02.02.2014 Views

View PDF - Australasian Plant Pathology Society

View PDF - Australasian Plant Pathology Society

View PDF - Australasian Plant Pathology Society

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

APPS 2009<br />

<strong>Plant</strong> Health Management:<br />

An Integrated Approach<br />

29 September – 1 October 2009<br />

Newcastle City Hall<br />

ISBN 978‐0‐646‐52919‐6<br />

Contents<br />

Welcome........................................................................................... 2<br />

Conference Organising Committee................................................... 2<br />

Sponsors ........................................................................................... 3<br />

Exhibitors.......................................................................................... 4<br />

Conference information ................................................................... 5<br />

General information ......................................................................... 5<br />

Social program.................................................................................. 6<br />

The McAlpine lecture........................................................................ 8<br />

Keynote biographies......................................................................... 9<br />

Venue map...................................................................................... 10<br />

Program .......................................................................................... 12<br />

Oral abstracts.................................................................................. 19<br />

Poster abstracts ............................................................................ 123<br />

List of posters ............................................................................... 124


Welcome<br />

On behalf of the Local Organising Committee welcome to<br />

Newcastle and the 17th <strong>Australasian</strong> <strong>Plant</strong> <strong>Pathology</strong> <strong>Society</strong><br />

Conference, an event that marks the 40th (or Ruby) anniversary<br />

of the <strong>Australasian</strong> <strong>Plant</strong> <strong>Pathology</strong> <strong>Society</strong>. It provides us with a<br />

good opportunity to reflect on the achievements of our<br />

profession over four decades of unprecedented discovery about<br />

the nature and management of plant disease. It is also a time to<br />

ponder the directions of our profession amidst the challenges<br />

posed by emerging and persistent plant diseases, food security,<br />

climate change, water shortages, rising atmospheric carbon<br />

dioxide levels, bioterrorism, consumer safety and preferences,<br />

and the opportunities presented to agriculture and horticulture<br />

by biofuels, phytomedicines and leisure activities.<br />

The conference theme ‘<strong>Plant</strong> Health Management: an<br />

integrated approach’ addresses these challenges from three<br />

angles—fundamental discovery, the application of these<br />

discoveries to practical problems and the adoption of research.<br />

Local and international keynote speakers have been invited to<br />

challenge you with their perspectives on the big questions in<br />

plant pathology. Many of you will have already been challenged<br />

by, and enjoyed, the supporting program of workshops and field<br />

trips.<br />

Newcastle is a bustling, historic, post‐industrial seaside city<br />

boasting exciting cultural activities, superb beaches, and other<br />

nearby attractions including the Hunter Valley, Barrington Tops<br />

National Park and more superb coastal scenery. Please take<br />

time to enjoy the location, catch up with friends and colleagues,<br />

meet new ones, and return home invigorated, wiser and happy.<br />

Conference Organising<br />

Committee<br />

• David Guest, Convenor<br />

• Rosalie Daniel<br />

• Robert Park<br />

• Peter Magee<br />

• Nerida Donovan<br />

• Len Tesoriero<br />

• Angus Carnegie<br />

• Chris Steel<br />

• Gavin Ash<br />

Workshop Convenors<br />

Microbial ecology—concepts and techniques for disease control<br />

—Kerry Everett<br />

Tree <strong>Pathology</strong> Workshop<br />

—André Drenth and Angus Carnegie<br />

Magical Mystery Vegetable Tour<br />

—Len Tesoriero and Nerida Donovan<br />

Biology and management of organisms associated with bunch<br />

rot diseases of grapes—Chris Steel<br />

David Guest<br />

Conference Convenor, APPS 2009<br />

Conference Secretariat<br />

Conference Logistics*<br />

PO Box 6150<br />

Kingston ACT 2604<br />

02 6281 6624 [ph]<br />

02 6285 1336 [fx]<br />

0448 576 105 [mobile]<br />

conference@conlog.com.au<br />

www.apps2009.org.au<br />

*acting as agent for APPS<br />

2 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Sponsors<br />

The Local Organising Committee gratefully acknowledges the support of our sponsors:<br />

Conference sponsors<br />

Grains Research and Development<br />

Corporation<br />

HAL<br />

Gold sponsor<br />

APPS<br />

Silver sponsor<br />

Nufarm Australia and BASF<br />

Welcome Reception<br />

Cooperative Research Centre for<br />

National <strong>Plant</strong> Biosecurity<br />

International speaker and<br />

post‐conference tour sponsor<br />

Grape and Wine Research and<br />

Development Corporation<br />

Keynote speaker<br />

Forest and Wood Products Australia<br />

Limited<br />

Lunch, Day 2<br />

Agrichem<br />

Supporters<br />

<strong>Plant</strong> Health<br />

Australia<br />

Lomb Scientific<br />

AusVeg<br />

The Crawford Fund<br />

Mars<br />

The University of<br />

Sydney<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 3


Exhibitors<br />

APPS<br />

Nufarm<br />

The <strong>Australasian</strong> <strong>Plant</strong> <strong>Pathology</strong> <strong>Society</strong> is dedicated to the<br />

advancement and dissemination of knowledge of plant pathology and its<br />

practice in Australasia. Australasia is interpreted in the broadest sense to<br />

include not only Australia, New Zealand and Papua New Guinea, but also<br />

the Indian, Pacific and Asian regions. Although the <strong>Society</strong>’s activities are<br />

mainly focused on the <strong>Australasian</strong> region, many of the activities of our<br />

members are of international importance and significance.<br />

The <strong>Society</strong> was founded in 1969. Our members represent a broad range<br />

of scientific interests, including research scientists, teachers, students,<br />

extension professionals, administrators, industry and pest management<br />

personnel.<br />

FOR MORE INFORMATION:<br />

Dr Peter Williamson<br />

Business Manager<br />

APPS Inc<br />

Telephone (07) 4632 0467<br />

Facsimile (07) 46378326<br />

www.appsnet.org<br />

Nufarm Australia Limited and the link with BASF Australia Limited.<br />

Nufarm has become a successful crop protection company based in<br />

Australia but now with global activities that place it at number eight in<br />

the global ranking of agrochemical companies. The Nufarm head office is<br />

based at Laverton North in Victoria.<br />

In 2004 Nufarm entered into an agreement with BASF Australia Limited<br />

for Nufarm to market and develop BASF products within Australia. BASF<br />

has an excellent record for developing new horticultural products ,<br />

especially the discovery of new fungicides.<br />

For further information on the Nufarm/BASF range of products contact:<br />

doug.wilson@au.nufarm.com<br />

FOR MORE INFORMATION:<br />

Doug Wilson<br />

R&D Projects Co‐ordinator<br />

Nufarm Australia Limited<br />

Telephone (03) 9282 1427<br />

Facsimile (03) 9282 1022<br />

Mobile 0427 806 386<br />

e‐mail: doug.wilson@au.nufarm.com<br />

Leica Microsystems<br />

Leica Microsystems is a leading global designer and producer of<br />

innovative high‐tech precision optics systems for the analysis of<br />

microstructures.<br />

It comprises 11 manufacturing facilities in eight countries, sales and<br />

service companies in 20 countries and an international network of<br />

dealers; the company is also represented in over 100 countries and the<br />

international headquarters are based in Wetzlar, Germany.<br />

Leica Microsystems is one of the market leaders in each of the fields of<br />

microscopy, confocal laser scanning microscopy, microscope software,<br />

specimen preparation and medical equipment. The company<br />

manufactures a broad range of products for numerous applications<br />

requiring microscopic imaging, measurement and analysis. It also offers<br />

system solutions in the areas of life science, including biotechnology and<br />

medicine, as well as the science of raw materials and industrial quality<br />

assurance.<br />

Specific to this conference, we will be displaying automated compound<br />

and stereoscopic microscopes highlighting our imaging systems using<br />

Montage, multifocus 3D imaging and Web Module allowing viewing and<br />

analysis of images from a remote station via internet.<br />

FOR MORE INFORMATION:<br />

1800 625 286 [ph]<br />

www.leica‐microsystems.com<br />

Leica Microsystems Pty Ltd<br />

Unit 3, 112‐118 Talavera Road<br />

NORTH RYDE NSW 2113<br />

4 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Conference information<br />

Registration desk<br />

The registration desk is located in the Concert Hall Foyer of City<br />

Hall. Please direct any questions you may have regarding<br />

registration, attendance, accommodation or social functions to<br />

the staff at this desk. The registration desk will be open during<br />

the following hours:<br />

Monday 28 September 1730–1900 (Newcastle Art Gallery)<br />

Tuesday 29 September 0800–1900<br />

Wednesday 30 September 0800–1730<br />

Thursday 1 October 0800–1730<br />

The registration desk can be contacted during these hours on<br />

0448 576 105.<br />

Name badges<br />

Your name badge is your entry to all sessions, exhibition, lunches<br />

and morning and afternoon teas. Please wear it at all times.<br />

Catering<br />

Morning and afternoon teas and lunches will be held in the<br />

Banquet Room, which is located on the ground floor of City Hall.<br />

Lunches will be served as an informal stand‐up buffet. We have<br />

arranged for special meals to be prepared for those delegates<br />

who have pre‐registered their special requirements. These meals<br />

will be available from the designated buffet stations during meal<br />

breaks. Please see a member of the banquet staff for assistance.<br />

Program changes<br />

The conference organisers cannot be held responsible for any<br />

program changes due to external or unforeseen circumstances.<br />

Please check the program board located outside the Concert Hall<br />

for any changes to sessions.<br />

Speakers preparation area<br />

A speaker preparation area is located in the Concert Hall Foyer<br />

of City Hall and will be open during the following hours:<br />

Monday 28 September 1730–1900 (Newcastle Art Gallery)<br />

Tuesday 29 September 0800–1900<br />

Wednesday 30 September 0800–1730<br />

Thursday 1 October 0800–1600<br />

All speakers must take their presentation to the speaker<br />

preparation area a minimum of four hours prior to their<br />

presentation, or the day before if presenting at a morning<br />

session. Speakers are also requested to assemble in their session<br />

room 15 minutes before the commencement of the session, to<br />

meet with their session chair and to familiarise themselves with<br />

the room and the audiovisual equipment.<br />

Noticeboard<br />

A noticeboard will be maintained adjacent to the registration<br />

desk showing program changes, messages and other<br />

information. Please check the board regularly for updates.<br />

Mobile phones<br />

As a courtesy to speakers and other delegates, please ensure<br />

that all mobile phones are switched off during sessions.<br />

Participant list<br />

The participant list has been included in the conference satchel.<br />

Those delegates who have indicated on their registration form<br />

that they do not wish to have their name and organisation<br />

appear on the participant list have not been included.<br />

General information<br />

Useful telephone numbers<br />

TAXIS<br />

Newcastle Taxis 13 33 00<br />

HOTELS<br />

Crowne Plaza Newcastle 4907 5065<br />

Travelodge Newcastle 4926 3777<br />

Ibis Newcastle 4925 2266<br />

PUBLIC TRANSPORT<br />

Buses 13 15 00<br />

www.newcastlebuses.info/timetables.htm<br />

AIRLINES<br />

Qantas 13 13 13<br />

Virgin Blue 13 67 89<br />

Jetstar 13 15 38<br />

Brindabella Airlines 1300 66 88 24<br />

Eating out in Newcastle<br />

Newcastle has a great food scene, with eateries to suit all<br />

budgets. There are four main dining precincts to explore in the<br />

inner city:<br />

• Darby Street in Cooks Hill (5–10 min walk from City Hall). A<br />

diverse, friendly, relaxed bohemian precinct. Darby Street<br />

has a vibrant cafe culture, and a good selection of<br />

restaurants, pubs and take away outlets.<br />

• Honeysuckle and the Harbour waterfront (5–10 min walk<br />

from City Hall). Down at the waterfront you will find cafes,<br />

bars and restaurants, with wonderful views across the<br />

wharves. The foreshore promenade offers a great way to<br />

walk off dessert!<br />

• Beaumont Street in Hamilton (10 min drive from City Hall).<br />

There is a strong Mediterranean focus along Beaumont<br />

Street, with many sidewalk cafes and a thriving pub‐scene.<br />

• The Junction (25 min walk / 5 min drive from City Hall). An<br />

upmarket shopping precinct with a smattering of first‐class<br />

restaurants and cafes to relax in.<br />

Dining options closest to City Hall are:<br />

• Civic Precinct, which has a few coffee shops and sandwich<br />

bars<br />

• Honeysuckle and Derby Street, which both have a great<br />

selection of sit‐down cafes, bars and restaurants.<br />

For further information on places to eat in Newcastle please visit<br />

www.eatlocal.com.au/.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 5


Things to do in Newcastle<br />

CAROLE FRAZERS WALKS AND TALKS<br />

Discover what makes Newcastle unique and discover<br />

Newcastle’s best hidden treasures with Carole Frazer’s Walks<br />

and Talks. You may be surprised that Newcastle has many<br />

fascinating walks in and around Newcastle city. Included are the<br />

spectacular Bogey Hole, Leadlight Tower and historic houses, Art<br />

Gallery and cultural buildings. All walks include commentary on<br />

many city topics and especially of local Newcastle history. There<br />

are a number of different types of walks you can do catering for<br />

a diverse range of areas and interesting locations. Prices start at<br />

$10 per person for a 1 hour tour. For more information, log onto<br />

www.walks‐talks.com.au or phone Carole on 02 4952 1537<br />

BLACKBUTT RESERVE<br />

Blackbutt Reserve provides nature trails, wildlife exhibits,<br />

children’s playgrounds and recreational facilities. It is the perfect<br />

place for a relaxing family picnic or to explore the wonders of<br />

nature. Wildlife Exhibits open 9.00 am to 5.00 pm every day of<br />

the year. Picnic and recreation facilities open from 7.00 am to<br />

5.00 pm and entry is free. For more information log onto<br />

www.ncc.nsw.gov.au/discover_newcastle/blackbutt_reserve<br />

NEWCASTLES FAMOUS TRAM<br />

Everything about Newcastle’s Famous Tram is unique. Built from<br />

scratch in 1994, the tram is a genuine replica of the original<br />

Newcastle working tram, which was in service in 1923.<br />

Newcastle’s Famous Tram is a very novel and nostalgic way to<br />

visit the historical city of Newcastle. The Newcastle tour is a 45<br />

minute tour of our city, beaches and historical sites. A full<br />

commentary is provided. This service in the heart of Newcastle<br />

reveals to its passengers the beauty of the city and beach areas<br />

as well as an astonishing blend of history and current changes to<br />

the city lifestyle. Detailed information is provided about many<br />

historic sites. The tour is great value at $12 an adult and $6 and<br />

operates weekday tours from Newcastle’s Railway Station and<br />

the Crown Plaza in Wharf Road. The Tram operates at 11.00 am<br />

and 1.00 pm, with a special pick up at the Brewery Wharf Road<br />

at 12.55 pm, and during school holidays between 10.00 am<br />

11.00 am 12 noon and 1.00 pm, but please ring to confirm<br />

operating times. No service on weekends or public holidays. For<br />

more information log onto www.famous‐tram.com.au<br />

DARBY ST PRECINCT<br />

Conveniently situated and only 5 minutes from Newcastle<br />

Harbour and Foreshore, Darby St Precinct offers a diverse,<br />

friendly, and relaxed cosmopolitan destination. Consisting of<br />

over 20 cafes, outdoor dining and cosy retreats, including some<br />

award winning restaurants that boast the fine cuisine with<br />

friendly prices. Shoppers look out for unique fashion boutiques,<br />

art and gift galleries. You also have photography studios, homewares,<br />

everyday living, music and professional services. For more<br />

information log onto www.darbystreet.com.au<br />

Social program<br />

Welcome Reception<br />

Monday 28 September 2009<br />

5.30 pm – 7.00 pm<br />

Venue: Level 1, Newcastle Region Art Gallery, 1 Laman Street,<br />

Newcastle (opposite City Hall)<br />

Dress: Conference attire/neat casual<br />

Marking the opening of the conference, drinks and canapés will<br />

be served in the Newcastle Region Art Gallery. The welcome<br />

reception will give you the opportunity to register early and<br />

catch up with friends.<br />

Poster, Wine and Cheese Night<br />

Tuesday 29 September 2009<br />

6.00 pm – 7.00 pm<br />

Venue: Banquet Room and Concert Hall, Newcastle City Hall<br />

Dress: Conference attire/neat casual<br />

Cost: Included in full conference registration. $25 for extra<br />

attendees or other registration categories. If you wish to attend,<br />

please check with the registration desk staff if there are still<br />

tickets avaible.<br />

Conference Dinner<br />

Wednesday 30 September 2009<br />

7.00 pm (for 7.30 pm start) until late<br />

Venue: Auditorium 1, Newcastle Panthers Club, corner King and<br />

Union Streets, Newcastle (5 minutes walk from City Hall)<br />

Theme: Ruby—celebrating the 40th Anniversary of the APPS<br />

Dress: Smart casual (wear something ruby)<br />

Cost: Included in full conference registration. $110 for extra<br />

attendees or other registration categories. If you have not<br />

indicated on your registration form that you would like to<br />

attend, please see the registration desk staff to find our if there<br />

are still places or tickets available for purchase.<br />

It is not every day we turn 40. Come and help us celebrate our<br />

Ruby Anniversary at the Newcastle Panthers Club. You are<br />

assured of a night of great food, great wines, fun dancing and<br />

excellent company. Let’s paint Newcastle Ruby.<br />

Beach Party<br />

Thursday 1 October 2009<br />

6.30 pm – 10.30 pm<br />

Venue: Newcastle Surf Life Saving Club<br />

Dress: Casual<br />

Cost: The cost of the Beach Party is not included in registration<br />

fees. Cost to all delegates and guests is $55.. If you would like to<br />

attend, please check with the registration desk staff if there are<br />

still tickets avaible for purchase.<br />

Coach transfer: Departs from the front of City Hall, King Street at<br />

5.30 pm sharp and will return at 10.30 pm. Please be waiting at<br />

the front of City Hall at least 5 minutes before the scheduled<br />

departure time.<br />

6 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 7


The McAlpine lecture<br />

The invitation to present the McAlpine lecture to the biennial<br />

conference of the <strong>Australasian</strong> <strong>Plant</strong> <strong>Pathology</strong> <strong>Society</strong> is<br />

extended to an eminent scientist in recognition of their<br />

significant contribution to <strong>Australasian</strong> plant pathology. The<br />

lecture is named after Daniel McAlpine, considered to be the<br />

father of plant pathology in the <strong>Australasian</strong> region. His most<br />

notable contributions were to study wheat rust following the<br />

1889 epidemic, to classify and describe Australian smuts, and to<br />

recognise Ophiobolus graminis (now Gaeumannomyces<br />

graminis) as the cause of wheat take‐all. He also collaborated<br />

with Farrer on resistance to rust in wheat (John Randles 1994,<br />

Stanislaus Fish 1976). Daniel McAlpine also contributed<br />

extensively to vegetable pathology. It is therefore fitting that a<br />

plant pathologist with extensive experience and passion such as<br />

Phil Keane be asked to deliver the McAlpine lecture in 2009.<br />

1976 Dr Lilian Fraser, Department of Agriculture, NSW<br />

Disease of citrus trees in Australia—the first hundred years<br />

1978 Dr David Griffin, Australian National University, ACT<br />

Looking ahead<br />

1980 Mr John Walker, Department of Agriculture, NSW<br />

Taxonomy, specimens and plant disease<br />

1982 Professor Richard Matthews, The University of Auckland, NZ<br />

Relationships between plant pathology and molecular biology<br />

1984 Professor Bob McIntosh, University of Sydney, NSW, and<br />

Dr Colin Wellings, Department of Agriculture, NSW<br />

Wheat rust resistance: the continuing challenge<br />

1986 Dr Allen Kerr, Waite Agricultural Research Institute, SA<br />

Agrobacterium: pathogen, genetic engineer and biological<br />

control agent<br />

1989 Dr Albert Rovira, CSIRO Division of Soils, SA<br />

Ecology, epidemiology and control of take‐all, rhizotomies bare<br />

patch and cereal cyst nematode in wheat<br />

1991 Mr John Walker, Department of Agriculture, NSW<br />

<strong>Plant</strong>s, diseases and pathologists in Australasia—a personal<br />

view<br />

1993 Dr John Randles (University of Adelaide, SA<br />

<strong>Plant</strong> viruses, viroids and virologists of Australasia<br />

1995 Dr Ron Close, Lincoln University, NZ<br />

The ever changing challenges of plant pathology<br />

1997 Professor John Irwin, CRC Tropical <strong>Plant</strong> <strong>Pathology</strong>, Qld<br />

Biology and management of Phytophthora spp. attacking field<br />

crops in Australia<br />

1999 Dr Dorothy Shaw, Department of Primary Industries, Qld<br />

Bees and fungi with special reference to certain plant pathogens<br />

2001 Dr Alan Dube, South Australian Research and Development<br />

Institute, SA<br />

Long‐term careers in plant pathology<br />

2003 Dr Mike Wingfield, University of Pretoria, South Africa<br />

Increasing threat of disease to exotic plantation forests in the<br />

southern hemisphere<br />

2007 Dr Graham Stirling, Biological Crop Protection, Qld<br />

The impact of farming systems on soil biology and soil‐borne<br />

diseases: examples from the Australian sugar and vegetable<br />

industries, the case for better integration of sugarcane and<br />

vegetable production and implications for future research<br />

2009 Assoc Prof Phil Keane, La Trobe University, Vic<br />

Lessons from the tropics—the unfolding mystery of vascularstreak<br />

dieback of cocoa, the importance of genetic diversity,<br />

horizontal resistance, and the plight of farmers<br />

McAlpine lecturer 2009: Philip Keane<br />

Philip grew up in the wheat/sheep belt<br />

of rural South Australia and gained his<br />

Bachelor of Agricultural Science (Hons)<br />

at the Waite Agricultural Research<br />

Institute, University of Adelaide in<br />

1968.<br />

He was awarded a PhD at the<br />

University of Papua New Guinea in<br />

1972 for his studies of vascular‐streak<br />

dieback, a serious epidemic disease of<br />

cocoa. He described and named the<br />

pathogen, Oncobasidium theobromae,<br />

and remains the world authority on what is a particularly<br />

unusual vascular wilt disease. Not only was the pathogen a new<br />

species, but also a new genus within the Basidiomycetes.<br />

Philip taught at UPNG before taking up a lectureship at La Trobe<br />

University in 1975. His time at La Trobe has been supplemented<br />

with sabbatical periods in the USA and Central America, as well<br />

as extensive project‐related travel through PNG and Indonesia.<br />

Since returning to Australia in 1975 Philip maintained his interest<br />

in diseases of cocoa in South East Asia and Papua New Guinea,<br />

and in agricultural development and education in tropical<br />

countries. His approach is focussed on the farmer—from<br />

listening to farmers, evaluating their ideas, then translating his<br />

research to be used by the farmers. Philip also initiated research<br />

into fungal diseases of crop plants and eucalypts, and co‐edited<br />

the standard monograph on Eucalypt Pathogens and Diseases.<br />

He is involved in research on a range of big questions in plant<br />

pathology, including the nature of resistance to crop diseases,<br />

especially cereal rusts, plant disease epidemiology, the diversity<br />

of macrofungi and broad questions in plant ecology.<br />

Philip is an enthusiastic undergraduate teacher and has trained<br />

many local and international PhD students, many of whom will<br />

be attending this conference. He has made a special and unique<br />

contribution to plant pathology in Australia and neighbouring<br />

countries, and it is a great honour that he has accepted our<br />

invitation to present the McAlpine Lecture.<br />

2005 Dr Gretna Weste, University of Melbourne, Vic<br />

A long and varied fungal foray<br />

8 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Keynote biographies<br />

Barbara Christ<br />

Professor Barbara Christ, the current President of the American<br />

Phytopathological <strong>Society</strong>, is Senior Associate Dean in the College of<br />

