchet et al. in review) also enhances the value of
FEK in long-term studies for the Bahía Blanca Estuary, a scenario that is under the
global warming change and industrial development (Arias et al. 2013; Marcovecchio
et al. 2013; López Abbate et al. 2017). We found that the fishers have a wide knowl-
edge about the diet, migration, and reproduction of fishes as well as the geomor-
phology and long-term climatic condition of the estuary. Nevertheless, their
knowledge is dismissed in management programs, especially in the case of the
industrial and port companies that see them as “annoying” and do not embrace their
culture as they do with NGOs and scientists who represent an expert and imperialist
rationalist knowledge. FEK has demonstrated to be a valuable tool that could change
the modern epistemological and technocratic ways of management in human-­
impacted estuarine systems of developing countries, such as the Bahía Blanca
Estuary.
514 D. M. Truchet and M. B. Noceti

18.4  Future Perspectives and Recommendations

Nowadays, thousands of communities live in coastal areas around the world. Daily,
most of them cope with the consequences of degradation, pollution, overexploita-
tion, and overcapitalization of natural resources. These socio-environmental con-
flicts are an invitation to re-think the ways in which conservation and management
of estuarine and coastal areas have been developed. These problems are complex
and have not been solved nor achieved through reductionist, disciplinary, and neu-
tral knowledge models of science, where emotion and common sense are dissoci-
ated. This disassociation has deprived scientists of the valuable local knowledge
acquired through peoples’ emotions and experiences based on an intimate relation-
ship with the sea.
The new ways of management and resolution of environmental conflicts in estua-
rine and coastal systems need to be solved with more tangible and participative
strategies conceived from and for the communities that are attached to the environ-
ment they are part of. Thus, it is fundamental to generate critical research for man-
agement programs that integrate the human compound that is part of the nature and
promote the collective creation of knowledge, offering not only a sustainability
economy but also ecological and human justice for the most vulnerable sectors.
This is where small-scale artisanal fishers of the Bahía Blanca Estuary play a
role, because their livelihoods are in risk to be lost due to neoliberal policies that
favor big foreign companies, the port consortium, and protected areas that are the
last link for denying the identities connected with the sea. Also, as scientist, we must
recognize that rationalist ecological management did not consider their voices for
the conservation of this ecosystem. Therefore, these subjects and their cultures live
in a vulnerable scenario, where they do not take place in the arenas where manage-
ment decisions are taken. Now, our main goal lies in not taken for granted what the
different cultures inhabiting the estuary think; moreover, we need to work side by
side in order to create dialogists and participative science as well as new decolonial
paradigms in the management and conservation of this sacrificial estuarine territory.

Acknowledgments  Authors would like to thank to all the fishers of the Bahía Blanca Estuary and
their families and to Lic. Rocío M. Truchet (IHuCSo-Litoral, CONICET-UNL) for the photos of
the fishers and ports of the Bahía Blanca Estuary.

References

Allison EH, Perry AL, Badjeck M-C, Neil Adger W, Brown K, Conway KD, Halls AS,
Pilling GM, Reynolds JD, Andrew NL, Dulvy NK (2009) Vulnerability of national econo-
mies to the impacts of climate change on fisheries. Fish Fish 10(2):173–196. https://doi.
org/10.1111/j.1467-­2979.2008.00310
Álvarez MC, Ruiz G, Navia D, Cortes C (2017) La visualización femenina en la pesca artesanal:
transformaciones culturales en el sur de Chile. Polis, Revista Latinoamericana 16(46):175–191
18  Small-Scale Artisanal Fishers and Socio-environmental Conflicts in Estuarine… 515

Ambrose WG, Clough LM, Johnson JC, Greenacre M, Griffith DC, Carroll ML, Whiting A (2014)
Interpreting environmental change in coastal Alaska using traditional and scientific ecological
knowledge. Front Mar Sci 1:680. https://doi.org/10.3389/fmars.2014.00040
Anbleyth-Evans J (2018) Aggregate dredging impacts in South East England: improving ecologi-
cal health by integrating fisher ecological knowledge with scientific research. Mar Pollut Bull
55(9):34–64. https://doi.org/10.1016/j.marpolbul.2018.06.051
Anbleyth-Evans J, Lacy SN (2019) Feedback between fisher local ecological knowledge and sci-
entific epistemologies in England: building bridges for biodiversity conservation. Mar Stud
18:189–203. https://doi.org/10.1007/s40152-­019-­00136-­3
Arias AH, Fernández Severini MD, Delucchi F, Freije RH, Marcovecchio JE (2010) Persistent
pollutants monitoring in a South Atlantic coastal environment. Case study: the Bahía Blanca
Estuary. In: Ortiz AC, Griffin NB (eds) Pollution monitoring. Nova Science Publishers,
New York, pp 1–30
Arias AH, Piccolo MC, Spetter CV, Freije H, Marcovecchio JE (2013) Lessons from multi decadal
oceanographic monitoring at an estuarine ecosystem in Argentina. I J Environ Res 6(1):219–
234. https://doi.org/10.22059/ijer.2011.488
Bortolus A, Carlton JT, Schwindt E (2015) Reimagining South American coasts: unveiling the
hidden invasion history of an iconic ecological engineer. Div Distrib 21:1267–1283. https://
doi.org/10.1111/ddi.12377
Buzzi NS, Marcovecchio JE (2018) Heavy metal concentrations in sediments and in mussels
from Argentinean coastal environments, South America. Environ Earth Sci 77:321. https://doi.
org/10.1007/s12665-018-7496-1
Carman M (2011) Las trampas de la Naturaleza. Medio ambiente y segregación. Fondo de Cultura
Económica, Buenos Aires, p 285
Carman M, Gónzalez Carman V (2016) La fragilidad de las especies: tensiones entre biólogos y
pescadores artesanales en torno a la conservación marina. Entnográfica 20:411–438. Available
at: https://journals.openedition.org/etnografica/4333
Castellini DL, González-Castro M, Lopez Cazorla A, Díaz De Astarloa JM (2018) Abnormalities in
flatfishes of the South-West Atlantic Ocean. J Fish Biol 92:1225–1234. https://doi.org/10.1111/
jfb.13578
Charles AT (1992) Fishery conflicts. Mar Policy 16(5):379–393. https://doi.
org/10.1016/0308-­597x(92)90006-­b
Cheung WWL, Lam VWY, Sarmiento JL, Kearney K, Watson R, Zeller D, Pauly D (2010) Large-­
scale redistribution of maximum fisheries catch potential in the global ocean under climate
change. Glob Chang Biol 16(1):24–35. https://doi.org/10.1111/j.1365-­2486.2009.01995.x
Colautti D, Baigun C, Lopez Cazorla A, Llompart F, Molina JM, Suquele P, Calvo S (2010)
Population biology and fishery characteristics of the smooth-hound Mustelus schmitti in Anegada
Bay, Argentina. Fish Res 106(3):351–357. https://doi.org/10.1016/j.fishres.2010.09.004
Coulthard S, Johnson D, McGregor JA (2011) Poverty, sustainability and human wellbeing: a
social wellbeing approach to the global fisheries crisis. Glob Environ Chang 21(2):453–463.
https://doi.org/10.1016/j.gloenvcha.2011.01.003
De la Cadena M, Blaser M (2018) A world of many worlds. Duke University Press, Durham, p 224
Descola P (2003) Constructing natures: symbolic ecology and social practice. In: Descola
P, Pálsson G (eds) Nature and society. Anthropological perspectives. Taylor and Francis,
New York, pp 82–102
Descola P (2005) Par-delà la nature et culture. Bibliothèque des sciences humaines, Gallimard,
Paris, p 624
Descola P, Pálsson G (2003) Introduction. In: Descola P, Pálsson G (eds) Nature and society.
Anthropological perspectives. Taylor and Francis, New York, pp 1–22
DuBois C, Zografos C (2012) Conflicts at sea between artisanal and industrial fishers: inter-­
sectoral interactions and dispute resolution in Senegal. Mar Policy 36(6):1211–1220. https://
doi.org/10.1016/j.marpol.2012.03.007
516 D. M. Truchet and M. B. Noceti

Ellen RF (2003) The cognitive geometry of nature: a contextual approach. In: Descola P, Pálsson
G (eds) Nature and society. Anthropological perspectives. Taylor and Francis, New  York,
pp 103–124
FAO (2017a) Promover la igualdad de género y el empoderamiento de las mujeres en la pesca y la
acuicultura, p. 12. Available at: http://www.fao.org/3/a-­i6623s.pdf
FAO (2017b) Towards gender-equitable small-scale fisheries governance and development  – a
handbook. In: Biswas N (Ed) Support of the implementation of the voluntary guidelines for
securing sustainable small-scale fisheries in the context of food security and poverty eradica-
tion, Rome, Italy, p. 154. Available at: http://www.fao.org/3/a-­i7419e.pdf
Fernández Severini MD, Villagrán DM, Buzzi NS, Chatelain Sartor G (2019) Microplastics in
oysters (Crassostrea gigas) and water at the Bahía Blanca Estuary (Southwestern Atlantic):
an emerging issue of global concern. Reg Stud Mar Sci 32:100829. https://doi.org/10.1016/j.
rsma.2019.100829
García-Quijano CG, Valdés-Pizzini M (2015) Ecosystem-based knowledge and reasoning in tropi-
cal, multispecies, small-scale fishers’ LEK: what can fishers LEK contribute to coastal ecologi-
cal science and management? In: Fischer J, Jorgensen J, Josupeit H, Kalikoski D, Lucas C (eds)
Fishers’ knowledge and the ecosystem approach to fisheries: applications, experiences and
lessons in Latin America. Publisher: FAO Fisheries and Aquaculture Technical Paper, Rome,
p 591
Giarraca N, Teubal M (2013) Las actividades extractivas en la Argentina. In: Giarraca N, Teubal M
(eds) Actividades extractivas en expansión. Reprimarización de la economía en la Argentina?
Antropofagia, Buenos Aires, pp 19–44
Girado A (2012) Resistencias y conflictos socioambientales en Tandil. Un estudio de caso Revista
Sociedad y Equidad 4. http://www.sye.uchile.cl/index.php/RSE/article/view/20930/22123
Gonçalves PH, Rego AE, Medeiros P (2019) “There was a virgin forest here; it was all woods”:
local perceptions of landscape changes in Northeastern Brazil. Ethnobio Conserv 8:4. https://
doi.org/10.15451/ec2019-­01-­8.04-­1-­17
González Carman V, Carman M (2018) A coexistence of paradigms: understanding human–envi-
ronmental relations of fishers involved in the bycatch of threatened marine species. Conserv
Soc 16(2):205–216. https://doi.org/10.4103/cs.cs_17_45
Harper S, Dirk Z, Hauzer M, Pauly D, Sumaila U (2013) Women and fisheries: contribu-
tion to food security and local economies. Mar Policy 39:56–63. https://doi.org/10.1016/j.
marpol.2012.10.018
Heredia Chaz E (2014) De la Responsabilidad a la Contaminación Social Empresaria: la ingeni-
ería social del Polo Petroquímico de Bahía Blanca. Degree in History Thesis, Universidad
Nacional del Sur, Available at: http://repositoriodigital.uns.edu.ar/bitstream/123456789/528/1/
Heredia%20Chaz,%20Emilce.%20Tesina.pdf
Ingold T (2000) The perception of the environment. Essays on livelihood, dwelling and skill.
Routledge, New York, p 465
Ingold T (2003) The optimal forager and economic man. In: Descola P, Pálsson G (eds) Nature and
society. Anthropological perspectives. Taylor and Francis, New York, pp 25–44
Johannes RE (1993) Integrating traditional ecological knowledge and management with environ-
mental impact assessment. In: Inglish JT (ed) Traditional ecological knowledge: concepts and
cases. International Development Research Centre, Ottawa, pp 33–40
Johannes RE, Freeman MMR, Hamilton RJ (2000) Ignore fishers’ knowledge and miss the boat.
Fish Fish 1:257–271
Johnson DS (2018) The values of small-scale fisheries. In: Johnson DS, Acott TG, Stacey N,
Urquhart J (eds) Social wellbeing and the values of small-scale fisheries. MARE Publication
Series, Springer, Cham, pp 1–22
Latour B (2004) Politics of nature. How to bring the sciences into democracy? Harvard University
Press, Cambridge, MA/London, p 320
Leite MCF, Gasalla MA (2013) A method for assessing fishers’ ecological knowledge as a practi-
cal tool for ecosystem-based fisheries management: seeking consensus in Southeastern Brazil.
Fish Res 145:43–53. https://doi.org/10.1016/j.fishres.2013.02.013
18  Small-Scale Artisanal Fishers and Socio-environmental Conflicts in Estuarine… 517

Llompart FM, Colautti DC, Baigún CRM (2017) Conciliating artisanal and recreational fisheries in
Anegada Bay, Argentina. Fish Res 190:140–149. https://doi.org/10.1016/j.fishres.2017.01.011
London S, Rojas ML, Ibáñez MM, Scordo F, Huamantinco Cisneros MA, Bustos ML, Perillo
GME, Piccolo MC (2017) Characterization of an artisanal fishery in Argentina using the social-­
ecological systems frameworks. I J Commons 11:11–69. https://doi.org/10.18352/ijc.534
López Abbate MC, Molinero JC, Guinder VA, Perillo GME, Sommer U, Spetter CV, Marcovecchio
JE (2017) Time-varying environmental control of phytoplankton in a changing estuarine sys-
tem. Sci Total Environ 609:1390–1400. https://doi.org/10.1016/j.scitotenv.2017.08.002
Lopez Cazorla A (1987) Contribución al conocimiento de la ictofauna marina del área de Bahía
Blanca. Doctoral Thesis, Universidad Nacional de la Plata, pp. 247
Lopez Cazorla A (2004) Peces del estuario de Bahía Blanca. In: Píccolo MC, Hoffmeyer MS (eds)
El ecosistema del estuario de Bahía Blanca. Bahía Blanca, Instituto Argentino de Oceanografía
(IADO, CONICET-UNS), Bahía Blanca, pp 191–220
Lugones B (1883) Una excursión al sur, el puerto de Bahía Blanca. La Nación, Buenos Aires
14:3753
Luna López L (2011) Estudio de caso sobre el rol de las mujeres en comunidades pesqueras del
Golfo de Fonseca, Marcovia, Choluteca. Degree Thesis, Escuela Agrícola Panamericana,
Zamorano, Honduras, p. 33. Available at: https://bdigital.zamorano.edu/bitstream/11036/408/1/
IAD-­2011-­T016.pdf
Machado Aráoz H (2010) Imperialismo ecológico y racismo ambiental. Una lectura EcoBiopolítica
sobre las Industrias extractivas en el Sur. Aportes Científicos desde Humanidades 8:1897–1911
Marcovecchio JE, Botté SE, Delucchi F, Arias AH, Fernández Severini MD, de Marco SG,
Tombesi N, Andrade S, Ferrer L, Freije HR (2008) Pollution processes in Bahía Blanca estua-
rine environment. In: Neves R, Barreta J, Mateus M (eds) Perspectives on integrated coastal
zone management in South America. IST Press, Lisbon, pp 303–316
Marcovecchio JE, Buzzi NS, Tartara MN, Spetter CV, Simonetti P (2013) Potential effects of cli-
mate changes on the marine ecosystem stability: assessment of the water quality. In: Arias AH,
Menéndez C (eds) Marine ecology in a changing world. CRC Press, London, pp 1–43
McGregor A (2018) Foreword: social wellbeing and the values of small-scale fisheries. In: Johnson
DS, Acott TG, Stacey N, Urquhart J (eds) Social wellbeing and the values of small-scale fisher-
ies. MARE Publication Series, Springer, Cham
Merlinsky G (2013) La cuestión ambiental en la agenda pública. In: Merlinsky G (ed) Cartografías
del conflicto ambiental en Argentina. Buenos Aires, CLACSO-CICCUS, pp 19–60
Messias MA, Alves TIP, Melo CM, Lima M, Rivera-Rebella C, Rodrigues DF, Madi RR (2019)
Ethnoecology of Lutjanidae (snappers) in communities of artisanal fisheries in Northeast
Brazil. Ocean Coast Manag 181:104866. https://doi.org/10.1016/j.ocecoaman.2019.104866
Nahuelhual L, Saavedra G, Blanco G, Wesselink E, Campos G, Vergara XV (2018) On super fish-
ers and black capture: images of illegal fishing in artisanal fisheries of southern Chile. Mar
Policy 95:36–45. https://doi.org/10.1016/j.marpol.2018.06.020
Nahuelhual L, Saavedra G, Mellado MA, Vergara XV, Vallejos T (2019) A social-ecological trap
perspective to explain the emergence and persistence of illegal fishing in small-scale fisheries.
Mari Stud. https://doi.org/10.1007/s40152-­019-­00154-­1
Neis B (1998) Fishers’ ecological knowledge and the identification and management of localized
populations of Atlantic Cod (Gadus morhua) in Northeast Newfound-land. In: von Herbin H,
Kornfield II, Tupper M, Wilson J (eds) The implications of localized fisheries stocks. NRAES,
Ithaca, pp 217–238
Neis B, Schneider DC, Felt L, Haedrich RL, Fischer J, Hutchings JA (1999) Fisheries assessment:
what can be learned from interviewing resource users? Can J Fish Aquat Sci 756:1949–1963
Noceti MB (2013) Tejedores de redes, pescadores y anécdotas que se desvanecen. Miradas antrop-
ológicas en torno a saberes y modo de vida de pescadores artesanales del sudoeste bonae-
rense Conference presente at: VII Jornadas Santiago Wallace de Investigación en Antropología
Social. Available at: http://www.11caas.org/conf-­cientifica/comunicacionesDocGetfile.php
518 D. M. Truchet and M. B. Noceti

Noceti MB (2017) ¿Reserva, puerto o ría? Conflicto socioambiental en el estuario de Bahía Blanca,
Argentina. Etnografías Contemporáneas 3(4):64–91
Noceti MB (2018) Maritimidad vs Maritimización. Ontologías y Territorialidades en disputa en el
sudoeste bonaerense, Argentina. Conference Presented at the Universidad Autónoma de Nuevo
León. Instituto de Investigación en Ciencias Sociales, Monterrey, Mexico, p 46
Oliveira OM, Silva VL (2012) ¿Pescadoras o “mujeres de pescador”? La regulación jurídica de
la pesca artesanal en Brasil y el problema del no reconocimiento del trabajo profesional de
las pescadoras en Santa Catarina-Brasil. Investigación y género, inseparables en el presente
y en el futuro. Conference presented at: IV Congreso Universitario Nacional “Investigación y
Género”, Sevilla, pp. 1327–1352
Pálsson G (2003) Human-environmental relations: orientalism, paternalism and communalism.
In: Descola P, Pálsson G (eds) Nature and society. Anthropological perspectives. Taylor and
Francis, New York, pp 63–81
Peron F, Rieucau J (1996) La maritimité aujourd’hui. Ed. L’Harmattan, Paris, p. 102
Quist LM (2019) Fishers’ knowledge and scientific indeterminacy: contested oil impacts in
Mexico’s sacrifice zone. Mar Stud 18:65–76. https://doi.org/10.1007/s40152-­018-­0123-­7
Rodríguez S (2009) “A veces las mujeres también entramos al mar”. La pesca de camarón en
Machalilla. In: Paulson S, Poats S, Argüello M (eds) Huellas de género en el mar, el parque y el
páramo. EcoCiencia, Corporación Grupo Randi Randi y Abya Yala, Quito, Ecuador, pp 12–33
Rojas Herrera S (2018) Significaciones identitarias asignadas al mar desde las mujeres que habitan
territorios marinos costeros en Costa Rica y Nicaragua. Doctoral Thesis, Universidad Nacional
de Costa Rica, Costa Rica, p  216. Available at: https://www.repositorio.una.ac.cr/bitstream/
handle/11056/14797/Rojas%20Herrera%2C%20Silvia%20Elena.pdf?sequence=1&isAllowe
d=y
Sack R (1986) Human territoriality: its theory and history. Cambridge University Press, Cambridge,
p 272
Sievanen L (2014) How do small-scale fishers adapt to environmental variability? Lessons from
Baja California, Sur, Mexico. Mar Stud 13:9. https://doi.org/10.1186/s40152-­014-­0009-­2
Silvano RAM, Begossi A (2016) From ethnobiology to ecotoxicology: Fisher’s knowledge on
trophic levels as indicators of bioaccumulation in tropical marine and freshwater fishes.
Ecosystems 19:1310–1324. https://doi.org/10.1007/s10021-­016-­0002-­2
Spetter CV, Buzzi NS, Fernández EM, Cuadrado DG, Marcovecchio JE (2015) Assessment of the
physicochemical conditions sediments in a polluted tidal flat colonized by microbial mats in
Bahía Blanca Estuary (Argentina). Mar Pollut Bull 9(2):491–505. https://doi.org/10.1016/j.
marpolbul.2014.10.008
Svampa M (2012) Pensar el desarrollo desde América Latina. Available at: http://maristellas-
vampa.net/archivos/ensayo56.pdf
Thykjaer V, Rodrigues L, Haimovici M, Cardoso LG (2019) Long-term changes in fishery
resources of an estuary in southwestern Atlantic according to local ecological knowledge.
Fisheries Manag Ecol. https://doi.org/10.1111/fme.12398
Tola F (2011) Reflexiones dislocadas. Facultad de Filosofía y Letras  – UBA, Asociación Civil
Rumbo Sur, Buenos Aires, Argentina, p 208
Truchet DM (2018) De espaldas al mar: conflictividad socioecológica en el Estuario de Bahía
Blanca (Buenos Aires, Argentina). La pesquería artesanal frente a las políticas neoextractivis-
tas. Papeles del Centro de Investigaciones 8(19):9–26
Truchet DM, Noceti MB, Villagrán DM, Truchet RM (2020a) Vindicating fishers’ ecological
knowledge in a scenario of climate change and industrial development: the Bahía Blanca
Estuary, Southwestern Atlantic Ocean. Mar Stud (in revision)
Truchet DM, Truchet RM, Noceti MB (2020b) Roles y relaciones de género en contextos de
pesca artesanal: una reconstrucción a partir de las narrativas orales de varones pescadores del
Estuario de Bahía Blanca. Revista de Estudios Sociales y Marítimos 13(16):64–86
Tuan Y-F (1974) Topophilia: a study of environmental perception, attitudes, and values. Prentice-­
Hall, Englewood Cliffs, p 260
18  Small-Scale Artisanal Fishers and Socio-environmental Conflicts in Estuarine… 519