Agricultural Sciences and Professor of <strong>Plant</strong> <strong>Pathology</strong> at<br />

Pennsylvania State University in the United States. Her research is<br />

focused on potato breeding and disease management, including<br />

basic research into understanding the inheritance of disease<br />

resistance as well as extension. Her research includes developing<br />

and releasing new varieties adapted for Pennsylvania growing<br />

conditions, developing disease‐resistant potato germplasm.<br />

examining the genetic variability and biology of potato pathogen<br />

populations, developing methods to detect and forecast potato<br />

diseases, developing integrated pest management strategies for<br />

potatoes in Pennsylvania, and evaluating new fungicides for efficacy<br />

against potato diseases.<br />

André Drenth<br />

Dr André Drenth is a Principal <strong>Plant</strong> Pathologist from the University<br />

of Queensland, and founder and Leader of the Tree <strong>Pathology</strong><br />

Centre which is a joint initiative between the University of<br />

Queensland and Queensland Primary Industries and Fisheries. André<br />

studied <strong>Plant</strong> Breeding and <strong>Pathology</strong> at Wageningen University and<br />

Cornell University, USA. André was Research Program Leader in the<br />

CRC for Tropical <strong>Plant</strong> Protection dealing with a large number of<br />

Tropical diseases. His ability to deliver practical outcomes from basic<br />

research in plant pathology is well recognised internationally. André<br />

has been involved in research on plant pathogens for nearly 20 years<br />

and has published widely on a range of plant diseases with a special<br />

focus on Phytophthora.<br />

Adrienne Hardham<br />

Professor Adrienne Hardham works in the Research School of<br />

Biology at the Australian National University. The main focus of her<br />

research is on cellular and molecular mechanisms responsible for<br />

the infection of plants by Phytophthora and rust fungi and the<br />

plant’s defence response to pathogen invasion.<br />

Greg Johnson<br />

Dr Greg Johnson is President of the <strong>Australasian</strong> <strong>Plant</strong> <strong>Pathology</strong><br />

<strong>Society</strong> (APPS) 2007–2009 and Secretary General of the International<br />

<strong>Society</strong> for <strong>Plant</strong> <strong>Pathology</strong> (ISPP) 2006–2013. Greg has had over 20<br />

years’ experience in development assistance in tropical horticulture<br />

and postharvest R&D collaboration with developing countries in Asia<br />

and the Pacific, and over 30 years’ experience in plant pathology<br />

practice, diagnostic advice and publishing on tropical and temperate<br />

plants and crops. His especial interest is postharvest diseases of<br />

mangoes. Greg currently operates a Canberra‐based consultancy,<br />

Horticulture 4 Development, that builds upon Greg’s background in<br />

managing a portfolio of projects and activities in postharvest<br />

technology, horticulture and crop protection with the Australian<br />

Centre for International Agricultural Research (ACIAR) in Asia and<br />

the Pacific. His recent activities have included an overview of the<br />

vegetable sector in tropical Asia and reviewing issues and priorities<br />

for postharvest disease management in mangoes.<br />

Eun Woo Park<br />

Professor Eun Woo Park is Dean of the College of Agriculture and<br />

Life Sciences in Seoul National University, Korea. Major research<br />

areas are epidemiology of airborne diseases with special emphasis<br />

on modeling and forecasting disease development, and applications<br />

of various information technologies to implement disease<br />

management strategies.<br />

Dov Prusky<br />

Professor Dov Prusky is Deputy Director Research and Development<br />

with the Agricultural Research Organization, Israel. He is also active<br />

in research in the Department of Postharvest Science of Fresh<br />

Produce of the ARO Technology and Storage of Agricultural Products<br />

Institute. Dov is currently Chair of the ISPP Postharvest Diseases<br />

Subject Matter committee. Dov’s research Interests include:<br />

• understanding the basic processes underlying the interactions<br />

between fruits and pathogenic fungi<br />

• studying biochemical and molecular mechanisms that are<br />

controlled by fungal virulence and fruit resistance factors<br />

• using transformation‐mediated gene disruption to create<br />

strains of pathogenic fungi that are specifically mutated in their<br />

ability to make cell‐wall degrading enzymes and other<br />

pathogenicity factors. These mutants are tested for their ability<br />

to cause disease and to elicit defense responses<br />

• studying the biochemical basis for modulation of pathogenicity<br />

factor affecting the transcription expression of nitrogen<br />

metabolism, ammonia secretion and the effect on the<br />

modulation of local pH<br />

• reduction of postharvest losses in deciduous and subtropical<br />

fruits.<br />

Robert Seem<br />

Robert C Seem has spent his 34‐year career as professor of plant<br />

pathology at Cornell University’s Agricultural Experiment Station in<br />

Geneva, New York. He specialises in the epidemiology of fruit and<br />

vegetable diseases. Robert also served in the station administration<br />

for 14 years. During this time he was instrumental in the<br />

development of the Cornell Agriculture and Food Technology Park<br />

Corporation, where he continues to serve a president of the board.<br />

Mike Wingfield<br />

Professor Michael Wingfield was born in South Africa. He graduated<br />

with BSc Hons (Natal) and MSc (Stellenbosch) degrees then<br />

completed his PhD in <strong>Plant</strong> <strong>Pathology</strong> (University of Minnesota),<br />

specialising in forest pathology and forest entomology. He returned<br />

to South Africa to establish the Tree Protection Co‐operative<br />

Programme (TPCP) at the University of the Free State, and in 1998<br />

established the Forestry and Agricultural Biotechnology Institute<br />

(FABI) at the University of Pretoria. FABI is now the Centre of<br />

Excellence in Tree Health Biotechnology. He is also an alumnus of<br />

the Harvard Business School Advanced Management Programme.<br />

Celeste Linde<br />

Celeste Linde investigates the population genetics of, for example,<br />

cereal pathogens, the influence of wild or weedy hosts on pathogen<br />

populations and their evolution of virulence. Her main focus has<br />

been with Rhynchosporium secalis, causing barley scald.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 9


Venue map<br />

Level 1<br />

Level 2<br />

10 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Level 3<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 11


Program<br />

Monday 28 September<br />

1400–<br />

1700<br />

1730–<br />

1800<br />

Registration open<br />

WELCOME RECEPTION<br />

Ground Floor Foyer, Newcastle Region Art Gallery<br />

Newcastle Region Art Gallery<br />

Tuesday 29 September<br />

0800–<br />

1900<br />

Registration open<br />

Concert Hall Foyer<br />

0800 ARRIVAL TEA AND COFFEE Concert Hall Foyer<br />

0830 Conference opening Concert Hall<br />

Prof David Guest, University of Sydney<br />

0845 Presidential address—‘Shield the young harvest from devouring blight’—Charles Darwin, Joseph Banks, Concert Hall<br />

Thomas Knight and wheat rust: discovery, adventure, and ‘getting the message out’<br />

Dr Greg Johnson, Horticulture 4 Development, ACT<br />

0930 Keynote address—The relevance of plant pathology in food production Concert Hall<br />

Dr André Drenth, Tree <strong>Pathology</strong> Centre, The University of Queensland and Primary Industries and Fisheries<br />

1030 MORNING TEA Banquet Room<br />

Session 1A<br />

Disease management<br />

Room: Concert Hall<br />

Chair: Andrew Miles<br />

1100 An integrated approach to<br />

husk spot management in<br />

macadamia<br />

Dr Olufemi Akinsanmi, The<br />

University of Queensland and<br />

Primary Industries and<br />

Fisheries, Qld<br />

1120 Application methods of<br />

phosphonate to control<br />

Phytophthora pod rot and<br />

stem canker on cocoa<br />

Dr Peter McMahon, La Trobe<br />

University, Vic<br />

1140 Botrytis bunch rot control<br />

strategies in cool climate<br />

viticultural regions of Australia<br />

and New Zealand<br />

Dr Jacqueline Edwards,<br />

Department of Primary<br />

Industries, Vic<br />

Session 1B<br />

Disease surveys<br />

Room: Cummings Room<br />

Chair: Eileen Scott<br />

Why Australia needs a<br />

coordinated national<br />

diagnostic system<br />

Ms Jane Moran, Department<br />

of Primary Industries, Vic<br />

Development of a soil DNA<br />

extraction and quantitative<br />

PCR method for detecting two<br />

Cylindrocarpon species in soil<br />

Ms Chantal Probst, Lincoln<br />

University, NZ<br />

Bananas in Carnarvon—good<br />

news for growers in survey for<br />

quarantine plant pests and<br />

pathogens<br />

Dr Sarah Collins, Department<br />

of Agriculture and Food WA<br />

Session 1C<br />

Soilborne diseases<br />

Room: Hunter Room<br />

Chair: Nerida Donovan<br />

Can investment in building up<br />

soil organic carbon lead to<br />

disease suppression in<br />

vegetable crops?<br />

Dr Ian Porter, Department of<br />

Primary Industries, Vic<br />

Evaluation of soil health<br />

indicators in the vegetable<br />

industry of temperate<br />

Australia<br />

Ms Robyn Brett, Department<br />

of Primary Industries, Vic<br />

Rhizoctonia AG2.1 and AG3 in<br />

soil—competition or<br />

synergism?<br />

Dr Tonya Wiechel,<br />

Department of Primary<br />

Industries, Vic<br />

Session 1D<br />

Virology<br />

Room: Newcastle Room<br />

Chair: John Randles<br />

Towards universal detection of<br />

Luteoviridae<br />

Miss Anastasija Chomic,<br />

Lincoln University, NZ<br />

Massive parallel sequencing of<br />

small RNAs to identify plant<br />

viruses and virus‐induced small<br />

RNAs<br />

Dr Robin MacDiarmid, The<br />

New Zealand Institute for <strong>Plant</strong><br />

and Food Research Ltd, NZ<br />

Chickpea chlorotic stunt virus,<br />

an important virus of coolseason<br />

food legumes in Asia<br />

and North Africa and<br />

potentially in Australia<br />

Dr Safaa Kumari, International<br />

Center for Agricultural<br />

Research in the Dry Areas,<br />

Syria<br />

1200 LUNCH Banquet Room<br />

Editor’s meeting (1200–1400)<br />

Waratah Room<br />

Student mentor lunch (1200–1320)<br />

Mulumbinba Room<br />

1320 Keynote address—Emerging frontiers in forest pathology Concert Hall<br />

Prof Mike Wingfield, Forestry and Agricultural Biotechnology Institute, University of Pretoria, South Africa<br />

12 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Session 2A<br />

Forest pathology/native<br />

Room: Concert Hall<br />

Chair: Angus Carnegie<br />

1410 Variability in pathogenicity of<br />

Quambalaria pitereka on<br />

spotted gums<br />

Mr Geoffrey Pegg, The<br />

University of Queensland/<br />

Primary Industries and<br />

Fisheries, Qld<br />

1430 Movement of pathogens<br />

between horticultural crops<br />

and endemic trees in the<br />

Kimberleys<br />

Ms Monique Sakalidis,<br />

Murdoch University, WA<br />

1450 Pathogenicity of Phytophthora<br />

multivora to Eucalyptus<br />

gomphocephala and<br />

E. marginata<br />

Dr Treena Burgess, Murdoch<br />

University, WA<br />

1510 Microscopy of progressive<br />

decay of fungi isolated from<br />

Meranti tree canker<br />

Dr Erwin Erwin, University of<br />

Mulawarman, Indonesia<br />

Session 2B<br />

Soilborne disease<br />

Room: Cummings Room<br />

Chair: Peter McGee<br />

Optimising conditions to<br />

investigate gene expression in<br />

pathogenic Streptomyces using<br />

RT‐qPCR<br />

Dr Tonya Wiechel,<br />

Department of Primary<br />

Industries, Vic<br />

Fusarium oxysporum and<br />

Pythium associated with<br />

vascular wilt and root rots of<br />

greenhouse cucumbers<br />

Mr Len Tesoriero, NSW<br />

Department of Primary<br />

Industries<br />

Fusarium oxysporum f. sp.<br />

fragariae: a main component<br />

of strawberry crown and root<br />

rots in Western Australia<br />

Dr Hossein Golzar,<br />

Department of Agriculture and<br />

Food WA<br />

Evaluation of resistant<br />

rootstocks for control of<br />

Fusarium wilt of watermelon in<br />

Nghe An Province, Vietnam.<br />

Prof Lester Burgess, The<br />

University of Sydney, NSW<br />

Session 2C<br />

Epidemiology<br />

Room: Hunter Room<br />

Chair: Chris Steel<br />

Bunch rot diseases and their<br />

management<br />

Prof Turner Sutton, NC State<br />

University, USA<br />

Inoculum and climatic factors<br />

driving epidemics of Botrytis<br />

cinerea in New Zealand and<br />

Australian vineyards<br />

Dr Rob Beresford, The New<br />

Zealand Institute for <strong>Plant</strong> and<br />

Food Research Limited, NZ<br />

Infection of apples by<br />

Colletotrichum acutatum in<br />

New Zealand is limited by<br />

temperature<br />

Dr Kerry Everett, The New<br />

Zealand Institute for <strong>Plant</strong> and<br />

Food Research Limited, NZ<br />

Epidemiology of walnut blight,<br />

caused by Xanthomonas<br />

arboricola pv. juglandis, in<br />

Tasmania, Australia<br />

Dr Katherine Evans, University<br />

of Tasmania<br />

Session 2D<br />

Disease management<br />

Room: Newcastle Room<br />

Chair: Robert Magarey<br />

Sugarcane smut—disease<br />

development and mechanism<br />

of resistance<br />

Dr Shamsul Bhuiyan, BSES<br />

Limited, Qld<br />

Dissemination of biological<br />

and chemical fungicides by<br />

bees onto Rubus and Ribes<br />

flowers<br />

Dr Monika Walter, The New<br />

Zealand Institute for <strong>Plant</strong> and<br />

Food Research Limited, NZ<br />

Current studies on divergence<br />

and management of pepper<br />

yellow leaf curl disease<br />

Indonesia<br />

Dr Sri Hidayat, Bogor<br />

Agricultural University,<br />

Indonesia<br />

Fungicide resistance in cucurbit<br />

powdery mildew<br />

Dr Chrys Akem, Primary<br />

Industries and Fisheries, Qld<br />

1530 AFTERNOON TEA Banquet Room<br />

1600 Keynote address—Population genetic analyses of plant pathogens: new challenges and opportunities Concert Hall<br />

Dr Celeste Linde, Research School of Biology, College of Medicine, Biology and Environment, Australian National University<br />

Session 3A<br />

Population genetics<br />

Room: Concert Hall<br />

Chair: Andre Drenth<br />

1640 Genetic diversity of Botryosphaeria parva<br />

(Neofusicoccum parvum) in New Zealand<br />

vineyards<br />

Mr Jeyaseelan Baskarathevan, Lincoln<br />

University, NZ<br />

1700 Anthracnose disease of chili pepper—<br />

genetic diversity, pathogenicity and<br />

breeding for resistance<br />

A/Prof Paul Taylor, The University of<br />

Melbourne, Vic<br />

1720 The diversity of Colletotrichum infecting<br />

lychee in Australia<br />

Ms Jay Anderson, Primary Industries and<br />

Fisheries, Qld and University of<br />

Queensland<br />

1740 Variation in Phytophthora palmivora on<br />

cocoa in Papua New Guinea<br />

Ms Josephine Saul Maora, PNG Cocoa<br />

Coconut Institute<br />

Session 3B<br />

Modelling and crop loss assessment<br />

Room: Cummings Room<br />

Chair: Ian Porter<br />

Spore traps for early warning of smut<br />

infestations in Australian sugarcane crops<br />

Dr Rob Magarey, BSES Limited, Qld<br />

Software‐assisted gap estimation (SAGE)<br />

for measuring grapevine leaf canopy<br />

density<br />

Mr Gareth Hill, The New Zealand Institute<br />

for <strong>Plant</strong> and Food Research Limited, NZ<br />

Evaluation of the efficacy of Brassica spot<br />

TM<br />

models for control of white blister in<br />

Chinese cabbage<br />

Mr Desmond Auer, Department of<br />

Primary Industries, Vic<br />

Evaluating an infection model of prune<br />

rust to improve the management of<br />

disease for almond and prune growers<br />

Mr Peter Magarey, South Australian<br />

Research and Development Institute<br />

Session 3C<br />

Disease management<br />

Room: Hunter Room<br />

Chair: Shane Hetherington<br />

Management of white blister on vegetable<br />

brassicas with irrigation and varieties<br />

Dr Elizabeth Minchinton, Department of<br />

Primary Industries, Vic<br />

Alternative screening methods for<br />

sugarcane smut using natural infection<br />

and tissue staining<br />

Dr Shamsul Bhuiyan, BSES Limited, Qld<br />

Interruption of cool chain and strawberry<br />

fruit rot by leak‐causing fungi Rhizopus<br />

species<br />

Dr Monika Walter, The New Zealand<br />

Institute for <strong>Plant</strong> and Food Research<br />

Limited, NZ<br />

Enhancing Papua New Guinea smallholder<br />

cocoa production through greater<br />

adoption of integrated pest and disease<br />

management<br />

Mr Yak Namaliu, PNG Cocoa Coconut<br />

Institute<br />

1800 DRINKS AND POSTERS Banquet Room and Concert Hall<br />

1830–<br />

2030<br />

Council of <strong>Society</strong> meeting<br />

Waratah Room<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 13


Wednesday 30 September<br />

0800–<br />

1730<br />

Registration open<br />

Concert Hall Foyer<br />

0800 ARRIVAL TEA AND COFFEE Concert Hall Foyer<br />

0830 Keynote address—Molecular cytology of Phytophthora‐plant interactions Concert Hall<br />

Prof Adrienne Hardham, <strong>Plant</strong> Cell Biology Group, Research School of Biology, The Australian National University<br />

Session 4A<br />

<strong>Plant</strong> pathogen interactions<br />

Room: Concert Hall<br />

Chair: David Guest<br />

0910 Gene expression changes<br />

during host‐pathogen<br />

interaction between<br />

Arabidopsis thaliana and<br />

Plasmodiophora brassicae<br />

Mrs Arati Agarwal,<br />

Department of Primary<br />

Industries, Vic<br />

0930 Hairpin RNA derived from viral<br />

NIa gene confers immunity to<br />

wheat streak mosaic virus<br />

infection in transgenic wheat<br />

plants<br />

Mr Muhammad Fahim, CSIRO<br />

<strong>Plant</strong> Industry, and Australian<br />

National University, ACT<br />

0950 Characterising inositol<br />

signalling pathways in<br />

Phytophthora spp. for future<br />

development of selective<br />

antibiotics<br />

Mr Dean Phillips, Deakin<br />

University, Vic<br />

1010 Systemic acquired resistance—<br />

a new addition to the IPM<br />

clubroot toolbox?<br />

Dr Caroline Donald,<br />

Department of Primary<br />

Industries, Vic<br />

Session 4B<br />

Disease surveys<br />

Room: Cummings Room<br />

Chair: Sandra Savocchia<br />

Prevalence and pathogenicity<br />

of Botryosphaeria lutea<br />

isolated from grapevine<br />

nursery materials in New<br />

Zealand<br />

Ms Regina Billones, Lincoln<br />

University, NZ<br />

Infection and disease<br />

progression of Neofusicoccum<br />

luteum in grapevine plants<br />

Mr Nicholas Amponsah,<br />

Lincoln University, NZ<br />

Carbohydrate stress increases<br />

susceptibility of grapevines to<br />

Cylindrocarpon black foot<br />

disease<br />

Miss Dalin Dore, Lincoln<br />

University, NZ<br />

Botryosphaeria spp. associated<br />

with bunch rot of grapevines in<br />

south‐eastern Australia<br />

Ms Nicola Wunderlich, Charles<br />

Sturt University, NSW<br />

Session 4C<br />

Epidemiology<br />

Room: Hunter Room<br />

Chair: Greg Johnson<br />

Honey bees— do they aid the<br />

dispersal of Alternaria radicina<br />

in carrot seed crops?<br />

Mr Rajan Trivedi, Lincoln<br />

University, NZ<br />

Translating research into the<br />

field: meta‐analysis of field pea<br />

blackspot severity and yield<br />

loss to extend model<br />

application for disease<br />

management in Western<br />

Australia<br />

Dr Moin Salam, Department of<br />

Agriculture and Food WA<br />

Development of a model to<br />

predict spread of exotic wind<br />

and rain borne fungal pests<br />

Dr Moin Salam, Department of<br />

Agriculture and Food WA<br />

Psyllid transmission of<br />

Huanglongbing from naturally<br />

infected Shogun mandarin to<br />

orange jasmine<br />

Dr Rantana Sdoodee, Prince of<br />

Songkla University, Thailand<br />

Session 4D<br />

Prokaryotic pathogens<br />

Room: Newcastle Room<br />

Chair: Lucy Tran‐Nguyen<br />

Transmission of 'Candidatus<br />

Phytoplasma australiense' to<br />

Cordyline and Coprosma<br />

Dr Ross Beever, Landcare<br />

Research, NZ<br />

Australian grapevine yellows<br />

phytoplasma found in<br />

symptomless shoot tips after a<br />

heat wave in South Australia<br />

Mr Peter Magarey, South<br />

Australian Research and<br />

Development Institute, SA<br />

Association of Phytoplasmas<br />

with papaya crown yellows<br />

(PCY) disease—a new disease<br />

of papaya in Northern<br />

Mindanao, Philippines<br />

Ms Regina Billones, Del Monte<br />

Phils Inc, Philippines<br />

Phytoplasma diseases in citrus<br />

orchards of Pakistan<br />

Dr Shazia Mannan, COMSATS<br />

Institute of Information<br />

Technology, Pakistan<br />

1030 MORNING TEA Banquet Room<br />

1100 Keynote address—Mechanisms modulating fungal attack in postharvest pathogen interactions and Concert Hall<br />

their modulation for improved disease control<br />

Prof Dov Prusky, Department of Postharvest Science of Fresh Produce, Agricultural Research Organization, Israel<br />

Session 5A<br />

<strong>Plant</strong> pathogen interactions<br />

Room: Concert Hall<br />

Chair: Rosalie Daniel<br />

1140 ABA‐dependant signalling of PR genes and<br />

potential involvement in the defence of<br />

lentil to Ascochyta lentis<br />

Dr Rebecca Ford, The University of<br />

Melbourne, Vic<br />

1200 Fundamental components of resistance to<br />

Phytophthora cinnamomi: using model<br />

system approaches<br />

Prof David Cahill, Deakin University, Vic<br />

1220 Genes involved in hypersensitive cell death<br />

responses during Fusarium crown rot<br />

infection in wheat<br />

Dr Jill Petrisko, University of Southern<br />

Queensland, Qld<br />

Session 5B<br />

Disease surveys<br />

Room: Cummings Room<br />

Chair: Aaron Maxwell<br />

Fishing For Phytophthora across Western<br />

Australia’s water bodies<br />

Dr Daniel Hüberli, Murdoch University,<br />

WA<br />

Incidence of fungi isolated from grape<br />

trunks in New Zealand vineyards<br />

Mr Dion Mundy, The New Zealand<br />

Institute for <strong>Plant</strong> and Food Research<br />

Limited, NZ<br />

Isolation and characterisation of strains of<br />

Pseudomonas syringae from waterways of<br />

the Central North Island of New Zealand<br />

Dr Joel Vanneste, The New Zealand<br />

Institute for <strong>Plant</strong> and Food Research<br />

Limited, NZ<br />

Session 5C<br />

Chemical control<br />

Room: Hunter Room<br />

Chair: Len Tesoriero<br />

Evaluation of fungicides to manage<br />

brassica stem canker<br />

Ms Lynette Deland, South Australian<br />

Research and Development Institute, SA<br />

Evaluation of spray programs for powdery<br />

mildew management in greenhouse<br />

cucumbers<br />

Dr Kaye Ferguson, South Australian<br />

Research and Development Institute, SA<br />

The incidence of copper resistant bacteria<br />

in Australian pome and stone fruit<br />

orchards<br />

Dr Chin Gouk, Department of Primary<br />

Industries, Vic<br />

14 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


1240 AGRICHEM LUNCH Banquet Room<br />

1340 Poster session Banquet Room and Concert Hall<br />

1430 AFTERNOON TEA Banquet Room<br />

1500 McAlpine lecture—Lessons from the tropics—the unfolding mystery of vascular‐streak dieback Concert Hall<br />

of cocoa, the importance of genetic diversity, horizontal resistance, and the plight of farmers<br />

Assoc Prof Phil Keane, Department of Botany, La Trobe University, Vic<br />

1600 AGM Concert Hall<br />

1730 Close of day<br />

1900 CONFERENCE DINNER Newcastle Panthers Club<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 15


Thursday 1 October<br />

0700 Regional Councillor’s meeting Waratah Room<br />

0700 CHAIRMAN’S BREAKFAST Mulumbinba Room<br />

0800–<br />

1730<br />

Registration open<br />

Concert Hall Foyer<br />

0800 ARRIVAL TEA AND COFFEE Concert Hall Foyer<br />

0830 Keynote address—Translating research into the field: how it started, how it is practised and Concert Hall<br />

how we carry out grape powdery mildew research<br />

Dr Bob Seem, Cornell University, USA<br />

0910 GRDC book launch: Mr James Clarke, Grains Research and Development Corporation Concert Hall<br />

Session 6A<br />

Cereal pathology 1<br />

Room: Concert Hall<br />

Chair: Mark Sutherland<br />

0925 Stem rust race Ug99: international<br />

perspectives and implications for Australia<br />

Dr Colin Wellings, The University of<br />

Sydney, NSW<br />

0945 Mitigating crop losses due to stripe rust in<br />

Australia: integrating pathogen<br />

population dynamics with research and<br />

extension programs<br />

Dr Colin Wellings, The University of<br />

Sydney, NSW<br />

1005 Impact of sowing date on crown rot losses<br />

Dr Steven Simpfendorfer, Department of<br />

Primary Industries, NSW<br />

1025 Symptom development and pathogen<br />

spread in wheat genotypes with varying<br />

levels of crown rot resistance<br />

Dr Cassandra Malligan, Queensland<br />

Primary Industries and Fisheries<br />

Session 6B<br />

Quarantine and exotic pathogens<br />

Room: Cummings Room<br />

Chair: Suzy Perry<br />

Development of an eradication strategy<br />

For exotic grapevine pathogens<br />

Dr Mark Sosnowski, South Australian<br />

Research and Development Institute, SA<br />

Green grassy shoot disease of sugarcane,<br />

a major disease in Nghe An Province,<br />

Vietnam<br />

Dr Rob Magarey, BSES Limited, Qld<br />

Molecular detection of Mycosphaerella<br />

fijiensis in the leaf trash of ‘Cavendish’<br />

banana<br />

Dr Seona Casonato, The New Zealand<br />

Institute for <strong>Plant</strong> and Food Research<br />

Limited, NZ<br />

Optimising responses to incursions of<br />

exotic plant pathogens<br />

Dr Mike Hodda, CSIRO Entomology, ACT<br />

Session 6C<br />

Alternatives to chemical control<br />

Room: Hunter Room<br />

Chair: Carolyn Blomley<br />

The influence of soil biotic factors on the<br />

ecology of Trichoderma biological control<br />

agents<br />

Prof Alison Stewart, Lincoln University, NZ<br />

Understanding Trichoderma bioinoculants<br />

in the root system of Pinus<br />

radiata<br />

Mr Pierre Hohmann, Lincoln University,<br />

NZ<br />

A bioassay to screen Trichoderma isolates<br />

for their ability to promote root growth in<br />

willow<br />

Mr Mark Braithwaite, Lincoln University,<br />

NZ<br />

Biofumigation for reducing Cylindrocarpon<br />

spp. in New Zealand vineyard and nursery<br />

soil<br />

Ms Carolyn Bleach, Lincoln University, NZ<br />

1045 MORNING TEA Banquet Room<br />

Session 7A<br />

Cereal pathology 2<br />

Room: Concert Hall<br />

Chair: Colin Wellings<br />

1100 Crown rot of winter cereals: integrating<br />

molecuar studies and germplasm<br />

improvement<br />

Prof Mark Sutherland, University of<br />

Southern Queensland, Qld<br />

1120 Infection of wheat tissues by Fusarium<br />

pseudograminearum<br />

Mr Noel Knight, University of Southern<br />

Queensland, Qld<br />

1140 Monitoring sensitivity to Strobilurin<br />

fungicides in Blumeria graminis on wheat<br />

and barley in Canterbury, New Zealand<br />

Dr Suvi Viljanen‐Rollinson, The New<br />

Zealand Institute for <strong>Plant</strong> and Food<br />

Research Limited, NZ<br />

1200 Cross inoculation of crown rot and<br />

Fusarium head blight isolates of wheat<br />

Mr Philip Davies, University of Sydney,<br />

NSW<br />

Session 7B<br />

Quarantine and exotic pathogens<br />

Room: Cummings Room<br />

Chair: Nerida Donovan<br />

Twenty years of quarantine plant disease<br />

surveillance on the island of New Guinea:<br />

key discoveries for Australia and PNG<br />

Mr Richard Davis, Australian Quarantine<br />

and Inspection Service, Qld<br />

The importance of reporting suspect exotic<br />

or emergency plant pests to your State<br />

Department of Primary Industry<br />

Dr Sophie Peterson, <strong>Plant</strong> Health<br />

Australia, ACT<br />

The use of sentinel plantings in forest<br />

biosecurity; results from mixed eucalypt<br />

species trails in South‐East Asia and<br />

Australia<br />

Dr Treena Burgess, Murdoch University,<br />

WA<br />

Methyl bromide alternatives for<br />

quarantine and pre‐shipment and other<br />

purposes—future perspectives<br />

Ms Janice Oliver, Office of the Chief <strong>Plant</strong><br />

Protection Officer, ACT<br />

Session 7C<br />

Alternatives to chemical control<br />

Room: Hunter Room<br />

Chair: Alison Stewart<br />

Fruit extracts of Azadirachta indica<br />

induces systemic acquired resistance in<br />

tomato against Pseudomonas syringae pv<br />

tomato<br />

Dr Prabir Paul, Amity University, India<br />

Fungal foliar endophytes induce systemic<br />

protection in cacao seedlings against<br />

Phytophthora palmivora<br />

Ms Carolyn Blomley, The University of<br />

Sydney, NSW<br />

Effectiveness of the rust Puccinia<br />

myrsiphylli in reducing populations of the<br />

invasive plant bridal creeper in Australia<br />

Dr Louise Morin, CSIRO Entomology, ACT<br />

Evaluation of essential oils and other plant<br />

extracts for control of soilborne pathogens<br />

of vegetable crops<br />

Ms Cassie Scoble, Department of Primary<br />

Industries and La Trobe University, Vic<br />

1230 LUNCH Banquet Room<br />

16 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


1330 Keynote address—Use of grid weather forecast data to predict rice blast development in Korea Concert Hall<br />

Prof Eun Woo Park, College of Agriculture and Life Sciences, Seoul National University, Korea<br />

1410 Investigating the impact of climate change on plant diseases<br />

Dr Jo Luck, Department of Primary Industries, Vic<br />

1430 Impact of climate change in relation to blackleg on oilseed rape and blackspot on field pea in Western Australia<br />

Dr Moin Salam, Department of Agriculture and Food, WA<br />

1450 Approaches to training in plant pathology capacity building projects in developing countries<br />

Prof LW Burgess, University of Sydney, NSW<br />

1510 Increasing global regulations on fumigants stimulates new era for plant protection and biosecurity<br />

Dr Ian Porter, Department of Primary Industries, Vic<br />

1530 AFTERNOON TEA Banquet Room<br />

1600 Keynote address—A world of possibilities Concert Hall<br />

Dr Barbara Christ, The Pennsylvania State University, USA<br />

1630 Incoming presidential address<br />

Dr Caroline Mohammed, School of Agricultural Science, University of Tasmania<br />

1645 Awards<br />

1700 Close of day<br />

1730 Bus leaves Civic Centre for Beach Party<br />

1800 BEACH PARTY Newcastle Surf Life Saving Club<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 17


Oral abstracts<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 19


President’s address<br />

‘Shield the young harvest from devouring blight’—Charles Darwin, Joseph Banks,<br />

Thomas Knight and wheat rust: discovery, adventure, and ‘getting the message out’<br />

G.I. Johnson{ XE "Johnson, G.I." }<br />

Horticulture 4 Development, PO Box 412, Jamison ACT 2614 Australia Email: greg.johnson@velocitynet.com.au<br />

1969: The year of the first moon landing (20 July 1969), the<br />

Woodstock Festival in upstate New York (15–18 August 1969),<br />

and (coinciding the last day of Woodstock) the beginning of the<br />

<strong>Australasian</strong> <strong>Plant</strong> <strong>Pathology</strong> <strong>Society</strong> (first AGM at 41st ANZAAS<br />

Meeting, Adelaide (18 August 1969) (Purss 1994)). All had a<br />

lengthy gestation and challenges along the way. All have<br />

changed the world!<br />

In the 17th President’s Address to the <strong>Australasian</strong> <strong>Plant</strong><br />

<strong>Pathology</strong> <strong>Society</strong>, David Guest (2001), noted: ‘I became a plant<br />

pathologist because the mechanisms organisms use to<br />

communicate fascinate me’. Well, I became a plant pathologist<br />

because I am gardener at heart. But I have learned along the<br />

way that communication is a critical issue—not only the<br />

communication amongst and between microorganisms and<br />

plants, but also that between plant pathologists, farmers,<br />

politicians and communities. And, communication that is timely,<br />

inspiring and, (preferably) accurate, often yields the most<br />

favourable outcomes.<br />

In this paper, I will explore some of the early communication<br />

relating to plant disease, particularly wheat rusts. I refer to<br />

Erasmus and Charles Darwin, Joseph Banks, Thomas Knight, and<br />

some pioneering Australian researchers, and the roles of<br />

conferences, publications and newspapers, to highlight how<br />

‘getting our message out’ was as important in the 19th and early<br />

20th centuries as it is now. And, finally, I will consider how a<br />

scientific society in the 21st century still has relevance and the<br />

potential to change the world.<br />

20 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


The relevance of plant pathology in food production<br />

André Drenth{ XE "Drenth, A." }<br />

Tree <strong>Pathology</strong> Centre, The University of Queensland and Primary Industries and Fisheries, Indooroopilly, Queensland 4068 Australia<br />

DrenthA@dpi.qld.gov.au<br />

<strong>Plant</strong>s are our only true primary producers of food, fibre and fuel<br />

through the process of photosynthesis. The objective of plant<br />

science in general is to understand the principles and processes<br />

involved in plant growth and reproduction and the objective of<br />

crop science concerns the productivity of our crops.<br />

Keynote address<br />

Over a period of 35 years from 1960 to 1995 the world food<br />

production doubled while the world population more than<br />

doubled from 2.5 billion to 5.6 billion. The present world<br />

population is 6.7 billion and expected to grow to 9 billion by<br />

2050. Agricultural production needs to increase 2.3% a year just<br />

to meet global food demand. At present we increase it by 1.5% a<br />

year. Thus the challenge for Agriculture is to double the global<br />

food production over the next three decades. In addition to<br />

meeting the challenge for food production Agriculture is also<br />

expected to provide renewable fuel. It should be clear that<br />

society needs Agriculture now more than ever before.<br />

The objective of the discipline of plant pathology is to reduce the<br />

impact of diseases on the production of plants for food, fibre<br />

and fuel. <strong>Plant</strong> pathology is an important biological science and<br />

arose out of need during times of famine, poor food security and<br />

large scale crop losses. Despite clearly defined objectives and a<br />

proud history of achievements many plant pathologists would<br />

agree with the statement ‘<strong>Plant</strong> pathology in relation to its<br />

importance to humanity continues to be a grossly underfunded<br />

discipline’ (Strange and Scott, 2005, Annual Review of<br />

Phytopathology). In order to ascertain the significance of our<br />

relatively small discipline we must recognise and document past<br />

contributions, identify and understand future challenges, and be<br />

actively working on tomorrow’s problems. The aim of my<br />

presentation is to address the following three questions:<br />

• What have been the contributions and impacts of plant<br />

science and more specifically plant pathology on food<br />

production?<br />

• What are the key challenges with regards to plant<br />

production which will enable Agriculture to feed a growing<br />

world population in the future?<br />

• What role can our discipline of plant pathology play in<br />

feeding the world?<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 21


Session 1A—Disease management<br />

An integrated approach to husk spot management in macadamia<br />

O.A. Akinsanmi{ XE "Akinsanmi, O.A." } and A. Drenth<br />

Tree <strong>Pathology</strong> Centre, The University of Queensland and Primary Industries and Fisheries, Queensland, 80 Meiers Road Indooroopilly,<br />

4068 Qld, Australia<br />

INTRODUCTION<br />

Husk spot, caused by Pseudocercospora macadamiae is a major<br />

fungal disease of macadamia in Australia (3). Husk spot occurs<br />

only in eastern Australia, costing over $10 million in lost<br />

productivity if the disease is not adequately controlled.<br />

P. macadamiae infects macadamia husks on which it continually<br />

produces inoculum (4), the infection causes premature<br />

abscission of diseased fruit, thus, resulting in extensive yield<br />

losses and reduced kernel quality. Application of fungicide is<br />

currently the only effective method for controlling husk spot (1).<br />

However, several factors including the upsurge in organic<br />

farming, the need for sustainable management practices, and<br />

possible development of fungicide resistant fungal strains and<br />

lack of quantitative information on levels of disease resistance in<br />

varieties require development of integrated management<br />

strategies for controlling husk spot. Systematic studies were<br />

performed to improve husk spot control through the provision of<br />

alternative fungicide, biological control options, effective<br />

cultural practices and diagnostic characters for disease resistant<br />

cultivars.<br />

MATERIALS AND METHODS<br />

In order to evaluate the efficacy of different fungicides and<br />

biocontrol options against husk spot, both laboratory and field<br />

trials were established. Macadamia trees treated with different<br />

fungicide products at three field sites in south east Queensland<br />

and northern New South Wales were evaluated for husk spot<br />

incidence and severity, from onset of visual symptoms to kernel<br />

maturity (2). The area under disease progress curves of the<br />

treatments were compared against each other and with the<br />

untreated control using analysis of variance. The activities of five<br />

Trichoderma species against P. macadamiae were assessed<br />

either singly or in combination with each other and bacteria<br />

(Bacillus subtilis and Pseudomonas fluorescence) in laboratory<br />

experiments. Five characters of macadamia varieties with<br />

varying incidence of husk spot were evaluated as possible<br />

diagnostic characters for husk spot resistant varieties and<br />

discriminant analysis was performed using the identified<br />

diagnostic characters to partition 12 macadamia varieties to<br />

husk spot resistance.<br />

stomata per unit area and number of lesion per fruit classified<br />

macadamia varieties into various resistance groups.<br />

Nut harvested<br />

100%<br />

80%<br />

60%<br />

40%<br />

20%<br />

0%<br />

Poor Moderate Good Very good<br />

None Once Twice<br />

Number of spray applications<br />

Figure 1. Proportion of quality of total macadamia kernel produced from<br />

trees that received varying number of spray applications.<br />

ACKNOWLEDGEMENTS<br />

We acknowledge the University of Queensland, Primary<br />

Industries and Fisheries Queensland, Australian Macadamia<br />

<strong>Society</strong> Ltd., Horticulture Australia Ltd. and Nufarm Australia Pty<br />

Ltd for the funding for this project.<br />

REFERENCES<br />

1. Akinsanmi, O.A., Miles, A.K., and Drenth, A. 2007. Timing of<br />

fungicide application for control of husk spot caused by<br />

Pseudocercospora macadamiae in macadamia. <strong>Plant</strong> Dis. 91:1675–<br />

1681.<br />

2. Akinsanmi, O.A., Miles, A.K., and Drenth, A. 2008. Alternative<br />

fungicides for controlling husk spot caused by Pseudocercospora<br />

macadamiae in macadamia. Australas. <strong>Plant</strong> Pathol. 37:141–147.<br />

3. Beilharz, V., Mayers, P.E., and Pascoe, I.G. 2003. Pseudocercospora<br />

macadamiae sp. nov., the cause of husk spot of macadamia.<br />

Australas. <strong>Plant</strong> Pathol. 32 (2):279–282.<br />

4. Miles, A.K., Akinsanmi, O.A., Sutherland, P.W., Aitken, E.A.B., and<br />

Drenth, A. 2009. Infection, colonisation and sporulation by<br />

Pseudocercospora macadamiae on macadamia fruit. Australas.<br />

<strong>Plant</strong> Pathol. 38 (1):36–43.<br />

RESULTS AND DISCUSSION<br />

Results of field trials showed that pyraclostrobin conferred<br />

significantly (P < 0.05) better protection than trifloxystrobin and<br />

also had somewhat similar efficacy as a tank‐mixture of<br />

carbendazim and copper against husk spot incidence and<br />

severity. The use of pyraclostrobin in rotation with tank mixture<br />

of carbendazim and copper would play a key role in the<br />

management of fungicide resistance in the industry. The<br />

reduction of copper usage would also provide additional benefits<br />

to the macadamia industry. Frequency or number of fungicide<br />

spray applications influenced total kernel quality (Fig. 1). In vitro<br />

volatile inhibition trials showed that growth rate of<br />

P. macadamiae was inhibited by 8% in mixed cultures of T. viride<br />

and T. harzianum, and by 5% in the mixed culture of T. koningii<br />

and T. harzianum but no mycoparasitism was observed in the<br />

hyphal interaction experiments. Our results showed that<br />

significant differences exist in the reaction of macadamia<br />

varieties to husk spot. The discriminant analysis on the disease<br />

incidence and severity, prevalence of stick‐tights, number of<br />

22 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Application methods of phosphonate to control Phytophthora pod rot and stem<br />

canker on cocoa<br />

A. Purwantara, A. Wahab, Y. Imron, V.K.M. Dewi, P.J. McMahon{ XE "McMahon, P.J." }, S. Lambert, P.J. Keane, D.I. Guest<br />

Presenting author: Department of Botany, La Trobe University, Bundoora, Victoria 3086 Email: peter.mcmahon@latrobe.edu.au<br />

INTRODUCTION<br />

Stem canker and Phytophthora pod rot (PPR) or black pod<br />

caused by Phytophthora palmivora are among the most serious<br />

diseases of cocoa in Sulawesi, Indonesia, causing high yield<br />

losses for farmers. Potassium phosphonate (phosphite) has<br />

previously been demonstrated to effectively control canker and<br />

PPR in Papua New Guinea (1). To test the effectiveness of<br />

phosphonate in Sulawesi against Phytophthora diseases and to<br />

compare methods of application of the chemical, two<br />

experimental trials were conducted on cocoa farms in Sulawesi,<br />

Indonesia.<br />

Session 1A—Disease management<br />

METHODS<br />

1 The effect of trunk‐injected phosphonate on stem canker and<br />

PPR in Ladongi, Southeast Sulawesi. Fifty 10 year‐old hybrid<br />

cocoa trees were injected annually with 16 g a.i. phosphonate<br />

(Agrifos 600; Agrichem), fifty with water and fifty left untreated.<br />

For 4.5 years following the initial injection, trees were scored<br />

each month for canker severity and monitored for PPR and CPB<br />

incidence (% ripe pods affected). Treatments were compared by<br />

one‐way ANOVA followed by either the Bonferroni or the<br />

Games‐Howell post‐hoc tests.<br />

2 The effect of phosphonate on stem canker applied by three<br />

differing methods. In a blocked trial with four replicates, 2‐yearold<br />

grafted cocoa trees naturally infected with Phytophthora<br />

stem canker were treated with phosphonate either by trunk<br />

injection, bark painting (combined with Pentrabark; Agrichem),<br />

or implants?, or left untreated and were then scored monthly (as<br />

in Experiment 1) for five months for canker lesion size.<br />

RESULTS AND DISCUSSION<br />

1 Trunk injection. Phosphonate injection cured stem canker<br />

within four months of the initial injection (Fig. 1). Over a 2.5 year<br />

period, phosphonate significantly decreased cumulative PPR<br />

incidence (Table 1), while PPR incidence did not differ between<br />

the untreated and water‐injected trees indicating that the<br />

injection procedure itself did not influence these results. In both<br />

the control and phosphonate‐treated trees, cycles of PPR<br />

infection occurred in the wet season with PPR incidence<br />

fluctuating from less than 30% to greater than 75% (data not<br />

shown). These might have been due to variations in rainfall, the<br />

regular removal of infected pods in the experiment or natural<br />

cycles of sporulation and infection. CPB incidence did not differ<br />

significantly between treatments (data not shown).<br />

Figure 1. Mean canker scores for phosphonate‐injected trees (solid line<br />

and squares) and water‐injected (broken line, open diamond symbols)<br />

assessed from 2002 to 2006. Only trees that initially had canker<br />

infections are included. Trees were injected in June 2002, December<br />

2002 and then every year until the end of 2006.<br />

Table 1. Mean cumulative incidences of PPR in ripe pods from April 2004<br />

to December 2006.<br />

Treatment Mean PPR (%)<br />

Untreated<br />

43.4 ± 1.9 a<br />

Water‐injected<br />

40.2 ± 2.1 a<br />

Phosphonate‐injected<br />

26.3 ± 1.8 b<br />

Mean incidences were calculated from the total number of ripe pods and<br />

infested/infected pods harvested from each tree in the 2.5 years. Each treatment<br />

had 50 trees (replicates). Means (± SE) within columns followed by the same letter<br />

are not significantly different (P


Session 1A—Disease management<br />

Botrytis bunch rot control strategies in cool climate viticultural regions of Australia<br />

and New Zealand<br />

J. Edwards{ XE "Edwards, J." } 1 , D. Riches 1 , K.J. Evans 2 , R.M. Beresford 3 , G.N. Hill 3 , P.N. Wood 4 and D.C. Mundy 5<br />

1 Department of Primary Industries, 621 Burwood Highway, Knoxfield, Victoria 3180.<br />

2 Tasmanian Institute of Agricultural Research, University of Tasmania, New Town Research Laboratories, 13 St Johns Avenue, New Town,<br />

Tasmania 7008.<br />

3 <strong>Plant</strong> and Food Research, Mt Albert Research Centre, Private Bag 92‐169 Auckland, NZ<br />

4 <strong>Plant</strong> and Food Research, Hawke’s Bay Research Centre, Private Bag 1401, Havelock North, Hastings 4157, NZ<br />

5 <strong>Plant</strong> and Food Research, Marlborough Wine Research Centre, P.O. Box 845, Blenheim 7240, NZ<br />