Vasconcellos M, Diegues A, Kalikosli D (2011) Coastal fisheries of Brazil. In: Salas S, Chuepadgee
R, Charles A, Seijoo JC (eds) Coastal fisheries of Latin America and the Cari Bahía Blanca
Estuaryan, FAO Fisheries and Aquaculture Technical Paper. No. 544. FAO, Rome, pp 73–116
Viveiros de Castro E (1996) Os pronomes cosmológicos e o perspectivismo amerindio. Mana
2(2):115–144
Wagner L (2016) Problemas ambientales y conflicto social en Argentina. Movimientos socioambi-
entales en Mendoza. La defensa del agua y el rechazo a la megaminería en los inicios del Siglo
XXI. Doctoral Thesis, Universidad Nacional de Quilmes, Argentina. Available at: http://ridaa.
unq.edu.ar/handle/20.500.11807/192
Yory C (2013) El desarrollo territorial integrado: una estrategia sustentable de construcción social
del territorio, en el contexto de la globalización, a partir del concepto de Topofilia. Universidad
Piloto de Colombia, Bogotá, p 388
Yory C (2015) La construcción social del hábitat como estrategia de integración social. Universidad
Piloto de Colombia, Bogotá, p 434
Zukowski S, Curtis A, Watts RJ (2011) Using fisher local ecological knowledge to improve man-
agement: the Murray crayfish in Australia. Fish Res 110(1):120–127. https://doi.org/10.1016/j.
fishres.2011.03.020
Chapter 19
Estuarine Environmental Monitoring
Programs: Long-Term Studies

Jorge E. Marcovecchio, Sandra E. Botté, Silvia G. De Marco,


Andrea Lopez Cazorla, Andrés H. Arias, Mónica Baldini,
María Amelia Cubitto, Sandra M. Fiori, Ana L. Oliva, Noelia La Colla,
Gabriela Blasina, Juan Manuel Molina, Pia Simonetti, Analía V. Serra,
Vanesa L. Negrín, Ana C. Ronda, and Marcelo T. Pereyra

19.1  What Is an Environmental Monitoring Program?

A continuous observation and control system of measures and evaluations for a


defined purpose is called “monitoring.” This is an important tool within the impact
assessment process and in any vigilance and control program (Pali and Swaans
2013; Valle Junior et al. 2015). There is currently a strong consensus that environ-
mental monitoring is not an end-point in itself but an essential step in environmental
management processes (Stelzenmüller et al. 2013). Taking into account the previ-
ously mentioned concepts, the importance of monitoring within different processes

J. E. Marcovecchio (*)
Instituto Argentino de Oceanografía (IADO – CONICET/UNS). CCT CONICET Bahía
Blanca, Bahía Blanca, Argentina
Universidad Tecnológica Nacional, Facultad Regional Bahía Blanca (UTN-BHI),
Bahía Blanca, Argentina
Facultad de Ingeniería, Universidad FASTA (FI-UFASTA), Mar del Plata, Argentina
Academia Nacional de Ciencias Exactas, Físicas y Naturales (ANCEFN),
Ciudad Autónoma de Buenos Aires, Argentina
e-mail: jorgemar@iado-conicet.gob.ar
S. E. Botté · A. L. Cazorla · S. M. Fiori · G. Blasina · J. M. Molina · V. L. Negrín
A. C. Ronda
Instituto Argentino de Oceanografía (IADO – CONICET/UNS). CCT CONICET Bahía
Blanca, Bahía Blanca, Argentina
Departamento de Biología, Bioquímica y Farmacia, Universidad Nacional del Sur (DBByF-­
UNS), Bahía Blanca, Argentina
S. G. De Marco
Facultad de Ingeniería, Universidad FASTA (FI-UFASTA), Mar del Plata, Argentina
Departamento de Biología, Facultad de Cs.Exactas y Naturales, Universidad Nacional de Mar
del Plata (UNMdP), Mar del Plata, Argentina

© Springer Nature Switzerland AG 2021 521


S. M. Fiori, P. D. Pratolongo (eds.), The Bahía Blanca Estuary,
https://doi.org/10.1007/978-3-030-66486-2_19
522 J. E. Marcovecchio et al.

of human activity can be observed. In addition, and as it is rightly mentioned, it is a


fundamental tool within all that development or procedure that is desired to be car-
ried out in a controlled and safe way (Collins et al. 2012).
The monitoring programs consist of carrying out a permanent surveillance of a
natural system, controlling the state and evolution of its variables, quantifying the
variations that occur, and identifying the reasons that generated them. There are dif-
ferent work strategies, both spatial and temporal, and can be developed either exclu-
sively based on field data or based on experimental data and even combining both
(Lindenmayer and Likens 2010). When these programs are carried out for a long
period of time (e.g., decades), the set of information they provide supports a solid
basis on which it is possible to accurately characterize the structure and operation of
the system under study (Biber 2013). As long as the continuity of monitoring is
maintained, it can be argued that as the program has more seniority (therefore more
information), it is easier to identify a significant anomaly or variation in any of the
parameters studied, taking into account that the distribution of “normal” values will
​​
be very dense, and therefore those that move away from that model will be easily
distinguishable (Gray and Shimshack 2011).
Another type of alternative work is the application of environmental specimen
banks (ESBs) that can be very useful, since they allow the establishment of environ-
mental monitoring networks (Snyder et al. 2013; Hsu and Zomer 2014) with the
objective of early detection of changes in the parameters evaluated (i.e., increasing
concentrations of several contaminants). This process constitutes a development of
real-time monitoring (Viana et al. 2010; Fang et al. 2014), which results from enor-
mous application in management and control processes. In this case, and after hav-
ing developed a monitoring program for a period long enough to have the system’s

A. H. Arias
Instituto Argentino de Oceanografía (IADO – CONICET/UNS). CCT CONICET Bahía
Blanca, Bahía Blanca, Argentina
Departamento de Química, Universidad Nacional del Sur (DQ-UNS),
Bahía Blanca, Argentina
M. Baldini
Departamento de Agronomía, Universidad Nacional del Sur (DA-UNS),
Bahía Blanca, Argentina
M. A. Cubitto
Departamento de Biología, Bioquímica y Farmacia, Universidad Nacional del Sur (DBByF-­
UNS), Bahía Blanca, Argentina
A. L. Oliva · N. La Colla · P. Simonetti · A. V. Serra
Instituto Argentino de Oceanografía (IADO – CONICET/UNS). CCT CONICET Bahía
Blanca, Bahía Blanca, Argentina
M. T. Pereyra
Departamento de Química, Universidad Nacional del Sur (DQ-UNS),
Bahía Blanca, Argentina
19  Estuarine Environmental Monitoring Programs: Long-Term Studies 523

operation characterized, copies of one or several key species of the studied environ-
ment are stored and used as a permanent reference to compare values of the evalu-
ated parameter obtained at different times (Küster et al. 2015).
A particular study situation is the use of environmental indicators (e.g., bioindi-
cators, geoindicators, biomarkers, etc.) that allow monitoring of the system through
variations recorded in a species, type of sediment, molecules, etc. (Martínez-Haro
et al. 2015). It is a very practical and dynamic method, but it requires a very deep
and consolidated knowledge of the evaluated environment. The incorrect, ill-­
founded, or inconsistent application of these techniques can lead to false conclu-
sions and subsequently to unnecessary or inappropriate management measures
(Brewin et al. 2015).
In the particular case of water quality control methods, it is fair to say that this
type of monitoring has often been seen as a practical exercise by those who initiate
a program. The underlying philosophy of why monitoring should be done tends to
be neglected. Wondering why, however, leads to an essential step in formulating the
objectives for which the information is obtained. The choice of how to monitor is
then limited, considering that only some data sampling and processing actions allow
the objectives of the information to be achieved (Behmel et al. 2016). In fact, both
the choice of data analysis methods and statistical methods and procedures on how
to make observations should be made before sampling. In addition, the concept of
water quality must be defined a priori in terms of indicators that can be observed and
quantified (Hounslow 2018).

19.1.1  Long-Term Monitoring Programs

In the specialized bibliography, it is possible to find numerous definitions and


conceptual frameworks about the long-term monitoring program (LMP) process,
and in this sense and for the purposes of this review, we will consider the LMP as
“repeated empirical measurements based on field work, collected in continuously,
and subsequently analyzed for at least 10 years” (Goldsmith 2012).
In some viewpoints, the scientific value of monitoring tends to be underesti-
mated, without knowing that the structuring of an LMP necessarily requires excel-
lent scientific research work (Luthardt 2010; Wolf et al. 2013). In principle it should
be noted that good science and therefore good LMPs begin with good questions.
Some of the considerations that we must keep in mind to ask “good questions” are
the following:
• Think critically.
• Build strong conceptual models that represent the functioning of the ecosystems
studied (i.e., Grace et al. 2010).
• Evaluate real questions (de Oliveira Roque et al. 2018) on policies relevant to
management (Voget-Kleschin 2013).
524 J. E. Marcovecchio et al.

• Promote open dialogue between scientists and management leaders (Lawton


2007; Likens et al. 2009).
• Critically evaluate the design and opportunism of the proposed study.
In this regard, it is important to clarify that new visits (or revisits) to a site after
a prolonged absence are not considered as long-term monitoring ones (i.e., Smith
et al. 2007), as well as to eventual measurements discontinued over time.
Depending on its structure and design, we can differentiate three main types of
monitoring programs:
1. Passive or curiosity-driven monitoring. This is the monitoring devoid of specific
questions or underlying design of the study and with little or no purpose other
than curiosity. Its usefulness is very limited to address environmental problems
or to discover how the world works, because it is not driven by hypotheses and
because it lacks management interventions or different experimental treatments
that facilitate the scientific understanding of these things in response to ecosys-
tems to natural or human disturbances.
2. Mandatory or directed monitoring. In this type of monitoring, environmental
data is collected as stipulated by some requirement of government or political
legislation (i.e., a directive on monitoring the climate or river flow, etc.). The
usual thing is that they have enormous quality requirements and that they must
follow strict protocols to ensure the validity of the information collected. The
design of this type of monitoring does not attempt to identify or understand the
mechanism that produces a change in an ecosystem or an entity but only to
detect its existence.
3. Question-driven monitoring. This is monitoring guided by a conceptual model
and a rigorous study design. The use of a conceptual model will typically result
in a priori predictions that can be tested as part of program development. Often,
this type of learning is characterized by strong contrasting management inter-
ventions (Johnson 2012). In statistical language, these types of studies could be
called “studies with longitudinal interventions” (i.e., Lindenmayer et al. 2008).
Therefore, these approaches can lead to generating a strong predictive capacity
and allow the research team to raise new questions. Thus, a monitoring program
can evolve as the key questions change. Predictive ability can be of great value
to researchers, resource managers, and decision makers, in contrast to the simple
trend lines obtained from many targeted surveillance programs (Possingham
et al. 2012).

19.2  Structure of a Long-Term Monitoring Program

It is important to be clear that an LMP is not simply limited to a large collection of


data and its corresponding storage, but that it must fulfill a sequential series of steps
that constitute a monitoring cycle. The greater the number of monitoring cycles that
the LMP has, the greater its efficiency, the easier the interpretation of the generated
19  Estuarine Environmental Monitoring Programs: Long-Term Studies 525

Fig. 19.1  General scheme of a long-term monitoring program

information will be, the anomalies will be identified with greater speed and cer-
tainty, the evaluation tasks will be simplified, and the obtained conclusions to be
those that arrive will allow better management measures to be taken by the corre-
sponding authorities (Smith et al. 2012; Converse and Royle 2012).
A tentative scheme of an LMP would be as follows: In Fig. 19.1, there are some
properties of LMPs that distinguish them particularly and that frame many of their
advantages. The main ones of these distinctive characteristics that deserve to be
highlighted are:
• Based on good questions, which evolve
• Based on conceptual models
• Specified by active interlocutors
526 J. E. Marcovecchio et al.

• Developed by well-led solid work teams


• Sustained by adequate and continuous financing
• Generators of frequent use of information
• Reservoirs of intense scientific productivity
• Qualified through appropriate validation processes of data quality and calibra-
tion of field and laboratory methods
When the long-term monitoring programs are executed according to the scheme
mentioned above, the data generated and the reached conclusions are fundamentally
valuable for many purposes, including:
• Document and provide benchmarks against which changes in conditions or end-
points can be evaluated (i.e., Mayne and Zapico-Goñi 2017).
• Evaluate ecological responses to natural or experimental disturbances (Seidl
et al. 2016).
• Detect and quantify changes in the structure and function of the ecosystem
(Wrona et al. 2016).
• Identify ecological “surprises” (unexpected events) (i.e., Zhan et al. 2006).
• Establish evidence-based principles that support environmental legislation (i.e.,
laws that control the levels of pollutants in air and water) (Davies and Mazurek
2014).
• Generate significant new questions about ecological dynamics (Persson et  al.
2009).
• Provide empirical data to test ecological theory (McClure et al. 2016) and devel-
opment models, such as numerical simulation models (Neves et al. 2008).

19.3  A Little History of the Monitoring Programs

Although for a long time scientific works were carried out basically in the field, with
structures similar to monitoring programs, it can be considered as a pioneer case
that carried out by Jack Pearce in 1967, when he completed a benthic research pro-
gram in the marine basin from New York, with a view to determining the conse-
quences of dredging materials and sewage sludge in that marine area for more than
2 years. The research objects for that initial phase were about 100 species of benthic
and demersal organisms, a half-dozen toxic metal footprint, an indicator microor-
ganism (Escherichia coli), and the levels of organic matter in the sediments affected
by solid waste. In the early 1970s, reports were published with the data obtained,
including distributions of trace metals in sediments, levels of bacteria in and around
landfills, data on sick fish, and the main missing (or severely affected) benthic spe-
cies, like some amphipods, in and around the discharge operations’ centers
(Pearce 1998).
At approximately the same time, oil spills produced in different parts of the
planet (i.e., the English Channel, San Juan de Puerto Rico harbor, or in front of
Santa Barbara, CA, among others) drove the development of monitoring programs
19  Estuarine Environmental Monitoring Programs: Long-Term Studies 527

to evaluate the effects and consequences on the affected natural systems. A few
years later, scientists were reporting the arrival of agricultural pesticides to the
Monterey Canyon Bay, based on activities developed in the great valleys of
California. Others pointed to some ports as heavily polluted areas (i.e., London, in
the UK or Boston, Seattle, and Washington in the USA). These studies, and many
other reports developed in that period, significantly claimed long-term surveillance.
Soon national and international workshops and study groups were convened to dis-
cuss the strategies for monitoring and evaluation of pollutants and outline models
for the development of monitoring programs in both aquatic and terrestrial systems
(US National Academy of Sciences 1975, 1977). By 1978, scientific concern
focused on the preparation of long-term programs and the creation of specimen
banks, as well as the standardization of procedures for the collection and handling
of samples.
In the following decade, several national research and monitoring programs were
carried out, and global evaluations began; among others, the United Nations
Environment Program (UNEP) began to transfer the data and facilitated studies and
reports on the state of the marine environment (GESAMP 1990).
These programs were evolving and being improved, and many efforts were
mutating and focused on other previously ignored aspects, such as the measurement
of ecological, genetic, pathological, and/or physiological values that allowed to
identify the changes produced when coastal and marine life began to be impacted by
pollutants and physical pressures. Many of the measurements used as tools were
outlined in McIntyre and Pearce (1980), considering that they turned out to be more
realistic indicators of habitat degradation than conventional measurement and the
evaluation of the presence of toxins in biota, sediments, or water.

19.4  A Historical Turning Point: The Mussel Watch Program

The Mussel Watch Program was initially proposed as an alternative strategy to


understand the quality of marine waters (Goldberg 1975). Bivalves had been recog-
nized for their ability to concentrate four groups of pollutants from the waters within
the system they inhabit: (i) transuranic elements; (ii) halogenated hydrocarbons
(i.e., PCBs, DDTs); (iii) petroleum hydrocarbons; and (iv) heavy metals (Bryan
1973; Boyden 1974). These substances, mobilized by human activities, are intro-
duced through air and/or water to the marine environment and, subsequently, can
generate deleterious effects on human populations or marine life. The initial design
of this program involved a little more than 100 sampling stations in coastal waters
of the USA, analyzed quarterly and originally funded during the first 3 years
(1976–1978) by the US Federal Environmental Protection Agency (US EPA)
(Goldberg et al. 1978).
This concept of sentinel organism or Mussel Surveillance (Goldberg et al. 1978;
Phillips 1980) is a method for assessing the current state of chemical contamination
528 J. E. Marcovecchio et al.

of coastal waters and identifying geographic areas of interest and new or renewed
topics for scientific research.
The theoretical framework that allows the use of the common mussel (Mytilus
sp.), several oyster species (i.e., Crassostrea sp., Ostrea sp.), and other bivalves as
sentinel organisms includes the following premises (Goldberg et al. 1978; Phillips
1980; Farrington et al. 1983):
• Bivalves are cosmopolitan (widely distributed geographically). This feature min-
imizes the problems inherent in comparing the data between very different spe-
cies, with different life histories and types of relationships within their habitat.
• They are sedentary and are therefore more useful than mobile species as integra-
tors of chemical state of pollution in a given area.
• They can concentrate many chemicals with factors of 102–105 with respect to the
contents of seawater in their habitat. This significantly facilitates the measure-
ment of trace elements.
• When measuring the chemicals in bivalves, an evaluation of the biological avail-
ability of the compounds studied is being achieved.
• Unlike fish and crustaceans, bivalves have a very low or undetectable activity of
enzyme systems that metabolize many xenobiotics (i.e., polycyclic aromatic
hydrocarbons (PAHs) or polychlorinated biphenyls (PCBs), among others). This
allows more precise assessments of the magnitude of xenobiotic contamination
in the habitat of the bivalves studied.
• There are many local populations of bivalves that are relatively stable and large
enough to be repeatedly sampled, providing data on short-term and long-term
temporary changes in the concentrations of pollutants.
• They survive in pollution conditions that could often severely reduce or elimi-
nate other species.
• They can be transplanted from anchor sites and successfully maintained in dif-
ferent coastal intertidal waters, even with very different conditions. This allows
the identification of different toxic distribution processes between different areas,
detoxification mechanisms, recovery times, etc.
• They are marine species that are used for human food in almost all cultures and
ethnicities, which assign them high commercial value throughout the world.
Therefore, the measurement of chemical contamination in these species is of
interest to public health.
After that, Goldberg himself (1986) commented that another additional advan-
tage of this work strategy is that the biological half-lives of contaminants in bivalves
can last up to periods of months. There is, therefore, a spatial and temporal integra-
tion of pollutant levels in body tissues. Therefore, the use of bivalves offers advan-
tages over seawater and sediments for the determination of the contaminant, where
the integration period can vary from short (1 day or less) to long (usually 1 year or
plus). This situation is very advantageous for the design of surveillance programs.
This excellent evaluation tool began to spread rapidly throughout the world, i.e.,
since 1979 a Mussel Watch Program has been carried out in bivalves of the French
coastal zone: the National Network for the Observation of the Quality of the Marine
19  Estuarine Environmental Monitoring Programs: Long-Term Studies 529

Environment (Réseau National d’Observation, RNO). Concentration levels of met-


als (Hg, Cd, Pb) and organic pollutants (PAHs, PCBs, DDT, DDD, and DDE) in
mussel soft tissues and oysters sampled quarterly have been determined. The French
Ministry of Environment has funded this monitoring program to determine levels of
pollution in French coastal waters, using the mentioned organisms as quantitative
indicator species. The planning and implementation have been coordinated by the
French Research Institute for the Exploration of the Sea (IFREMER) (Claisse
1989). This program allowed to summarize the results of 10-year monitoring of
mercury, cadmium, lead, polycyclic aromatic hydrocarbons (PAHs), polychlori-
nated biphenyls (PCBs), and dichloro-diphenyl-chloroethanes (p, p ′DDT, p, p′
DDD, p, p ′DDE) in the soft tissues of the mollusks studied: Mytilus edulis, Mytilus
galloprovincialis, and Crassostrea gigas. The network consisted of 110 sampling
stations along the entire length of the French coast, and samples were taken in
February, May, August, and November of each year.
Despite the wide acceptance achieved by the concept of Mussel Surveillance
(Goldberg et al. 1978), the use of sentinel organisms (mainly bivalve such as Mytilus
sp.) for monitoring the concentration of pollutants in coastal areas, and as an indica-
tor of its bioavailability, demonstrated through the establishment of such programs
both nationally and internationally, an important problem was observed.
Unfortunately, the most useful species, mussels and oysters, do not occur frequently
in tropical waters, where pollution monitoring strategies were also required, and
this was inconvenient for the implementation of global surveillance programs. To
overcome this problem, a research program was carried out in the Todos os Santos
Bay region (Bahia, Brazil) (Porte et al. 1990), which basically highlighted two cen-
tral aspects: (i) the evaluation of the geographical distribution of hydrocarbons in
edible bivalves of the bay’s intertidal and (ii) the selection of organisms that indicate
organic pollution in the tropical regions of the coast of Brazil. Particular emphasis
was placed on the use of chemical markers for the recognition of the origin of
hydrocarbons, considering that marine organisms can concentrate not only a variety
of anthropogenic hydrocarbons in their tissues but also biogenic ones obtained
through their diet (Albaigés et al. 1987). This was the first study of this type con-
ducted in the South Atlantic. From the obtained results, it could be concluded that
the concept of surveillance through the Mussel can also be applied in tropical areas
using alternative bivalve species such as sentinel organisms. In that sense, prelimi-
nary data indicate that Anomalocardia brasiliana gives a satisfactory response to
exposure to local changes in the concentration of pollutants. The bivalves selected
for this study were Anomalocardia brasiliana, Protothaca pectorina, Lucina pecti-
nata, and Macoma constricta (Porte et al. 1990).
The rapid increase in trade in chemical substances within developing Asian
countries implied increased production and use of toxic products such as organo-
chlorines (OCs) and organic tin compounds (BTs), with a potential increase in the
exposure of beings, humans and wildlife, to those substances. Consequently, envi-
ronmental problems associated with contamination by compounds such as BTs and
OCs became of great concern (Tanabe et al. 2000). Considering that bivalves, par-
ticularly mussels, have been used as bioindicators for the monitoring of toxic sub-
530 J. E. Marcovecchio et al.

stances in coastal waters due to their own characteristics previously mentioned


(Phillips 1980; Goldberg 1986), it was decided to apply them to assess this
­conflicting environmental situation. For this case, the green mussel Perna viridis
has been chosen, considering that it has a wide geographical distribution within the
Asia-­Pacific region and is also recognized as a seafood with high commercial value
and intensively consumed by the populations of the region (Tanabe et al. 2000). The
monitoring studies developed include a part of the Asia-Pacific Mussel Watch
Program (APMWP) region, which aims to control marine pollution in this region
using mussel as a bioindicator (Tanabe 1994). APMWP was under the umbrella of
the International Mussel Watch Program, which mainly involved a coastal monitor-
ing program with bivalves as sentries (mussels and oysters as bioindicators), in
order to determine the quality of marine coastal waters. The initial phase of this
program was carried out in South and Central America during the period from 1991
to 1993 and revealed serious contamination by organochlorine insecticides in third
world countries within this region (Sericano et al. 1995). The second phase intended
to cover the three zones of the mentioned region: Asia-Pacific, Northwest Pacific,
and South Pacific. On the other hand, in addition to the traditional evaluation of
persistent compounds such as PCBs, DDT and their metabolites, and HCH and
HCB, other compounds of environmental interest such as organically bound metals
(Cu and Sn), and highly toxic compounds (dioxins, furans, and coplanar PCBs)
were also included to be monitored within this second phase. The APMWP project
was a collaborative work of scientists from the Philippines, Thailand, India, and
Japan, coordinated by Prof. Shinsuke Tanabe of Ehime University (Matsuyama,
Japan) (Tanabe et al. 2000).
This environmental strategy to assess the quality of aquatic ecosystems has dem-
onstrated to be extremely successful and has been fully applied all over the world.
Briefly, several of the developed studies on the last decades are mentioned, i.e.,
Minier et al. (2006) in the Seine estuary, France; Frontalini and Coccioni (2008) in
the Adriatic Sea coast, Italy; Ogata et al. (2009) within the five continents; Fang
et  al. (2009) in coastal areas of Hong Kong; Looi et  al. (2013) in the Strait of
Malacca, Malaysia; and Bat and Özkam (2019) in the Turkish Black Sea coast,
among many others, including the extense and very detailed review by Beyer et al.
(2017) on this topic.