INTRODUCTION<br />

Botrytis bunch rot can cause major losses in wine grapes, while<br />

negligible damage may occur in some seasons, even in the<br />

absence of control measures. We evaluated the effectiveness of<br />

different disease control strategies in five cool climate viticulture<br />

regions in the context of determining disease risk during the<br />

season (1) and economic outcomes for the grower.<br />

MATERIALS AND METHODS<br />

Standard measures of yield, weather, canopy density, bunch<br />

exposure, latent Botrytis cinerea incidence at pre‐bunch closure<br />

(PBC), and botrytis development post‐veraison were recorded at<br />

23 sites in Southern Tasmania, Southern Victoria and New<br />

Zealand between 2006–08 (Table 1). Untreated plots and<br />

combinations of treatments including fungicide timing, canopy<br />

modification (leaf plucking and shoot thinning), and inoculum<br />

reduction by removing bunch trash, were used to generate a<br />

wide range of harvest botrytis severities.<br />

RESULTS<br />

The efficacy of early season (5%, 80% cap‐fall), mid season (peasize,<br />

PBC), late season (veraison, pre‐harvest) fungicide<br />

applications and combination treatments varied considerably<br />

across the different trial sites and seasons. Only 8 out of 23 trials<br />

exceeded 3% botrytis severity at harvest (a level at which many<br />

wineries impose price penalties) on unsprayed vines. Fungicide<br />

applications reduced disease below the 3% threshold in only four<br />

of them (Table 1). Leaf plucking was as effective as the fungicide<br />

treatments in six New Zealand and one Victorian trial, but not in<br />

two other trials. Leaf removal combined with fungicide<br />

applications was the most effective treatment, reducing botrytis<br />

severity to below the 3% threshold even in trials where severity<br />

was >8% in the untreated plots. The experimental inoculum<br />

reduction treatment (bunch trash removal) did not reduce<br />

botrytis severity in these trials.<br />

Tactical decisions about the need for individual fungicide sprays<br />

or canopy management actions depend on the season, canopy<br />

density and bunch exposure, and underlying risk for the region.<br />

The data collected in these trials is being used to develop a<br />

botrytis risk model (1) to assist growers in making decisions for<br />

botrytis management in future.<br />

Table 1. Treatment effects on botrytis severity at harvest for trials<br />

conducted 2006–08 in Australia and New Zealand.<br />

Year<br />

Mean botrytis severity at harvest (%)<br />

Location and<br />

Leaf<br />

Variety 1 Control Spray program 2 pluck<br />

Spray<br />

+leaf pluck<br />

06–07 NZ (A) ‐ SB 25.3 7.7 (mid) ‐ ‐<br />

“ NZ (HB) ‐ SB 4.5 2.1 2.6 0.3<br />

“ NZ (HB) ‐ CH 9.6 5.9 ns 2.2 1.7<br />

“ NZ (M) ‐ SB1 0.2 0.1 ns(early) ‐ ‐<br />

NZ (M) ‐ SB2 0.6 ‐ ‐ ‐<br />

“ Tas ‐ SB 2.5 0.9 (mid) ‐ ‐<br />

“ Tas ‐ R 0.6 0.8 ns ‐ ‐<br />

“ Vic ‐ CH1 0.0 0.0 ns ‐ ‐<br />

“ Vic ‐ SB 0.1 0.0 ns ‐ ‐<br />

“ Vic ‐ CH2 0.0 0.0 ns ‐ ‐<br />

07–08 NZ (A) ‐ SB 1.6 0.1 (early) 0.7 0.3<br />

“ NZ (HB) ‐ SB 8.4 3.5 3.4 0.4<br />

“ NZ (HB) ‐ CH 9.5 1.9 1.9 0.3<br />

“ NZ (M) ‐ SB1 2.1 1.5 (mid/early) 1.1 0.2<br />

NZ (M) ‐ SB2 0.8 ‐ ‐ ‐<br />

“ NZ (M) ‐ SB3 5.3 0.6 ‐ ‐<br />

“ Tas ‐ SB 2.7 1.6 ns (mid) 2.2 1.9<br />

“ Tas ‐ R1 6.2 2.6 (mid) ‐ ‐<br />

“ Tas ‐ R2 1.9 0.8 ns (mid) ‐ ‐<br />

“ Tas ‐ CH 3.6 2.8 (late) ‐ ‐<br />

“ Vic ‐ CH1 2.0 0.0 2.1 0.0<br />

“ Vic ‐ SB 2.3 0.1 0.4 0.6<br />

“ Vic ‐ CH2 0.8 0.0 0.6 0.0<br />

DISCUSSION<br />

Assuming that 3% botrytis severity is a threshold level at which<br />

growers may suffer a price penalty on their crop, many of our<br />

trial sites did not reach the 3% severity threshold even when left<br />

untreated. Botrytis severity at harvest is strongly correlated with<br />

the length of the ripening period (1) and Auckland and Hawke’s<br />

Bay were most prone to serious botrytis epidemics, followed by<br />

Marlborough and Southern Tasmania, with Victoria the lowest<br />

risk. These results highlight the need for benefit‐cost analyses, as<br />

the full cost of a spray program (fuel, pesticide, labour, water,<br />

etc) may not always be recouped, particularly in drier viticultural<br />

regions like Victoria.<br />

Leaf plucking was as effective as a complete fungicide program<br />

at sites with dense canopies. The combined effects of leaf<br />

removal and fungicide application gave the lowest levels of<br />

botrytis, but risk of sunburn and effects on fruit quality are also<br />

important when considering leaf plucking for botrytis control.<br />

1<br />

M=Marlborough, HB=Hawke’s Bay, A=Auckland, SB=Sauvignon Blanc, R=Riesling,<br />

CH=Chardonnay<br />

2 Full fungicide program or the most effective fungicide timing (shown in brackets).<br />

ns—fungicide treatments not significantly different to untreated at P ≤ 0.05<br />

ACKNOWLEDGEMENTS<br />

GWRDC, New Zealand Winegrowers and our respective agencies<br />

for funding this research.<br />

REFERENCES<br />

1. Beresford and Hill (2008) Predicting in‐season risk of botrytis bunch<br />

rot in Australian and New Zealand vineyards. ‘Breaking the mould:<br />

a pest and disease update’, Australian <strong>Society</strong> of Viticulture and<br />

Oenology Seminar Proceedings, Mildura 24 July 2008.<br />

24 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Development of a soil DNA extraction and quantitative PCR method for detecting two<br />

Cylindrocarpon species in soil<br />

INTRODUCTION<br />

C.M. Probst{ XE "Probst, C.M." }, M.V. Jaspers, E.E. Jones and H.J. Ridgway<br />

Faculty of Agriculture and Life Sciences, Lincoln University, PO Box 84, Lincoln 8150, New Zealand<br />

Cylindrocarpon black foot disease has been identified worldwide<br />

as a common cause of vine death in nurseries and in young<br />

vineyards (1), especially in sites converted from orchards or<br />

replanted from grapes. This indicates that the existing soil<br />

conditions may have contributed to the disease, although very<br />

little information is available regarding the survival of the<br />

pathogens in soil. In New Zealand, the three Cylindrocarpon<br />

species equally responsible for black foot disease on grapevines<br />

are C. destructans, C. macrodidymum and C. liriodendri (1).<br />

Recently, quantitative PCR (qPCR) has begun to supersede soil<br />

dilution plating as a method for precise determination of soil<br />

inoculum levels (2). In this research program, qPCR was<br />

optimised to test large soil samples for presence of two<br />

Cylindrocarpon species.<br />

MATERIALS AND METHODS<br />

Fungal isolates. The Lincoln University Culture Collection<br />

provided isolates of the three Cylindrocarpon species, which had<br />

been obtained from symptomatic New Zealand grapevines.<br />

Single spore colonies of 10 randomly selected isolates per<br />

species were grown on potato dextrose agar (PDA) at 20°C in the<br />

dark.<br />

RESULTS<br />

The primer pairs were specific for each of the two species and no<br />

cross reactivity was observed. They produced a 300 bp and 197<br />

bp product for C. macrodidymum and C. liriodendri, respectively.<br />

They were also suitable for a range of genetically diverse<br />

isolates. Each of the primers was able to detect as little as 30 pg<br />

DNA in a standard PCR and 3 pg DNA in a nested PCR. The qPCR<br />

could detect pure DNA at the same level as the nested PCR. For<br />

spore suspensions, the qPCR was able to detect as little as 100<br />

spores in soil. The time course experiment showed that for each<br />

species, less than half of the nuclei remained 1 week after<br />

infesting the soil with conidia. However, at this time, the DNA<br />

could still be visualised on agarose gels of the qPCR products<br />

(Fig 1). After 2, 3 and 6 weeks, the DNA could not be detected.<br />

M ng Weeks<br />

30 3 0.3 0.03 0.003 0 0 1 2<br />

Session 1B—Disease surveys<br />

Soil Infestation. A mixed conidium suspension (10 6 conidia /mL)<br />

was obtained for each species using three isolates. They were<br />

used to infest 5 L pots of soil, three replicates per species, with a<br />

final concentration of 10 5 conidia per gram of soil. Controls were<br />

treated with a similar quantity of water. The 5 L pots were sunk<br />

into, and level with, the ground in the Lincoln University<br />

Vineyard. Two 15 g soil samples were taken from each pot for<br />

DNA extraction at 0, 1, 2, 3 and 6 weeks after set up. All three<br />

species were inoculated into each pot to ensure detection<br />

specificity<br />

DNA extraction. For pure culture extraction, a small plug of<br />

mycelium was transferred from a 5 d old PDA culture to potato<br />

dextrose broth (PDB) and incubated for 7 d at room<br />

temperature, in 12 h light/ dark. The mycelium was harvested<br />

and stored at ‐80°C until DNA extraction using a PureGene DNA<br />

extraction kit (Qiagen). Spectrophotometry was used to quantify<br />

the genomic DNA. DNA was also extracted from 10 6 conidia of<br />

each isolate in a similar way.<br />

For DNA extraction from soil samples, 10 g of soil was placed in a<br />

250 mL bottle with 45 mL water containing 0.01% agar. The<br />

bottles were shaken for 10 min and left to stand for 10 min. The<br />

supernatant was put into two 15 mL tubes, and centrifuged at<br />

4000 x g for 15 min. The pellets were combined and centrifuged<br />

again, with the final pellet being used for DNA extraction using a<br />

PowerSoil kit (MO BIO Laboratories Inc.). Genomic DNA quality<br />

was assessed by visualisation on a 1% agarose gel.<br />

PCR amplification. Species specific primers for the β− tubulin<br />

region of C. macrodidymum and C. liriodendri, donated by Dr<br />

Lizel Mostert (University of Stellenbosch, South Africa), were<br />

used in qPCR with SYBR green chemistry and a Rox internal<br />

standard on an ABI Prism 7700 sequence detector. After the<br />

qPCR, the products were separated by electrophoresis on a 1.5%<br />

agarose gel and visualised under UV light.<br />

Figure 1. 1.5% agarose gel of products amplified during qPCR with the<br />

primers specific for C. liriodendri.<br />

DISCUSSION<br />

The results show that the species specific primers used in a qPCR<br />

system could detect as little as 3 pg of pure DNA, which is<br />

equivalent to 30 Cylindrocarpon macroconidia. When spores<br />

were recovered from soil a sensitivity of 100 spores was<br />

achieved. Following soil inoculation, DNA of fewer than 50% of<br />

the conidium nuclei in the soil was detected after 1 week.<br />

Additional unpublished data has indicated that in the soil<br />

environment, the 4‐celled conidia were often converted to 2<br />

chlamydospores (uninucleate). It is possible that the >50%<br />

decrease observed is a reflection of that conversion, rather than<br />

the death of conidia. Research is continuing to investigate the<br />

population dynamics of these Cylindrocarpon species in soil.<br />

Future research will include the development of C. destructans<br />

specific primers.<br />

ACKNOWLEDGEMENTS<br />

We gratefully acknowledge Lizel Mostert for providing the<br />

primers sequences and Winegrowers New Zealand and TECNZ<br />

for funding this project.<br />

REFERENCES<br />

1. Halleen, F., Fourie, P.H. and Crous P.W. (2006). A review of black<br />

foot disease of grapevine. Phytopathologia Mediterranea 45: S55‐<br />

S67.<br />

2. Kernaghan, G., Reeleder, R.D. and Hoke, S.M.T. (2007).<br />

Quantification of Cylindrocarpon destructans f. sp. Panacis in soils<br />

by real‐time PCR. <strong>Plant</strong> <strong>Pathology</strong> 56: 508–516.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 25


Session 1B—Disease surveys<br />

Why Australia needs a coordinated national diagnostic system<br />

J.R. Moran{ XE "Moran, J.R." } 1 , J.H. Cunningtonv, F.J. Macbeth 2 , C.C. Murdoch 1 , A.M. Kelly 3 , D. Hailstones 4 , S.A. Peterson 5 , B. Hall 6 , S.<br />

Perry 7 , M. Williams 8<br />

1 Department of Primary Industries, Victoria, Private Bag 15, Ferntree Gully DC, 3156, VIC<br />

2 Office of the Chief <strong>Plant</strong> Protection Officer, Department of Agriculture, Fisheries and Forestry, GPO Box 858, Canberra, 2601, ACT<br />

3 NSW Department of Primary Industries, PMB 8 Camden, 2570, NSW<br />

4 CRC Diagnostics Research Program, PMG 8, Camden, 2570, NSW<br />

5 <strong>Plant</strong> Health Australia, 5/4 Phipps Close, Deakin, 2600, ACT<br />

6 South Australian Research and Development Institute, GPO Box 397, Adelaide, 5001, SA<br />

7 Department of Employment, Economic Development and Innovation, GPO Box 46, Brisbane, 4001, Qld<br />

8 Department of Primary Industries and Water, 13 St John’s Ave, New Town, 7008, Tasmania<br />

INTRODUCTION<br />

The accurate and rapid diagnosis of plant pests and diseases<br />

underpins all management activities aimed at preventing the<br />

establishment of exotic pests and is a critical component of<br />

surveillance activities.<br />

<strong>Plant</strong> pest diagnostic services also provide the foundation of<br />

endemic pest management through the timely identification of<br />

plant pests, if left uncontrolled, can disrupt agricultural plant<br />

production and restrict market access for agricultural produce.<br />

Rapid identification also supports quarantine processes, such as<br />

maintaining pest free areas that allow access to domestic and<br />

international markets.<br />

Diagnostic services are delivered by a range of agencies across a<br />

dispersed geography and climate range. The majority of<br />

diagnostic services are provided by state agencies with some<br />

being delivered by the Australian Government, commercial<br />

diagnostic laboratories, CSIRO, and the Universities (1). Services<br />

are provided on an ad‐hoc, commercial and nationally<br />

coordinated basis as required. Diagnostic operations are often<br />

performed in conjunction with collaborative research activities.<br />

The development of a national diagnostic system for plant pests<br />

and diseases has been the subject of discussion for over 10<br />

years. A number of reviews have been conducted of the national<br />

capability (1, 2) and issues of standards and accreditation have<br />

been addressed by the Subcommittee on <strong>Plant</strong> Health Diagnostic<br />

Standards (SPHDS) (3). More recently SPHDS has been given the<br />

responsibility to develop a National Diagnostic Strategy for plant<br />

health. This strategy will complement the National <strong>Plant</strong> Health<br />

Strategy that will be released this year.<br />

This paper reports on the activities of four diagnostic<br />

laboratories, their findings and implications for national<br />

biosecurity.<br />

METHODS AND RESULTS<br />

Data was collected from state government diagnostic<br />

laboratories in Tasmania, South Australia, Victoria and NSW.<br />

Data was also provided by the Office of the Chief <strong>Plant</strong><br />

Protection Officer (OCCPO).<br />

Between January 2006 and May 2009, 45 new pest or diseases<br />

have been brought to the attention of the OCPPO and have<br />

warranted action by the Consultative Committee on Exotic <strong>Plant</strong><br />

Pests. Of these 16 were fungi, 13 were invertebrates, 11 were<br />

virus or phytoplasmas and 5 were bacteria.<br />

interest is that 43% of these investigations arose from ad hoc<br />

samples submitted to the laboratory from industry. Most of<br />

these were ornamentals. The diagnostic investigations revealed<br />

15 new hosts of pathogens already present in Australia, three<br />

new pathogen records for Victoria, seven new pathogen records<br />

for Australia and one new undescribed species.<br />

Data from the other states proved difficult to obtain and to<br />

analyse. Diagnostic activity was not centrally coordinated and<br />

details were not easily accessed due to a lack of readily<br />

searchable databases. South Australian laboratories conducted<br />

in excess of 90 investigations into suspect emergency plant pest,<br />

NSW 88, Queensland 63 and Tasmania 20.<br />

The numbers of investigations conducted by DPI Vic have<br />

increased from 16 in 2006 to 64 in 2008. Similarly in NSW<br />

numbers increased from 19 in 2006 to 25 in the first five months<br />

of 2009. This is in part due to a better understanding by the<br />

diagnostic laboratory of reporting obligations under the<br />

Emergency <strong>Plant</strong> Pest Response Deed (4).<br />

DISCUSSION<br />

The data presented here demonstrates the increasing level of<br />

activity and importance of diagnostic activity nationally. It also<br />

highlights the benefits of a central coordination of diagnostic<br />

services within an agency. Those agencies without a centrally<br />

coordinated diagnostic service find it difficult to ascertain the<br />

level of diagnostic activity that underpins their states<br />

biosecurity. This shortcoming will be overcome as Australia<br />

works towards developing a National Diagnostic System.<br />

ACKNOWLEDGEMENTS<br />

To all the diagnosticians in Australia who have contributed to the<br />

data presented here.<br />

REFERENCES<br />

1. Moran JR and Muirhead IF, 2002, Assessment of the current status<br />

of the human resources involved in diagnostics for plant insect and<br />

diseases pests. A report prepared for <strong>Plant</strong> Health Australia.<br />

2. Anon, 2003, Developing a world class plant pathology diagnostic<br />

network. <strong>Plant</strong> Health Australia Workshop Proceedings<br />

3. Williams, MA, Hall, BH, Plazinski, J, Gray, P, Moran, JR, Stevens, P,<br />

Perry, S. (2009) Subcommittee on <strong>Plant</strong> Health Diagnostic<br />

Standards. APPS 2009.<br />

4. Peterson, SA, Macbeth, FJ, 2009, The importance of reporting<br />

suspect exotic emergency plant pests to your state department of<br />

primary industry, APPS 2009.<br />

In the same time the Victorian DPI diagnostic laboratory<br />

conducted investigations into 133 suspect emergency plant<br />

pests. Some of these investigations arose from detections<br />

interstate and consequent national surveillance activities. Of<br />

26 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Bananas in Carnarvon—good news for growers in survey for quarantine plant pests<br />

and pathogens<br />

S.J. Collins{ XE "Collins, S.J." } A , A.E. Mackie A , J.H. Botha A , V.A.Vanstone A , J.M. Nobbs B and V.M. Lanoiselet A<br />

A<br />

Department of Agriculture and Food Western Australia, South Perth, 6151, Western Australia<br />

B<br />

SARDI <strong>Plant</strong> and Soil Health, GPO Box 397, Adelaide, 5001, South Australia<br />

INTRODUCTION<br />

The banana industry in Carnarvon, Western Australia, is unique<br />

among other Australian banana growing areas. While bananas are<br />

typically grown in tropical to sub‐tropical regions where rainfall is<br />

abundant, Carnarvon has an arid‐desert climate and growers depend<br />

on year round irrigation. Under these unique conditions, plant pests<br />

and pathogens common to other Australian banana growing areas<br />

may not be successful. However, pathogens and pests more suited<br />

to local conditions could potentially thrive. A survey to identify<br />

potential quarantine risks and exotic viruses, bacteria, fungi, insects<br />

and nematodes was initiated by HortGuard® and the Carnarvon<br />

Banana Producers Committee.<br />

METHODS<br />

From the approx. 55 banana growing properties in the Carnarvon<br />

area, 15 were selected for assessment based on a range in<br />

production and management practices, yields and years in<br />

production. Survey activities comprised three main areas, each with<br />

specific sampling methods.<br />

Nematodes. From each property, soil and roots were collected from<br />

ten trees using methods adapted from Pattison et al. (1). Sample<br />

trees were chosen randomly within older blocks where nematode<br />

numbers were likely to be higher. Roots were examined for<br />

symptoms. Nematodes were extracted from roots and soil in a mist<br />

chamber over 5 days, then quantified and identified.<br />

Virus, bacteria and fungi. A minimum of 100 plants per property<br />

were inspected. Leaves, pseudostems, suckers and bunches were<br />

visually assessed for symptoms of disease, mechanical and<br />

physiological damage. Symptoms were described in terms of type,<br />

quantity and severity. The most severely infected leaves were<br />

incubated in moist trays under natural light for 2–7 days. For<br />

identification, fruiting bodies on leaves were examined<br />

microscopically, on half strength Potato Dextrose Agar and on Water<br />

Agar. As no symptoms were observed for viruses, further testing was<br />

not conducted.<br />

Invertebrates. At each property 2–4 blocks of different aged plants<br />

were assessed using sweep netting and direct observation. Sweep<br />

net samples were collected from approx. 200 sweeps of the main<br />

canopy and from 100x10 minute sweeps near the plantation floor.<br />

Direct observation (10x hand lens) focused on invertebrates found<br />

on leaves, fruit, flowers and the plantation floor.<br />

RESULTS AND DISCUSSION<br />

Nematodes. No nematodes of quarantine significance were<br />

identified. Root Knot Nematode (RKN, Meloidogyne sp.) and Spiral<br />

Nematode (Helicotylenchus multicinctus) were identified from both<br />

the roots and soil (Table 1). Spiral Nematode and RKN were present<br />

in all samples. Root Lesion Nematode (Pratylenchus sp.) was<br />

identified from the roots, but not the soil in only one sample (2.7/g<br />

dry root). Root symptoms of both RKN and Spiral Nematode were<br />

observed. Typical symptoms of Burrowing Nematode (Radopholus<br />

similis) were absent, and this species was not identified from any<br />

sample.<br />

These nematode levels would not be regarded as a production<br />

constraint in tropical areas. In Carnarvon, where plants did not have<br />

well developed root systems, it is possible that Spiral and Root Knot<br />

Nematodes may have a greater impact (T. Pattison, pers. comm.,<br />

2009). In tropical areas, Spiral Nematode is of secondary importance<br />

to Burrowing Nematode. However, in areas where temperature and<br />

rainfall conditions are limiting, R. similis is rare, and H. multicinctus is<br />

the major nematode problem which can cause severe damage and<br />

decline in bananas.<br />

Table 1. Spiral and Root Knot Nematode (RKN) densities extracted from<br />

roots and soil.<br />

Nematodes/g dry root Nematodes/200g dry soil<br />

Property Spiral RKN Spiral RKN<br />

1 473.0 25.9 533.1 58.2<br />

2 0 56.6 8.5 357.6<br />

3 162.2 67.4 183.7 164.4<br />

4 142.9 32.0 1418.9 54.2<br />

5 638.9 9.7 892.4 29.6<br />

6 530.4 83.1 236.5 19.4<br />

7 254.8 22.8 245.6 18.9<br />

8 149.5 39.0 571.6 89.1<br />

9 458.1 21.7 577.2 25.5<br />

10 344.5 22.1 1141.9 26.8<br />

11 38.4 279.9 59.4 80.2<br />

12 675.5 1.3 640.6 7.7<br />

13 415.0 188.3 404.7 11.4<br />

14 433.8 3.8 450.1 8.4<br />

15 12.5 200.9 0 167.3<br />

Virus, bacteria and fungi. No diseases of quarantine significance<br />

were identified. From the 20 samples collected, secondary fungal<br />

pathogens were identified from 14, including Alternaria sp.,<br />

Stemphyllium sp., Penicillium sp. and Aspergillus sp. Colletotrichum<br />

sp. was identified from leaf spots on one sample. Pleospora<br />

herbarum was identified from one sample of a leaf spot with dark<br />

margin and bleached centre. P. herbarum is a cosmopolitan fungus<br />

that is occasionally a weak parasite known to cause leaf spots.<br />

Invertebrates. No pests of quarantine significance were identified.<br />

Pest levels were low over the entire area. Seventy per cent of the<br />

surveyed properties use biocontrol agents, which may account for<br />

the low pest numbers. Large numbers of spiders, which are<br />

beneficial in reducing pest invertebrates, were found on all<br />

properties. These included typical wheel‐web spiders (Araneidae)<br />

such as Eriophora spp. as well as Sparassidae, Lycosidae and<br />

Salticidae.<br />

CONCLUSION<br />

No exotic pests or diseases were identified. Only low levels of fungal<br />

pathogens and invertebrate pests were detected. The unique<br />

environmental conditions of Carnarvon, as well as its isolation from<br />

other banana growing areas, may contribute to this finding.<br />

Although Burrowing Nematode (R. similis) was not detected, Spiral<br />

and Root Knot Nematodes may pose a potential production<br />

constraint to bananas in this area.<br />

ACKNOWLEDGEMENTS<br />

Thanks to the Carnarvon Banana HortGuard ® Committee and the<br />

APC Carnarvon Banana Producers Committee for supporting and<br />

funding this work. D. Parr and S. Lawson assisted with sample<br />

collection. H. Hunter, X. Zhang, L. De Brincat, C. Wang, M. You and N.<br />

Eyres processed, and identified samples in the laboratory.<br />

REFERENCES<br />

1. Pattison T, Stanton J, Treverrow N, Lindsay S, Campagnolo D (2000)<br />

Managing Banana Nematodes. 2nd ed. Qld Horticulture Institute,<br />

DPIQ.<br />

Session 1B—Disease surveys<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 27


Session 1C—Soilborne diseases<br />

Can investment in building up soil organic carbon lead to disease suppression in<br />

vegetable crops?<br />

INTRODUCTION<br />

I. Porter{ XE "Porter, I." } A , R. Brett A , S. Mattner A , C. Hall A , R. Gounder A , N. O’Halloran B , P. Fisher B and J. Edwards A<br />

A Biosciences Research Division, Knoxfield Centre, Private Bag 15, Ferntree Gully Delivery Centre, 3156<br />

B Future Farming Systems Division, Tatura Centre, 3616, DPI Victoria<br />

Over the past few years, there has been increasing recognition<br />

of soil as an important non‐renewable asset that needs to be<br />

managed well and looked after. Practices to improve soil health<br />

are increasingly being recommended such as building up soil<br />

organic carbon by the addition of green manures, animal<br />

manures, organic mulches, composts, etc.<br />

Soilborne diseases have traditionally been controlled with the<br />

use of chemicals such as fumigants and fungicides. These options<br />

are being withdrawn worldwide for human health and<br />

environmental reasons. Alternative methods of control such as<br />

the use of biofumigant crops and strategic nutrient application<br />

are now being used more widely by industries. However, what is<br />

the penalty cost of using these methods when conditions are<br />

unfavourable for disease? Can improvements in soil health such<br />

as building up organic carbon be used to create disease<br />

suppressive conditions on vegetable farms?<br />

MATERIALS AND METHODS<br />

Trials were set up to examine the impact of disease<br />

management practices and soil organic amendments on crop<br />

productivity and profitability at two vegetable farms in southern<br />

Victoria.<br />

At site 1, with a history of clubroot of brassica (Plasmodiophora<br />

brassicae), chemical fumigation, biofumigation, fungicides,<br />

nitrate and slow release ammonium fertilisers and organic soil<br />

conditioners (chicken manures, silage and composted green<br />

waste) were applied to broccoli grown during winter when<br />

conditions were unfavourable for disease.<br />

Long‐term trials were also established at both sites to compare<br />

the effect of organic amendments with the standard practice of<br />

metham fumigation on crop yield, profitability and soil health<br />

characteristics. Treatments wer applied in autumn and spring of<br />

2008. Crop rotations included lettuce, broccoli, endive, leeks and<br />

celery.<br />

Overhead irrigation, base fertiliser, insecticide and herbicides<br />

were applied as required, according to local grower practice.<br />

Trial designs were randomised complete blocks with treatments<br />

replicated 4 times. Yield data were analysed using ANOVA.<br />

Effects on soil biological, chemical and physical characteristics<br />

were measured but data analysis is still under way.<br />

RESULTS AND DISCUSSION<br />

Tailored site‐specific nutrient applications were able to provide<br />

equivalent yields (Figure 1) and profit as soil fumigation, with<br />

potentially better benefits to the environment. The<br />

biofumigants, Voom and Fumifert, and composted organic mulch<br />

gave yields equivalent to standard grower practice. Fluazinam, a<br />

pesticide registered for clubroot control, resulted in lower yields<br />

than standard grower practice due to its tendency to cause a<br />

delay in crop maturity.<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

Fluaz.+GBA+CaNO3<br />

3rd harvest<br />

2nd harvest<br />

1st harvest<br />

Rustica fertilzer<br />

Perlka<br />

Fluazinam<br />

Broccoli Yields (kg/plot) Lamattina, 2008<br />

GBA<br />

Green Org. Compost<br />

Grower practice<br />

Voom<br />

Fumifert<br />

Alzon<br />

Metham 800 L/ha<br />

Metham 425 L/ha<br />

GBA+CaNO3<br />

Figure 1. Short term study: Marketable yield of broccoli (3 harvests)<br />

grown with different soil treatments in winter 2008 at Boneo, Victoria,<br />

LSD = 1.24.<br />

Broccoli yield (kg)<br />

10<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

Standard grower - low<br />

Metham - low<br />

Average Total Broccoli Yield<br />

Standard grower - high<br />

Composted mulch - low<br />

Metham - high<br />

Silage - low<br />

Composted mulch - high<br />

Treatment<br />

Chicken Manure - low<br />

Chicken Manure - high<br />

Silage - high<br />

Yield cut 4<br />

Yield cut 3<br />

Yield cut 2<br />

Yield cut 1<br />

Figure 2. Long term study: Marketable yield of broccoli grown with<br />

different organic amendments during winter 2008.<br />

Organic amendments, although producing moderate yields<br />

(Figure 1), reduced grower profit in the short term study, but<br />

increased yields and maintained profit in the long term study<br />

(Figure 2). The trials also confirmed better water and nutrient<br />

use efficiency and biological indicators are showing an increase<br />

in biological activity which leads to better soil resilience and<br />

possible disease suppression. Current and future trials have been<br />

set up to evaluate the long term effect of repeated treatments,<br />

particularly with respect to disease suppression. A cost/benefit<br />

model is also being developed to help growers determine the<br />

benefits of soil health for different crop production systems.<br />

ACKNOWLEDGEMENTS<br />

Horticulture Australia Ltd, DPI Victoria and the Vegetable<br />

Industry funded this research.<br />

28 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Evaluation of soil health indicators in the vegetable industry of temperate Australia<br />

INTRODUCTION<br />

R.W. Brett{ XE "Brett, R.W." }, R.K. Gounder, S.W. Mattner, C.R. Hall and and I.J. Porter<br />

Biosciences Research Division, DPI Victoria, Private Bag 15, Ferntree Gully Delivery Centre, 3156, Vic<br />

Soil health benchmarking trials commenced in the temperate<br />

vegetable industry in 2007 as part of a national program to<br />

better understand soil health and its impact on vegetable<br />

production efficiency and soilborne diseases.<br />

The aim of the benchmarking trials was to identify soil biological,<br />

physical and chemical tests that could be used to detect changes<br />

in soil caused by various farm management practices. Ultimately,<br />

the study aims to use these indicators to identify good soil<br />

health strategies that suppress soilborne pathogens.<br />

MATERIALS AND METHODS<br />

In 2007, 14 commercial production sites and three nonproduction<br />

(“control”) sites were selected at two vegetable<br />

farms on sandy soil at Cranbourne, Victoria.<br />

Olsen P (mg/kg)<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

optimal P<br />

TS3<br />

PS5<br />

TSNC<br />

PS4<br />

PS3B<br />

PS2<br />

PS3A<br />

PS1<br />

TS2<br />

TS1<br />

TS43b<br />

TS43a<br />

TS33a<br />

TS43d<br />

TS33b<br />

TS43c<br />

TS4<br />

Session 1C—Soilborne diseases<br />

Sites that were at different stages of production (table 1) were<br />

compared with each other and with non‐production sites. A<br />

suite of biological, chemical and physical tests were evaluated<br />

for their robustness as indicators of soil health under a range of<br />

farm practices (1).<br />

Table 1. Site selection for benchmarking studies<br />

Site name<br />

TS1<br />

TS2<br />

TS4<br />

PS1, PS3A<br />

PS2, PS3B<br />

PS4<br />

TS33a, TS43a,c<br />

TS33b, TS43b,d<br />

PS5, TS3, TSNC<br />

Crop stage at sampling<br />

6 week spinach<br />

Fallow<br />

Celery tansplant<br />

Leeks residue incorporated<br />

Leeks mature<br />

Kohl rabi residue incorporated<br />

Fumigated, lettuce transplant<br />

Lettuce transplant<br />

Non‐production<br />

RESULTS AND DISCUSSION<br />

Biological tests. Populations of free‐living nematodes were<br />

quantified as an indicator of a site’s previous history and<br />

potential resilience following disturbance. Populations and<br />

community structure of free‐living nematodes were affected by<br />

soil disturbances such as tillage, fumigation, or residue<br />

incorporation, but tended to recover over time. These data<br />

indicate that although tillage, fumigation and crop residue<br />

management can stimulate flushes of bacterial colonisation,<br />

populations of microbial organisms may stabilise during crop<br />

growth. This was more evident for crops of longer (leeks) rather<br />

than shorter (lettuce) duration. These observations are<br />

important when considering the resilience of soils and long term<br />

sustainability of vegetable production.<br />

Chemical tests. Available phosphorus was always higher in<br />

production compared with non‐production sites, irrespective of<br />

cropping history (Figure 1). Combined with wide‐ranging levels<br />

of potassium and sulphur measured in many cropped sites, these<br />

results suggest inappropriate fertiliser application may be<br />

common in vegetable production. The impact of this on control<br />

of soilborne diseases could be important and will be considered<br />

in future studies.<br />

Figure 1. Phosphorus levels were higher than required for vegetable<br />

production at cropped sites compared with non‐production sites,<br />

irrespective of cropping history.<br />

CONCLUSION<br />

On the sandy Cranbourne soils, nematode communities<br />

responded rapidly to changing soil conditions caused by farm<br />

practices and are a good indicator of past soil management, and<br />

potentially a good indicator of soil resilience and soil health.<br />

Some chemical measures such as available phosphorus,<br />

potassium and sulphur were useful in identifying poor fertiliser<br />

management and therefore poor soil health practices.<br />

Additional chemical and biological measures such as nitrogen,<br />

carbon and soil‐borne disease thresholds are being evaluated in<br />

current field trials to identify indicators that show how altered<br />

farm management can improve suppression of soilborne<br />

diseases.<br />

The indicators of soil health that prove to be robust predictors of<br />

the relationship between on‐farm practices, crop yield and soil<br />

health will be further developed so that growers can improve<br />

nutrient, water and disease management on farm. The<br />

information will become part of a national database which will<br />

enable growers to better manage farm inputs and improve the<br />

sustainability and soil health of the vegetable industry in<br />

temperate Australia.<br />

ACKNOWLEDGEMENTS<br />

This work was funded by Horticulture Australia Ltd, the Victorian<br />

Vegetable Industry and the Victorian DPI. We thank the<br />

vegetable farmers who provided land and resources to conduct<br />

these trials.<br />

REFERENCES<br />

1. Porter, I., Brett, R., and Mattner, S. (2007) Management of soil<br />

health for sustainable vegetable production. Horticulture Australia<br />

Ltd VG06090 Final report.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 29


Session 1C—Soilborne diseases<br />

INTRODUCTION<br />

Rhizoctonia AG2.1 and AG3 in soil—competition or synergism?<br />

T.J. Wiechel{ XE "Wiechel, T.J." } and N.S. Crump<br />

Biosciences Research Division, DPI Victoria, Knoxfield Centre, Private Bag 15, Ferntree Gully Delivery Centre 3156<br />

Rhizoctonia solani causes stem and stolon canker, which delay<br />

and reduce emergence, as well as black scurf on tubers. AG2.1<br />

and AG3 are the most dominant anastomosis groups isolated<br />

from potato plants (1). The interactions between AG2.1 and 3<br />

have not been fully examined. Strains within the one species<br />

that co‐occur in a habitat (soil) and may have similar resource<br />

requirements may be expected to compete with each other.<br />

Competition between fungi has been categorised as either 1)<br />

colonisation of unoccupied habitat or 2) colonisation of habitat<br />

that is already occupied (2). This study used radish as a model<br />

system to investigate whether AG2.1 and AG3 compete with<br />

each other in soil, or act synergistically to produce disease.<br />

MATERIALS AND METHODS<br />

Soil inoculation. An isolate of AG2.1 and AG3 (both originally<br />

from potato) were inoculated into soil in combination at various<br />

rates: 1—1 plate fungal mycelium, 0.5—half plate fungal<br />

mycelium, 0.25 quarter plate fungal mycelium per 8 kg soil<br />

(Table 1). Nine radish seeds were planted per pot with 5<br />

replicate pots, and grown at 17°C in a growth room (cool) or<br />

27°C in the glasshouse (warm). Emergence and pruning of the<br />

radish seedlings was assessed weekly for 5 weeks. <strong>Plant</strong>s were<br />

harvested and assessed for yield, after 4 weeks (warm<br />

conditions), or 8 weeks (cool conditions).<br />

There was a good negative relationship between AG2.1 soil<br />

inoculum treatments and radish emergence (R 2 ‐0.9).<br />

Emergence at 7 days<br />

10<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

lsd 0.5217 p


INTRODUCTION<br />

Towards universal detection of Luteoviridae<br />

A. Chomic{ XE "Chomic, A." } A , M. Pearson B and K. Armstrong A<br />

A<br />

Bio‐Protection Research Centre, Lincoln University, PO Box 84, Lincoln 7647, New Zealand<br />

B School of Biological Sciences, University of Auckland, PB 92019, Auckland, New Zealand<br />

Luteoviridae is a plant virus family of 26 species which causes<br />

yield losses in cereals, potatoes and other economically<br />

important crops worldwide. Accurate detection and fast<br />

identification are important components of controlling the<br />

spread of luteoviruses and reducing yield losses.<br />

3. C1F1/C2R3 detected ScYLV species.<br />

None of the primer combinations detected PEMV‐1 or CtRLV.<br />

Session 1D—Virology<br />

PCR based detection is often the method of choice for<br />

Luteoviridae as it is more sensitive than serological methods [1]<br />

which frequently fail to detect infection due to the low<br />

concentration of luteoviruses in plants. Since the currently<br />

available luteovirus primers are mostly species specific and work<br />

under different PCR conditions, universal primers are desirable.<br />

Following PCR amplification sequencing may be required to<br />

confirm virus identity, often resulting in several days delay. A<br />

possible alternative is the use of melt curve analysis (MCA)<br />

which requires less time than sequencing and can be preformed<br />

in most real‐time PCR equipment. MCA is already used for<br />

identification of some animal and plant viruses.<br />

We tested 7 primers designed to target the most conserved<br />

regions of the Luteoviridae genomes and which possess high<br />

homology to over 75% of Luteoviridae species. We also tested<br />

the suitability of MCA in identification of Luteoviridae species.<br />

MATERIALS AND METHODS<br />

Virus isolates. Thirty luteovirus isolates representing 15 species<br />

were obtained from New Zealand and overseas (either as<br />

infected freeze dried plant material or as cDNA). These included<br />

all 5 species from the Luteovirus genus, 8 of the 9 species from<br />

the Polerovirus genus (except CYDV‐RPS), PEMV‐1 (the only<br />

species from the Enamovirus genus) and CtRLV which is not<br />

assigned to a genus.<br />

PCR. Reactions were performed using three combinations of<br />

primers (7 primers in total):<br />

1. C1F1/C1F2/C1R1/C1R2 (129bp and 148bp amplicons)<br />

2. C2F1/C2F2/C2R3 (68bp amplicon)<br />

3. C1R1/C2R3 (75bp amplicon)<br />

All primers have a low level of degeneracy and amplify a<br />

fragment from the coat protein gene. Amplification products<br />

were sequenced to confirm their identity.<br />

Figure 1. Amplification of Luteoviridae with primers C2F1, C2F2 and<br />

C2R3. Lanes: 1 – BWYV, 2 – BMYV, 3 – BClV, 4 – BLRV, 5 – PLRV, 6 – TuYV,<br />

7 – CABYV, 8 – CYDV‐RPV, 9 – SbDV, 10 – Negative (water), M – 1kB<br />

Ladder (NEB).’<br />

The primers are homologous to 6 other Luteoviridae species<br />

(SPLSV, BYDV‐SGV, BYDV‐GPV, BYDV‐RMV, GRAV and TVDV) but<br />

amplification of those species is yet to be tested.<br />

Melt curve analysis showed that at least 5 Luteoviridae species<br />

can potentially be distinguished by different melt temperatures;<br />

but exact temperatures are yet to be determined.<br />

ACKNOWLEDGMENTS<br />

We acknowledge Dave Saul and Karin Farreyrol (Univ. Auckland)<br />

for the original primer design. We thank J.Fletcher, C.Delmiglio,<br />

O.LeMaire, M.Stevens, B.Wintermantel, S.Saumtally, T.van<br />

Antwerpen and S.Fuentes for kindly supplying virus isolates and<br />

MAF Bio‐Security New Zealand, in particular J.Tang, for technical<br />

assistance. This work was supported by MAF Biosecurity New<br />

Zealand and Better Border Biosecurity.<br />

REFERENCES<br />

1. D’Arcy, C.J., L.L. Domier, and L. Torrance, Detection and Diagnosis<br />

of Luteoviruses, in The Luteoviridae, H.G. Smith, and Barker, H.,<br />

Editor. 1999, CABI Publishing.<br />

Melt Curve Analysis. Real‐time PCR was performed using the<br />

same three primer combinations as above, SYBR GreenER I<br />

fluorescent dye and an ABI PRISM ® 7000 real‐time cycler. The<br />

melting of amplification products was performed at a rate of<br />

0.2ºC/min.<br />

RESULTS AND DISCUSSION<br />

Between them the three combinations of primers detected all 13<br />

of the the Luteovirus and Polerovirus species tested:<br />

1. C2F1/C2F2/C2R3 primers detected nine species (Fig.1).<br />

2. C1F1/C1F2/C1R1/C1R2 primers detected BYDV‐PAV, BYDV‐<br />

MAV and BYDV‐PAS.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 31


Session 1D—Virology<br />

Massive parallel sequencing of small RNAs to identify plant viruses and virus‐induced<br />

small RNAs<br />

R.M. MacDiarmid{ XE "MacDiarmid, R.M." } A , D. Cohen A , A.G. Blouin A and L.J. Collins B<br />

A<br />

The New Zealand Institute for <strong>Plant</strong> and Food Research Ltd, 120 Mount Albert Road, Auckland, New Zealand<br />

B<br />

Allan Wilson Centre for Molecular Ecology and Evolution and Institute of Molecular BioSciences, Massey University, Palmerston North,<br />

Massey University, Palmerston North, New Zealand<br />

INTRODUCTION<br />

Small RNAs are short (20–25 nt) RNAs that can guide the<br />

degradation of target mRNA or alternatively inhibit their<br />

translation. The net result of either process is the lack of<br />

functionality of the target RNA. In plants, small RNAs are<br />

generated from two distinct sources: the plant genome or<br />

infecting viruses 1 . The intrinsic small RNAs encoded by a plant<br />

are mainly composed of microRNAs that are processed from<br />

longer RNA transcripts and target other plant‐encoded mRNAs.<br />

Small interfering (si) RNAs are derived from an infecting virus<br />

and target that same virus RNA for degradation, thus forming<br />

the basis of a highly malleable, sequence‐specific defence<br />

mechanism.<br />

Recent advances in sequencing technologies now permit the<br />

economic identification of millions of RNA sequences from a<br />

single sample of plant tissue. Knowledge of the siRNA sequences<br />

in a plant infected with an unknown virus provides a potential<br />

tool for identification of the virus from which the sequences are<br />

derived. This process is aimed at use at the border by Biosecurity<br />

New Zealand.<br />

Figure 1. Venn diagram illustrating the pools of small RNA sequences<br />

derived from uninfected and infected tissue.<br />

MATERIALS AND METHODS<br />

<strong>Plant</strong> and virus samples. Nicotiana occidentialis plants grown at<br />

20°C with 10 hours low light per day were either mockinoculated<br />

or inoculated from leaf samples infected with one of<br />

four known viruses or from six plants infected with unknown<br />

viruses. Small RNAs were isolated and sequenced using the<br />

Genome Analyser (Illumina®) at the Allan Wilson Centre<br />

Genome Sequencing Service.<br />

Bioinformatic programs. Data were matched to known targets<br />

using ELAND (Illumina®), or formed into contiguous sequences<br />

using Velvet 2 or Edena 3 before matching to virus sequences in<br />

publicly available sequence databases using Basic Local<br />

Alignment Search Tool (BLAST) 4 .<br />

RESULTS AND DISCUSSION<br />

We have sequenced small RNAs from both uninfected and virus<br />

infected samples, resulting in over 22 million sequences<br />

(~500,000 unique sequences). The unique sequences present in<br />

uninfected tissue were subtracted in silico from the infected<br />

sequence pool to identify the small RNAs specific to uninfected<br />

tissue and those common to both uninfected and virus‐infected<br />

tissues (Figure 1). The remaining sequences that were present<br />

only in the virus‐infected tissue were then used either to match<br />

to a known infecting virus (Figure 2) or to develop a method to<br />

identify unknown virus infectious agents also present in the<br />

tissue. To increase search specificity, contiguous sequences were<br />

generated before BLAST analyses.<br />

Subtraction in silico of the infecting virus siRNAs from the<br />

sequences found only in virus‐infected plants left a large pool of<br />

small RNAs (~240,000 unique sequences) that were<br />

(presumably) not of viral origin. The identity of these novel<br />

plant‐derived small RNA sequences present in response to virus<br />

infection is being investigated and will be discussed.<br />

Figure 2. Alignment of small interfering RNAs to the genome of a<br />

potexvirus. The potexvirus genome (five arrows representing open<br />

reading frames) is aligned with the ~4,000 unique, cognate siRNA<br />

sequences (dashes).<br />

ACKNOWLEDGEMENTS<br />

This research is funded by the New Zealand Foundation for<br />

Research, Science and Technology (C06X0710).<br />

REFERENCES<br />

1. Mlotshwa S, Pruss GJ, Vance V. (2008) Small RNAs in viral infection<br />

and host defense. Trends <strong>Plant</strong> Sci. 13(7): 375–82.<br />

2. Zerbino DR, Birney E. (2008) Velvet: algorithms for de novo short<br />

read assembly using de Bruijn graphs. Genome Res. 18(5):821–9.<br />

3. Hernandez D, Francois P, Farinelli L, Osteras M, Schrenzel J. (2008)<br />

De novo bacterial genome sequencing: millions of very short reads<br />

assembled on a desktop computer. Genome Res. 18:802–809.<br />

4. Altschul SF, Madden TL, Schäffer AA, Zhang J, Zhang Z, Miller W. &<br />

Lipman DJ. (1997) Gapped BLAST and PSI‐BLAST: a new generation<br />

of protein database search programs. Nucleic Acids Res. 25:3389–<br />

3402.<br />

32 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Chickpea chlorotic stunt virus, an important virus of cool‐season food legumes in Asia<br />

and North Africa and potentially in Australia<br />

Safaa G. Kumari{ XE "Kumari, S.G." } A , Nouran Attar A , H. Josef Vetten B and Joop van Leur c<br />

A International Center for Agricultural Research in the Dry Areas (ICARDA), P.O. Box 5466, Aleppo, Syria<br />

B Julius Kuehn Institute, Federal Research Centre for Cultivated <strong>Plant</strong>s (JKI), Messeweg 11/12, 38104 Braunschweig, Germany<br />

C New South Wales Department of Primary Industries, Tamworth Agricultural Institute, 4 Marsden Park Road, Calala NSW 2340, Australia<br />