19.5  Monitoring Programs in Argentina

A history of monitoring programs in different aquatic environments from Argentina


exists, although those of long-term monitoring are very scarce. In general, existing
programs are linked either to large urban centers or to environments conditioned by
some particularity that potentially hinders their functionality.
19  Estuarine Environmental Monitoring Programs: Long-Term Studies 531

One of the environments that historically has received successive programs for
monitoring the quality of its waters, sediments, and/or organisms is the Río de la
Plata, although these programs have had different coordination, management styles,
and objectives over time. So, the works presented by AGOSBA/SHIN (1997) in
which the environmental quality (physicochemical characteristics, b­ acteriology,
nutrients, heavy metals, pesticides) in water and surface sediments of the south
coastal strip of the Río de la Plata can be mentioned. This report synthesized the
information of 3 years of work. In the same way, it is possible to mention numerous
works of research groups that studied different environmental phenomena and
reached very valuable conclusions. So, among others, the work of the Administrative
Commission of the Río de la Plata (CARP 1989), which includes water quality data
and some contaminant data, can be mentioned; or by Janiot et al. (1991) determin-
ing organochlorine pesticides in sediments of the Argentine coast and frontal zone
of the Río de la Plata. Subsequently, works by other groups, such as those by
Colombo et al. (2006) measuring concentrations and flows of aromatic hydrocar-
bons in sediments of the Río de la Plata; by Ronco et  al. (2008) who presented
information on a “screening” of contaminants in sediments of tributaries of the
southwestern area of the​​ Río de la Plata; by Lombardi et  al. (2010) determining
trace metal levels in a large fish species (Prochilodus lineatus) from Río de la Plata
waters; by López et  al. (2013) studying the microbiological pollution in surface
waters of Reconquista River; by Sathicq et al. (2015) researching on the influence
of El Niño phenomenon on environmental condition of Río de la Plata estuary; or
by Castañé et al. (2015) assessing the water quality of Luján River near the metro-
politan area of Buenos Aires city, among others. This environment is a clear exam-
ple of systems that have a lot of information generated at different times over an
extended period but cannot be considered a long-term monitoring program, because
the data series has different origins, discontinuities in its execution, and different
work methodologies in several stages of its concretion.
Another environment currently under monitoring is the Uruguay River. Given
the problem of the existence of algal blooms occurring within this river and the pos-
sible appearance of cyanobacteria that can affect the localities using this water
resource for the production and consumption of drinking water, as well as recre-
ational and tourist use, organisms of the involved provinces signed two agreements,
one in 2008 and the last in 2010, with a duration of 2 years, through which perma-
nent controls of the water quality of the mentioned river were carried out. Together,
the Institute of Water and Environment from Corrientes Province (ICAA) as well as
the regulatory agency Administration of Current Sanitary Works (AOSC) carried
out permanent monitoring of the Uruguay River since December 2008, establishing
as sampling stations four locations in the province: Garruchos, Paso de los Libres,
São Tomé, and Monte Caseros, determining parameters in situ and in the physico-
chemical, bacteriological, and protistological laboratory. The study allowed to
determine early alerts for the implementation of contingency plans in the water
532 J. E. Marcovecchio et al.

treatment plants, before the presence of cyanobacteria, as well as to characterize


effluents and to have truthful information regarding the loading capacity and self-­
purification within the Uruguay River (ICAA, 2019).
Many other environments from Argentina are eventually under surveillance
through environmental monitoring programs which are carried out following differ-
ent design and application models, i.e., Río Hondo Reservoir, Santiago del Estero
Province (Secretaría de Recursos Hídricos de la Nación 2007); monitoring on San
Roque Reservoir, Córdoba Province (Crema et  al. 2014, Instituto Nacional del
Agua); in the scope of the National Parks of Argentina (APN, Administración de
Parques Nacionales, 2016); or Integrated Monitoring Program of the Matanza-­
Riachuelo Basin/authority of the Matanza-Riachuelo Basin, ACUMAR, Buenos
Aires Province (ACUMAR 2019), among others.
Moreover it is important to highlight that the abovementioned monitoring pro-
grams are not included within the long-term monitoring ones, considering they are
not continuously developed a long time and their designs are not the required as to
be considered so.

19.6  A
 Nice Study Case: Long-Term Monitoring Within
the Bahía Blanca Estuary, in Argentina

The existing information on the estuary of Bahía Blanca is very wide, and its history
goes back to the second half of the nineteenth century, including reports by Darwin
(1845) on the distribution and characterization of salts present in the Salitral de la
Vidriera. In any case, it is important to point out that in its first stage, the informa-
tion is fragmented and dispersed and has not been obtained following a systematic
protocol. In spite of the existence of several works that included measurements of
physicochemical parameters of the estuary water, only in 1974 a systematic
biweekly sampling started at both Ingeniero White and Puerto Cuatreros ports (in
the inner area of ​​the estuary) and continues up to the present. The results obtained
within this work program have allowed to define the interior area of the estuary as
“vertically homogeneous” and “hypersaline” on the occasion of hot and dry sum-
mers (Freije and Gayoso 1988; Marcovecchio and Freije 2004; Freije et al. 2008).
In addition, the influence of the precipitation produced not only on the tributary
basin but also on the bay itself on salinity values was also identified (Freije and
Marcovecchio 2004).
The identification of the behavior and natural distribution of the structural (i.e.,
temperature, salinity, pH, alkalinity), ecophysiological (i.e., inorganic nutrients,
dissolved oxygen, organic matter, photosynthetic pigments), and physicochemical
parameters of the system, determined during a sufficiently long period (as is the
case in this study), allow us to fully characterize its baseline operating conditions,
as well as to quickly detect deviations from normal behavior attributable to human
activities carried out within the region.
19  Estuarine Environmental Monitoring Programs: Long-Term Studies 533

19.6.1  Is 40 Years a Sufficiently Long Monitoring Period?

There is a history of long-term global research programs on the distribution of


chemical substances, including pollutants, in large geographic areas on various con-
tinents; among them, the International Mussel Watch Program (Goldberg et  al.
1978; Farrington and Tripp 1995) or the Worldwide Persistent Organochlorine
Compounds’ Monitoring Program (Tanabe et al. 1987, 1994; Tanabe and Tatsukawa
1991) have been mentioned, which they undertook to assess global trends in the
distribution of some compounds in coastal marine systems around the world.
Unlike these, the study program developed in the inner zone of the Bahía Blanca
Estuary includes the generation of a long time series of data obtained with high
frequency; in fact, and since December 1974, biweekly samplings were carried out
at two points in the aforementioned region, and they were even weekly throughout
much of the indicated period (Marcovecchio et al. 2010). The systematic develop-
ment of this study in the Bahía Blanca Estuary, carried out over the last four decades,
has allowed the generation of a database, which allows to base some concepts that
explain the operation and the physical-chemical condition of the estuary. On the
other hand, the long data series makes it possible to differentiate deviations due to
the natural variability of the samples from those produced by the incidence of
human activities which separates the results from their normal distribution. This
study program has allowed to significantly characterize both the environmental con-
dition and the functioning of the Bahía Blanca Estuary.
In addition, and since the middle 1990s, a complementary monitoring program
required by Bahía Blanca municipal government has started within Bahía Blanca
Estuary, directed to fully characterize the environmental quality within the inner
area of the estuary. This study includes not only the chemical aspect previously
mentioned but also the surveillance of biological features (i.e., microbiology, biol-
ogy of fishes as well as benthic communities within Bahía Blanca Estuary) which
allow optimizing the obtained results.
Both monitoring programs continue to operate simultaneously and have proven
to be an excellent tool to support decision-making on environmental issues within
this system. The following is a brief summary of the main results obtained through
these long-term monitoring programs:

19.6.2  Chemical Aspects

The distribution of the physical-chemical parameters of seawater, observed both in


the stage prior to the strong industrialization of this region (1970s) and in the later
period, shows that the system has not undergone significant changes that implied
variations in its operation. In this sense, it is possible to comment on the relative
stability shown by the structural parameters (i.e., temperature and salinity) as well
as the trends of the ecophysiological parameters (i.e., nutrients, dissolved oxygen,
534 J. E. Marcovecchio et al.

photosynthetic pigments) sustained over time and observed over more than 40-year
evaluation. Furthermore, the fact that the estuary fully conserves its synchronicity,
by repeating with natural frequencies the natural cycles of the elements (i.e., N, P,
Si) which sustain biological production (as well as its power and magnitude), must
be highlighted. The horizontal distribution of these parameters did not present sig-
nificant spatial variations, and the vertical one was totally homogeneous, which
allows to characterize the water body as non-stratified (Marcovecchio and Freije
2004; Freije et al. 2008). All this indicates that the human activities that take place
in the internal zone of the Bahía Blanca Estuary have not generated significant
changes within the basal physical-chemical condition of the system, which main-
tains its properties’ and parameters’ distribution trends similar to the corresponding
historical records (Marcovecchio and Freije 2013).
On the other hand, when the distribution of potentially pollutant substances was
evaluated within this environment, a different situation could be recorded. Most of
the studied pollutants were determined at clearly detectable levels in the estuary,
including heavy metals, hydrocarbons, organochlorine pesticides, and microplastic
particles. These contaminants were basically found in the sediments of the system,
although some (i.e., heavy metals, PAHs) were also determined in the tissues of
organisms, suspended in the particulate matter, and dissolved in water. Temporal
distribution trends of these compounds were scattered, i.e., while trace metals have
showed a decreasing trend since 1980s and up to now (IADO 1997, 1999, 2002,
2006, 2008, 2010, 2012, 2014, 2016, 2018), microplastic particles have been just
recently determined and appeared to be increasing within the system (Arias et al.
2019; Ronda et al. 2019).
So the joint application of both monitoring programs has allowed to characterize
the BB estuary as an environment that functions properly and without problems
from a biogeochemical point of view, with a marked synchronicity in the production
of nutrients as well as a corresponding high level of primary production. However,
the presence and distribution of different types of potentially toxic substances in its
waters, sediments, and organisms have also been verified, although without reach-
ing critical levels for human health and ecosystem’s one.

19.6.3  Microbiological Aspects

Marine and coastal resources represent strategic assets of great importance to diver-
sify and improve regional economies. The settlement of large cities and important
industries is also common in coastal areas, with the potential that they entail to
generate pollution processes.
The increasing levels of pollution of estuaries and bays constitute an increasing
risk to public health. They affect marine productivity and diversity and, at the same
time, raise costs for tourism and aquaculture.
In the Bahía Blanca Estuary, coastal pollution comes through diffuse sources
such as continental rain runoff and point sources such as sewage and industrial dis-
19  Estuarine Environmental Monitoring Programs: Long-Term Studies 535

charges. The last of them have increased in recent decades, due to intense human
activity related to urban and industrial development and, above all, due to the con-
stant and increasing dumping of raw or insufficiently treated sewage. For decades,
the city’s sewage effluents have been dumped, insufficiently treated through the
Primera Cuenca collector (2500 m3.h−1). Toward the end of September 2008, the
discharge of the Tercera Cuenca sewage liquid treatment was added with a tipping
flow of 200 m3.h−1 (Streitenberger and Baldini 2010). This was designed to carry out
secondary treatment of effluents and is located in the innermost area of the estuary.
Bacterial Indicators  The number and type of bacteria present in a natural ecosys-
tem are generally in balance (homeostasis). When environmental conditions are
altered, changes occur in the bacterial community (Cabezalí et al. 1995). Therefore
the bacterial populations present in marine and coastal ecosystems constitute an
excellent early warning system, since they respond to changes in the environment,
through qualitative or quantitative changes, more quickly than the rest of the biota.
Hence, any environmental alteration can be detected in its initial stages, if carefully
and appropriately chosen bacterial indicators are used.
Bacteria are key participants in the flow of heterotrophic energy and mineraliza-
tion of organic matter. They actively contribute to the self-cleaning capacity of the
environment, that is, they are directly involved in the removal and recycling
processes.
Both in the water column and in the sediment, heterotrophic bacteria efficiently
transform dissolved organic matter into particulate one, leaving this critical resource
available for the rest of the food chain. The intensity and nature of the exchanges
carried out by the microorganisms strongly depend on the qualitative and quantita-
tive distribution of their various communities and their dynamics in the different
ecological niches in which they are located. In this way, differences in the abun-
dance of bacterial populations in different areas can be attributed to existing concen-
trations of easily assimilable organic matter.
The advantage of using biological indicators is that they reflect the cumulative
impact of all stressors over a period. As such, they are a unique measure of the
response of the environment and provide information that cannot be collected in any
other way.
Coastal waters are the final receptacle for most of the waste caused by human
activity on land. Although man’s organic residues must be integrated into the bio-
geochemical cycles of ecosystems, it is not reasonable to believe that increasing
amounts of organic matter and bacteria can be safely introduced. This input of mate-
rial influences macro- and microscopic trophic chains, rapidly stimulating the
microbial growth of indigenous groups and contributing new populations, associ-
ated with the type of substances that reach the environment.
It is a known fact that enteric infections in the population are favored by multiple
deficiencies in basic sanitation and in the excreta final deposition systems. This
contributes to maintain a level of human fecal contamination, particularly in aquatic
ecosystems (Cabezalí et  al. 1995). Epidemiological studies carried out since the
1950s demonstrate the relationship between fecal contamination of recreational
536 J. E. Marcovecchio et al.

waters and the adverse effect on human health including the appearance of gastro-
intestinal symptoms and eye, ear, nasal, respiratory, and skin infections (Wade et al.
2006; Soller et al. 2010).
Escherichia coli is the most routinely used microbial compliance parameter
for confirming fecal (though not necessarily pathogen) contamination of the environ-
ment, and their use in water quality legislation around the world demonstrates their
widespread utility (Oliver et al. 2016). It is assumed that their behavior is similar to
that of other bacteria of the same origin when they are released into the environment
(Anderson et al. 2005). Its detection alerts about the possible presence of pathogenic
intestinal microorganisms, with the consequent hygiene-health risk for the population.
In the internal area of the estuary, since years, are recorded higher population
densities than those internationally recommended for primary contact waters
(≤126 CFU.100 mL−1) (USEPA 2003). The most compromised areas coincide with
the sewage discharges (Primera and Tercera Cuenca) and with Ingeniero White Port.
The fluctuations in the bacteria indicator counting show that the Tercera Cuenca
sewage treatment plant fails to stabilize its operation; therefore there is a continuous
contribution of effluents to the Bahía Blanca Estuary with parameters outside the
acceptable range. The sustained bacteriological counts, the constant presence of
fecal indicators, and the tendency to accumulate in sediments in the surrounding
areas are evidence of the impact that the system has been suffering in recent years
(Streitenberger and Baldini 2010; Pierini et al. 2012).
In Bahía Blanca Estuary waters, terrestrial heterotrophic bacteria (THB) densi-
ties fluctuate between 102 and 103 CFU.mL−1, depending on the continental influ-
ences that the different zones receive. For example, the proximity to the coast
determines the presence of THB coming from the continent and arriving at the estu-
ary transported by the wind, the runoff produced by the rains, or the water courses
that flow into the estuary. Marine heterotrophic bacteria always exceed THB by an
order of magnitude (104–105 CFU.mL−1) and with little spatial and temporal fluctua-
tion. These are native bacteria that are better adapted to environmental conditions
and that compete more successfully for nutrients. They are metabolically versatile,
being able to use a wide range of N and P sources for their growth (Kirchman 2000).
Hydrocarbon-degrading microorganisms are ubiquitous in the world’s oceans,
and biodegradation mediated by indigenous microbial communities is the ultimate
fate of the majority of oil hydrocarbon that enters the marine environment (Leahy
and Colwell 1990; Prince 2010; Atlas and Hazen 2011). Hydrocarbon-degrading
bacteria represent the first line of defense against oil pollution in the marine envi-
ronment. In response to the complexity of hydrocarbon compounds found in petro-
leum deposits, diverse marine microorganisms have evolved with an equal
complexity of metabolic pathways to take advantage of hydrocarbons as a rich car-
bon and energy source (Kostka et al. 2011, 2014). So it becomes significant to know
whether microbial degraders of oil are present in water and sediments of the area to
be studied. Enumeration of petroleum-degrading microorganisms is important to
assess the magnitude of oil pollution that has occurred. So they are powerful indica-
tors of the impacts that the environment has received or is receiving. As it was
indicated above, the Bahía Blanca Estuary has several sources of pollutions such as
19  Estuarine Environmental Monitoring Programs: Long-Term Studies 537

continental rain runoff and point sources such as sewage and industrial discharges.
During several years quantification of hydrocarbon-degrading bacteria in superficial
sediments of Principal Channel of the estuary was carried out. These counts have
been able to indicate the sites most affected by the impact of hydrocarbons, and
counts greater about 104 (MPN.g−1) were detected in major part of the sampled sites.
These values indicated the impact of port activity and municipal sewages in the
entire studied area. However, in recent years, the highest values have been coinci-
dent with the site where the insufficiently treated municipal sewages are dumped.
So the application of the monitoring program within Bahía Blanca Estuary
allowed to identify quite significant points. Throughout the years, bacteriological
studies demonstrated the significant anthropic impact received by the Principal
Channel (Fig. 2.1, Chap. 2), as well as the urgent need for all the sewage effluents
using the estuary as a receiving body to be properly treated and controlled according
to current legislation. This was evidenced by the increased amount of bacteria indi-
cating fecal contamination in the innermost stations after the establishment of a
sewage treatment plant (2008), which worked poorly for a long time. It is of crucial
importance to ensure the proper functioning of the Tercera Cuenca sewage treat-
ment plant, as it is located in a particularly vulnerable area of the estuary with a
notable tendency to accumulate contaminants. The waters and sediments of the
areas near the sewage outlet showed the least variability in them, indicating the
continuity and impact produced by municipal effluents without treatment.
The hydrocarbon-degrading bacteria counts indicated the existence of hydrocar-
bons’ impact and their accumulations in the sediments in studied area, especially
those affected by municipal sewage.
The results show the important impact of the dumping of insufficiently treated
sewage effluents on the studied area of the Bahía Blanca Estuary. This situation
generates an accumulation of biological and chemical contaminants in the sedi-
ments, which can be transformed into reservoirs of potentially dangerous microor-
ganisms. In turn, there is a probability that they will be resuspended in the water
column, due to weather conditions or dredging activities. If this situation persists,
not only the ecosystem is put at risk but also the health of the population that uses
the area for different purposes.
International experience has shown that the most effective results in maintaining
adequate quality for recreational purposes are achieved with rigorous control of
discharges and not with mere speculation about the self-cleaning power of receiving
courses. The dumping of pollutants into the aquatic environment and the degree of
impact they produce on the receiving body are technically controllable.

19.6.4  Biology of Fishes

Increased anthropic activities in estuaries negatively impact water quality and


aquatic fauna, producing changes that can be observed in feeding ecology, loss of
spawning and recruitment grounds, and reduction of biodiversity, which in turn
538 J. E. Marcovecchio et al.

affect the ecosystem as a whole (López Rojas and Bonilla Rivero 2000; Whitfield
and Elliott 2002; Eddy 2005).
Feeding is one of the most important factors that control fish populations, directly
affecting abundance, growth, mortality, and migrations (Sánchez and Prenski 1996).
Traditionally, feeding ecology studies imply the analysis of fish gut content. The infor-
mation produced by these studies has been key to our understanding of trophic rela-
tionships between species, as well as on the impact of fish feeding on other components
of the aquatic communities (Hynes 1979, in Escalante 1987). It is therefore desirable
to generate knowledge on estuarine trophic links to be able to better understand how
anthropic disturbances can impact fish and other estuarine species (Elliot et al. 2007).
Fish are useful bioindicators and can provide relevant information on habitat
quality evaluations (Hynes 1979, in Escalante 1987); however, not all fish species
are suitable for this purpose, as many show migratory behaviors, meaning that they
will migrate outside of estuaries and be exposed to other stressors, alien to the estu-
ary itself. It is for that reason that in the present study, the fish species employed to
assess water quality were juveniles of Cynoscion guatucupa (Lopez Cazorla 2000;
Sardiña and Lopez Cazorla 2005a), Micropogonias furnieri (Sardiña and Lopez
Cazorla 2005b), and Mustelus schmitti (Lopez Cazorla 1987, 2004), which remain
in the estuary during their first year. Additionally, Ramnogaster arcuata, a fish spe-
cies that completes its entire life cycle inside the waters of the estuary, was also used
(Lopez Cazorla and Sidorkewicj 2009).
Starting in 2015, a biological monitoring of the abovementioned fish species has
been seasonally conducted in two sites within Bahía Blanca Estuary. The general
objective was to help understand how the fish species utilize the impacted areas of the
estuary, in order to use the obtained knowledge to inform management plans and con-
servation actions. The specific objectives laid out to achieve the aforementioned gen-
eral objectives were: a) generate size frequency and age composition distributions, b)
estimate length-weight relationships and growth type, c) evaluate feeding intensity,
and d) describe the feeding ecology and selectivity by species, season, and size class.
The biological monitoring allows us to follow the evolution of the changing bio-
logical parameters of the populations of the fish species under study throughout
time and space and establish links and relationships with the results obtained by
chemical analysis of muscle tissue of these species.