INTRODUCTION<br />

Chickpea chlorotic stunt virus (CpCSV), a proposed new member<br />

of the genus Polerovirus (family Luteoviridae) was first described<br />

in Ethiopia in 2006 (1) and has since been reported from Eritrea<br />

(4), Syria (3), Egypt, Morocco and Sudan (2). It naturally infects<br />

many legume crops (e.g., chickpea, lentil, field pea, faba bean) as<br />

well as some leguminous weeds and four wild non‐legume plant<br />

species (1, 2, 3, 4). Typical symptoms of CpCSV‐infected plants<br />

are leaf rolling, yellowing and stunting (Figure 1‐A). CpCSV is a<br />

phloem‐limited virus that is present in very low concentrations<br />

and transmitted only by aphids (Aphis craccivora Koch.) in a<br />

persistent manner (1, 3).<br />

This study reports the use of a few monoclonal antibodies to<br />

CpCSV (1, 2) for detecting different CpCSV isolates from 8<br />

countries in Asia and North Africa.<br />

MATERIALS AND METHODS<br />

Sample collections, diagnostic techniques and reagents used—A<br />

total of 3265 food legume samples showing yellowing/stunting<br />

symptoms were collected from 8 countries (Azerbaijan, China,<br />

Eritrea, Ethiopia, Lebanon, Syria, Tunisia and Yemen) and tested<br />

for the presence of CpCSV using the following three mixtures of<br />

CpCSV monoclonal antibodies (MAbs) in tissue‐blot<br />

immunoassay (TBIA): M‐I = 1‐1G5 + 1‐3H4 + 1‐4B12; M‐II = 5‐2B8<br />

+ 5‐3D5; and M‐III= 5‐5B8 (1, 2). In addition, over 500 TBIA blots<br />

from chickpea, faba bean and field peas collected in northern<br />

NSW, Australia, for different types of symptoms were processed<br />

with the CpCSV MAbs.<br />

RESULTS AND DISCUSSION<br />

Figure 1‐B shows how CpCSV is detected in infected plants by<br />

using TBIA.<br />

A<br />

B<br />

Figure 1. (A) Symptoms produced<br />

on a chickpea plant infected with<br />

chickpea chlorotic stunt virus<br />

(CpCSV); (B) TBIA detection of<br />

CpCSV in a stem blot from an<br />

infected chickpea plant (top) as<br />

compared to that from a noninfected<br />

plant (bottom).<br />

Table 1 summarises the reactions of the MAbs with the samples<br />

from 8 countries in Asia and North Africa. Results obtained<br />

placed the tested samples in four groups: group A comprised<br />

1218 samples (from Azerbaijan, Lebanon, Syria, Tunisia and<br />

Yemen) reacting with MAbs M‐II and M‐III; group B contained<br />

254 samples (from Azerbaijan, China, Eritrea, Ethiopia and<br />

Tunisia) reacting only with MAb M‐I; group C included 77<br />

samples (from Azerbaijan and Ethiopia) reacting with MAbs M‐I<br />

and M‐II; and group D consisted of 38 samples (from Ethiopia<br />

and Tunisia) reacting only with MAb M‐II. The presence of CpCSV<br />

was confirmed in a representative number of samples from each<br />

group and country by RT‐PCR using specific primer sets. This is<br />

the first record of CpCSV in Azerbaijan, China, Lebanon, Tunisia<br />

and Yemen.<br />

Testing of the Australian samples gave CpCSV‐positive reactions<br />

for several chickpea, faba bean and field pea samples, which<br />

tested negative to antisera specific for other luteoviruses. These<br />

findings require reconfirmation, but are an indication that CpCSV<br />

(or a closely related non‐described virus) may be present in<br />

Australia. Sequence analysis of RT‐PCR amplicons is in progress<br />

and will shed more light on the relatedness among the<br />

aforementioned groups and to the two major CpCSV strains<br />

proposed by Abraham et al. (2).<br />

Table 1. TBIA reactions of different luteovirus isolates with three groups<br />

of monoclonal antibodies raised against CpCSV<br />

No. of<br />

samples<br />

No. of TBIApositive<br />

TBIA reaction with<br />

different CpCSV MAbs*<br />

Country/Crop tested samples M‐I M‐II M‐III<br />

Eritrea<br />

Chickpea 211 32 + ‐ ‐<br />

Tunisia<br />

Chickpea 711 6 + ‐ ‐<br />

8 ‐ + ‐<br />

335 ‐ + +<br />

Faba bean 127 8 ‐ + +<br />

Azerbaijan<br />

Chickpea 320 11 ‐ + +<br />

2 + + ‐<br />

153 + ‐ ‐<br />

Lentil 86 1 + ‐ ‐<br />

Ethiopia<br />

Chickpea 129 70 + + ‐<br />

29 ‐ + ‐<br />

1 + ‐ ‐<br />

Lentil 9<br />

5 + + ‐<br />

1 ‐ + ‐<br />

Faba bean 190 4 + ‐ ‐<br />

Fenugreek 80 55 + ‐ ‐<br />

China<br />

Faba bean 15 2 + ‐ ‐<br />

Syria<br />

Chickpea 579 460 ‐ + +<br />

Faba bean 362 279 ‐ + +<br />

Lentil 78 46 ‐ + +<br />

Pea 21 12 ‐ + +<br />

Yemen<br />

Pea 35 19 ‐ + +<br />

Lebanon<br />

Chickpea 32 3 ‐ + +<br />

Faba bean 232 43 ‐ + +<br />

Pea 35 1 ‐ + +<br />

Lentil 13 1 ‐ + +<br />

* M‐I= 1‐1G5+1‐3H4+1‐4B12; M‐II= 5‐2B8+5‐3D5; M‐III= 5‐5B8 (1, 2)<br />

REFERENCES<br />

1. Abraham AD, Menzel W, Lesemann DE, Varrelmann M, Vetten HJ<br />

(2006) Chickpea chlorotic stunt virus: A new polerovirus infecting<br />

cool‐season food legumes in Ethiopia. Phytopathology 96, 437–<br />

446.<br />

2. Abraham, AD, Menzel W, Varrelmann M, Vetten HJ (2009)<br />

Molecular, serological and biological variation among chickpea<br />

chlorotic stunt virus isolates from five countries of North Africa and<br />

West Asia. Archives of Virology 154 (Published online, April 5, 2009)<br />

3. Asaad NY, Kumari SG, Kassem AH, Shalaby A, Al‐Chaabi S, Malhotra<br />

RS (2009) Detection and characterization of chickpea chlorotic<br />

stunt virus in Syria. Journal of Phytopathology 157 (Published<br />

online, April 15, 2009).<br />

4. Kumari SG, Makkouk KM, Loh M, Negassi K, Tsegay S, Kidane R,<br />

Kibret A, Tesfatsion Y (2008) Viral diseases affecting chickpea crop<br />

in Eritrea. Phytopathologia Mediterranea 47, 42–49.<br />

Session 1D—Virology<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 33


Keynote address<br />

INTRODUCTION<br />

Emerging frontiers in forest pathology<br />

Michael J. Wingfield{ XE "Wingfield, M.J." }<br />

Forestry and Agricultural Biotechnology Institute, University of Pretoria, Pretoria 0002, South Africa<br />

If one uses the first text book on the topic as starting point, the<br />

study of tree diseases is a little more than 100 years old. Thus,<br />

Robert Hartig’s fundamental text published in 1874 and<br />

translated into English in 1894 as “Textbook of the diseases of<br />

trees” provided the first firm foundation for forest pathology. It<br />

is interesting, that it was not long after the publication of<br />

Hartig’s forest pathology text book that chestnut blight caused<br />

by Cryphonectria parasitica was first found outside its native<br />

range in the United States in 1904. At the time, this might have<br />

been considered co‐incidental, but it was clearly linked to<br />

growing trade, particularly between northern hemisphere<br />

countries and particularly including wood and wood products.<br />

Thus, outbreaks of devastating diseases such as white pine<br />

blister rust, Dutch elm disease and others caused by non‐native<br />

pathogens first emerged more or less during the same period of<br />

time.<br />

The study of tree diseases and the field of forest pathology have<br />

grown firmly during the course of the last Century. The causal<br />

agents of diseases of unknown aetiology have been discovered,<br />

new diseases have been recorded and the biology of the biology<br />

of many important tree pathogens has been elucidated. A suite<br />

of outstanding text books have been produced both for the<br />

teaching environment as well as for diagnosticians. Furthermore,<br />

we have celebrated the careers of many outstanding forest<br />

pathologists, working in many different parts of the world and<br />

contributing substantially to our understanding of tree<br />

pathogens and the diseases that they cause.<br />

While forest pathology has become a well‐established field of<br />

study and outstanding forest pathologists practice in many<br />

countries, there are causes for concern. In this regard, new and<br />

in some cases very serious tree diseases continue to appear<br />

regularly, both in natural forests and in plantation environments.<br />

This implies that in many situations, we are failing to manage the<br />

global threats to forests and forestry due to diseases. Ironically,<br />

there are also indications, at least in some countries that<br />

positions for forest pathologists and educators in this field are<br />

decreasing, rather than increasing in number.<br />

NEW TREE PATOHGEN TRENDS<br />

The threats relating to new diseases emerging from the<br />

accidental movement of pathogens from one area to another is<br />

well‐recognised. This is also a category of disease that continues<br />

to increase, despite global efforts to manage the movement of<br />

plants and plant products. A worrying trend that appears to be<br />

increasing is the adaptation of pathogens to new hosts.<br />

Apparent host and vector shifts, at levels previously unexpected<br />

are occurring for host specific pathogens. Novel pathogens are<br />

also emerging through hybridisation. The processes leading to<br />

host/vector shifts and hybridisation leading to the evolution of<br />

so‐called “new pathogens” are pathogens” are poorly<br />

understood and they deserve increased attention.<br />

FOREST PATHOLOGY OPPORTUNITIES<br />

New tree diseases and new categories of tree disease are<br />

emerging and this is a trend that is likely to continue to grow.<br />

This will also bring new and exciting challenges for forest<br />

pathologists. New tools and technologies continue to become<br />

available for research and these will clearly also enhance the<br />

depth and breadth of tree pathology investigations.<br />

Identification of tree pathogens and disease diagnosis is a field<br />

that has changed markedly during the course of the last two<br />

decades. The emergence of DNA‐based tools for these purposes<br />

is now commonly used and this will be increasingly true in the<br />

future. Many tree pathogens that have been treated as single<br />

species are now known to represent suites of cryptic taxa and<br />

this is already having a substantial impact on for example<br />

quarantine regulations. DNA based techniques have likewise<br />

made it possible to recongnise hybrids between species and they<br />

have shifted our understanding of pathogen population biology<br />

to new and exciting levels. Furthermore, diagnostics have<br />

become much more rapid and also more accurate. These are all<br />

areas that will grow substantially in the foreseeable future.<br />

For many years, quarantine efforts to restrict the movement of<br />

pathogens, was based on lists of names of undesirable<br />

organisms. It is now widely recognised that such lists have many<br />

flaws, many that have also emerged from the more<br />

sophisticated taxonomy and population biology tools now<br />

available. The focus has clearly shifted to understanding<br />

pathways that allow pathogens to move globally. Many<br />

opportunities remain to be explored in this domain and exciting<br />

new strategies are likely to emerge in the future<br />

In order to understand the importance and impact of climate<br />

change on tree health, there will clearly need to be a greater<br />

focus on team research. The importance of integrating the<br />

efforts of scientists from different disciplines has long been<br />

recognised. Yet there are disappointingly few examples of tree<br />

health research conducted by formally constituted and<br />

supported multidisciplinary teams.<br />

CONCLUSION<br />

Forest pathology is more important today than it has ever been<br />

in the past. The threat of tree diseases in their many different<br />

manifestations has increased steadily ever since they were first<br />

formally recognised. This trend is likely to continue to grow. The<br />

challenge for forest pathologists will be to fight for a fair share of<br />

biological science budgets for tree health research. This funding<br />

can then be expended to capture the many new and exciting<br />

opportunities and technologies that can contribute substantially<br />

to improving the current and future health of the world’s<br />

forests.<br />

Climate change is one of the most important issues facing the<br />

world and it will clearly also impact on tree health. Very little<br />

focused research has been conducted on the likely impact of<br />

climate change on tree diseases. This is a complex area of study<br />

but it is also one that will require the attention of forest<br />

pathologists globally.<br />

34 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


INTRODUCTION<br />

Variability in pathogenicity of Quambalaria pitereka on spotted gums<br />

G.S. Pegg{ XE "Pegg, G.S." } A , A.J. Carnegie B , M.J. Wingfield C , A. Drenth A<br />

A<br />

Tree <strong>Pathology</strong> Centre, The University of Queensland/Primary Industries and Fisheries, Queensland, Australia, 4068<br />

B<br />

Forest Resources Research, NSW Department of Primary Industries, New South Wales, Australia, 2119<br />

C<br />

Forestry and Agricultural Biotechnology Institute, University of Pretoria, South Africa, 0002<br />

Quambalaria shoot blight (QSB) is a serious disease affecting the<br />

expanding eucalypt plantation estate in subtropical and tropical<br />

eastern Australia (1,2). Quambalaria pitereka has been isolated<br />

from foliage, stems and woody tissue of species in the genera<br />

Corymbia, Blakella and Angophora in Australia (1,2,3).<br />

Quambalaria pitereka affects the new flush of Corymbia foliage<br />

causing spotting, necrosis and distortion of expanding leaves and<br />

green stems.<br />

The aim of this investigation was to identify if variation in<br />

pathogenicity levels existed between isolates of Q. pitereka<br />

collected from a range of species and regions in NSW and<br />

Queensland.<br />

MATERIALS AND METHODS<br />

Isolates of Q. pitereka were collected from spotted gum<br />

plantations and Corymbia species trials in north Queensland<br />

(NQ), south east Queensland (SEQ) and northern New South<br />

Wales (NSW). Isolates used were: Q107 from C. torelliana (NQ),<br />

Q147 from C. citriodora (NQ), Q151, Q152 from C. variegata<br />

(SEQ) and Q191, Q200 from C. variegata (NSW).<br />

Spotted gum seedlings (1/2 sibling Woondum provenance) and<br />

cuttings were grown in steam sterilised soil mix and fertilised<br />

with slow release Osmocote ® (Native Trees) as required and<br />

irrigated twice a day for 10 minutes each using overhead<br />

sprinklers. Glasshouse temperatures were maintained at 25–<br />

28°C during the day and 20–22°C overnight. Seedlings were<br />

grown for three months prior to inoculation with Q. pitereka.<br />

QSB Score (x/100)<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

Q107 Q147 Q151 Q152 Q191 Q200<br />

Quambalaria pitereka isolate<br />

Figure 1. Comparison of pathogenicity of a range of isolates of<br />

Q. pitereka on spotted gum, Woondum provenance.<br />

DISCUSSION<br />

Variability in pathogenicity between isolates of Q. pitereka has<br />

not previously been studied. This preliminary study using isolates<br />

from a range of regions identifies only a small degree of<br />

variability. More extensive studies are needed to investigate the<br />

significance of pathogen virulence on disease development,<br />

especially in relation to selecting for disease resistance. Further<br />

studies are required to investigate Q. pitereka population<br />

genetics and potential variability in isolate virulence.<br />

ACKNOWLEDGEMENTS<br />

We thank Queensland Primary Industries Innovation and<br />

Biosecurity Program Investment, Forest <strong>Plant</strong>ations Queensland,<br />

Integrated Tree Cropping, Forest Enterprises Australia and<br />

Forests NSW for providing the necessary funding for this<br />

research.<br />

Session 2A—Forest pathology/native<br />

Isolates of Q. pitereka were obtained from single lesions and<br />

grown on Potato Dextrose Agar (PDA) for 2 to 3 weeks in the<br />

dark at 25°C. A spore suspension (1x10 6 spores/ml) was obtained<br />

by washing plates with sterile distilled water (SDW) to which two<br />

drops of Tween 20 had been added prior to inoculation.<br />

Seedlings were inoculated using a fine mist spray (2.9 kPa<br />

pressure) generated by a compressor driven spray gun (Iwata<br />

Studio series 1/6 hp; Gravity spray gun RG3, Portland, USA), to<br />

the upper and lower leaf surfaces of the seedlings until runoff<br />

was achieved. All seedling trees were covered with plastic bags<br />

immediately after inoculation to maintain high humidity levels<br />

and to increase the period of leaf wetness. Bags were removed<br />

after 48 hours and plants watered using overhead irrigation<br />

systems twice a day for a period of 10 minutes. Sub‐samples of<br />

the spore suspension applied to the trees were placed onto PDA<br />

and incubated at 25°C for 48 hours to ensure that the spores<br />

were viable.<br />

REFERENCES<br />

1. Carnegie AJ, 2007a. Forest health condition in New South Wales,<br />

Australia, 1996–2005. I. Fungi recorded from eucalypt plantations<br />

during forest health surveys. <strong>Australasian</strong> <strong>Plant</strong> <strong>Pathology</strong> 36, 213–<br />

24.<br />

2. Pegg GS, O’Dwyer C, Carnegie AJ, Wingfield MJ, Drenth A 2008.<br />

Quambalaria species associated with plantation and native<br />

eucalypts in Australia. <strong>Plant</strong> <strong>Pathology</strong> 57, 702–14.<br />

3. Simpson JA, 2000. Quambalaria, a new genus of eucalypt<br />

pathogens. <strong>Australasian</strong> Mycologist 19, 57–62.<br />

Each treatment was replicated six times with a single inoculation<br />

per tree and assessed for disease incidence and severity 14 days<br />

later. A QSB score was then calculated to compare levels of<br />

pathogenicity between isolates. A comparison between isolates<br />

was done using ANOVA (Stat<strong>View</strong>®).<br />

RESULTS<br />

Significant variability in pathogenicity was observed between<br />

isolates. Isolate Q152 was significantly more virulent than all<br />

other isolates.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 35


Session 2A—Forest pathology/native<br />

Movement of pathogens between horticultural crops and endemic trees in the<br />

Kimberleys<br />

INTRODUCTION<br />

T.I. Burgess A , J.D. Ray B , M.L. Sakalidis{ XE "Sakalidis, M.L." } A , V. Lanoiselet C , G.E.StJ. Hardy A<br />

A<br />

School of Biological Sciences and Biotechnology, Murdoch University, Murdoch, 6150, Australia<br />

B<br />

Australian Quarantine and Inspection Service, NAQS/OSP, Marrara NT, 0812<br />

C Department of Agriculture and Food , Baron‐Hay Court , South Perth, 6151, WA<br />

Recently a survey of endophytes associated with boabs<br />

(Adansonia gregorrii) and associated tree species in the<br />

Kimberleys, Western Australia has resulted in the description of<br />

seven new species in the Botryosphaeriaceae (Pavlic et al. 2008).<br />

Additionally several common species of Lasiodiplodia,<br />

(L. theobromae, L. pseudoptheobromae and L. parva) were also<br />

isolated as endophytes of endemic tree species.<br />

Concurrently, surveys in the Ord River Irrigation Area (ORIA)<br />

have revealed Mangiferum indica trees showing symptoms of<br />

dieback and cankers. In this project we isolated, identified and<br />

determined the pathogenicity of fungi associated with these<br />

cankers.<br />

MATERIALS AND METHODS<br />

Isolation and identification. Fungi were isolated from cankers<br />

using standard techniques. Due to the similarity in morphological<br />

features among the Botryosphaeriaceae, molecular<br />

identification was conducted by extracting DNA, amplifying and<br />

sequencing the ITS and EF1‐α gene regions (Burgess et al. 2005)<br />

and conducting phylogenetic analysis to identify cryptic species.<br />

Pathogenicity trails. Trials were conducted using unripe but<br />

mature Kensington pride mangoes The mangoes were washed in<br />

water then submerged in 1.5% NaOH (Bleach) solution for 2<br />

minutes to disinfest. Mangoes were then allowed to air dry and<br />

were stored overnight at temp 28–33°C. Mangoes were<br />

inoculated with 11 fungal isolates (9 replicates per isolate), by<br />

wounding with the tip of a sterile pipette and immediately<br />

placing a colonised agar plug mycelia side down over the wound.<br />

After 6 days the lesions were measured.<br />

RESULTS AND DISCUSSION<br />

Seven species from the Botryosphaeriaceae were isolated from<br />

mangoes in the ORIA. Three of these species, L. theobromae,<br />

N. ribis and N. dimidiatum are known pathogens of mangoes<br />

causing either cankers or post‐harvest disease. Four species,<br />

P. adansoniae, N. novaehollandia, L. pseudotheobromae and<br />

L. parva have not been reported previously from mangoes, but<br />

are commonly found as tree endophytes in the surrounding<br />

region.<br />

Table 1. Species of Botrysphaeriaceae isolated as endophytes of endemic<br />

trees in the Kimberleys and causing disease in mangoes in Kununurra<br />

Species Mangoes Trees<br />

Pseudofusicoccum adansoniae #<br />

Pseudofusicoccum kimberleyense #<br />

Pseudofusicoccum ardesiacum #<br />

Lasiodiplodia theobromae<br />

Lasiodiplodia pseudotheobromae<br />

Lasiodiplodia parva<br />

Lasiodiplodia margaritacea #<br />

Lasiodiplodia crassispora<br />

Dothiorella longicollis #<br />

Fusicoccum ramosum #<br />

Neoscytalidium novaehollandia #<br />

Neoscytalidium dimidiatum<br />

Neofusicoccum ribis<br />

# denotes species newly described by Pavlic et al. 2008<br />

Common endophytes and latent pathogens of mangoes,<br />

Neofusicoccum mangiferae, N. parvum and Botryosphaeria<br />

dothidea, were not isolated in the ORIA. In addition N. ribis was<br />

isolated very rarely. The most commonly isolated pathogens<br />

were L. theobromae, P. adansoniae and the two Neoscytalidium<br />

spp. interestingly these fungi are also common in the<br />

surrounding endemic vegetation. It appears that many of the<br />

pathogens of mangoes in the ORIA have moved onto the trees<br />

from the surrounding environment rather than arriving with<br />

nursery stock. The differences in the species found in the ORIA<br />

compared with other mango growing regions could be due to<br />

the geographic isolation of the region.<br />

REFERENCES<br />

1. Burgess TI, Barber PA, Hardy GESJ, 2005. Botryosphaeria spp.<br />

associated with eucalypts in Western Australia including<br />

description of Fusicoccum macroclavatum sp. nov. <strong>Australasian</strong><br />

<strong>Plant</strong> <strong>Pathology</strong>. 34: 557–567.<br />

2. Pavlic D, Barber PA, Hardy GESJ, Slippers B, Wingfield MJ, Burgess<br />

TI, 2008. Seven new species of the Botryosphaeriaceae discovered<br />

on baobabs and other native trees in Western Australia. Mycologia.<br />

100: 851–866.<br />

All tested species were pathogenic to mangoes, with<br />

Lasiodiplodia theobromae causing the largest lesion followed by<br />

the Neoscytalidium spp. P. adansoniae caused the smallest<br />

lesions.<br />

36 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Pathogenicity of Phytophthora multivora to Eucalyptus gomphocephala and<br />

E. marginata<br />

P.M. Scott 1 , P. Barber 1 , T. Jung 1 , B.L. Shearer 1,2 , G.E. Hardy, T.I. Burgess{ XE "Burgess, T.I." } 1<br />

1 Centre of Phytophthora Science and Management, School of Biological Sciences and Biotechnology, Murdoch University, Australia<br />