19.6.5  Benthic Communities

Biodiversity inventories and monitoring programs have increased markedly world-


wide over the last decade in response to concerns about extinctions and the sustain-
ability of natural ecosystems. In general, inventories are conducted to determine the
distribution and composition of wildlife and wildlife habitats in areas where such
information is lacking, and monitoring is typically used to understand the rates of
change or the effects of management practices on wildlife populations and habitats
19  Estuarine Environmental Monitoring Programs: Long-Term Studies 539

(Morrison et al. 2008). Making inventories and monitoring biodiversity are crucial
to identifying the natural processes and human activities that affect ecosystems.
Benthic communities are considered to be good indicators of ecosystem health
because of their sedentary nature and longevity, providing long-term exposure to
toxic substances, and also due to the presence of diverse taxa that can respond to
multiple types of man-made perturbations (Pearson and Rosenberg 1978). In addi-
tion, benthic communities play an important role in the regulation of ecosystem
processes; they are central to the maintenance of the links between benthic and
pelagic systems. Moreover, carbon, oxygen, and nutrient cycling and the decompo-
sition of dead matter or waste materials are important processes driven by the ben-
thic species(Snelgrove et al. 1997).
The evaluation of benthic communities was included in the general environmen-
tal monitoring program of the Bahía Blanca Estuary during the last decade. Since
then the inventory of macrofaunal invertebrates in the Principal Channel and other
minor channels, affected by the discharge of urban and industrial effluents, has been
carried out as well as monitoring of benthic communities associated with the port
area. It took many years to complete the inventory, due to the large size of the area
surveyed and the limited taxonomic knowledge of the species identified (see Chap.
9 for the description of the invertebrate communities). Sampling for the biodiversity
inventory was carried out along the Principal Channel at the same sites where the
annual measurements of the physical and chemical variables of the water column
and sediments are taken periodically. In addition, environmental samples were
taken in the minor channels together with the biological samples (IADO 2014,
2016, 2018). The comprehensive analysis of the biological and environmental data
allowed the elaboration of a distribution map of the main biological assemblages,
their characteristic habitats, and the main risks that affect the local biodiversity. The
biological communities associated with minor channels have low diversity and bio-
mass, and only a few polychaete species are dominant in the areas most affected by
the discharge of sewage effluents (IADO 2018).
The introduction of exotic species is considered as one of the main threats to
biodiversity at different scales, and the extent is causing serious and persistent
changes in the marine and coastal environments (Bax et al. 2003; Souza et al. 2009).
A generalized consensus points to prevention through the management of vectors
and pathways as being the most effective line of defense against invasive marine
species (Molnar et  al. 2008). The monitoring of benthic communities associated
with the port areas enables early detection of new exotic species since the most
common pathway for invasive marine species includes shipping, ballast, and/or
fouling. Monitoring of the port areas in the Bahía Blanca Estuary indicates that the
artificial structures of this sector are dominated by few exotic species, mainly bar-
nacles, most of which have not been detected elsewhere (IADO 2014, 2016).
However, the Pacific oyster, which was detected thanks to port monitoring, has
accelerated its expansion and abundance in recent years, generating serious environ-
mental problems.
540 J. E. Marcovecchio et al.

Consequently, and keeping in mind the previously commented results, several


concluding points should be pointed out:
• Periodic monitoring of port areas is an effective and low-cost tool to alert about
the entry and establishment of non-native species.
• The development of a management plan is necessary for recently established and
expanding exotic species, such as the Pacific oyster.
• It is recommended to update the inventory of the benthic communities along the
main channel every 5 years.
• It is recommended to start monitoring the communities associated with the minor
channels in order to evaluate the environmental changes associated with the
industrial and urban discharge areas.

19.7  Concluding Comments

The application of monitoring programs, particularly the long-term ones, has dem-
onstrated to be an excellent and quite efficient tool not only to evaluate environmen-
tal quality within aquatic systems but also to assess their evolution along times and
changes within their biological, physical, and chemical conditions, as well as to
consider the possible and sustainable exploitation of their natural resources.
Sustaining this type of programs in the long term allows to observe consolidated
tendencies of the environmental condition of the systems and helps to suggest better
management decisions to the corresponding authorities. In this sense, the example
of the BBE case study fully indicates the mentioned concept.

References

ACUMAR (Autoridad de Cuenca Matanza-Riachuelo) (2019) Plan Integral de Saneamiento


Ambiental (PISA). http://www.acumar.gob.ar/monitoreo-­ambiental/ (retrieved Nov.07/2019)
AGOSBA/SHIN (Administración General de Obras Sanitarias de la Pcia. de Buenos Aires/Servicio
de Hidrografía Naval) (1997) Calidad de las aguas de la franja costera sur del Río de la Plata
(San Fernando – Magdalena). Informe Consejo Permanente para el Monitoreo de la Calidad de
las Aguas de la Franja Costera Sur del Rio de la Plata, Buenos Aires (Argentina), 173 p
Albaigés J, Farran A, Soler M et al (1987) Accumulation and distribution of biogenic and pollutant
hydrocarbons, PCBs and DDT in tissues of Western Mediterranean fishes. Mar Environ Res
22:1–18. https://doi.org/10.1016/0141-­1136(87)90078-­X
Anderson KL, Whitlock JE, Harwood VJ (2005) Persistence and differential survival of fecal indi-
cator bacteria in subtropical waters and sediments. Appl Environ Microbiol 71:3041–3048.
https://doi.org/10.1128/AEM.71.6.3041-­3048.2005
APN – Administración de Parques Nacionales (2016) Reglamento para la Evaluación de Impacto
Ambiental en la Administración de Parques Nacionales. Ministerio de Ambiente y Desarrollo
Sustentable, Administración de Parques Nacionales, Ley N°22351
19  Estuarine Environmental Monitoring Programs: Long-Term Studies 541

Arias AH, Ronda A, Oliva AL et al (2019) Evidence of microplastic ingestion by fish from the
Bahía Blanca estuary in Argentina, South America. B Environ Contam Tox 102(6):750–756.
https://doi.org/10.1007/s00128-­019-­02604-­2
Atlas RM, Hazen TC (2011) Oil biodegradation and bioremediation: a tale of the two worst spills
in U.S. Environ Sci Technol 45:6709–6715. https://doi.org/10.1021/es2013227
Bat L, Özkam EY (2019) Heavy metal levels in sediments of the Turkish Black Sea Coast.
Oceanography and Coastal Informatics: Breakthroughs in Research and Practice: 86–107, IGI
Global. https://doi.org/10.4018/978-­1-­5225-­7308-­1.ch004
Bax N, Williamson A, Agüero M et al (2003) Marine invasive alien species: a threat to global bio-
diversity. Mar Policy 27:313–323. https://doi.org/10.1016/S0308-­597X(03)00041-­1
Behmel S, Damour M, Ludwig R et  al (2016) Water quality monitoring strategies  – a review
and future perspectives. Sci Total Environ 571:1312–1329. https://doi.org/10.1016/j.
scitotenv.2016.06.235
Beyer J, Green NW, Brooks S et al (2017) Blue mussels (Mytilus edulis spp.) as sentinel organ-
isms in coastal pollution monitoring: a review. Mar Environ Res 130:338–365. https://doi.
org/10.1016/j.marenvres.2017.07.024
Biber E (2013) The challenge of collecting and using environmental monitoring data. Ecol Soc
18(4):68–81. https://doi.org/10.5751/ES-­06117-­180468
Boyden CR (1974) Trace element content and body size in molluscs. Nature 251:311–314. https://
doi.org/10.1038/251311a0
Brewin RJW, de Mora L, Jackson T et al (2015) On the potential of surfers to monitor environmen-
tal indicators in the coastal zone. PLoS One 10(7):e0127706. https://doi.org/10.1371/journal.
pone.0127706
Bryan GW (1973) The occurrence and seasonal variation of trace metals in the scallops Pecten
maximus (L.) and Chlamys opercularis (L.). J Mar Biol Assoc UK 53:145–166. https://doi.
org/10.1017/S0025315400056691
Cabezali CB, Baldini MD, Cubitto MA et al (1995) Estudio bacteriológico de aguas marinas para
uso recreacional. Rev Argent Microbiol 27:115–122
CARP (Comisión Administradora del Río de la Plata) (1989) Estudios para la evaluación de la
contaminación del Río de la Plata. CARP/Servicio de Hidrografía Naval (Argentina)/Servicio
de Oceanografía, Hidrografía y Meteorología de la Armada (Uruguay). Informe de Avance.
Buenos Aires-Montevideo. 42 p
Castañé PM, Sánchez-Caro A, Salibián A (2015) Water quality of the Luján river, a lowland
watercourse near the metropolitan area of Buenos Aires (Argentina). Environ Monit Assess
187:645–659. https://doi.org/10.1007/s10661-­015-­4882-­y
Claisse D (1989) Chemical contamination of French coasts. The results of a ten years Mussel
Watch. Mar Pollut Bull 20:523–528. https://doi.org/10.1016/0025-­326X(89)90141-­0
Collins A, Ohandja D-G, Hoare D et al (2012) Implementing the Water Framework Directive: a
transition from established monitoring networks in England and Wales. Environ Sci Policy
17:49–61. https://doi.org/10.1016/j.envsci.2011.11.003
Colombo JC, Cappelletti N, Lasci J et al (2006) Sources, vertical fluxes, and equivalent toxicity of
aromatic hydrocarbons in coastal sediments of the Río de la Plata Estuary, Argentina. Environ
Sci Tecnol 40:734–740. https://doi.org/10.1021/es051672y
Converse SJ, Royle JA (2012) Dealing with incomplete and variable detectability in multi-year,
multi-site monitoring of ecological populations. In: Gitzen RA, Millspaugh JJ, Cooper AB,
Licht DS (eds) Design and analysis of long-term ecological monitoring studies. Cambridge
Univ. Press, Cambridge, UK. Ch.19: 426–442 (586 pp). ISBN: 978-0-521-19154-8
Crema N, Fernández A, Pacharoni F et al (2014) Monitoreo de la calidad del agua del embalse
San Roque. Efecto de contaminantes por descargas de efluentes. https://www.ina.gob.ar›ifrh-­
2014›Eje2›2.08.pdf Accessed on 14 Nov 2019
Darwin CR (1845) Journal of researches into the geology and natural history of the various
countries visited by HMS Beagle under the command of Captain Fitzroy, from 1832 to 1836.
Murray Ed, London, 738 p
542 J. E. Marcovecchio et al.

Davies JC, Mazurek J (2014) Pollution control in the United States: evaluating the system.
Resources for the Future Publ, New York. 59 pp. ISBN 0-915707-8-8
de Oliveira RF, Uehara-Prado M, Valente-Neto F et al (2018) A network of monitoring networks
for evaluating biodiversity conservation effectiveness in Brazilian protected areas. Perspect
Ecol Conser 16:177–185. https://doi.org/10.1016/j.pecon.2018.10.003
Eddy FB (2005) Ammonia in estuaries and effects on fish. J Fish Biol 67:1495–1513. https://doi.
org/10.1111/j.1095-­8649.2005.00930.x
Elliott, M., Whitfield, A. K., Potter, I. C., Blaber, S. J., Cyrus, D. P., Nordlie, F. G., & Harrison, T.
D. (2007). The guild approach to categorizing estuarine fish assemblages: a global review. Fish
and fisheries, 8(3), 241–268.
Escalante A (1987) Alimentación de Bryconamericus iheringi y Jenynsia lineata lineata
(Osteichthyes) en Sierra de la Ventana (Argentina). Anales del Museo de Historia Natural de
Valparaíso 18:101–108
Fang JKH, Au DWT, Wu RSS et al (2009) The use of physiological indices in rabbitfish Siganus
oramin for monitoring of coastal pollution. Mar Pollut Bull 58:1229–1244
Fang S, Xu LD, Zhu Y et al (2014) An integrated system for regional environmental monitoring
and management based on Internet of Things. IEEE T Ind Electron 10(2):1596–1605. https://
doi.org/10.1109/TII.2014.2302638
Farrington JW, Tripp BW (1995) International Mussel Watch Project. Initial implementation
phase: final report. NOAA Technical Memorandum NOS-ORCA 95, 136 p
Farrington JW, Goldberg ED, Risebrough RW et  al (1983) U.S.  Mussel Watch 1976–1978: An
overview of the trace-metal, DDE, PCB, hydrocarbon and artificial radionuclide data. Environ
Sci Technol 17:490–496. DOI: https://doi.org/10.1021/es00114a010
Freije RH, Gayoso AM (1988) Producción primaria en el estuario de Bahía Blanca. Informes sobre
Ciencias del Mar – UNESCO 47:112–114
Freije RH, Marcovecchio JE (2004) Oceanografía química del estuario de Bahía Blanca. En:
Piccolo MC, Hoffmeyer MS (eds) El ecosistema del estuario de Bahía Blanca, Instituto
Argentino de Oceanografía (IADO–CONICET/UNS), Bahía Blanca (Argentina), Cap. 8:69–
78. ISBN 987-9281-96-9
Freije RH, Spetter CV, Marcovecchio JE et al (2008) Water chemistry and nutrients of the Bahía
Blanca Estuary. In: Neves R, Baretta J, Mateus M (eds) Perspectives on integrated coastal zone
management in South America. Part B: from shallow water to the deep fjord: the study sites,
IST Scientific Publishers, Lisbon. Chapter 23:243–256. ISBN: 978-972-8469-74-0
Frontalini F, Coccioni R (2008) Benthic foraminifera for heavy metal pollution monitoring: a case
study from the Central Adriatic Sea coast of Italy. Estuar Coast Shelf Sci 76:404–417. https://
doi.org/10.1016/j.ecss.2007.07.024
GESAMP (1990) GESAMP: the state of the marine environment. UNEP Regional Seas Reports
and Studies No115. Group of Experts on Scientific Aspects of Marine Pollution, Nairobi, 111 p
Goldberg ED (1975) The Mussel Watch – a first step in global marine monitoring. Mar Pollut Bull
6:111–114
Goldberg ED (1986) The Mussel Watch Concept. Environ Monit Assess 7:91–103. https://doi.
org/10.1007/BF00398031
Goldberg ED, Bowen VT, Farrington JW et al (1978) The Mussel Watch. Environ Conserv 5:101–
125. https://doi.org/10.1017/S0376892900005555
Goldsmith B (ed) (2012) Monitoring for conservation and ecology, vol 3. Springer Science &
Business Media, Heidelberg. 396 p. ISBN: 978-0-412-35600-1
Grace JB, Anderson TM, Olff H et al (2010) On the specification of structural equation models for
ecological systems. Ecol Monogr 80(1):67–87. https://doi.org/10.1890/09.0464.1
Gray WB, Shimshack JP (2011) The effectiveness of environmental monitoring and enforce-
ment: a review of the empirical evidence. Rev Environ Econ Policy 5(1):3–24. https://doi.
org/10.1093/reep/req017
Hounslow A (ed) (2018) Water quality data: analysis and interpretation, 1st edn. CRC Press–
Taylor & Francis Publ. Co, Boca Raton. 416 pp. ISBN: 9780203734117
19  Estuarine Environmental Monitoring Programs: Long-Term Studies 543

Hsu A, Zomer A (2014) Environmental performance index. Wiley StatsRef: Statistics Reference
Online, 1–5. https://doi.org/10.1002/9781118445112.stat03789.pub2
IADO (1999) Programa de monitoreo de la calidad ambiental de la zona interior del estuario de
Bahía Blanca. Informe Final, Instituto Argentino de Oceanografía: 25 + 68 pp. Available in:
http://www.bahiablanca.gov.ar/cte/index.html
IADO (2002) Programa de monitoreo de la calidad ambiental de la zona interior del estuario de
Bahía Blanca. Informe Final, Instituto Argentino de Oceanografía: 76 pp. Available in: http://
www.bahiablanca.gov.ar/cte/index.html
IADO (2006) Programa de monitoreo de la calidad ambiental de la zona interior del estuario de
Bahía Blanca. Informe Final, Instituto Argentino de Oceanografía: 82 pp. Available in: http://
www.bahiablanca.gov.ar/cte/index.html
IADO (2008) Programa de monitoreo de la calidad ambiental de la zona interior del estuario de
Bahía Blanca. Informe Final, Instituto Argentino de Oceanografía: 107 pp. Available in: http://
www.bahiablanca.gov.ar/cte/index.html
IADO (2010) Programa de monitoreo de la calidad ambiental de la zona interior del estuario de
Bahía Blanca. Addenda al Informe Final 2009, Instituto Argentino de Oceanografía: 92  pp.
Available in: http://www.bahiablanca.gov.ar/cte/index.html
IADO (2012) Programa de monitoreo de la calidad ambiental de la zona interior del estuario de
Bahía Blanca. Informe Final, Instituto Argentino de Oceanografía: 122 pp. Available in: http://
www.bahiablanca.gov.ar/cte/index.html
IADO (2014) Programa de monitoreo de la calidad ambiental de la zona interior del estuario de
Bahía Blanca. Informe Final, Instituto Argentino de Oceanografía: 242 pp. Available in: http://
www.bahiablanca.gov.ar/cte/index.html
IADO (2016) Programa de monitoreo de la calidad ambiental de la zona interior del estuario de
Bahía Blanca. Informe Final, Instituto Argentino de Oceanografía: 228 pp. Available in: http://
www.bahiablanca.gov.ar/cte/index.html
IADO (2018) Programa de monitoreo de la calidad ambiental de la zona interior del estuario de
Bahía Blanca. Informe Final, Instituto Argentino de Oceanografía: 364 pp. Available in: http://
www.bahiablanca.gov.ar/cte/index.html
IADO (Instituto Argentino de Oceanografía, CONICET/UNS) (1997) Programa de monitoreo de
la calidad ambiental de la zona interior del estuario de Bahía Blanca. Informe Final, Instituto
Argentino de Oceanografía: 65 pp. Available in: http://www.bahiablanca.gov.ar/cte/index.html
ICAA (Instituto Correntino del Agua y del Ambiente) (2019) Monitoreo permanente del río
Uruguay. http://icaa.gov.ar/ Accessed on 11 Nov 2019
Janiot LJ, Orlando AM, Roses OE (1991) Niveles de plaguicidas clorados en el Río de la Plata.
Acta Farmacológica Bonaerense 10:15–23
Johnson DH (2012) Monitoring that matters. In: Gitzen RA, Millspaugh JJ, Cooper AB et al (eds)
Design and analysis of long-term ecological monitoring studies. Cambridge University Press,
Cambridge. Chapter 3:54–73, 586 p. ISBN: 978-0-521-19154-8
Kirchman DL (2000) Uptake and regeneration of inorganic nutrients by marine heterotrophic bac-
teria. In: Kirchman DL (ed) Microbial ecology of the oceans. Wiley & Sons, Birmingam. 528
p. ISBN: 978-0-470-04344-8
Kostka JE, Prakash O, Overholt WA et al (2011) Hydrocarbon-degrading bacteria and the bacte-
rial community response in Gulf of Mexico Beach sands impacted by the Deepwater Horizon
oil spill. Appl Environ Microbiol 77(22):7962–7974. https://doi.org/10.1128/AEM.05402-­11
Kostka JE, Teske AP, Joye SB et al (2014) The metabolic pathways and environmental controls
of hydrocarbon biodegradation in marine ecosystems. Front Microbiol 5:471. https://doi.
org/10.3389/fmicb.2014.00471
Küster A, Becker PR, Kucklick JR et  al (2015) The international environmental specimen
banks—let’s get visible. Environ Sci Pollut R 22:1559–1561. https://doi.org/10.1007/
s11356-­013-­2482-­3
Lawton JH (2007) Ecology, politics and policy. J Appl Ecol 44:465–474. https://doi.
org/10.1111/j.1365-­2664.2007.01315.x
544 J. E. Marcovecchio et al.

Leahy JG, Colwell RR (1990) Microbial degradation of hydrocarbons in the environment.


Microbiol Rev 54(3):305–315
Likens GE, Walker K, Davies P et  al (2009) Ecosystem science: toward a new paradigm for
managing Australia’s inland aquatic ecosystems. Mar Freshw Res 60:271–279. https://doi.
org/10.1071/MF08188
Lindenmayer DB, Likens GE (2010) The science and application of ecological monitoring. Biol
Conserv 143:1317–1328. https://doi.org/10.1016/j.biocon.2010.02.013
Lindenmayer DB, Cunningham RB, McGregor C et  al (2008) The changing nature of bird
populations in woodland remnants as a pine plantation emerges: results from a large-scale
“natural experiment” of landscape context effects. Ecol Monogr 78:567–590. https://doi.
org/10.1890/07-­0945.1
Lombardi PE, Peri SI, Verrengia Guerrero NR (2010) Trace metal levels in Prochilodus lineatus
collected from the La Plata River, Argentina. Environ Monit Assess 160:47–59. https://doi.
org/10.1007/s10661-­008-­0656-­0
Looi LJ, Aris AZ, Johari WLW et al (2013) Baseline metals pollution profile of tropical estuar-
ies and coastal waters of the Straits of Malacca. Mar Pollut Bull 74(1):471–476. https://doi.
org/10.1016/j.marpolbul.2013.06.008
Lopez Cazorla A (1987) Contribución al conocimiento de la ictiofauna marina del área de Bahía
Blanca. Tesis Doctoral Universidad Nacional de La Plata, Argentina
Lopez Cazorla A (2000) Age structure of the population of weakfish Cynoscion guatucupa (Cuvier)
(Pisces: Sciaenidae) in the Bahía Blanca waters, Argentina. Fish Res 46:279–286. https://doi.
org/10.1016/S0165-­7836(00)00152-­1
Lopez Cazorla A (2004) Peces del estuario de Bahía Blanca. In: Piccolo MC, Hoffmeyer MS (eds)
El ecosistema del estuario de Bahía Blanca. Instituto Argentino de Oceanografía (CONICET-­
UNS), Bahía Blanca, pp 191–201. (233 p). ISBN 987-9281-96
Lopez Cazorla A, Sidorkewicj N (2009) Some biological parameters of Jenyns’s sprat Ramnogaster
arcuata (Pisces: Clupeidae) in South-Western Atlantic waters. Mar Biodivers Rec 2(e127):1–8.
https://doi.org/10.1017/S1755267209001286
López Rojas H, Bonilla Rivero AL (2000) Anthropogenically induced fish diversity reduction
in Lake Valencia Basin, Venezuela. Biodivers Conserv 6:757–765. https://doi.org/10.102
3/A:1008945813101
López OCF, Duverne LB, Mazieres JO et  al (2013) Microbiological pollution of surface water
in the upper-middle basin of the Reconquista river (Argentina): 2010–2011 monitoring. Int J
Environ Health Res 6(3):276–289. https://doi.org/10.1504/IJENVH.2013.054087
Luthardt V (2010) Monitoring of ecosystems: two different approaches – long-term observation
vs success control. In: Müller F, Baessler C, Schubert H, Klotz S (eds) Long-term ecological
research – between theory and application. Springer Sci Publ, Dordrecht. Chapter 22:317–325,
(475 p). ISBN: 978-90-481-8781-2
Marcovecchio JE, Freije RH (2004) Efectos de la intervención antrópica sobre sistemas marinos
costeros: el estuario de Bahía Blanca. Anales de la Academia Nacional de Ciencias Exactas,
Físicas y Naturales, Buenos Aires (Argentina) 56:115–132
Marcovecchio JE, Freije RH (eds) (2013) Procesos Químicos en Estuarios, Editorial de
la Univ Tecnológica Nacional (EdUTecNe), Buenos Aires (Argentina). E-book. ISBN
978-987-1896-16-5
Marcovecchio JE, Freije RH, Popovich CA et  al (2010) Long-term observational system for
oceanographic studies within Bahía Blanca estuary (Argentina): state of the art and perspec-
tives. AQUASHIFT: life in warming waters, Christian Albrecht’s University, Kiel
Martinez-Haro M, Beiras R, Bellas J et al (2015) A review on the ecological quality status assess-
ment in aquatic systems using community based indicators and ecotoxicological tools: what
might be the added value of their combination? Ecol Indic 48:8–16. https://doi.org/10.1016/j.
ecolind.2014.07.024
19  Estuarine Environmental Monitoring Programs: Long-Term Studies 545

Mayne J, Zapico-Goñi E (eds) (2017) Monitoring performance in the Public Sector future  –
Directions from international experience. Routledge – Taylor & Francis Group, London, p 52.
ISBN 13: 978-1-4128-0632-9
McClure ML, Hansen AJ, Inman RM (2016) Connecting models to movements: testing connectiv-
ity model predictions against empirical migration and dispersal data. Landsc Ecol 31:1419–
1432. https://doi.org/10.1007/s10980-­016-­0347-­0
McIntyre AD, Pearce JB (1980) Biological effects of marine pollution and the problems of moni-
toring. ICES Rapports et Procès-Verbaux des Réunions 179, Copenhagen (Denmark), 346 p
Minier C, Abarnou A, Jaouen-Madoulet A et  al (2006) A pollution-monitoring pilot study
involving contaminant and biomarker measurements in the Seine estuary, France, using
zebra mussels (Dreissena polymorpha). Environ Toxicol Chem 25(1):112–119. https://doi.
org/10.1897/05-­161r.1
Molnar JL, Gamboa RL, Revenga C et al (2008) Assessing the global threat of invasive species to
marine biodiversity. Front Ecol Environ 6(9):485–492. https://doi.org/10.1890/070064
Morrison ML, Block WM, Strickland MD et  al (2008) Inventory and monitoring stud-
ies. In: Morrison ML, Block WM, Strickland MD et  al (eds) Wildlife study design, 2nd
edn. Series on Environmental Management, Springer, New  York, 238 p. https://doi.
org/10.1007/978-­0-­387-­75528-­1_7
Neves R, Baretta J, Mateus M (eds) (2008) Perspectives on integrated coastal zone management in
South America. IST Scientific Publ, Lisbon. 679 p. ISBN: 978-972-8469-74-0
Ogata Y, Takada H, Mizukawa K et al (2009) International Pellet Watch: Global monitoring of
persistent organic pollutants (POPs) in coastal waters. 1. Initial phase data on PCBs, DDTs,
and HCHs. Mar Pollut Bull 58:1437–1446. https://doi.org/10.1016/j.marpolbul.2009.06.014
Oliver DM, Porter KDH, Pachepsky YA et al (2016) Predicting microbial water quality with mod-
els: over-arching questions for managing risk in agricultural catchments. Sci Total Environ
544:39–47. https://doi.org/10.1016/j.scitotenv.2015.11.086
Pali P, Swaans K (2013) Guidelines for innovation platforms: facilitation, monitoring and evalua-
tion. ILRI Manual 8, Nairobi. 42 p. ISBN 92-9146-311-6
Pearce JB (1998) A short history of marine environmental monitoring. Mar Pollut Bull 37(1–2):1–2
Pearson TH, Rosenberg R (1978) Macrobenthic succession in relation to organic enrichment and
pollution of the marine environment. Oceanogr Mar Biol: Ann Rev 16:229–311
Persson I-L, Nilsson MB, Pastor J et  al (2009) Depression of belowground respiration rates at
simulated high moose population densities in boreal forests. Ecology 90:2724–2733. https://
doi.org/10.1890/08-­1662.1
Phillips DJH (ed) (1980) Quantitative Biological Indicators. Their use to monitor trace metal and
organochlorine pollution. Applied Science Publ, Barking. 356 p. ISBN: 0853348847
Pierini JO, Streitenberger ME, Baldini MD (2012) Evaluación de la contaminación fecal en el
estuario de Bahía Blanca (Argentina) aplicando un modelo numérico. Rev Biol Mar Oceanogr
[online] 47:193–202. https://doi.org/10.4067/S0718-­19572012000200003
Porte C, Barceló D, Tavares TM et al (1990) The use of the Mussel Watch and molecular marker
concepts in studies of hydrocarbons in a tropical bay (Todos Los Santos, Bahia, Brazil). Arch
Environ Contam Toxicol 19:263–274. https://doi.org/10.1007/BF01056096
Possingham HP, Fuller RA, Joseph LN (2012) Choosing among long-term ecological monitoring
programs and knowing when to stop. In: Gitzen RA, Millspaugh JJ, Cooper AB et  al (eds)
Design and analysis of long-term ecological monitoring studies. Cambridge Univ. Press,
Cambridge. Chapter 23:498–508 (586 p). ISBN: 978-0-521-19154-8
Prince RC (2010) Bioremediation of marine oil spills. In: Timmis KN (ed) Handbook of hydrocar-
bon and lipid microbiology. Springer-Verlag, Berlin. 2618–2626 p. ISBN 978-3-540-77587-4
Ronco A, Peluso L, Jurado M et al (2008) Screening of sediment pollution in tributaries from the
southwestern coast of the Río de la Plata Estuary. Lat Am J Sedimentol Basin Anal 15:67–75.
ISSN 1669 7316
546 J. E. Marcovecchio et al.

Ronda AC, Arias AH, Oliva AL et al (2019) Synthetic microfibers in marine sediments and surface
seawater from the Argentinean continental shelf and a Marine Protected Area. Mar Pollut Bull
149:110618. https://doi.org/10.1016/j.marpolbul.2019.110618
Sánchez F, Prenski LB (1996) Ecología trófica de peces demersales en el Golfo San Jorge. Revista
de Investigación y Desarrollo Pesquero 10:57–71. ISSN: 0325-6375
Sardiña P, Lopez Cazorla A (2005a) Trophic ecology changes of the whitemouth croaker,
Micropogonias furnieri (Pisces: Sciaenidae), in South-Western Atlantic waters. J Mar Biol
Assoc UK 85:405–413. https://doi.org/10.1017/S0025315405011331h
Sardiña P, Lopez Cazorla A (2005b) Feeding habits of the juvenile striped weakfish, Cynoscion
guatucupa Cuvier 1830, in Bahía Blanca estuary (Argentina): seasonal and ontogenetic
changes. Hydrobiologia 532(1):23–38. https://doi.org/10.1007/s10750-­004-­8769-­0
Sathicq MB, Bauer DE, Gómez N (2015) Influence of El Niño Southern Oscillation phenom-
enon on coastal phytoplankton in a mixohaline ecosystem on the southeastern of South
America: Río de la Plata estuary. Mar Pollut Bull 98(1–2):26–33. https://doi.org/10.1016/j.
marpolbul.2015.07.017
Secretaría de Recursos Hídricos de la Nación (2007) Programa de monitoreo Embalse Río Hondo
(Diciembre de 2007). http://www.hidricosargentina.gov.ar/InformeMonitoreo_cap5.pdf
Seidl R, Spies TA, Peterson DL et al (2016) Searching for resilience: addressing the impacts of
changing disturbance regimes on forest ecosystem services. J Appl Ecol 53:120–129. https://
doi.org/10.1111/1365-­2664.12511
Sericano JL, Wade TL, Jackson T et al (1995) Trace organic contamination in the Americas: an over-
view of the US National Status & Trends and the International “Mussel Watch” Programmes.
Mar Pollut Bull 31:214–225. https://doi.org/10.1016/0025-­326X(95)00197-­U
Smith TB, Purcell J, Barino JF (2007) The rocky intertidal biota of the Florida Keys: fifty-two
years of change after Stephenson and Stephenson (1950). Bull Mar Sci 80:1–19
Smith DR, Lei Y, Walter CA et  al (2012) Incorporating predicted species distribution in adap-
tive and conventional sampling designs. In: Gitzen RA, Millspaugh JJ, Cooper AB et al (eds)
Design and analysis of long-term ecological monitoring studies. Cambridge Univ. Press,
Cambridge., Chapter17:381–396 (586 p). ISBN: 978-0-521-19154-8
Snelgrove PVR, Blackburn TH, Hutcjing P et al (1997) The importance of marine biodiversity in
ecosystem process. Ambio 26:578–583
Snyder EG, Watkins TH, Solomon PA et al (2013) The changing paradigm of air pollution moni-
toring. Environ Sci Technol 47:11369–11377. https://doi.org/10.1021/es4022602
Soller JA, Schoen M, Bartrand T et al (2010) Estimated human health risks from exposure to rec-
reational waters impacted by human and non-human sources of faecal contamination. Water
Res 44:4674–4691
Sousa R, Gutiérrez JL, Aldridge DC (2009) Non-indigenous invasive bivalves as ecosystem engi-
neers. Biol Invasions 11:2367–2385. https://doi.org/10.1007/s10530-­009-­9422-­7
Stelzenmüller V, Breen P, Stamford T et al (2013) Monitoring and evaluation of spatially managed
areas: a generic framework for implementation of ecosystem based marine management and its
application. Mar Policy 37:149–164. https://doi.org/10.1016/j.marpol.2012.04.012
Streitenberger ME, Baldini MD (2010) Deterioro de un área recreacional por efectos del volcado
de líquidos cloacales. Rev Argent Microbiol 42:307–310
Tanabe S (1994) International mussel watch in Asia-Pacific. Mar Pollut Bull 28(9):518–526.
https://doi.org/10.1016/0025-­326X(94)90057-­4
Tanabe S, Tatsukawa R (1991) Persistent organochlorines in marine mammals. In: Jones KC (ed)
Organic contaminants in the environment  – environmental pathways and effects. Elsevier
Science Publ, New York. Chapter 8: 275–289 (348 p). ISBN-I3: 978-94-010-8424-6
Tanabe S, Kannan N, Subramanian AN et  al (1987) Highly toxic coplanar PCBs: occurrence,
source, persistency and toxic implications to wildlife and humans. Environ Pollut 47:147–163.
https://doi.org/10.1016/0269-­7491(87)90044-­3
19  Estuarine Environmental Monitoring Programs: Long-Term Studies 547

Tanabe S, Iwata H, Tatsukawa R (1994) Global contamination by persistent organochlorines and


their ecotoxicological impact on marine mammals. Sci Total Environ 154:163–177. https://doi.
org/10.1016/0048-­9697(94)90086-­8
Tanabe S, Prudente MS, Kan-Atireklap S et al (2000) Mussel watch: marine pollution monitoring
of butyltins and organochlorines in coastal waters of Thailand, Philippines and India. Ocean
Coast Manag 43:819–839. https://doi.org/10.1016/S0964-­5691(00)00060-­0
U.S.  National Academy of Sciences (1975) Assessing potential ocean pollutants: a report of
the study panel on assessing potential ocean pollutants. US National Academy of Sciences,
Washington DC, 438 p
U.S. National Academy of Sciences (1977) Environmental monitoring, a report to the U.S.EPA
from the Study Group on Environmental Monitoring. US National Academy of Sciences,
Washington DC, 181 p
USEPA (U.S. Environmental Protection Agency) (2003) Bacterial water quality standards for rec-
reational waters (freshwater and marine waters) Status Report. US Environmental Protection
Agency publication N° EPA- 823-R-03-008. Office of Water, Washington, DC, 76 p
Valle Junior RF, Varandas SGP, Sanches Fernandes LF et  al (2015) Multi criteria analysis for
the monitoring of aquifer vulnerability: a scientific tool in environmental policy. Environ Sci
Policy 48:250–264. https://doi.org/10.1016/j.envsci.2015.01.010
Viana IG, Aboal JR, Fernández JA et al (2010) Use of macroalgae stored in an environmental spec-
imen bank for application of some European framework directives. Water Res 44:1723–1744.
https://doi.org/10.1016/j.watres.2009.11.036
Voget-Kleschin L (2013) Large-scale land acquisition: evaluating its environmental aspects against
the background of strong sustainability. J Agric Environ Ethics 26(6):1105–1126. https://doi.
org/10.1007/s10806-­013-­9448-­9
Wade TJ, Calderón RL, Sams E (2006) Rapidly measured indicators of recreational water qual-
ity are predictive of swimming-associated gastrointestinal illness. Environ Health Perspect
114:24–28
Whitfield AK, Elliott M (2002) Fishes as indicators of environmental and ecological changes
within estuaries: a review of progress and some suggestions for the future. J Fish Biol 61:229–
250. https://doi.org/10.1111/j.1095-­8649.2002.tb01773.x
Wolf B, Lindenthal T, Szerencsits M et al (2013) Evaluating research beyond scientific impact –
how to include criteria for productive interactions and impact on practice and society. GAIA
22(2):104–114. https://doi.org/10.14512/gaia.22.2.9
Wrona FJ, Johansson M, Culp JM et al (2016) Transitions in Arctic ecosystems: ecological impli-
cations of a changing hydrological regime. Eur J Vasc Endovasc Surg 121:650–674. https://doi.
org/10.1002/2015JG003133
Zhan X, Li MY, Zhang Z, Goossens B et  al (2006) Molecular censusing doubles giant panda
population estimate in a key nature reserve. Curr Biol 16:451–452. https://doi.org/10.1016/j.
cub.2006.05.042
Chapter 20
Environmental Education: Mud and Salt
Classrooms

Cristina Sanhueza and Paola Germain

20.1  Introduction

By the end of the 1960s and the beginning of the 1970s, the impact of human actions
on the environment became the focus of attention for government agencies and enti-
ties worldwide (Zabala and García 2008). Issues such as air and water pollution, the
loss of environments, and the extinction of species began to concern society, and
educational actions were taken into account as possible ways, among others, to
reduce the planet deterioration. Thus, in 1972 the United Nations convened the
Stockholm Conference, considered the first world environment forum, where the
environmental problems generated by the prevailing style of development, dispro-
portionate industrialization, and accelerated population growth were considered
(Zabala and García 2008). This conference has among its achievements the creation
of the United Nations Environment Programme (UNEP 1972) and the Stockholm
Declaration (1972) that establishes in its 19th principle: “Education in environmen-
tal matters, for the younger generation as well as adults, giving due consideration to
the underpriviliged, is essential in order to broaden the basis for an enlightened
opinion and responsible conduct by individuals, enterprises and communities in
protecting and improving the environment in its full human dimension….” This
principle is considered the formal beginning of environmental education, although,
previously, there were initiatives from nongovernmental organizations (NGO) and
educators.
There is a lot of definitions of environmental education, which is also character-
ized by its heterogeneity of practices, but all promote some kind of change and have
the action in common, that is, it is education for action (Melillo et al. 2015), and it
is oriented toward the training of active and committed citizens in building a more

C. Sanhueza (*) · P. Germain


GEKKO – Grupo de Estudios en Conservación y Manejo, Departamento de Biología,
Bioquímica y Farmacia, Universidad Nacional del Sur (UNS), Bahía Blanca, Argentina

© Springer Nature Switzerland AG 2021 549


S. M. Fiori, P. D. Pratolongo (eds.), The Bahía Blanca Estuary,
https://doi.org/10.1007/978-3-030-66486-2_20
550 C. Sanhueza and P. Germain

just, democratic, and caring society. It is a pedagogical-political practice that tends


toward a critical analysis of the socio-environmental reality in which its transforma-
tion in favor of responsible human development prevails (González-Gaudiano and
Puente-Quintanilla 2010). This is why environmental education must be incorpo-
rated into educational systems worldwide.

20.2  Environmental Education in Latin America

As in the rest of the world, the first steps in environmental education in Latin
America took place in the 1960s. Groups linked to popular education incorporated
an environmental vision in the conception of development, although at the begin-
ning they considered human beings as mere predators of nature. On the other hand,
ecological education incorporated a systemic understanding that interrelated social
and natural processes (Tréllez Solís 2006). Social problems such as poverty and
inequality, widespread in Latin America, and the need to make contributions and
involve communities in improving their living conditions crossed environmental
education. Thus, Latin America laid the foundations for participatory approaches
that were subsequently collected at the international level. Besides the initial con-
servationist proposals, it adds fundamental elements to involve the communities in
the processes toward the improvement of the environmental situation, including
natural, social, and economic components at the local level (Tréllez Solís 2006).
In the 1980s, a lot of NGOs related to conservation and environmental manage-
ment emerged. Educational material was generated, with theoretical and reflective
support for participatory action and with regional environmental thought. Different
topics are discussed: the model of development, its impact on ecosystems and its
link with the population impoverishment, the complex causality of the environmen-
tal problems, and the necessary understanding of the articulation of social and natu-
ral processes. All of this, in order to find concrete solutions to environmental
problems, related to a new organizational capacity of society as a whole, based on
the cultural values of communities, on popular creativity, and in its innovative
potential (Tréllez Solís 2006).
In the 1990s, Latin American governments began to include environmental edu-
cation in the curriculum of the different educational levels of formal education and
to develop educational materials. However, teacher training is scarce, and in prac-
tice, most of the countries tend to equate traditional natural science instruction with
the teaching of environmental education, and teachers use traditional methods and
approaches (Arias-La Forgia 1994). On the other hand, on many occasions, environ-
mental education does not represent an institutional interest but is the result of the
interest of local actors, who promote innovative initiatives on their own (González-­
Gaudiano & Puente-Quintanilla 2010).
Besides possessing great biological diversity, Latin American countries are mul-
ticultural societies, where indigenous or mestizo people cultures as well as peoples
of African and European ancestry coexist. Environmental education must integrate
20  Environmental Education: Mud and Salt Classrooms 551

the different knowledge, generating new pedagogical possibilities, especially


because the knowledge of indigenous peoples and people of African ancestry
transcends the anthropocentric perspective of colonial European cultural heritage
(García-Campos 2019).
Community interest in environmental issues has been increasing, which repre-
sents a great opportunity for environmental education. It must be a priority for polit-
ical authorities in charge of the formal education system, who should be more
actively involved. Considering humanity as part of nature, exchanging knowledge,
generating proposals for community participation, incorporating artistic and
awareness-­raising activities, and highlighting the importance of ethics and reflec-
tion are some of the aspects to be addressed, to involve the community in actions of
nature conservation and improvement of the quality of life.

20.3  Environmental Education in Argentina

In Argentina, in 2002, the General Law of the Environment (N° 25,675) was passed.
This law establishes the minimum budgets for the achievement of a sustainable and
adequate management of the environment, the preservation and protection of bio-
logical diversity, and the implementation of sustainable development in Argentina.
It also establishes a general framework for information and participation in environ-
mental matters, responsibility for environmental damage, and environmental educa-
tion. In several subsections, it mentions environmental education as a tool and a
means to achieve certain goals. For instance, Article 14 says: Environmental educa-
tion is the basic instrument to generate in citizens values, behaviors and attitudes
that are consistent with a balanced environment, aim at the preservation of natural
resources and their sustainable use, and improve the quality of life of the popula-
tion. In article 15, it refers to environmental education as a continuous and perma-
nent process, subjected to constant updating that, as a result of the orientation and
articulation of different educational disciplines and experiences, should facilitate
the comprehensive perception of the environment and the development of an envi-
ronmental conscience. The competent authorities must coordinate with the Federal
Councils for the Environment (COFEMA), Culture (CFC), and Education (CFE),
the implementation of plans and programs in the education systems, formal and
non-formal. The jurisdictions, depending on the determined basic contents, will
implement the respective programs or curricula through the pertinent norms.
Despite having a legal framework, environmental education in the different edu-
cational spaces is insufficient, low, or nonexistent. When it exists, it mostly refers to
global environmental problems, and little is known at a regional or local level. This
may be due to different factors, like the lack of information or the ignorance about
regional or local natural, cultural, and social environment. Many teachers have not
had such training and information, which is transferred to their students. This
unknowledge of the local environment leads teachers to address global topics such
as pollution or garbage under a general approach, without knowing the true values
552 C. Sanhueza and P. Germain

of the environment, the interactions that take place in it, the relation with human
beings, and the factors that threaten the health of that environment, among others.
The current way of life, the technology, the rush, the comfort, and the consumer-
ism have created a wall between nature and people. The societies forget or ignore
the essential, the relationship of our ancestors with nature, our relationship with the
environment, and that we are part of it. The rush leads to immediacy, loss of obser-
vation, contemplation, and search of answers through our means.
In a global context of loss of environments and biodiversity, and distance from
the natural environments of which we are a part, particularly by people living in
cities, environmental education is an essential tool for nature conservation, promot-
ing a change of attitudes toward the environment and a life in harmony with it. As
Leonardo Boff (1995) says “We do not live on Earth. We are Earth, part of Earth.
Between living and inert beings, between the atmosphere, the oceans, the moun-
tains, the earth’s surface, the biosphere and the anthroposphere, interrelationships
prevail. There is no addition of all these parts, but organicity between them.”

20.4  Environmental Education in the Bahía Blanca Estuary

Environmental education can be approached from three perspectives: community,


systemic, and interdisciplinary. This allows to achieve a comprehensive environ-
mental education (Álvarez 2004).
From a community perspective, it aims to form environmental attitudes and val-
ues to reduce the crisis and transform the predatory attitude of the human being
(Álvarez 2004). The environmental problems and their causes must be studied and
analyzed from a local to a global perspective. At first, start from the local problem
and from daily realities of the community, facing individuals with local environ-
mental realities, and then move along regional or global situations. A clear example
of this approach in the Bahía Blanca Estuary is the problem of plastic garbage in the
sea. From this problem in the coastal area, we can see how this problem affects
coastal activities and artisanal fishing or how it affects marine species. It is not nec-
essary to work with that issue on a global level, in distant and foreign seas, since it
is a local problem where the community could play an active role that can be known
and acted on the local environment. Many NGOs conduct coastal garbage censuses
and coastal cleanings, activities that involve the community and can generate aware-
ness and responsible attitudes toward the environment.
Environmental education must develop in all individuals of all ages in a com-
munity the capacity for critical observation, understanding, and responsibility
toward the environment. These skills must be acquired by all social actors. From a
community perspective, everyone should know the situation of the environment
where they live.
Following the community approach, to achieve a comprehensive environmental
education, the focus must be not only on environmental problems but also on the
value of the environment “without problems,” the environment in balance, fulfilling
20  Environmental Education: Mud and Salt Classrooms 553

all its ecological, social, and cultural functions with its values positively impacting
the daily life of the local community.
All environmental problems necessarily have a systemic constitution, under-
standing the system as “a set of elements related to each other and to the environ-
ment, which constitute a specific integral formation” (Álvarez 2004). Therefore,
understanding the environment as a system in which its elements are interrelated is
a fundamental characteristic of the environmental dimension, where the compo-
nents of the system are integrated into the physical, biotic, economic, and sociocul-
tural environment, added to the group of beliefs, values, techniques, and worldview,
shared by the members of the community (Álvarez 2004). This approach allows to
take into account the different dimensions that cross a certain environmental topic,
recognizing the interactions that exist between them and valuing each element that
makes up these dimensions. An example from this perspective in the Bahía Blanca
Estuary is the environmental conflict described later in this text, in Sect. 20.5. In that
conflict, a great diversity of dimensions such as ecological, biological, social, cul-
tural, economic, and social representations about the estuary of the community were
revealed. All these dimensions and the relationships between them were interpel-
lated by the activity, the dredging, that would be carried out in the environment. It
was a case in which it would be decided whether or not to protect aquatic birds,
marine fauna, or the ecosystem dynamic against job opportunities, but also it
required a comprehensive analysis of all dimensions, its elements, and interrela-
tions. In many arguments and opinions about this conflict, the estuary was
approached from a systemic perspective. However, more simplistic discourses were
also exists, in which the environment was summarized in only one dimension or one
element, leaving aside the systemic vision.
Interdisciplinarity represents a set of interrelated disciplines and with defined
relationships, so their activities do not occur in isolation, dispersed nor divided. The
articulation of the different disciplines allows a global understanding of a process
and then performing the analysis and solution of a particular problem (Álvarez 2004).
The incorporation of this interdisciplinary approach both in educational practice
and in the treatment of environmental conflicts presupposes the realization of a
planning that achieves an adequate organization of the analysis, which allows
understanding the complex structure of the environment. The interaction of its phys-
ical, biological, social, and cultural aspects is demonstrated, as well as providing a
clear awareness of the political, economic, and ecological interdependence of the
environment.
In addition to the three approaches to environmental education discussed above,
there is the symbolic vision of the environment. That refers to the different mean-
ings that social groups have on the environment, which vary depending on cultural
contexts (Pedroza and Argüello 2002). The coastline of the Bahía Blanca Estuary
has undergone multiple transformation processes that have led to a loss of biodiver-
sity and parts of cultural heritage. At the same time, port and industrial development
practically hinder people’s access to the coast, creating a kind of “collective amne-
sia” regarding the city’s marine and coastal character. This local natural environ-
ment is generally unknown by the new generations. This has led to its transformation
554 C. Sanhueza and P. Germain

into a city facing away from the sea for many years. Frequently, the inhabitants of
the area ignore the environmental values that the estuary contains, its scenic beauty,
and its importance for regional development. The idea that it is a “lost” place from
the point of view of contamination is also frequent. This symbolic vision of loss and
deterioration does not promote attitudes of caring for the environment.