2 Science Division, Department of Environment and Conservation, Australia<br />

INTRODUCTION<br />

Since the early 1990s there has been a significant decline of<br />

E. gomphocephala, and more recently E. marginata, in the tuart<br />

forest in tuart woodland in Yalgorup National Park SW Western<br />

Australia, although no satisfactory aetiology has been<br />

established to explain the decline. Characteristics of the canopy<br />

dieback and decline distribution are reminiscent of other forest<br />

declines known to involve Phytophthora soil pathogens and<br />

indicate that a Phytophthora species may be involved in the<br />

decline. In 2007 isolates of Phytophthora multivora, recently<br />

described by (1), were recovered from rhizosphere soil of<br />

declining or dead trees of Eucalyptus gomphocephala and<br />

E. marginata. For E. gomphocephala and E. marginata, the<br />

pathogenicity of P. multivora was tested: ex situ on seedlings<br />

using a soil infestation method; and in situ on stems using an<br />

under bark infestation method.<br />

MATERIALS AND METHODS<br />

Ex situ soil inoculation trial: The roots of E. gomphocephala<br />

seedlings, grown in neutral coarse river sand, were infested with<br />

a vegetable juice—vermiculite medium colonised with five<br />

isolates of P. multivora isolated across the Yalgorup decline and<br />

two isolates of P. cinnamomi; while the roots of E. marginata<br />

seedlings were infested with one isolate of P. multivora as<br />

described (2). After one year the roots of infested seedlings were<br />

scanned and the lengths of roots of different diameters were<br />

calculated using the WINRHIZO Pro V 2007d software (Reagent<br />

Instruments, Québec, Canada). Isolates were recovered from the<br />

roots of all infested seedlings.<br />

In situ under bark inoculation trial: The stems of less than 1.5 m<br />

tall E. gomphocephala saplings, naturally growing on site in<br />

Yalgorup National Park, were under bark inoculated with five<br />

isolates of P. multivora; while saplings of E. marginata were<br />

inoculated with one isolate of P. multivora as described (3).After<br />

nine weeks saplings were harvested and lesion lengths<br />

calculated.<br />

RESULTS<br />

Ex situ soil inoculation trial: Preliminary results from the ex situ<br />

soil infestation trial indicate that E. gomphocephala seedlings<br />

treated with P. multivora isolate Pm1 and Pm2 and P. cinnamomi<br />

isolates Pc1 and Pc2, had significantly less roots between 0–2<br />

mm in diameter compared to the control (Figure 1). Eucalyptus<br />

gomphocephala seedlings infested with P. multivora isolates<br />

Pm3, Pm4 and Pm5 did not have significantly less roots<br />

compared to the control across any size class. Eucalyptus<br />

marginata seedlings infested with P. multivora isolate Pm1, did<br />

not have significantly less roots compared to the control across<br />

any size class.<br />

In situ under bark inoculation trial: Saplings of E. gomphocephala<br />

and E. marginata inoculated with all P. multivora isolates had<br />

significantly longer lesions compared to the control. When<br />

harvested the average lesion length on P. multivora inoculated<br />

E. gomphocephala and E. marginata seedlings was 13.6 mm and<br />

90.5 mm respectively.<br />

Root length (cm)<br />

20000<br />

16000<br />

12000<br />

8000<br />

4000<br />

0<br />

Con Pm1 Pm2 Pm3 Pm4 Pm5 Pc1 Pc2<br />

Isolates<br />

Figure 1. Length of live roots (mean ± SE) of E. gomphocephala seedlings<br />

for roots 0–2 mm in diameters Con, control; Pm 1–5, P. multivora<br />

isolates 1–5; Pc 1–2 P. cinnamomi isolates.<br />

DISCUSSION<br />

The significant reduction in root diameter in E. gomphocephala<br />

seedlings after infestation with isolates Pm1 and Pm1 indicates<br />

that P. multivora is a pathogen of E. gomphocephala under<br />

glasshouse conditions and may be a significant soil pathogen to<br />

E. gomphocephala in the field. The variation in pathogenicity of<br />

P. multivora isolates used in the soil infestation trial suggests<br />

that there is variation in the pathogenicity of P. multivora<br />

isolates within the field.<br />

The significant lesion lengths measured in E. gomphocephala and<br />

E. marginata sapling inoculated with P. multivora isolates<br />

confirms that P. multivora is a pathogen to both host species<br />

under conditions where P. multivora can colonise the vascular<br />

tissue in the field. The lesion lengths indicate that P. multivora<br />

can be especially E. marginata saplings.<br />

The variation in pathogenicity observed between isolates and<br />

species in both soil infestation and under bark inoculation trials<br />

suggests that P. multivora can be significantly aggressive to both<br />

E. gomphocephala and E. marginata; although further research<br />

is needed to understand the population dynamics of the<br />

pathogen and its impact within the tuart decline in Yalgorup<br />

National Park.<br />

REFERENCES<br />

1. Scott PM, Burgess TI, Barber PA, Shearer BL, Stukely MJC, Hardy<br />

GESJ, Jung T (2009) Phytophthora multivora sp. nov., a new species<br />

recovered from declining Eucalyptus, Banksia, Agonis and other<br />

plant species in Western Australia. Persoonia.<br />

2. Jung T, Blaschke H, Neumann P (1996) Isolation, identification and<br />

Pathogenicity of Phytophthora species from declining oak stands.<br />

European Journal of Forest <strong>Pathology</strong> 26, 253–272.<br />

3. Shearer BL, Michaelsen BJ, Somerford PJ (1988) Effects of isolate<br />

and time of inoculation invasion of secondary phloem of Eucalyptus<br />

spp. and Banksia grandis by Phytophthora spp. <strong>Plant</strong> Disease 72,<br />

121–126.<br />

Session 2A—Forest pathology/native<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 37


Session 2A—Forest pathology/native<br />

INTRODUCTION<br />

Microscopy of progressive decay of fungi isolated from meranti tree canker<br />

E. Erwin{ XE "Erwin, E." } A , Won‐Joung Hwang B , Shuhei Takemoto C and Yuji Imamura D<br />

A Forestry Faculty, University of Mulawarman, Jln. Ki Hajar Dewantara No.1, Samarinda 75123, Indonesia<br />

B Institute of Wood Technology, Akita Prefectural University, 11‐1 Kaieizaka, Noshiro, Akita 016‐0876, Japan<br />

C <strong>Plant</strong> <strong>Pathology</strong> Research Team, National Institute of Fruit Tree Science, Tsukuba, Ibaraki 305‐8605, Japan<br />

D Research Institute for Sustainable Humanosphere (RISH), Kyoto University, Uji, Kyoto 611–0011, Japan<br />

White rot basidiomycetes are especially important in the forest<br />

ecosystem because they are the only fungi capable of degrading<br />

all cell wall components (cellulose, lignin, hemicelluloses) of<br />

wood. Micromorphological aspects of two main types of white<br />

rot, selective delignification and simultaneous rot, have been<br />

distinguished [1]. Light red meranti (Shorea smithiana) and<br />

yellow meranti (Shorea gibossa) trees which growing in one area<br />

of natural dipterocaprs in Kalimantan, Indonesia inhabited by<br />

white‐rot fungi, Schizophyllum commune [2] and Phlebia<br />

brevispora [3], respectively. It is of interest that S. commune and<br />

P. brevispora in the decayed wood of standing S. smithiana and<br />

S. gibbosa and have been causing simultaneous decay.<br />

This study was performed to confirm their progressive decay<br />

patterns in artificial laboratory conditions (in vitro).<br />

MATERIALS AND METHODS<br />

Fungal strain and decay test procedure. The decay fungi isolated<br />

from decayed wood of S. smithiana and S. gibbosa cankerous<br />

trees, were genetically identified by their ITS sequence as<br />

Schizophyllum commune [2] and Phlebia brevispora [3],<br />

respectively. Twelve sound wood‐blocks (20 x 20 mm in crosssection<br />

x 10 mm in length) obtained from the uninfected<br />

portions of the stem disks were inoculated with the identified<br />

fungi, and incubated in accordance with the JIS K 1571 soil‐block<br />

test procedure [4].<br />

Microscopic observations. Various stages of the decay wood<br />

were examined using light and scanning electron microcopy. Six<br />

exposure times were analyzed: 2, 4, 6, 8, 10 and 12 weeks.<br />

RESULTS<br />

After 12 weeks of exposure in the laboratory decay test, wood<br />

blocks of S. smithiana that had been inoculated with S. commune<br />

fungus sustained an average weight loss of 1.82%, whereas P.<br />

breviospora decayed S. gibbosa wood more aggressively than S.<br />

commune (Table 1).<br />

Table 1. Weight loss in S. smithiana and S. gibbosa wood infected with S.<br />

commune and P. brevispora, respectively.<br />

Weight loss percentage<br />

Incubation period<br />

(Mean±SE)<br />

(weeks) S. smithiana S. gibbosa<br />

2 0.42 ± 0.32 0.91 ± 0.10<br />

4 0.50 ± 0.44 2.24 ± 0.60<br />

6 0.54 ± 0.50 5.02 ± 1.03<br />

8 0.60 ± 0.62 8.23 ± 1.22<br />

10 0.80 ± 0.40 11.80 ± 5.15<br />

12 1.82 ± 0.40 12.34 ± 2.76<br />

Figure 1. Decay wood caused by fungus for 12 weeks incubation. (A)<br />

Hyphal branches of S. commune fungus passed through pits in vessels of<br />

S. smithiana (arrows). Bar 5 μm; (B) Erosion channels in parenchyma<br />

cells adjacent to infected vessels of S. gibbosa (arrow). Bar 10 μm.<br />

DISCUSSION<br />

Slight erosion of wood cell walls in S. smithiana over 12 weeks’<br />

incubation was classified as the early stage of simultaneous<br />

decay, and showed a similar pattern to that observed in<br />

naturally decayed wood samples.<br />

P. brevispora reduced S. gibbosa wood weight by 0.91–12.34%<br />

and produced progressive simultaneous decay over 2–12 weeks’<br />

incubation in vitro. The first 6 weeks of incubation was classified<br />

as the early stages decay, in which pit erosion and slight erosion<br />

of cell walls facilitated hyphal between cells. Numerous and<br />

conspicuous holes as well as erosion troughs in cell walls, which<br />

were found at the end of 8 weeks’ incubation, showed that an<br />

intermediate stage of decay had occurred. Furthermore,<br />

complete degradation of wood cell components, termed the<br />

advanced stage of decay, was found in some areas of wood<br />

blocks after 12 week’s incubation.<br />

REFERENCES<br />

A<br />

1. Blanchette RA (1984) Screening wood decayed by white‐rot fungi<br />

for preferential lignin degradation. Appl Environ Microbiol 48:647–<br />

653<br />

2. Erwin, Takemoto S, Hwang WJ, Takeuchi M, Itoh T, Imamura Y<br />

(2008) Anatomical characterization of decayed wood in standing<br />

light red meranti and identification of the fungi isolated from the<br />

decayed area. Journal of Wood Science 5:233–241<br />

3. Erwin, Hwang WJ, Takemoto S and Imamura, Y (2007)<br />

Micromorphological changes of wood and decay fungi in yellow<br />

meranti (Shorea gibossa) stem canker. 57th Annual Meeting of the<br />

Japan Wood Research <strong>Society</strong>. Hiroshima, August 8–10.<br />

4. JIS K 1571 (2004) Test methods for determining the effectiveness of<br />

wood preservatives and their performance requirements. Japanese<br />

Industrial Standard (JIS), Japanese Standards Association, Tokyo.<br />

B<br />

38 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Optimising conditions to investigate gene expression in pathogenic Streptomyces<br />

using RT‐qPCR<br />

INTRODUCTION<br />

T.J. Wiechel{ XE "Wiechel, T.J." }, R.C. Ates and N.S. Crump<br />

Biosciences Research Division, DPI Victoria, Knoxfield Centre, Private Bag 15, Ferntree Gully Delivery Centre 3156<br />

Pathogenic Streptomyces species cause common scab of potato.<br />

The pathogen produces a phytotoxin, thaxtomin A, that induces<br />

a host response that creates the corky symptoms of common<br />

scab that can be superficial, raised, netted or deep pitted. QPCR<br />

is being increasingly used as a method for mRNA quantification<br />

but only a relatively few studies have reported the use of RTqPCR<br />

for plant pathogens. This study aimed at optimising total<br />

RNA isolation from pathogenic Streptomyces spp. and using RTqPCR<br />

conditions to quantify thaxtomin A transcripts.<br />

Ct value<br />

Session 2B—Soilborne disease<br />

MATERIALS AND METHODS<br />

Pathogenic Streptomyces strain. Streptomyces scabies strain S29<br />

was grown on yeast malt extract agar (YME) at 27°C. After one<br />

week S29 was inoculated into 3 types of growth media (yeast<br />

malt extract broth YME, thaxtomin defined medium broth TDM,<br />

oatmeal broth OM) both with and without cellobiose (1). The<br />

inoculated media were incubated at 27°C on a rotating shaker<br />

for one week then filtered through cheese cloth to separate the<br />

Streptomyces cultures from the media. The cultures were used<br />

fresh or stored at ‐20°C until ready to extract RNA.<br />

RNA extraction. RNAqueous 4 PCR (Ambion) for isolation of<br />

DNA‐free RNA was used to extract total RNA from the<br />

Streptomyces cultures according to the manufacturers<br />

instructions.<br />

cDNA synthesis and quantification. Total RNA was treated with<br />

DNaseI and ABI High capacity cDNA reverse transcription kit was<br />

used to transcribe RNA into cDNA according to the<br />

manufacturers instructions. The resulting cDNA was quantified<br />

using Nanodrop ND‐1000 Spectrophotometer.<br />

cDNA dilutions and qPCR. The cDNA was diluted to the following<br />

concentrations 500, 250, 125 and 62.5 ng/5 µL and replicated 3<br />

times. QPCR was performed in 0.2 mL tubes in a Rotor Gene<br />

3000 machine (Corbett Research). Each 25 µL reaction consisted<br />

of 50 ng template cDNA, 1x SYBR qPCR Super Mix (Invitrogen)<br />

0.1 µM of primers TxTATq1 and TxTATq2 (2). The thermal cycle<br />

protocol was 50°C for 2 min, 95°C for 2 min and 45 cycles of 95°C<br />

for 15 sec and 60°C for 60 sec and Melt cycle of 60–95°C in 1°C<br />

intervals. Negative controls without cDNA template were run<br />

with every assay to rule out contaminations.<br />

RESULTS<br />

The RNA isolation method RNAqueous 4 PCR provided a high<br />

quantity of high integrity RNA. The RNA was transcribed into<br />

cDNA using the ABI High capacity cDNA reverse transcription kit<br />

and gave good quality cDNA. The transcript levels of thaxtomin A<br />

were quantified using SYBR qPCR with primers specific to<br />

thaxtomin A. A standard curve was constructed from the means<br />

of the serially diluted cDNA replicates (Figure 1). The small<br />

standard errors with these replicates demonstrated that the<br />

assay is reproducible and highly robust. The transcript levels of<br />

thaxtomin A were greater in media amended with cellobiose<br />

than without, with oatmeal broth having the highest transcript<br />

levels as expressed by Ct value (Figure 2).<br />

Concentration ng/ 5 µL<br />

Figure 2. Standard curve obtained from serially diluted cDNA. Ct values<br />

are the means of 3 replicates. Error bars represent standard error of the<br />

means.<br />

Fluorescence<br />

OM+<br />

YME-<br />

TDM+<br />

OM-<br />

+ve<br />

YME+<br />

TDM-<br />

Figure 3. Amplification curves for amended media. OM oatmeal, OM+<br />

oatmeal with cellobiose, TDM thaxtomin defined medium, TDM+<br />

thaxtomin defined medium with cellobiose, YME yeast malt extract,<br />

YME+ yeast malt extract with cellobiose.<br />

DISCUSSION<br />

This proof of concept study optimised the RNA extraction<br />

method from pathogenic Streptomyces and RT‐ qPCR conditions<br />

to quantify thaxtomin A transcripts. The levels of thaxtomin A<br />

varied depending on the growth media used, with oatmeal plus<br />

cellobiose giving the highest quantity. The next stage of this<br />

research will be to use these optimum conditions to study gene<br />

expression of pathogenic Streptomyces spp.<br />

ACKNOWLEDGEMENTS<br />

This project was facilitated by HAL in partnership with the<br />

processing potato industry, funded by the processing potato levy<br />

and voluntary contributions from industry partners. The<br />

Australian Government provided matched funding for HAL’s R&D<br />

activities. DPI Victoria contributed funding to this project.<br />

REFERENCES<br />

1. Johnson EG, Joshi MV, Gibson DM, Loria R (2007) Cellooligosaccharides<br />

released from host plants induce pathogencity in<br />

scab‐causing Streptomyces species. Physiological and Molecular<br />

<strong>Plant</strong> <strong>Pathology</strong> 71:18–25.<br />

2. Crump NS (2005) Prediction and molecular detection of soilborne<br />

pathogens of potato. Final Report HAL PT01019.<br />

-ve<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 39


Session 2B—Soilborne disease<br />

Fusarium oxysporum and Pythium species associated with vascular wilt and root rots<br />

of greenhouse cucumbers<br />

INTRODUCTION<br />

L.A. Tesoriero{ XE "Tesoriero, L.A." } A , L.M. Forsyth A and L.W. Burgess B<br />

A NSW Department of Primary Industries, EMAI, PMB 8, Camden, 2570, NSW<br />

B The University of Sydney, 2006, NSW<br />

Fusarium wilt of cucumber caused by F. oxysporum f.sp.<br />

cucumerinum (F.o.cuc) and root rots caused by Pythium species<br />

have a worldwide distribution. F. oxysporum f. sp. radiciscucumerinum<br />

(F.o.r‐cuc) has more recently been described in<br />

Europe and North America causing stem and root rots of<br />

greenhouse cucumbers (1, 2). F.o.r‐cuc was shown to be<br />

genetically distinct from F.o.cuc (3).<br />

RESULTS AND DISCUSSION<br />

Pythium isolates were assigned to nine taxonomic groups. Seven<br />

conformed to named species: P. aphanidermatum (Edson)<br />

Fitzpatrick, P. coloratum Vaartaja, P. irregulare Buisman, P.<br />

mamillatum Meurs, P. oligandrum Drechsler, P. spinosum<br />

Sawada, and P. ultimum Trow var. ultimum Trow. The remaining<br />

two taxa were placed in Pythium ‘Group T’ and the Pythium<br />

species ‘HS Group’ (4).<br />

Pythium species are commonly associated with roots from a<br />

wide botanical host range (4). Member species and isolates can<br />

be aggressive pathogens, weakly pathogenic, secondary<br />

invaders, saprophytes or hyperparasites.<br />

Except for the one previous study of Fusarium wilt in Australian<br />

greenhouse cucumbers (5), the causes and severity of wilt, stem<br />

and root rots are unknown.<br />

Preliminary surveys of greenhouse cucumbers in crops grown in<br />

the peri‐urban districts of Sydney revealed severe vascular wilt,<br />

stem and root rots associated with heavy crop losses. The<br />

purpose of this study was to identify potential agents<br />

responsible for these diseases and to determine their<br />

significance across major Australian production areas.<br />

MATERIALS AND METHODS<br />

Greenhouse cucumber crops were surveyed across major<br />

Australian production areas in NSW, Qld., S.A. and W.A. between<br />

2001 and 2006. Cucumber plants were assessed in‐situ for<br />

stunting, wilting, discolouration of hypocotyl tissue, and pinkorange<br />

sporulation on stem lesions. Samples of diseased plants<br />

that were not permanently wilted were collected, transported to<br />

a laboratory and processed within 24 hours where possible.<br />

Growers freighted further samples directly to the laboratory.<br />

Sub‐samples of symptomatic roots and stems were washed<br />

clean of soil or media, surface‐sterilised in a hypochlorite<br />

solution and rinsed in sterile‐distilled water.<br />

Pythium species were isolated from roots and crown tissue that<br />

had been plated to potato carrot agar amended with 5 ppm<br />

pimaricin and 10 ppm rifampicin. Fusarium species were<br />

similarly isolated from roots and stem vascular tissue plated to<br />

¼‐strength potato dextrose agar amended with 100 ppm<br />

novobiocin. Plates were incubated at 25°C and examined for the<br />

presence of typical features of these genera under a compound<br />

light microscope. Pythium species were sub‐cultured and<br />

identified (4). Fusarium isolates were single‐spored and<br />

identified (6).<br />

Representative isolates of Pythium taxa were further<br />

characterised by sequence analysis of ITS rDNA regions and<br />

compared with published sequences (7).<br />

Fusarium isolates were compared with F.o.cuc and F.o.r‐cuc<br />

reference isolates from overseas by determining their vegetative<br />

compatibility groupings (VCGs), repPCR profiles and sequences<br />

of their β−tubulin, calmodulin and α−elongation factor genes.<br />

One unique VCG of F. oxysporum dominated in all production<br />

areas. It also had a unique repPCR profile while gene sequences<br />

aligned closely with F.o.cuc. A number of other isolates each had<br />

unique VCGs while they shared similar repPCR profiles. A subset<br />

of this group clustered with gene sequences of F.o.r‐cuc. Further<br />

VCG studies will be required for F.o.cuc isolates since the<br />

publication of four new groups from China (8).<br />

This study confirmed that F.o.cuc is the major subspecies of<br />

Fusarium oxysporum associated with greenhouse cucumber<br />

vascular wilt disease in Australia. Isolates consistent with F.o.rcuc<br />

were also identified as a minor population. Pathogenicity<br />

and host range studies completed the characterisation of<br />

representative Fusarium and Pythium isolates. Results of these<br />

studies will be reported elsewhere.<br />

ACKNOWLEDGEMENTS<br />

NSW DPI, the Australian Government and HAL Ltd funded this<br />

study. L. Gunn, S. Azzopardi, F. Lidbetter, C. Howard and S.<br />

Peterson assisted with molecular studies. Thanks also to B.<br />

Summerell and E. Liew for guidance and use of the laboratory at<br />

the RBG, Sydney.<br />

REFERENCES<br />

1. Vakalounakis, D.J. 1996. Root and stem rot of cucumber caused by<br />

Fusarium oxysporum f. sp. Nov., <strong>Plant</strong> Disease 80:313–316.<br />

2. Punja, Z.K. and Parker, P. 2000. Development of Fusarium root and<br />

stem rot, a new disease on greenhouse cucumber in British<br />

Columbia, caused by Fusarium oxysporum f.sp. radiciscucumerinum,<br />

Canadian Journal of <strong>Plant</strong> <strong>Pathology</strong> 22:349–363.<br />

3. Vakalounakis, D.J. and Fragkiadakis, G.A. 1999. Genetic diversity of<br />

Fusarium oxysporum isolates from cucumber: differentiation by<br />

pathogenicity, vegetative compatibility, and RAPD fingerprinting,<br />

Phytopathology 89:161–168.<br />

4. Plaats‐Niterink, A.J. van der. 1981. Monograph of the genus<br />

Pythium. Studies in Mycology 21, 242pp.<br />

5. Wicks, T.J., Volle, D. and Baker, B.T. 1978. The effect of soil<br />

fumigation and fowl manure on populations of Fusarium<br />

oxysporum f.sp. cucumerinum in glasshouse soil and on the<br />

incidence of cucumber wilt. Agricultural Record 5:5pp.<br />

6. Burgess L.W., Summerell B.A., Bullock S, Gott K.P. and Backhouse D<br />

1994. Laboratory Manual for Fusarium Research (3rd Edition) The<br />

University of Sydney, 133pp.<br />

7. Lévesque, C.A. and De Cock, A. 2004. Molecular phylogeny and<br />

taxonomy of the genus Pythium. Mycological Research 108:1363–<br />

1383.<br />

8. Vakalounakis, D.J, Wang, Z., Fragkiadakis, G.A., Skaracis, G. and Li,<br />

D‐B. 2004. Characterisation of Fusarium oxysporum isolates<br />

obtained from cucumber in China by pathogenicity, VCG and RAPD.<br />

<strong>Plant</strong> Disease 88:645–649.<br />

40 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Fusarium oxysporum f. sp. fragariae: a main component of strawberry crown and root<br />

rots in Western Australia<br />

INTRODUCTION<br />

H. Golzar{ XE "Golzar, H." } and D. Phillips<br />

Department of Agriculture and Food Western Australia, 3 Baron‐Hay Court South Perth WA 6151<br />