20.5  S
 ocial Conflicts That Triggered the Emergence
of “Guardians of the Estuary”

Bahía Blanca is no stranger to the problem of globalization of the environment, of


ignorance of the local natural heritage. Many of its citizens do not know its coasts.
This is undoubtedly the most serious aspect that threatens the conservation of the
environment’s biodiversity, history, and culture.
The estuary coast has undergone multiple transformation processes associated
both with the growth of the city and with the construction and development of
industries and port terminals, leaving a few coastal remnants in a natural state.
These processes lead to an impoverishment of biodiversity and the loss of parts of
the cultural heritage resulting from the historical interaction between humans and
their natural environment.
How is the wall that has been built between the citizens and the coastal environ-
ment, its nature, history, and culture demolished? The first step is undoubtedly the
public recognition of the importance of the environment, both from a practical point
of view, as a provider of resources and ecological services, and from the admiration,
aesthetic enjoyment, and responsibility of the entire community as custodian of the
regional nature.
The first step to begin to demolish that wall was in May 2011. That year rumors
of a dredging project for the extension of the Principal Channel in the Bahía Blanca
Estuary began to circulate. That extension would reach the area called Puerto
Cuatreros and included the construction of a dock, on the north bank, to receive
regasification ships and an interconnection gas pipeline. Puerto Cuatreros (Figs. 2.1
and 2.2; Chap. 2) is located in the innermost sector of the estuary, where small boats
circulate, and is not subject to deep dredging. This information generated concern
in neighbors and nongovernmental organizations, which began to make meeting and
learn about this initiative. Finally, in July 2011, the responsible for the project for-
malized the presentation of the proposal, and in September 2011, they communicate
it to the local media. Something that was not explicitly mentioned in the project, but
that would surely occur in the future, was that the deepening of the dredging to
Puerto Cuatreros would allow the development of new industries, nonexistent in the
area until this moment. For a year and a half, local people organized into an assem-
bly, which in January 2012 became the “Buenos Aires South Environmental
Assembly” (AABAS). A great variety of activities were launched to inform the
population, spread the value and importance of the ecosystem conservation, and
20  Environmental Education: Mud and Salt Classrooms 555

express the opposition to the project: natural interpretation walks in the terrestrial
area of the estuary, informative talks, debates, collection of signatures, bike riding,
murals, photo exhibitions, film cycles, interventions in different public events,
marches, recitals, celebrations for “World Wetlands Day,” bird watching, kayaking,
creation of the “From the mud” murga, teachers’ training, etc. At first, a few people
participated in the activities, but later they became massive, achieving a citizen par-
ticipation never seen before in an environmental cause. At the same time, profes-
sionals specialized in biological, ecological, environmental, port, and urban
planning issues from the National University of the South (UNS), the Argentine
Institute of Oceanography (IADO), and the National Technological University
(UTN), all academic institutions, expressed their disagreement with the project with
numerous reasons related to the environmental impact of this undertaking. This
commitment of academic institutions with an environmental-social problem was
also unprecedented. Finally, for economic reasons, the project was rejected, but the
spreading of the value of the Bahía Blanca Estuary continues, to be able to face
future threats that endanger its conservation. With no doubt, the way forward was
always environmental education, to transmit the value of the Bahía Blanca Estuary,
its biodiversity, its stories, and the interrelationships between nature and the city.

20.6  Guardians of the Estuary

To meet this challenge, the GEKKO Group (Group of Studies in Conservation and
Management), belonging to the Department of Biology, Biochemistry, and
Pharmacy of the National University of the South, organized in 2013 the course
workshop for the training of guides-interpreters of the Bahía Blanca Estuary, called
“Guardians of the estuary.” Through this project, the aim was to combine the work
of university teachers and students with environmental conservation NGOs and
residents of the towns of General Cerri, Ingeniero White, and Bahía Blanca in order
to develop educational materials and activities that promote the approach of com-
munity to natural environments, promoting the feeling of belonging to the environ-
ment, without losing sight of the environmental and social reality of the community.
In addition, those interested in developing educational and tourism ventures in con-
tact with nature, a genuine alternative for sustainable local development, were
trained. After this project, in 2014 a group of people interested in working in the
conservation, valuation, and dissemination of the biological, social, and cultural
wealth of the estuary is formed. This is how the NGO – Guardians of the estuary – in
Bahía Blanca was born.
Guardians’ different activities are carried out throughout the year, such as talks
at the different educational levels, outings of natural interpretation along the estuary
coast, awareness activities such as kayaking, walks to learn about the estuary envi-
ronment, trail development self-guided to discover firsthand the value of the envi-
ronment, as well as activities and dissemination material such as posters, brochures,
murals, and photographic exhibitions.
556 C. Sanhueza and P. Germain

Box 20.1 Estuarial Classrooms


Coastal marine areas generally undergo major anthropic changes that affect
the functioning of the ecosystem and its biodiversity and are also particularly
affected by climate change. On the other hand, they provide numerous eco-
system services that depend on their good state of conservation, such as the
control of coastal erosion, the production of food, and the possibility of carry-
ing out sports, tourism, and recreational activities. Thus, coastal marine envi-
ronments represent a great pedagogical opportunity for environmental
education.
In the next section, we will take a walk through the different classrooms
that we can find in the estuary: classrooms that are not only to visit with an
educational institution, in a formal education, but classrooms for the entire
community, inclusive classrooms, and action learning classrooms. Here
knowledge is within the reach of whoever wants to take it and whoever wants
to discover or build it; you just have to enter and enjoy those classrooms. They
are the appropriate space to achieve an integral, community, systemic, and
interdisciplinary environmental education.
Classrooms as Pharmacies: History of the Native Peoples
The territory that we walk today in the estuary has its history from the time of
the native peoples and from before them. Many of the plants and animals that
we see today in the estuary were the source of food, remedies, and survival of
the native peoples that inhabited this environment. Inquire how was the rela-
tionship between human beings and their natural environment in those times.
It offers us a great opportunity to rebuild the identity of the community with
its local environment and with its natural and cultural heritage. Many of the
plants that grow today in the Bahía Blanca Estuary are recognized by popular
wisdom for their medicinal uses, examples of which are Cyclolepis genistoi-
des, Schinus longifolius, and Jodina rhombifolia, among others. There are
also species like Sarcocornia perennis that was used and is still used today as
food (Sanhueza et al. 2014). The skin of some animals, such as guanaco, was
used for awnings and clothing. Some report that the clay from the estuary
coast was used to make vessels and pots (personal communication from a
renowned ceramist from the city).
Supply Classrooms
The sea as a provider of food and work continues to be of importance for
many families of artisanal fishers in the Bahía Blanca Estuary; currently 72
families live on this resource. In the past, many more families practiced this
art, as explained in Chap. 18 of this book; artisanal fishers have been protago-
nists of a strong socio-environmental conflict.
20  Environmental Education: Mud and Salt Classrooms 557

Knowing this rich source of food gives us the idea of an environment as a


supplier, as a supermarket, both for humans and for many other animals. The
importance of recognizing it as a finite resource, which must be cared for and
not exploited, provides us with the right space to learn, develop educational
programs, know the diversity that lives in the sea, know the dynamics of natu-
ral populations, know the interactions that take place there, recognize what
are the possible threats that put this resource at risk, and thus make decisions
to avoid its collapse.
This classroom also offers us the possibility of changing the paradigm and
thinking about the estuary from ecocentrism and not from anthropocentrism,
to think that the estuary and its coasts are not only a source of food for human
beings but also for thousands of species that inhabit or pass through it, such as
migratory birds. For several species of migratory birds, the Bahía Blanca
Estuary represents a key feeding site on their journey. If this environment
disappeared, they would be left without food to continue their journey.
Classrooms That Provide Shelter
Undoubtedly, all animals need a shelter, either to protect themselves from the
weather, to survive predators, to reproduce, or simply to rest. The estuary
seems to be a very hostile environment, a lot of wind, salt, and sun and little
water; it seems to be an unfriendly environment to take refuge, but it still turns
out to be a key place for many species. We can find caves, burrows, nests that
belong to crabs, polychaetes, guinea pigs, birds, ants, spiders, and snake; all
find the right conditions to make the estuary an ideal refuge.
Art Classrooms
Art and environmental education are two doors to wonder, to wonder at the
natural and social environment that surrounds the individual, and this is the
door to the question and therefore to meaningful learning. Art allows us to
perceive and get to know the environment and its socio-environmental prob-
lems in a new way. The artistic gaze allows us to discover new elements and
integrate other points of view. The German painter Paul Klee said that “Art
does not reproduce the visible. It makes it visible” (López Abril et al. 2017).
The NGO Guardians of the estuary used street art to make visible the great
biodiversity that exists in the estuary by creating collective murals in different
sectors of the city, making visible to the community the presence of flamin-
gos, dolphins, sea turtles, sharks, and wildlife that the community is unaware
of their existence (Figs. 20.2 and 20.3). Appealing to people’s astonishment in
order to bring them closer to the knowledge of the natural environment is a
powerful tool.
558 C. Sanhueza and P. Germain

Creativity and art allow us to perceive the elements of the landscape, as


well as the landscape as a whole, in a different way, activating different per-
spectives that integrate intellectual and emotional learning. The Bahía Blanca
Estuary has been a source of inspiration for literary works, drawings, paint-
ings, music, and ceramics, among other areas. Many of these artistic expres-
sions have been both individual and collective and have thus become
multipliers of knowledge about the estuary.
Classrooms of Different Spatial and Temporal Scales
The estuary is a natural space in which a large number of species live together,
occupying different environments (microenvironments and macro-­
environments) generating innumerable interactions between individuals of
the same species and between different species, as well as interactions with
the environment that surrounds them. Observing this space at different scales
will give us opportunities to discover different points of view.
Used points of view (Arengo et al. 2002) in environmental education offer
us a very powerful tool to perceive the impact that disturbances will have at
different spatial and temporal scales. If a bush falls, it probably will not affect
the flight of the Cathartes, but it will be a catastrophe for the firewood-gath-
erer who had his nest in it or the common yellow-toothed cavy with their
burrows underneath; the soil conditions will be modified generating windows
of opportunities for different plants species. The footsteps on crabs’ land trail
will be imperceptible to Olrog’s gull but a terrible impact on crabs. We can
also witness disturbances on a larger scale that affect many species, including
humans, such as the dredging that occurs in the estuary. Actions of daily life
can generate disturbances without people perceiving it, such as the waste that
is generated in homes or an external drain from the laundry or kitchen that
ends up in the estuary. Recognizing the different scales and their interactions
offers endless opportunities for learning and reflection on the action of the
human being on the environment.
Disturbance can be natural, such as a fire, a hurricane, and a flood, and
after a disturbance natural or anthropic, we can ask ourselves: What happens
next? Who does it affect? How affects? What were the impacts? How long
will it take for the environment to recover? We can work many of these ques-
tions to develop formal or non-formal environmental education activities.
The tide, seasonal changes, day, and night offer changing scenarios and
patterns that can be used as teaching and learning tools. We can inquire about
the consequences of disturbances, of temporal changes at different spatial
scales, or of different intensities.
20  Environmental Education: Mud and Salt Classrooms 559

Fig. 20.1  Educational walks of natural interpretation carried out by the NGO “Guardians of the
estuary” in the Bahía Blanca Estuary. (Photo by Guardians of the estuary)
560 C. Sanhueza and P. Germain

Fig. 20.3 (a) Collective green turtle mural made in a neighborhood of the city of Bahía Blanca.
(b) Collective mural of the estuary biodiversity made in the multiple use salon of the fishing club
in Puerto Cuatreros, General Daniel Cerri. (Photo by Guardians of the estuary)

The main objective of the group is to create a space where you can experiment
with all the senses, reinforcing the identity with the local environment. As a final
goal, it seeks to modify attitudes, acquiring new habits and achieving a change in
values both from the exercise of sensitivity and rationality, being part of current
historical social processes as active subjects and responsible for caring for the envi-
ronment. They work under the motto “Together we can become pillars of a para-
digm shift, turning our gaze towards nature and towards sustainable ways of life.”
Between the months of September to November 2018 and March to May and
September to November 2019, the NGO conducted guided tours of natural
20  Environmental Education: Mud and Salt Classrooms 561

Fig. 20.2  Collective murals made in the downtown area of the city of Bahía Blanca. Photos by
Guardians of the estuary

interpretation with different educational institutions of private and public manage-


ment of the initial, primary, and secondary levels, including the modalities of addi-
tional support needs and adults. In addition, the tertiary and university levels and
groups of PAMI (Integral Medical Care Program, health insurance for retirees and
pensioners) and open trips to all public were added. The total of visitors who enjoyed
this environment, learning and being part of it, were 2500 people, making one or two
weekly outings only (Fig. 20.1). If you take into account that the work carried out by
this NGO is completely voluntary, without a contribution from the state, neither from
the public nor the private sector, your work is highly valued and invites us to question
the role of the state in environmental education that it preaches in its legislation.
The NGOs that carry out environmental education activities in the estuary do not
meet the great demand of educational institutions and diverse groups that request a
guided visit. This realizes the need that exists in the community to know its local
environment in order to value it and thus preserve it (Box 20.1).
To learn more about the Guardians of the estuary and find out about the activities
they carry out, you can visit:
Facebook – Guardianes del Estuario-Bahía Blanca
Instagram – guardianesdelestuariobahia

20.7  Final Comment

The challenge is to achieve an environmental education that teaches how to produce


and consume responsibly, to live together to share life with other living beings, to
get out of the noise in which we live submerged, and to listen to nature, which rec-
ognizes, conserves, and thanks services that natural ecosystems provide us, to teach
us to be more sensible beings.
562 C. Sanhueza and P. Germain

References

Álvarez CO (2004) Educación ambiental a partir de tres enfoques: comunitario, sistémico e inter-
disciplinario. Revista Iberoamericana de Educación, ISSN-e 1681-5653, ISSN 1022-6508, Vol.
35, N°. Extra 1
Arengo N, Elfi Chaves M, Feinsinger P (2002) Guía metodológica para la enseñanza de ecología
en el patio de la escuela. Audobon Programa para America Latina y el Caribe
Arias-La Forgia A (1994) Environmental education in the school systems of Latin America and
the Caribbean. Working Papers 4. Agency for International Development. Washington, USA
Boff L (1995) Nueva era. La civilización planetaria: desafíos a la sociedad y al cristianismo. Ed.
Verbo Divino. España
García-Campos HM (2019) Environmental education from an intercultural approach: a glimpse
into Latin America. Southern Afr J Environ Educ 35. https://doi.org/10.4314/sajee.v35i1.12
González-Gaudiano EJ, Puente-Quintanilla JC (2010) El perfil de la educación ambiental en
América Latina y el Caribe: Un corte transversal en el marco del Decenio de la Educación para
el Desarrollo Sustentable. Pesquisa em Educação Ambiental 5:27–45
López Abril M, Vega M, Loren L (2017) El Arte como herramienta para la Educación Ambiental.
Boletín del Centro Nacional de Educación Ambiental
Melillo F, Belmes A, Priotto G et al (2015) Educación ambiental: Ideas y propuestas para docentes.
Nivel secundario. Secretaría de Ambiente y Desarrollo Sustentable de la Nación y Ministerio
de Educación de la Nación. Argentina
Pedroza R, Argüello F (2002) Interdisciplinariedad y transdisciplinariedad en los modelos de ense-
ñanza de la cuestión ambiental. Cinta de Moebio: Revista Electrónica de Epistemología de
Ciencias Sociales, ISSN-e 0717-554X, N°15
PNUMA. Programa de Naciones Unidas para el Medio Ambiente (1972) Resolución 2997/24
Sanhueza C, Germain P, Zapperi G et al (2014) Plantas nativas de Bahía Blanca y alrededores,
descubriendo su historia, belleza y magia. TELLUS
Stockholm Declaration (1972) United nations conference on the human environment.
Stockholm, Sweden
Tréllez Solís E (2006) Algunos elementos del proceso de construcción de la educación ambiental
en América Latina. Revista Iberoamericana de Educación 41:69–81
Zabala I, García M (2008) Historia de la Educación Ambiental desde su discusión y análisis en los
congresos internacionales. Rev Invest 63:201–218
Index

A Arctic skua (Stercorarius parasiticus),


Abiotic factors, 120, 121 337
Abiotic stress, 436 Arctocephalus australis, 365
Acartia tonsa, 96 behavior, 374
Acoela, 169 description, 373
Acoelomorpha, 167 distribution, 374
Acrostichum danaeifolium, 4 habitat, 374
Agreement on the Conservation of Small threats and conservation status, 375
Cetaceans of the Baltic and North Area for Bird Conservation (AICA), 486
Seas (ASCOBANS), 364 Arenicola marina, 235
Alitta succinea, 200 Argentina, 262, 263, 310, 336, 340, 350,
Allenrolfea sp., 453 551, 552
A. patagonica, 447, 450 Argentine Biogeographic Province, 11
Almirante Brown Maritime Park, 476 Argentine National Assessment, 417
American oystercatcher (Haematopus Argentine shelf, 11
palliatus), 346–348 Arid Diagonal, 8
Amphibalanus amphitrite, 201 Arid environments, 443, 444
Anaerobic decomposition processes, 190 Arroyo Los Gauchos Nature Reserve, 420
Analysis of variance (ANOVA), 64 “Arroyo Parejas” Spa, 27
Analytical quality (AQ), 64 Art classrooms, 557, 558, 560
Annelids, 172, 173 Artemesia longinaris, 97, 256, 257, 263, 264,
Anomalocardia brasiliana, 529 266, 479, 502, 509
Anoxia, 60 Asia-Pacific Mussel Watch Program
Anoxic conditions, 123 (APMWP) region, 530
Antarctic fur seal (Arctocephalus gazella), Asteraceae, 451
360, 368 Atmosphere, 345
Antarctic minke whales (Balaenoptera Atomic absorption spectroscopy (AAS), 64
bonaerensis), 370 Atriplex portulacoides, 458
Anthropogenic disturbance, 218 Atriplex undulata, 447, 450, 473
Anthropogenic impact, 220
Anthropogenic pressures, 155
Aquatic environment, 65 B
Aquic Ustifluvents, 443 Baccharis divaricata, 420
Aquisalids, 442 Bacillariophyceae, 117
Arctic fox (Vulpes lagopus), 401 Bacteria, 535

© Springer Nature Switzerland AG 2021 563


S. M. Fiori, P. D. Pratolongo (eds.), The Bahía Blanca Estuary,
https://doi.org/10.1007/978-3-030-66486-2
564 Index

Bahía Blanca district geographic zones, 101


Maldonado Channel, 24 geomorphic setting, 21
Petrochemical Pole, 25, 26 geomorphology, 21
Port of Ingeniero White, 26, 27 hinterland, 31
Puerto Cuatreros, 24 holoplankton, 96
Puerto Galván, 24, 25 hot and dry summers, 84
Sauce Chico River, 24 humid conditions, 22
urban effluent discharge treatment plant, 24 humid subtropical and semiarid climates, 1
Bahía Blanca Estuary, 120, 194, 264–266, hypersalinity, 84
327, 366–367, 404–405, 552–554 inland and coastal sabkhas, 6
algal settlement, 131–133 inorganic nutrients, 86
Argentine coast, 17 invasive species, 92
aridity, 6 La Vidriera Salt Flat, 21
artificial hard substrate, 131 light microscopy, 86
Atlantic coast, South America, 1, 2 location, 17
Atlantic South America, 1 macroalgal assemblages, 134, 136, 137
Bahía Falsa Channel, 19, 20 macroalgal biodiversity, 126
barrier-lagoon depositional systems, 4 macroscopic algae, 126, 133
Bermejo Island, 20 marine-dominated environment, 32
biogeographic classification, marine marine sediments, 22
environment, 10, 11 marsh vegetation, 6
biological invasions, 97 Merin lagoon, northern Uruguay, 4
biomass, 83 meroplankton, 97
biomass-specific productivity, 85 mesozooplankton community, 85,
Canal Principal, 31 91, 96, 97
cell’s cytoplasm, 94 microalgal taxa, 126
channel margins, 84 microbial mats, 126, 133
chlorophyll a, 91, 92 microbial plankton, 93
climate change, 21, 92 microphytobenthos, 126
coastal areas, 92 microscopic and macroscopic thalli, 126
coastal civilizations evolution, 31 microzooplankton community, 85, 86, 92
coastal system, 17 mid-Holocene transgressive stage, 22
coastal uses and environments nanoflagellates, 86, 92
Bahía Blanca District, 19 Napostá Grande Stream, 18
Coronel Rosales District, 20 nitrogenous compounds, 84
Colorado River, 6, 21, 22, 31 nitrogen-rich effluents, 85
commercial harbor activities, 91 North America, 1
community level, 94 nutrient and organic matter, 83
diatoms, 86 nutrients concentration, 92
dinoflagellates, 86 occupation and land use (see Occupation
diversity and composition, 84, 127–130 and land use, Bahía Blanca)
Dos Patos Estuary, 4 oligotrichs, 86, 91
drainage basin, 21 organisms, 96
ecology, 95 origin, 21
economic and sanitary concerns, 101 Patagonian Region, 6
ecosystem, 84 pelagic ecosystems, 99
ecosystem functioning, 93 pelagic environment, 94
effluents transport, 92 phagotrophic protists, 86, 90
El Rincón coastal system, 12 phototrophic and phagotrophic protistan
estuarine basin, 90 communities, 86
euryhaline species, 83 phytoplankton community, 84, 85, 91–93
faunal distribution, 84 plankton biomass, 83, 92
geographical location, 5 Principal Channel, 17, 18
Index 565

protistan plankton, 87–89 diversity, 113


protists, 84 epiphytic algae, 138
regional climate and vegetation patterns forms and body types, 115
(see Regional climate/vegetation habitats, 114, 115
patterns, central East Argentina) mat-forming macroalgae, 130,
Río de la Plata Estuary, 4–6 131, 138–140
sabkha, 6 metabolic processes, 114
S. alterniflora, 6 microalgae, 113
S. ambigua, 6 morpho-functional classification, 118–120
salinity, 83 organisms, 113
salt flats, 6, 18 photoautotrophs, 113
San Sebastián Bay, 6 taxonomic classification, 116–118
Sauce Chico River, 18 taxonomic lassification, 117
sea level, Holocene, 2–4, 18 Benthic communities, 155, 538–540
sediments, 91, 127 Benthic habitat mapping, 236
shear stress, 91 Benthic organisms, 154
solar radiation, 85 Benthic-pelagic coupling, 192–194
spatial gradients, 83 Big animals, 161
tidal cycle, 84 Biodiversity, 155, 199, 218, 538
Trinidad Island, 19 Biogenic mounds, 191
urban development, 92 Bioindicator, 161, 162
urban effluents, 93 Biological and fisheries data, 262
Vidriera Salt Flat, 19, 32 Biological communities, 539
water temperature, 92 Biological diversity, 550, 551
water turbidity, 90 Biological processes, 282
wind action, 22 Biomass, 220
wind patterns, 90, 92 Biomass patterns, 218
zooplankton, 96, 98–99 Biomitigation, 139, 140
Bahía Blanca Port Management Biomonitoring, 162
Consortium, 261 Biota, 64
Bahía Falsa Channel, 19, 20 Biotic interactions, 121–123
Balaenoptera acutorostrata, 360 Bioturbation, 130, 189–192
Balaenoptera musculus, 360 Blidingia marginata, 134, 136
Barnea lamellosa, 195 Blue carbon ecosystems, 489
Barnea truncata, 195, 196 Bottlenose dolphins (Tursiops truncatus),
Bathymetric variations, 237 361, 378
Benthic algae, 121 Brachidontes rodriguezii, 200, 201
abiotic factors, 120, 121 Breeding, 337, 338
adaptations mechanisms Brevoortia aurea, 285
biofilms and microbial mats, 125 Brown rats (Rattus norvegicus), 401
Cladophora, 124 Bryopsis plumosa, 136
Cladophorales, 124 Buccinanops deformis, 181
cyanobacteria, 125 Buenos Aires province, 419–422
estuaries, 123, 124 Burrowing crab (Neohelice granulata), 346
extracellular polymeric substances, 125 Bycatch, 309, 311, 314, 317
filamentous algae, 124
free-floating mats, 124
microorganisms, 125 C
microphytobenthic communities, 125 Cactaceae, 448
microphytobenthos, 125 Calidris canutus, 484
solar radiation, 126 Campylaimus arcuatus, 157
zonation patterns, 123 Campylaimus bonariensis, 157
biotic interactions, 121–123 Canal Principal, 36
566 Index

Cangrejal, 298 biological biodiversity, 486


Carbon sequestration, 362 communication, 487
Carcharhinus brachyurus, 486 comparative analysis, 487
Carcinus maenas, 122 context indicators, 485
Carnivores, 407 criteria and indicators, 485
Cephalothorax projects, 253 critical management activities, 488
Ceramiales, 138 ecological criteria, 485
Ceramium diaphanum, 134, 135 ecosystems, 487
Cetaceans, 364 ecosystem services, 488
Chaetoceros diadema, 94 education, 488
Chaetomorpha, 124 environmental education programs, 488
Chaetonotida, 174 environmental impact assessments, 487
Chaetophractus villosus, 406, 407, 413 environmental problems, 487
Charadrii, 328 environmental protection goals, 487
Charadriidae, 328, 329, 486 flora and fauna species, 487
Charadriiformes, 327, 328, 330–332, 335, indicators, 485, 486
337, 339 inter- and multidisciplinary work, 486
Charadrius falklandicus, 333 landscape level, 485
Chasmagnathus granulata, 183 legal protection, 488
Cheliped, 196 legislation, 487
Cheloniidae, 308 management, 484, 488
Chemical industry, 26 management effectiveness, 485
Chemosynthetic bacteria, 362 management plan, 487
Chenopodiaceae, 449 management practices, 487
Chlorarachniophytes, 118 methodology, 488
Chlorophyll a, 53, 60, 62, 63, 71 objectives and management
Chlorophyta, 118 planning, 488
City of the Seven Ports, 496 planning, 488
Cladophora surera, 135 planning variables, 486
Cladophorales, 124 pressures, 489
Climate Change Conference, 490 sustainable financial resources, 488
Climatology variables, 485
air-water interaction processes, 37 fishing resources, 470
annual temperature, 37 geo-environmental characteristics, 471–474
autumn, 38 geo-environmental units, 470
average wind speeds, 37 habitats, 470
Bahía Blanca Estuary, 37 human activities, 469
climate variability, 37, 38 human occupation, 475
El Niño events, 38 human pressure, 469
summer, 38 implementation, 470
water vs. atmosphere, 38 Latin America, 470, 471
wind, 37 management categories, 471
winter, 38 management measures, 469
Coastal ecosystems, 192, 257 marine resources, 470
Coastal habitats, 403 natural environments, 469
Coastal Nature Reserve, 477 preservation and conservation, 471
Coastal protected areas recreational activities and tourism, 470
coastal planning, 469 recreational and productive activities, 469
conservation management, 470 SDG, 489, 490
ecological interest, Northern Sector, transformations, 469
483, 484 Coastal salt flats, 437
environmental quality, 469–470 Coastal system, 17, 191
evaluation Coastal wetlands
Index 567

arid climates, 436 arthropods, 160


arid environments, 443, 444 Bahía Blanca Estuary, 163, 164
arid and semiarid coastal locations, 436 benthic organisms, 160
coastal landscape, 439–441 community, 165
commercial fish and shellfish species, 435 ecology, 163
ecosystems, 435, 436 harpacticoid species, 165
hydrology, 436 Harpacticoida, 160
intertidal zone, 436 marine interstitial sediment
landscape patterns, 441–443 communities, 160
plant communities, 444, 446–448 marine sediments, 160
relative sea level, 436 meiobenthos, 163
salt marsh plants, 436 meiofauna, 161
salt stress, 436 microfitobenthos, 165
sea level changes, 436–438 seasonal cyclical changes, 160
vegetation, 435 spectrum, 161
watersheds, 435 temporal pattern, 165
wetland system, 471 Corallinales, 138
Coleofasciculus chthonoplastes, 128, 129, 131 Corbula gibba, 223
Community, 551 Corbula patagonica, 220, 221, 223, 228
ecology, 218 Correlation analysis, 64
perspective, 552 Cortaderia spp., 421
Conepatus chinga, 418, 419, 421 Corvina negra, 288
Conservation, 550–552, 554–556 Corvina rubia, 287
Bahía Blanca Estuary, 339, 340 Corymorpha januarii, 101, 198
benthic invertebrates, 345 Coyote (Canis latrans), 399
biodiversity, 338 Crabeater seals (Lobodon carcinophaga), 360
coastal plant invasion, 342, 343 Crangon crangon, 122
contaminants, 349 Crustaceans, 229, 253, 256, 258, 260, 262,
diversity, 341 264, 266, 284
ecosystems, 338 Ctenomys australis, 403, 406, 411, 420
estuarine environments, 345 Ctenomys talarum, 403
habitats, 341 Cultural heritage, 553, 554
industrial and domestic discharges, 345 Cyanobacteria, 116, 531, 532
international and local communities, 350 Cyanoliseus patagonus, 511
invasive alien species, 339 Cyanophyta, 116
legal protection, 350 Cyclolepis genistoides, 447, 451, 452
management challenge, 342, 344 Cynoscion guatucupa, 258, 287, 293, 297
marine and estuarine ecosystems, 345 Cyrtograpsus altimanus, 180, 181, 234
microplastics, 349 Cyrtograpsus angulatus, 184, 348
natural processes, 341
petrochemical and industrial hub, 341
physicochemical changes, 341 D
pollutants, 345 Darwin fox (Lycalopex fulvipes), 408
populations, 342 Data Deficient, 365
seabirds, 339–341 Dendrobranchiata, 253
shorebird, 339, 340 Dermochelyidae, 308
stress factors, 341 Detritivores, 296
threats, 339 Deuterogonaria sp., 170
trace metals, 345–349 Diatoms, 117
Conservation of Sea Turtles (CIT), 309, 315 Didelphimorphia, 405
Convention on Biological Diversity, 470 Didelphis albiventris, 405
Convention on Wetlands, 489 Dinophyta, 117
Copepods, 98 Discaria americana, 420
568 Index

Dissolved gases, 52 homogeneous waters, 12


Dissolved oxygen (DO), 52, 53, 59, MPA, 12
60, 63, 443 oceanic conditions, 12
Distribution pattern, 59 Elephant seals (Mirounga leonina), 361
Diverse group, 220 Elymus athericus, 458
Diversity, 557 Endangered species, 315
Dogs (Canis familiaris), 424 Environmental conflicts, 553
Dolphins (Delphinus delphis), 361 Environmental conservation, 555
Domestic cats (Felis catus), 424 Environmental education
Dos Patos Estuary, 4 Argentina, 551, 552
Dules auriga, 279 Bahía Blanca Estuary, 552–554
Dusky dolphins (Lagenorhynchus challenge, 561
obscurus), 361 conference, 549
definition, 549
human actions, 549
E Latin America, 550, 551
Earth’s surface, 190 pedagogical-political practice, 550
Ecocentrism, 557 planet deterioration, 549
Ecological and biogeochemical processes, 192 social conflicts, 554, 555
Ecological education, 550 Environmental Education Program, 383
Ecophysiological parameters, 60–63, 533 Environmental quality index (EQI)
Ecosystem functions aggregate indices, 481
aerial photographs, 454 agricultural practices, 479
biogeochemical processes, 457 analysis, 479
carbon dioxide, 454 application, 482
carbon sequestration, 457 Bahía Blanca Estuary, 476
coastal waters, 457 Buenos Aires Province, 476
land cover changes, 456 capacity, 480
landscape evolution, 455 conservation species, 480
metals, 457 DPSIR model, 476
microbes, 455 environmental condition, 483
mineral uptake, 457 environmental indicators, 476
nutrients, 457 geo-environmental units, 478
phosphorus sequestration, 460 geographical area, 481
plant communities, 454 geomorphological units, 482
plant tissues, 457 human actions, 476
Principal Channel, 454 human activities, 477
shoreline transgression, 455 integrated management guidelines, 482
soil carbon, 457 IR resources, 478
soil organic material, 455 management plans, 477, 479
tidal inundation, 454 maximum and minimum values, 478
tidal saline wetlands, 454 maximum value, 481
tissues, 460 minimum value, 481
Ecosystem quality, 162 multi-use reserve, 482
Ecosystems, 419 natural environments, 478, 479
Ecotone, 444 naturalness index, 481
Ectocarpaceae, 137 NI values, 478
Educational material, 550 optimal environmental quality, 482
El Niño-Southern Oscillation (ENSO), 8 Patagonian region, 476
El Rincón coastal system, 12 physical environment, 476, 481
coastal waters, 12 pressure, 479
fish and zooplankton species, 12 pressure index, 477
high-salinity shelf waters, 12 Principal Channel, 478
Index 569

productive and industrial economic composition, 277–280


activities, 478 crabs, 297
solid waste deposit, 478 Cynoscion guatucupa, 287
strategic management guidelines, 482 distribution, 277–280
tourist activities, 480 ecology, 295
urban household waste deposits, 481 ecosystem, 296
variability, 479, 482 estuaries, 275
waste index, 479 abiotic variables, 292
Environmental specimen banks (ESBs), 522 Bahía Blanca Estuary, 292–294
Environmental strategy, 530 biology and life history, 292
Environmental stress, 162 estuarine-resident guilds, 292
Epibiosis, 138 food chains, 291
Epiphytic algae, 138 hypersaline conditions, 292
Escherichia coli, 536 migratory diadromous species, 290
Estuaries, 192, 470, 486, 490 species frequency, 291
aquatic environments, 52 extrinsic interactions, 296
Bahía Blanca Estuary, 54 fish community, 282
biodiversity, 52 fish species, 276, 281
biological processes, 52 food, 275
chemical functioning, 71, 72 habitat resources, 282
chemical/physicochemical parameters, 52 habitat uses, 290
environmental condition, 52 habitats, 275
environmental studies, 52 hydrographic conditions, 275
freshwater discharges, 54 hypotheses, 276
industrial nucleus, 54 inner and middle zones, 278
parameters, 52 inner zone, 278
physicochemical parameters, 52 mechanisms, 276
processes, 52 Micropogonias furnieri, 287, 288
temperature, 55, 57 morphological and behavioral
tidal channels, 54 constraints, 296
wastewater discharges, 54 Mustelus schmitti, 282, 284
Estuarine environments, 121 Myliobatis goodei, 284, 285
Estuarine water, 64 Odontesthes argentinensis, 286, 287
Eubalaena australis, 361 organisms, 296
Euglenophyta, 118 Paralichthys orbignyanus, 289, 290
Euhaline/myxohaline zones, 220 Pogonias cromis, 288, 289
Eurytemora americana, 96, 101, 102, 104 polychaetes, 297
Eustatic sea-level rise, 3, 4 Porichthys porosissimus, 283, 286
Exotic species, 539 Ramnogaster arcuata, 285, 286
Extracellular polymeric substances (EPS), 125 reproductive mode functional group, 295
richness-energy relationship, 276
salt marshes, 278
F straggler species, 294
Fabaceae, 452 trophic spectrum, 297
Filaments, 115 Fishers’ ecological knowledge
Fish assemblages (FEK), 511–513
artisanal fishery, 299, 300 Flooding Pampa, 10
Bahía Blanca Estuary, 277, 298 Flota amarilla, 264
biological systems, 276 Food web, 360
biotic and abiotic factors, 275 Foundation for Reception and Marine Animal
Brevoortia aurea, 285 Attendance and Environmental
chondrichthyan, 282, 298 Education (FRAAM), 316, 364,
community structure, 297 368, 383
570 Index

Franciscana Management Areas (FMAs), 376 activities, 555, 561


Franciscanas (Pontoporia blainvillei), 361 classrooms, 556–558, 560
Frankenia juniperoides, 447 community, 555
Freshwater supply, 155 local environment, 560
Fur seals (Arctocephalus australis), 361 social conflicts, 554, 555

G H
Gadus morhua, 122 Habitats, 114, 115, 420
Galictis cuja, 410, 417, 418, 421 Haematopus palliatus, 333
Gastrotrichs, 174 Halectinosoma parejae, 163, 165
GEKKO Group (Group of Studies in Halophytes, 191, 453
Conservation and Harbor seals (Phoca vitulina), 361
Management), 555 Harpacticoida, 158
Gelatinous plankton, 99 Hawksbill turtle (Eretmochelys
Gelidiales, 138 imbricata), 311
Gelidium sp., 134 Heating and cooling, 281
G. pusillum, 135 Heavy metals, 64–66, 377, 534
Generalized linear models (GLMs), 421 Heleobia australis, 180, 187–189, 200
Genidens barbus, 280, 294 Herbivores, 296
Genital system, 254 Herpailurus yagouaroundi, 407, 421
Geo-environmental units, 473 Heteromastus similis, 184
Geoffroea decorticans, 447, 452, 475 Heterostachys sp., 342, 453
Geoffroy’s cat (Leopardus geoffroyi), 408 H. ritteriana, 444, 451
Geographical spaces, 471 Himantopus mexicanus, 333
Geomorphology, Bahía Blanca Estuary Hirundo rustica, 484
bedforms, 36 Holocene, 438, 439, 455, 461
biological and physical interactions, 37 Hudsonian godwit (Limosa haemastica), 328
Canal Principal, 36 Human-driven process, 419
characteristics, wetlands and tidal Humid conditions, 22
courses, 34 Humpback whale (Megaptera novaeangliae),
erosional stage, 37 362, 369
marshes, 33 Hyalis argentea, 420
pioneer plants, 33 Hydrobates pelagicus, 425
Principal Channel, 36, 37 Hydrocarbons, 53, 64, 66, 67, 534
rivers, 35 compounds, 536
S. alterniflora, 32 deposits, 236
scour holes, 35, 36 Hydrochoerus hydrochaeris, 403, 412
sediments, 33, 37 Hydrochory, 344
Spartina marshes, 33, 37 Hydrodynamic and geomorphological
tidal channels, 32, 33 complexity, 219
tidal depressions diversity, 35 Hydrodynamic models, 344
tidal flats, 33, 35, 37 complex biological cycle, 262
Glacio-isostatic adjustment, 3 dispersal scales, 261
Glaucophyta algae, 116 environmental variabilities, 261
Glycera americana, 234 gravity effect, 260
Gonadosomatic index values (GSI), 284 habitat, 262
Grahamia bracteata, 447 high-density bathymetry, 261
Gray fox (Urocyon cinereoargenteus), 399 larvae dispersal, 261
Green turtle (Chelonia mydas), 311 measurement efforts, 260
Guanacos (Lama guanicoe), 406 MOHDIS model, 259
Guardians of the Estuary MOHID model, 260, 261
Index 571

numerical models, 260 coastal ecosystems


oceanography and behavior, 262 allochthonous energy subsidy, 401
penaeoid shrimp, 261 allochthonous marine materials, 398
physical-chemical parameters, 260 arthropods, 399
plankton, 262 carnivore species, 399
Principal Channel, 260 coastal foxes, 401
processes, 260 domestic grazers, 402
spatial and temporal domains, 260 ecological processes, 402
waves, 261 endogenous energy, 398
Hydrodynamics, 191 flow velocities, 398
Hydrologic conditions, 438 geographical area, 399
Hydroperiod, 436 gradual change, 399
Hyperbenthivores, 296 intertidal community structures, 401
Hyperphagia, 334 intertidal energy resources, 402
Hypersaline, 56 island ecosystems, 401
Hypertrophy, 334 littoral and supralittoral zone, 399
Hypothesis, 344 macrophytes, 399
Hypoxia, 60 mammals, 402
marine-based organic materials, 400
marine ecosystems, 399
I maritime mammals, 402
Index of trophic diversity (ITD), 162 mesopredators, 400, 401
Induced coupled plasma with optical oceanography and coastal
resolution (ICP-OES), 64 topography, 398
Industrial Park, 25 predatory arthropods, 399
Infauna, 122 salt and minerals, 402
Inorganic nutrients, 52, 53, 60, 62, 63 shorebirds, 399
Interdisciplinarity, 553 supralittoral zone, 399
Intermediate-field coastal regions, 3 terrestrial consumers, 398
International Union for Conservation of terrestrial/semi-terrestrial
Nature (IUCN), 485 arthropods, 398
Invasive exotic species (IES), 341 coastal urbanization, 423–425
Invertebrate hosts, 362 coastal zone, 397
Invertebrates, 254 ecological theory, 397
Irrigation, 24 ecosystem, 397
IUCN Red List of Threatened Species, 365 influence, 397
mammalian carnivores, 398
marine-derived food resources, 397
K maritime mammals, 403, 405, 407,
Kalyptorhynchia, 169, 170 408, 410
Killer whales (Orcinus orca), 369 Landscape patterns, 441–443
Kinorhynchs, 172 Large Marine Ecosystems (LMEs), 10
Kit fox (Vulpes macrotis), 399 Lari, 328
Köppen-Geiger climate classification, 7 Larus atlanticus, 474, 486
Larus dominicanus, 335
Last Glacial Maximum (LGM), 3
L Latin America, 550, 551
“La Vidriera” salt flat, 21, 340 Leatherback turtle (Dermochelys coriacea),
Laeonereis acuta, 181, 182, 184, 193, 311, 313
194, 239 Legal framework, 551
Lagoon-La Vidriera-Principal Channels, 19 Leopard seals (Hydrurga leptonyx), 360
Land mammals Leopardus colocolo, 408, 421
abiotic and biotic factors, 397 Leopardus geoffroyi, 416, 417
572 Index

Leptuca uruguayensis, 196, 197 coastal management and


Lepus europaeus, 406 conservation, 230
Limonium brasiliensis, 447 deposit-feeding organisms, 234
Liolaemus multimaculatus, 420 drag effects, 234
Lipophilic pollutants, 363 dredging activities, 234
Liriope tetraphylla, 101 ecosystem engineering roles, 236
“Liverpool of the South”, 496 ecosystem functions, 234
Loggerhead turtle (Caretta caretta), 311, 312 ecosystems services, 230
Long-term monitoring program (LMP), 523, estuaries, 230
524, 526 inner zone, 232, 234
Lycalopex gymnocercus, 415, 416 mechanisms, 233
Lycium chilense, 420 metal concentrations, 235
Lyncodon patagonicus, 408 mollusks, 236
nutrients, 236
organic matter, 234, 235
M organisms, 234
Macroalgae, 114, 122, 139 organization levels, 230
assemblages, 134, 136, 137 outer zone, 234
diversity, 133 Principal Channel, 232
Macrobenthic communities quantification, 233
acoustic approach, 236, 237 quantitative analysis, 230
atmosphere, 242 sandy sediments, 236
Bahía Blanca Estuary, 217 species-specific relationships, 233
benthic-feeding fish species, 216 subsurface organic particles, 235
benthic organisms, 243 suspension-feeders, 236
biogenic shallow gas, 242, 244 transport, 235
biogeochemical cycles, 241 gas-bearing sediments, 242, 243
biological effect, 242 gas deposits, 242
biomechanical properties, 215 high-contaminant loads, 216
biophysical processes, 216 human activities, 216, 240
biotic vs. abiotic factors, 215 impacts and conservation issues
characteristics, 216 anthropogenic activities, 237
coastal seafloor, 243 anthropogenic effluents, 238
ecological characteristics, 215 anthropogenic pressures, 237
ecological effects, 244 antioxidant enzymes, 239
ecological studies, 216 benthos and environmental
ecosystem services, 216 disturbances, 237
environmental monitoring program, 217 biodiversity, 239
environmental properties, 216 biofouling communities, 239
functional ecology channel, 238
abiotic conditions, 235 climate change, 237
benthic communities, 233 coastal environments, 239
benthic fauna, 234 detoxification, 239
biogeochemical processes, 235 ecosystem, 239
biogeochemistry, 235 El Rincón, 239
biological traits, 230–232, 234, 235 estuaries, 237
bioturbators, 235 estuarine system, 238
body-sized organisms, 234 human activities, 237
burrowing and deposit-feeding, hydrodynamic network, 238
235, 236 industrial wastewater, 238
characteristics, 232 inner zone, 240
Index 573

marine biota, 239 diverse group, 179


metals, 238 ecology, 195
molecular levels, 239 ecosystem services, 186, 202
monitoring program, 238 environmental changes, 186
non-native marine species, 239 environmental matrix, 202
organisms, 238, 240 environments, 180
petrochemical discharges, 238 epifaunal organisms, 202
political-economic influence, 240 feeding habits, 180
polychaetes, 238 herbivory, 199
Principal Channel, 238 human activities, 182
sectors, 240 hydrodynamic forces, 179
sediment, 238 Leptuca uruguayensis, 196, 197
socioeconomic variables, 240 Neohelice granulata, 185–187
wastewater effluent discharge, 239 organisms, 180, 199
marine ecosystem, 216 plankton community, 202
methane emissions, 241 polychaetes, 181, 182
methane gas, 241, 242 salt marshes, 186, 199
organic matter, 243, 244 sand flats, 180
patterns, 218–220 seawater, 199
polyps, 240 sediments, 199
population ecology, 216 shrimps, 179
sedimentation, 216 soft-bottom communities, 200
sediment grain size, 217 tidal flats, 179
seismo-acoustic methods, 241 water/sediment interface, 179
sessile fauna, 242 Macrobenthivores, 296
shallow gas, 241, 243 Macrobenthos, 162
species-specific relationships, 215 Macrodasyda, 174
structural and functional Macroinfauna
characteristics, 215 benthic organisms, 189
subtidal environments, 216 benthic-pelagic coupling, 192–194
surface deposit-feeders, 244 bio-stabilizing and -destabilizing
taxonomic and functional changes, 216 organisms, 189
taxonomic groups, 218–220 bioturbation, 190–192
variables, 215 erosion, 189
water salinity, 217 geomorphology, 190
Macrobenthic invertebrates, 190 mobility, 189
anthropogenic contaminants, 199 organisms, 190
Argentine coast, 200 salt marshes, 190
Barnea truncata, 195, 196 spatial and temporal scales, 190
beachrocks, 200 terrestrial and marine sediments, 189
benthic macrofaunal community, 180 tidal flats, 190
biological processes, 180 Macroinvertebrates, 237
biology, 195 Macrostomorpha, 169–171
biomass, 201 Macruronus magellanicus, 279, 280, 294
biotic factors, 180 Magallana gigas, 92, 97, 188, 200, 202, 203,
bioturbation activities, 180 480, 487
climate change, 203 Management effectiveness, 485
coastal ecosystems, 199 Mandatory/directed monitoring, 524
coastal regions, 199 Marine coastal protected areas (MPAs), 471
Corymorpha januarii, 198 Marine communities, 218
crab bioturbation, 199 Marine environments, 161
crabs, 179, 182 Marine invertebrates, 346
574 Index

Marine mammals Microbenthivores, 296


Bahía Blanca Estuary, 366–367, 370 Microenvironments, 189
cetaceans, 364 Microhabitats, 188
clinical examinations/gross necropsy, 368 Microorganisms, 437, 526
conservation Microphytobenthic algae, 123
animal health and welfare, 383 Microphytobenthic and macroalgal
anthropic disturbances, 382 communities, 121
apex predators, 385 Microphytobenthic communities, 114, 127
Bahía Blanca Estuary, 384 Microphytobenthos, 114, 125
cetaceans, 384 Microplastic particles, 534
ecosystems function, 384 Microplastics (MP), 53, 64, 69, 71, 72, 349
environments, 385 Micropogonias furnieri, 69, 258, 287, 288,
factors, 386 293, 297, 479, 502
feeding resources, 385 Microscale, 188
government agencies, 383 Microturbellarians, 156
institutions, 382 Migration patterns, 337
intervention, 383 Migratory birds, 557
national and provincial government Mixotrophs, 114
agencies, 382 Molecular machinery, 161
natural threats, 382 Mollusk shells, 131
plant communities, 385 Monitoring programs
spatial distribution, 382 alternative studies, 522, 527
transient species, 384 application, 522
consumers, 360, 361 aquatic and terrestrial systems, 527
ecosystems, 359 aquatic environments, 530
entanglement lesions, 364 behavior and natural distribution, 532
estuary, 380–382 benthic and demersal organisms, 526
food, 359 benthic communities, 538–540
food habits, 359 biological features, 533
fossil assemblages, 364 biology of fishes, 537, 538
habitats, 359, 362, 363 chemical aspects, 533, 534
health conditions, 365 chemical substances, 533
mortality, 364 coastal marine systems, 533
nutrient and material flux, 360, 362 data analysis methods, 523
oceanographic areas, 359 decision-making on environmental
pinnipeds, 364, 365 issues, 533
population parameters, 369 definition, 521
prey, 361 design and application models, 532
sentinels, 363, 364 environment, 531
skin and muscle tissues, 364 estuary, 532
Marine microscopic animals, 155 habitat degradation, 527
Marine organisms, 200 historical applications, 527–530
Marine shrimp, 253 human activities, 522, 533
Marine waters, 527 hypersaline, 532
Mean values assessment, 64 LMP, 523, 524, 526
Meidiama sp., 170 marine environment, 527
Meiobenthic animals, 161 microbiological aspects
Meiobenthic organisms, 155 advantage, 535
Meiofauna, 153–155, 161, 162 application, 537
Mephitidae family, 410 aquatic ecosystems, 535
Merluccius hubbsi, 262 bacterial populations, 535, 536
Metabolic functions, 114 coastal waters, 535
Microalgae, 113 diffuse sources, 534
Index 575

dumping, 537 National Action Programs for the


estuary, 536 Conservation, 383
food chain, 535 National Institute for Environmental Sciences
hydrocarbon-degrading (NIES), 64
microorganisms, 536 National Institute of Standards and
infections, 535 Technology (NIST), 64
levels of pollution, 534 National Ministry of Environment and
marine and coastal resources, 534 Sustainable Development, 316
municipal sewage, 537 Nauplius, 254
Principal Channel, 537 Near-field locations, 3
qualitative and quantitative Nematoda
distribution, 535 bacteria, 156
sediments, 537 Bahía Blanca Estuary, 158, 159
self-cleaning capacity, 535 biodiversity, 156, 157
sewage treatment plant, 537 buccal cavity, 156
THB, 536 community, 157, 160
waters, 537 deposit feeders, 157
Mussel Watch Program, 527–530 diverse group of animals, 156
natural systems, 522, 527 environmental variables, 157
physicochemical parameters, 532 estuary, 159
pollutants and physical pressures, 527 feeding habits, 157
practical and dynamic method, 523 food resources, 159
process, 522 harpacticoid copepods, 156
quality, 531 Harpacticoida, 158
screening, 531 intertidal muds, 156
standardization of procedures, 527 macrobenthos, 156
statistical methods, 523 meiobenthic inhabitants, 156
study cases, 532 meiobenthic marine nematodes, 156
systematic development, 533 meiofaunal taxa, 157
trace metals, 526 microambients, 158
water and surface sediments, 531 morphological and molecular
water quality control methods, 523 information, 156
water resource, 531 organic matter, 157
work strategies, 522 sand grain, 156
Monocorophium insidiosum, 243, 244 submerged surface, 156
Moose (Alces alces), 399 Nematodes, 155
Morphodynamics, 236 Nemerteans, 173, 174
Morpho-functional classification, 118–120 Neodalyellida, 169
Mugil liza, 279, 280 Neohelice granulata, 130, 131, 182–187,
Mussel Surveillance, 527 191
Mussel Watch Program, 527–530 Neosiphonia harveyi, 137
Mustelus schmitti, 282, 284, 294, 297 Nesomyinae, 425
Myliobatis sp., 294 Net primary productivity (NPP), 63
M. goodei, 279, 284, 285 Nitrate, 63
Myrsine parvifolia, 4 Nitrite, 63
Nitrogen, 71
Nongovernmental organizations (NGOs), 497,
N 507, 549, 554
Nannopus sp., 165 Northern Patagonian Frontal System, 11
Narrownose smooth-hound (Mustelus Notorynchus cepadianus, 375
schmitti), 509 Nutrient, 60
National Action Program, 310 Nutrient fluxes, 193
576 Index

O Petrogypsids, 443
Occupation and land use, Bahía Blanca pH, 52, 57, 59, 71
Coronel Rosales district, 27, 28 Phaeophyta, 117
environmental situations, 23 Phaeopigments, 53
fresh water, 23 Pholadidae, 195
industrial and railway occupation, 23 Photophores, 286
IslandS, 28 Photosynthetic pigments, 52, 53
Patagonia, 23 Photosynthetic stems, 444
port works, 23 Phragmites australis, 458
Principal Channel, 23 Physical oceanography, estuary
Villarino district, 23, 24 ebb currents, 43, 44
See also Bahía Blanca district ebbing condition, 41
Odontesthes argentinensis, 286, 287, 501 freshwater inputs, 38
Oil industry, 26 geomorphology, 40
Old Marine Plain (OMP), 440 harmonic components, 41, 42
Olindias sambaquiensis, 101, 103, 104 historical distribution, 39
Olrog’s gull (Larus atlanticus), 336, 343, longitudinal temperature distributions, 40
348, 508 mesotidal, 41
Omnivores, 296 NW and N winds, 42
Ontological perspectives, 496 Principal Channel, 41
Organic deposition, 192 quasi-stationary tidal wave, 41
Organic matter, 52, 71, 121, 192 Sauce Chico River, 38
Organisms, 130, 200 sewage discharges, 39
Organochlorine compounds (OCs), 69 tidal circulation, 43
Organochlorine pesticides (OCPs), 67, 534 tidal currents, 43
Organochlorines (OCs), 529 tidal cycle, 41, 45
Otaria flavescens, 365, 385, 502, 509 velocity profiles, tidal cycle, 45, 46
behavior, 372 wave types, 42
description, 370, 371 wind waves, 43
distribution, 371 winds, 42
habitat, 371 Physicochemical factors, 345
threats and conservation status, 372 Phytochemicals, 139
Phytoplankton, 121
blooms, 362
P cells, 193
Pacific Decadal Oscillation (PDO), 9 Pilayella littoralis, 122
Pampas fox (Lycalopex gymnocercus), 408 Piscivores, 296
Paralichthys orbignyanus, 289, 290, 292, 501 Pitar rostratus, 220
Particulate organic matter (POM), 53, 62, 63 Planktivores, 296
Passive/curiosity-driven monitoring, 524 Planosiphon nakamurae, 137
Patagonian region, 256 Plant associations, 441
Patagonian Shelf, 10, 11 Plant communities, 443, 444, 446–448
Patagonian Shelf Large Marine Ecosystem, 10 Plastic pollution, 349
Pathogen transmission, 363 Platyhelminthes, 166, 167
Peisos petrunkevitchi, 97, 257–259, 293, 297 Pleoticus muelleri, 255, 256, 262–264, 266,
Penaeoids, 254 502, 509
Penaeoidea, 253, 261 Pleurobranchaea maculata, 228
Penguins, 360 Podosira stelligera, 129
Perna viridis, 530 Pogonias cromis, 288, 289
Persistent Organic Pollutants (POP), 69 Polar bears (Ursus maritimus), 361
Pesticides, 53, 64, 67, 69, 72 Polychaetes, 218, 284
Petrochemical industry, 26 Polycyclic aromatic hydrocarbons
Petrochemical Pole, 25, 26 (PAHs), 66, 67
Index 577

Polysaccharide chains, 139 Regional climate/vegetation patterns, central


Polysiphonia abscissa, 134 East Argentina
Polysiphonia morrowii, 134, 135, 137 Arid Diagonal, 8
Pontoporia blainvillei, 368, 486 ENSO cycle, 8, 9
behavior, 377 Flooding Pampa, 10
description, 375 interannual rainfall variability, 8
distribution, 376 Köppen-Geiger climate classification, 7
habitat, 376 lands, 10
threats and conservation status, 377 Monte Region, 10
Porichthys porosissimus, 283, 286, 292, 297 Pampas Region, 8, 9
Port of Ingeniero White, 26, 27 PDO, 9
Porzana spiloptera, 475 rainfall, 8
Postlarvae, 254 Rolling Pampa, 10
Present Marine Plain (PMP), 440 SOI, 9
PRICTMA) (Regional Program of Southern Hemisphere, 8
Investigation and Conservation of temperate oceanic and cold desert
Argentine Sea Turtles), 310, 311 climates, 7
Primary production, 71 vegetation types, 9
Principal Channel, 36, 37, 45, 46, 218–221, Regional/local level, 551
439, 440, 472, 539, 554 Relative sea level, 436–439, 441, 455
Probability of observation (Prob “Ricardo Eliçabe” Oil Refinery, 25–26
Obs), 330–332 Ringed seal (Pusa hispida), 361
Procyonidae family, 399 Ringtail (Bassariscus astutus), 399
Promonotus, 170 Río de la Plata Estuary, 4–6, 10
Proseriata, 170, 171 Rolling Pampa, 10
Protected Provincial Natural Area, 383
Provincial Interest, 474
Pseudostomella, 174 S
Puerto Commercial Bahía Blanca, 26 Sabkha, 6
Puerto Coronel Rosales zone, 261 Salinity, 52–55, 71, 218, 220
Puerto Rosales area, 27 Salorthids, 442
Puma concolor, 408, 409, 414, 415, 421 Salsola soda (Amaranthaceae), 342, 344
Punctaria latifolia, 136 Salt marshes, 180, 182, 183, 186, 187,
Pyropia sp., 135 190, 199
aerenchyma, 448
anatomical changes, 448
Q cytoplasm, 453
Question-driven monitoring, 524 environmental stress, 448
Quinquelaophonte aestuarii, 163, 165 halophytes, 453
morphological adaptations, 448
oxygen availability, 448, 452
R plants, 121
Raccoon (Procyon lotor), 401 saline soils, 453
Rainfall/freshwater discharge, 437 soil salinity, 453
Ramnogaster arcuata, 258, 280, 285, 286, spatial distribution, 448
298 waves and storms, 448
Rapid Assessment Prioritization Protected Salt pans, 437
Area Management, 485 Salt wedge, 218
Recent Marine Plain (RMP), 440 “Sansinena” Industrial Meat Plant, 24
Red algae, 116, 117 Saraquita, 285
Red knot (Calidris canutus), 333 Sarcocornia sp., 444, 453
Red List of Threatened Species for Argentina, S. ambigua, 4, 6, 7, 342, 444, 446, 449
365, 369 S. magellanica, 7
Redoximorphic indicators, 443 S. perennis, 260, 298, 473, 474, 484, 556
578 Index

Satellite Monitoring Program, 317 Sexual maturity, 285


Schoenoplectus alterniflora, 6 Shallow bay systems, 121
Schoenoplectus californicus, 5 Shelf-Break Front, 11
Sciaenidae, 377 Shorebirds, 327
Scolopaci, 328 American oystercatcher, 334
Scolopacidae, 328, 329 Bahía Blanca Estuary, 329
Sea lions (Otaria flavescens), 361 conservation status, 330–332
Sea otters (Enhydra lutris), 360, 363 inland habitats, 328
Sea turtles mechanoreceptors, 329
ancestral cultures, 308 migration, 333, 334
in Argentina, 310 migratory pattern, 330–332
biology, 311–314 morphological variations, 328
Buenos Aires Province, 316 non-breeding season, 334
coastal artisanal fisheries, 309 patterns, 329
community, 316 populations, 334
conservation problems, 314 tidal flats, 328
ecological role, 308 wetlands, 334
factors, 317 white-rumped sandpiper, 333
feeding and development zone, 317 Shrimps, 254
fishermen communities, 316 Argentina, 262, 263
FRAAM foundation, 315 Artemesia longinaris, 256, 257
global warming, 309 Bahía Blanca Estuary, 264–266
governmental and nongovernmental eggs, 259
organizations, 309 estuarine and coastal ecosystems, 255
habitat, 317 estuary, 258
interdisciplinary and multidisciplinary fishing activities, 266
activities, 314 flood cycle, 258
juvenile, 317, 318 integrated coastal zone management, 267
marine ecosystem, 308 Lagrangian particles, 267
marine turtles, 317–320 larval location and developmental
meat and egg consumption, 309 stage, 258
national academic circuit, 314 meroplankton, 266
nutrients, 308 mesozooplankton, 259
satellite monitoring, 317 microscopic algae, 259
vertebrates, 308 migration process, 258
veterinary actions, 321–323 MOHID model, 266
Western South Atlantic, 315, 316 multidisciplinary analysis, 267
Seabirds, 327 numerical modelling, 267
breeding, 337, 338 patial and temporal distribution, 258
Charadriiformes, 335 Peisos petrunkevitchi, 257, 258
definition, 335 Pleoticus muelleri, 255, 256
feeding, 336 prawns, 258
flexibility, 336 zooplankton, 266
foraging, 336, 337 Silicate, 63
heterogeneous, 335 Single linear regression analysis, 64
oceanographic systems, 335 Skuas, 337
water, 336 Small-scale artisanal fishers
Seafloor, 236 academic environments, 493
Seaward margin, 437 ancestral culture and knowledge, 504
Sedimentary particles, 131 artisanal beach nets, 501
Sediments, 64, 155 artisanal fishing, 495, 497, 498
Sergestoids, 254 Bahía Blanca Estuary, 502, 514
Sergestoidea, 253, 254 biology, 495
Index 579

coastal ecosystems, 495 social resistance, 494


cognitive formation, 499 sociodemographic profile, 502
communities and cultures, 494 socio-ecological conflicts, 495
companies/enterprises, 498 socio-environmental conflicts, 494, 514
complexity, 494 topophilia developments, 496
conservation paradigms vulnerability, 494
Bahía Blanca Estuary, 509 Social and natural processes, 550
communalism, 506, 509 Social conflicts, 554, 555
development, 507 Socio-environmental conflicts, 493, 556
human compound, 506 Soft-bottom microalgal assemblages, 127–130
human ecology, 506 Soil-forming processes, 443
human-environment relations, 507, 509 South American fur seal (Arctocephalus
orientalism and paternalism, 506, australis), 386
509, 510 Southern elephant seal (Mirounga
port-petrochemical complex, 508 leonina), 365
prawn, 509 Southern Oscillation Index (SOI), 9
protected areas, 507 Southern Patagonian Front, 11
shrimp, 509 Southern right whale (Eubalaena
social movements, 508 australis), 369
social sciences, 506 Spartina spp., 187, 444, 458
socio-environmental conflicts, 507 S. alterniflora, 4, 6, 32, 187–189, 298, 441,
creative transformations, 494 444, 446, 457–460, 473, 484, 513
debate, 496 S. anglica, 458, 459
ecology, 495 S. densiflora, 4, 442, 446, 473, 474
economic cooperation agreements, 499 S. maritima, 458
economic refuge, 501 Spatial and long temporal scales, 363
environmental conflicts, 514 Spatial and seasonal dynamics, 280
environmental legislation, 495 Spatial and temporal scales, 558
environmental racism, 494 Spatial ecology, 419
estuary, 496 Spatial patterns, 121
ethnographic exploration, 497 Spatial variability, 219
FEK, 511–513 Spatial variations, 218
fishers’ conflict, 500 benthic communities, 223
gender roles, 505, 506 biomass, 221
geomorphology, 498 crustaceans, 229
industrial knowledge, 495 density, 221
intertidal zone, 501 diversity, 223
maritime-coastal resources, 496 ecological functions, 221
maritimity, 496 ecological patterns, 230
maritimization, 496 external zone, 223
natural resources, 514 geographical characteristics, 221
nature/environment, 493 inner zone, 223
neo-extractivism, 494 macrobenthic communities, 221
petrochemical complex, 498 myxohaline and euhaline sectors, 221
port system, 499 polychaetes, 223, 229
postharvest activities, 503 primary producers, 230
poverty, 494 sediments, 221
protected areas, 498 sewage and industrial effluents, 221
rapacity, 499 species, 224–228
reconversion plan, 502 taxonomic identification, 228
rights of nature, 495 temporal variations, 230
social and territorial demands, 494 wastewater discharges, 223
social fragmentation, 494 Spawning capacity, 256
580 Index

Spectrum, 162 diversity, 166, 168


Spheniscus magellanicus, 262 dominance indexes, 169
Sporobolus, 458 environment conditions, 170
Steller’s sea cows (Hydrodamalis gigas), 360 freshwater species, 167
Sterna hirundinacea, 335 industrial activities, 167
Sturnella defilippii, 475 light microscope, 166
Stylatula darwini, 182, 220, 234, 243 macrofaunal predators, 167
Suaeda divaricata, 448, 450, 451 macroturbellarians, 166
Submersed seagrass, 121 meiofaunal community, 167
Supralittoral macroalgae, 114–115 meiofaunal taxa, 171
Supralittoral zone, 442 microalgae, 167
Sus scrofa, 413, 414 mixotrophic behaviour, 167
Suspended particulate matter (SPM), 64 northern hemisphere, 167
Sustainable development, 551 Platyhelminthes, 166
Sustainable development goals (SDG), regeneration power, 166
489, 490 sediment destabilization, 167
Sympterygia bonapartii, 292 small-sized interstitial spaces, 170
Symsagittifera, 170 Turbidity, 53, 59, 71
Systemic constitution, 553 Tursiops truncatus gephyreus, 369
behavior, 378–380
description, 378
T distribution, 378
Tardigrades, 172 habitat, 378
Temperature, 52, 53, 55, 57, 71 threats and conservation status, 380
Terebellides totae, 223 Turtle-excluding devices (TED), 309
Terrestrial and aquatic systems, 345
Terrestrial heterotrophic bacteria (THB),
536 U
Textures, 236 Uca uruguayensis, 184
Thalassarche melanophris, 474 Ulva lactuca, 122, 136
Thalassiosira sp. Ulvophyceae, 135
T. curviseriata, 63, 93, 95 United Nations Environment Programme
T. minima, 63 (UNEP), 527, 549
T. pacifica, 94 Ursus americanus, 402
Thelycum, 254 US Federal Environmental Protection Agency
Thermic cycle, 55 (US EPA), 527
Tidal currents, 43
Tidal flats, 179, 180, 188, 190, 197
Tierra del Fuego Island, Argentina, 2 V
Trace metals, 53, 65, 66, 72, 345–349 Vascular plants, 437
Transformation processes, 554 Vidriera, 448
Transgressive stages, 439 Villarino district, 23, 24
Trinidad Island, 19 “Villarino Viejo” Place, 23
Trophic ecology, 295
Turbellarians
adhesive dual glands, 170 W
bacteria, 167 Waders, 328
Bahía Blanca Estuary, 168–171 Western Hemisphere Shorebird Reserve
biomass, 167 Network (WHSRN), 199–200, 334,
community, 167 471, 483
density, 167 Wetland communities, 437
Index 581

White-rumped sandpiper (Calidris X


fuscicollis), 328 Xenobiotics, 528
White sharks (Carcharodon carcharias), 361
Wild boar (Sus scrofa), 407
Wind waves, 43 Y
Winds, 37, 42 Young-of-year (YOY), 288
Wolves (Canis lupus), 400
“Wood” Island, 28
World Commission on Protected Areas Z
(WCPA), 485 Zooplanktivores, 296

You might also like