Root and crown rots are important diseases of strawberry crops<br />

worldwide. The fungi Phytophthora spp., Verticillium spp.,<br />

Fusarium spp., Gnomonia fragariae and Colletotrichum spp. have<br />

been reported as causal agents of strawberry crown and root<br />

rot, and they caused considerable yield reduction. In many cases<br />

the causes of root rot are referred to several pathogens of which<br />

Pythium spp., Rhizoctonia spp., Cylindrocarpon spp., Phoma spp.,<br />

Coniothyrium spp. and Fusarium spp. have been the most<br />

common in the root rot complex (1, 3). Fusarium oxysporum f.<br />

sp. fragariae has been reported in Queensland, Western<br />

Australia and Japan as an important pathogen of strawberry (2,<br />

3, 4).<br />

MATERIALS AND METHODS<br />

Field survey. During the surveys conducted a high incidence of<br />

strawberry death was observed in coastal districts up to 50 km<br />

north of Perth areas in 2005 and 2006. A total of 50 partially<br />

diseased and asymptomatic plants were randomly collected<br />

from ten major strawberry fields. Roots were carefully washed<br />

under running tap water and the crown of each plant was<br />

dissected lengthwise. Vascular discolouration of the crown and<br />

root was evident on some of the samples collected.<br />

Isolation method. Crowns and roots of diseased and<br />

asymptomatic plants were surface‐sterilised by immersion in a<br />

1.25% aqueous solution of sodium hypochlorite for 1 min, rinsed<br />

in sterile water and dried in a laminar flow cabinet. Out of 50<br />

samples 500 root and crown specimens with equal numbers of<br />

10 pieces per sample were selected randomly. Specimens were<br />

separately placed on potato dextrose agar (PDA), water agar and<br />

selective media (P10VPH and P10VP) and then incubated at 22 ±<br />

3°C. Emerged fungal colonies were sub‐cultured on carnation<br />

leaf agar, PDA and V‐8 juice agar and incubated at 25°C with a<br />

12 h dark and light cycle. Growth rate, colony morphology and<br />

morphological characteristics of the isolated fungi were<br />

determined.<br />

Pathogenicity test. The pathogenicity of 10 Fusarium isolates<br />

was tested on Fragaria × ananassa cv. Camarosa, Lycopersicon<br />

lycopersicum cv. Petula and Cucumis sativus (Lebanese<br />

cucumber) in a glasshouse experiment. Strawberry runners and<br />

4‐week‐old seedlings of tomato and cucumber were inoculated<br />

by dipping the roots in a spore suspension (10 5 spores/mL)<br />

before planting. Controls were dipped in tap water.<br />

RESULTS AND DISCUSSION<br />

During the surveys, a high incidence of decline and death of<br />

strawberries was observed. Mortality of Camarosa and Gaviota<br />

varieties of strawberry (Fragaria × ananassa) was between 0 and<br />

60% in some strawberry fields.<br />

recovery of the fungi (Table1) was indicated that Fusarium spp.<br />

(74%) was predominant while Phytophthora spp. (21%), Pythium<br />

spp. (23.5%), Phoma spp. (3%), Rhizoctonia spp. (9%),<br />

Colletotrichum spp. (1.5%) and Macrophomina spp. (16%) were<br />

minor components of root and crown rots of strawberries. In<br />

most cases combination of fungi were recovered from both root<br />

and crown strawberry samples tested. A culture of F. oxysporum<br />

f. sp. fragariae has been deposited in the culture collection of<br />

Department of Agriculture and Food Western Australia as WAC<br />

12708.<br />

Table 1. Average percentage of fungi associated with root and crown<br />

rots of Camarosa cultivar in 2005 and 2006<br />

2005 and 2006<br />

Total recovery %<br />

Recovered Fungi Diseased Healthy*<br />

Fusarium oxysporum f. sp. fragariae 74.0 10.0<br />

Phytophthora spp. 21.0 1.5<br />

Pythium spp. 23.5 9.0<br />

Phoma spp. 3.3 0.0<br />

Rhizoctonia spp. 9.3 2.0<br />

Colletotrichum spp. 1.5 0.0<br />

Macrophomina spp. 16.0 1.0<br />

Others 9.0 1.0<br />

* Asymptomatic plant<br />

Strawberry root and crown rots seem to be caused by diverse<br />

fungi. Determining the specific fungus or fungi causing crown<br />

and root rots is complicated, because the fungi isolated from<br />

diseased tissue may only be capable of causing disease under<br />

specific conditions or only in specific associations with other<br />

organisms.<br />

ACKNOWLEDGEMENT<br />

The authors would like to thank the Western Australia<br />

strawberry industry for financial support.<br />

REFERENCES<br />

1 D’Ercole N, Nipoti P, Manzali D (1989) Research on the root rot<br />

complex of strawberry plants. Acta Horticulturae 265, 497–506.<br />

2 Golzar, H, Phillips, D. and Mack S. 2007. Occurrence of strawberry<br />

root and crown rot in Western Australia. <strong>Australasian</strong> <strong>Plant</strong><br />

Disease Notes. 2, P. 145–147<br />

3 Maas JL (1998) ‘Compendium of strawberry diseases.’ 2nd edn.<br />

(APS Press: St Paul, MN)<br />

4 Winks BL, Williams YN (1965) A wilt of strawberry caused by a new<br />

form of Fusarium oxysporum. Queensland Journal of Agricultural<br />

and Animal Sciences 22, 475–479.<br />

Session 2B—Soilborne disease<br />

Of the 500 root and crown specimens, Fusarium spp. was<br />

consistently isolated from diseased plants. Fusarium isolates<br />

used were pathogenic on the strawberry runners but nonpathogenic<br />

on the tomato and cucumber plants tested. On the<br />

basis of morphological characteristics and pathogenicity tests, F.<br />

oxysporum f. sp. fragariae was identified as a main component<br />

of root and crown rots of strawberries. Average percentage<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 41


Session 2B—Soilborne disease<br />

Evaluation of resistant rootstocks for control of Fusarium wilt of watermelon in Nghe<br />

An Province, Vietnam<br />

V.T. Dau A , N.V. Dang B , D.H. Nguyen A , L.T. Pham A , T.T.M. Le A , H.T. Phan C , L.W. Burgess{ XE "Burgess. L.W." } D<br />

A Nghe An <strong>Plant</strong> Protection Sub‐Department, 19 Tran Phu St., Vinh, Nghe An, Vietnam<br />

B<br />

Syngenta, 16 3A St., Bien Hoa 2 Industrial Area, Dong Nai, Vietnam<br />

C National Institute of Medicinal Materials, 3B Quang Trung St, Hanoi, Vietnam<br />

D Faculty of Agriculture, Food and Natural Resources, The University of Sydney, Sydney, 2006, New South Wales, Australia<br />

INTRODUCTION<br />

Fusarium wilt of watermelon, caused by Fusarium oxysporum<br />

f.sp. niveum, has caused serious losses in watermelon crops in<br />

Nghia Dan District, Nghe An Province, Vietnam over the past two<br />

years, and is threatening the viability of the industry (1). We<br />

demonstrated that the local watermelon cultivar could be<br />

grafted successfully onto a resistant rootstock in 2008. In this<br />

paper we report on the evaluation of a range of potential<br />

resistant rootstocks in relation to graft compatibility and<br />

watermelon production and quality, and the assessment of three<br />

rootstock‐scion combinations under commercial conditions.<br />

MATERIALS, METHODS AND RESULTS<br />

We evaluated the local cultivar Bau trang (Lagenaria sinceraria)<br />

together with five hybrid cucurbits for use as resistant<br />

rootstocks. The hybrids, provided by Syngenta, Thailand, were<br />

Kazako, Carnivar and Bulrojangsaeng (S. sinceraria), and<br />

Emphasis and Argentaria (pumpkin). The simple grafting<br />

technique as refined by Syngenta, is efficient and can be taught<br />

quickly to farmers. The three basic tools used by our cooperating<br />

farmers are a razor blade, and home‐made scalpel and stylet<br />

(Figure 1).<br />

Figure 2. Successful graft of watermelon on resistant rootstock (A) and<br />

graft junction on mature plant (B).<br />

The Bulrojangsaeng hybrid was the best rootstock on all criteria<br />

but the farmers consider it too expensive. Consequently the Bau<br />

trang cultivar was used as a resistant rootstock in an initial 0.1ha<br />

field trial in soil with a history of severe watermelon wilt, in<br />

spring 2009. Two watermelon cultivars from Syngenta, Thuy Loi<br />

and Phu Dong, were used as scions. The grafted plants were not<br />

affected by Fusarium wilt and produced marketable fruit.<br />

Staff from Nghe An <strong>Plant</strong> Protection Sub‐Department and<br />

Syngenta held farmer field schools on Fusarium wilt and the<br />

grafting technique in March, 2009. A cooperating farmer planted<br />

an eight‐hectare summer crop of grafted seedlings in late April,<br />

2009. He used the Bau trang rootstock and two watermelon<br />

cultivars in his trial, Hac My Nhan and Dat Viet (Nong Viet<br />

Company), selections based on cost of seed. The results of this<br />

planting will be reported.<br />

Figure 1. Basic tools used for grafting watermelon onto Fusarium wilt<br />

resistant rootstock. Home made scalpel (top), stylet (middle) and razor<br />

blade (bottom)<br />

The grafting process involves excising the stem of the rootstock<br />

seedling immediately above the cotyledons, at the first true leaf<br />

stage. A narrow cone shaped slot is then made in the stem with<br />

the stylet. The upper part of the stem of a watermelon seedling<br />

(0.8 to 1.0cm long), at the cotyledon stage, is then removed with<br />

a razor blade, making an oblique transfer cut. The scion is then<br />

inserted in the slot in the rootstock stem. The grafting process<br />

takes about 10 to 15 seconds. The success rate is approximately<br />

80% or greater depending on experience. A successful graft is<br />

shown in Figure 2.<br />

DISCUSSION<br />

The use of watermelon grafted onto resistant rootstocks has<br />

been shown to be a successful approach for the prevention of<br />

Fusarium wilt in soils with high inoculum levels. The cost of seed<br />

in relation to yield, fruit quality and market acceptance in Hanoi<br />

will determine the preferred combinations of scions and<br />

rootstocks. The widespread use of tomatoes grafted onto<br />

resistant rootstocks for control of bacterial wilt in Vietnam (2)<br />

indicates the potential for using resistant rootstocks to prevent<br />

losses from a range of diseases caused by soil‐borne pathogens.<br />

ACKNOWLEDGEMENTS<br />

Financial support from the Australian Centre for International<br />

Agricultural Research (ACIAR CP/2002/115) is gratefully<br />

acknowledged.<br />

REFERENCES<br />

1. Dau VT, Burgess LW, Pham LT, Phan HT, Nguyen HD, Le TV, Nguyen<br />

DH (2009) First report of Fusarium wilt in watermelon in Vietnam.<br />

<strong>Australasian</strong> <strong>Plant</strong> Disease Notes 4, 1–3.<br />

42 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


INTRODUCTION<br />

Bunch rot diseases and their management<br />

T.B. Sutton{ XE "Sutton, T.B." }, J. Miranda Longland, K. Whitten Buxton, O. Anas<br />

Department of <strong>Plant</strong> <strong>Pathology</strong>, NC State University, Raleigh, NC 27695 USA<br />

Bunch rot diseases are of concern in all regions of the world<br />

where grapes are grown but nowhere are they more<br />

problematic than in moist and temperate growing regions.<br />

Among the most important bunch rot diseases in the<br />

southeastern United States are black rot, phomopsis, botrytis,<br />

bitter rot, ripe rot, and sour rot. This paper summarises a series<br />

of studies designed to gain a better understanding of the biology<br />

and epidemiology of the pathogens that cause bitter rot<br />

[Greeneria uvicola (Berk. & Curtis) Punith.] and ripe rot<br />

[homothallic and heterothallic strains of Colletotrichum<br />

gloeosporioides (Penz.) Penz. & Sacc. (teleomorph Glomerella<br />

cingulata (Stonem.) Spauld. & Schrenk) and C. acutatum J.H.<br />

Simmonds (teleomorph G. acutata Guerber & Correll)] and how<br />

to manage them.<br />

BIOLOGY AND EPIDEMIOLOGY OF G. UVICOLA AND<br />

COLLETOTRICHUM SPP.<br />

Fruit susceptibility to G. uvicola and Colletotrichum spp.<br />

through the growing season. From 2003–2007 a series of<br />

studies were conducted to determine the period during the<br />

growing season that fruit were susceptible to infection by G.<br />

uvicola and homothallic isolates of C. gloeosporioides. Fruit of<br />

Chardonnay, Merlot, and Cabernet Franc were susceptible to<br />

infection by G. uvicola from bloom until harvest but were most<br />

susceptible just prior to and during véraison. Although fruit were<br />

susceptible to C. gloeosporioides from bloom to harvest, the<br />

period of peak susceptibility was not as obvious. Chardonnay<br />

was most susceptible at bloom and véraison, Seyval blanc during<br />

bloom, post‐bloom, and véraison, and Cabernet Franc at postbloom,<br />

closing, véraison, and preharvest.<br />

Influence of temperature and leaf wetness on infection of fruit<br />

by G. uvicola. Detached fruit of V. vinifera (cv Chardonnay,<br />

Cabernet Franc, and Cabernet Sauvignon) were atomised with a<br />

conidial suspension of G. uvicola (10 5 spores/ml), placed in moist<br />

chambers and subjected to 14, 18, 22, 26, and 30°C for 6, 12, 18,<br />

or 24 h of wetting. Optimum conditions for infection were 6 or<br />

12 h of wetting and temperatures ranging from 22.4 to 24.6°C<br />

(mean=23.3°C) (1).<br />

Colletotrichum species in the diverse geographical regions of<br />

North Carolina and to examining differences based on host<br />

genera. The relative proportion of Colletotrichum spp. in a<br />

vineyard varied with the cultivar/species present and location<br />

(2).<br />

MANAGEMENT<br />

Effect of cane pruning. An experiment was conducted from<br />

2004–2006 on the cultivars Chardonnay, Merlot, and Cabernet<br />

Sauvignon in a 12‐year‐old vineyard in the Piedmont of North<br />

Carolina designed to evaluate the effect of cane pruning vs spur<br />

pruning on the incidence and severity of bitter rot and ripe rot.<br />

Cordons 1, 2, and 3‐years‐old were established during the course<br />

of the experiment and the incidence and severity of bitter rot,<br />

ripe rot and sour rot was compared to the 12‐year‐old spur<br />

pruned vines There was a significant reduction in the incidence<br />

and severity of bitter rot and ripe rot in the first year which<br />

carried over during the 3 years of the study. Cane pruning did<br />

not reduce the incidence of sour rot, but did reduce its severity.<br />

The disease management program for ripe rot and bitter rot in<br />

the southeastern US. The backbone of the disease management<br />

program in the southeastern US is cultural practices which are<br />

designed to reduce the initial inoculum and create an<br />

environment within the vine canopy less favorable for disease<br />

development. However, a fungicide program beginning at bloom<br />

and continuing until harvest is necessary to effectively manage<br />

summer bunch rot diseases. Captan, applied on a 10–14 day<br />

interval is the principle fungicide used from closing to harvest.<br />

REFERENCES<br />

1. Miranda, J.G. and Sutton, T.B. 2008. Factors affecting the infection<br />

of fruit of Vitis vinifera by the bitter rot pathogen, Greeneria<br />

uvicola. Phytopathology 98: 580–584.<br />

2. Whitten Buxton, K.R. and Sutton, T.B. 2008. Biology and<br />

epidemiology of Colletotrichum species associated with ripe rot of<br />

grapes. Phytopathology 98:S170.<br />

Session 2C—Epidemiology<br />

Relative susceptibility of cultivars to G. uvicola and<br />

Colletotrichum spp. Fruit of 38 grape cultivars or selections were<br />

evaluated for their susceptibility to G. uvicola while 35 cultivars<br />

or selections were tested for their susceptibility to a homothallic<br />

isolate of C. gloeosporioides in a laboratory assay<br />

There was a wide variation in the susceptibility of fruit to G.<br />

uvicola. Fruit of V. vinifera were more susceptible than the<br />

French American hybrids. V. aestavalis Cynthiana was the most<br />

resistant to G. uvicola (1). There was also a wide range in the<br />

susceptibility of cultivars and selections to C. gloeosporioides;<br />

however; there was not a clear difference between the relative<br />

susceptibility of V. vinifera and hybrids to C. gloeosporioides.<br />

Early harvest Cynthiana, Chardonnay, Merlot, Petit Syrah, and<br />

Pride were among the most susceptible cultivars to C.<br />

gloeosporioides.<br />

Population structure of Colletotrichum spp.associated with ripe<br />

rot of grapes. This study was conducted from 2004–2007 with<br />

the objectives of determining the population structure of<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 43


Session 2C—Epidemiology<br />

Inoculum and climatic factors driving epidemics of Botrytis cinerea in New Zealand<br />

and Australian vineyards<br />

R.M. Beresford{ XE "Beresford, R.M." } A , G.N. Hill A , K.J. Evans B , J. Edwards C , D. Riches C , P.N. Wood D and D.C. Mundy E<br />

A The New Zealand Institute for <strong>Plant</strong> & Food Research Limited (<strong>Plant</strong> & Food Research), Private Bag 92 169, Auckland 1142, New Zealand<br />

B Tasmanian Institute of Agr. Research, University of Tasmania, New Town Res. Laboratories, 13 St Johns Avenue, New Town, Tasmania<br />

7008<br />

C Department of Primary Industries, 621 Burwood Highway, Knoxfield, Victoria 3180<br />

D <strong>Plant</strong> & Food Research, Hawke’s Bay Research Centre, Private Bag 1401, Havelock North, Hastings 4157, New Zealand<br />

E <strong>Plant</strong> & Food Research, Marlborough Wine Research Centre, P.O. Box 845, Blenheim 7240, New Zealand<br />

INTRODUCTION<br />

Botrytis bunch rot (botrytis) reduces grape yield and wine quality<br />

in seasons when wet weather occurs during grape ripening.<br />

Grape growers need to be able to predict when there is a high<br />

risk of botrytis so that fungicide applications, vine canopy<br />

management and harvesting operations can be more effectively<br />

planned. This study investigated climatic predictors of harvest<br />

botrytis severity that were measured in non‐fungicide treated<br />

plots between key vine growth stages: 1) early season (flowering<br />

to pre‐bunch closure (PBC)), 2) mid season (PBC to beginning of<br />

ripening (veraison)), 3) late season (veraison to harvest).<br />

MATERIALS AND METHODS<br />

Weather data came from weather stations within 1–5 km of<br />

each vineyard trial site (1). Relationships between harvest<br />

botrytis severity (2) and rainfall, daily mean, minimum and<br />

maximum temperature and surface wetness duration were<br />

investigated. Surface wetness duration and temperature during<br />

wetness were summarised using the “Bacchus” model (3) from<br />

Hortplus TM (www.hortplus.com).<br />

RESULTS AND DISCUSSION<br />

There were strong regional associations between harvest<br />

botrytis severity and some climatic variables during some growth<br />

stage intervals (Figures 1–3). The figures show the 3% severity<br />

threshold at which wineries may impose price penalties for<br />

botrytis‐affected grapes. Amount of rainfall, even late in the<br />

growing season, was, surprisingly, a poor predictor of harvest<br />

botrytis severity. Further research is using the relationships<br />

found in this study, together with fungicide and vine<br />

management factors, to develop a botrytis risk prediction model.<br />

ACKNOWLEDGEMENTS<br />

This study was funded by New Zealand Winegrowers, the<br />

Australian Grape and Wine Research and Development<br />

Corporation, the N.Z. Foundation for Research, Science and<br />

Technology (C06X0810) and <strong>Plant</strong> and Food Research.<br />

REFERENCES<br />

1. Beresford RM, Hill GN 2008. Predicting in‐season risk of botrytis<br />

bunch rot in Australian and New Zealand vineyards. In: Breaking<br />

the mould—a pest and disease update. Proceedings of the ASVO<br />

seminar, Mildura, Victoria. Australian <strong>Society</strong> of Viticulture and<br />

Oenology Inc.: Adelaide, South Australia: 24–28.<br />

2 Beresford RM, Evans KJ, Wood PN, Mundy DC 2006. Disease<br />

assessment and epidemic monitoring methodology for bunch rot<br />

(Botrytis cinerea) in grapevines. New Zealand <strong>Plant</strong> Protection 59:<br />

355–360.<br />

3. Kim KS, Beresford RM, Henshall WR 2007. Prediction of disease risk<br />

using site‐specific estimates of weather variables. New Zealand<br />

<strong>Plant</strong> Protection 60: 128–132.<br />

Harvest botrytis severity (%)<br />

20<br />

15<br />

10<br />

5<br />

Auckland<br />

Hawke’s Bay<br />

Marlborough<br />

Tasmania<br />

Yarra Valley<br />

Bacchus index at<br />

3% severity= 0.57<br />

y = 38.287x - 19.009<br />

R 2 = 0.74<br />

0<br />

0.3 0.5 0.7 0.9<br />

Mean Bacchus index, early-season<br />

Figure 1. “Bacchus” index (surface wetness duration and temperature)<br />

was the most useful early‐season indicator of harvest botrytis severity.<br />

Harvest botrytis severity (%)<br />

20<br />

15<br />

10<br />

5<br />

y = 7E+07e -0.7139x<br />

R 2 = 0.81<br />

Auckland<br />

Hawke’s Bay<br />

Marlborough<br />

Tasmania<br />

Yarra Valley<br />

Mean max.temp. at<br />

3% severity= 23.8 o C<br />

0<br />

20 22 24 26<br />

Mean max. temp., early-season<br />

Figure 2. Mean daily maximum air temperature in the early part of the<br />

season was strongly inversely related to harvest botrytis severity.<br />

Harvest botrytis severity (%)<br />

20<br />

15<br />

10<br />

5<br />

Auckland<br />

Hawke’s Bay<br />

Marlborough<br />

Tasmania<br />

Yarra Valley<br />

Mean interval at<br />

3% severity= 40.8 days<br />

y = 0.005e 0.1567x<br />

R 2 = 0.87<br />

0<br />

0 20 40 60 80<br />

Mean late-season interval (days)<br />

Figure 3. A longer ripening period (late‐season interval) was associated<br />

with greater harvest botrytis severity. This regional trend is often<br />

reflected in individual vineyards, where a delay in harvest date to<br />

achieve higher grape sugar content is accompanied by increased risk of<br />

botrytis.<br />

44 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Infection of apples by Colletotrichum acutatum in New Zealand is limited by<br />

temperature<br />

K.R. Everett{ XE "Everett, K.R." }, O.E. Timudo‐Torrevilla, I.P.S. Pushparajah, R.W.A. Scheper, P.W. Shaw, T.M. Spiers, A. Ah Chee, J.T.<br />

Taylor, P. Wood, D.R. Wallis, M.A. Manning<br />

The New Zealand Institute for <strong>Plant</strong> and Food Research Limited, P.B. 92169, Mt Albert, Auckland, New Zealand<br />

INTRODUCTION<br />

Colletotrichum acutatum infects the surface of apples in New<br />

Zealand to cause small, 1–2 mm diameter dark spots, usually on<br />

the side of the fruit exposed to the sun, which can enlarge to<br />

cover the entire fruit surface with one or several orange,<br />

sporulating lesions. Infection eventually results in fruit drop.<br />

Symptoms express in summer. Little is known of the<br />

epidemiology of C. acutatum infecting apples, either in New<br />

Zealand or elsewhere (1). A study was conducted to investigate<br />

the epidemiology of this fungus on apples in New Zealand, to<br />

facilitate the design of strategies to achieve control with no<br />

residues.<br />

MATERIALS AND METHODS<br />

Laboratory inoculations. ‘Royal Gala’ apples were harvested<br />

monthly starting immediately after fruit set in November 2005<br />

until harvest in 2006. At each time, apples were woundinoculated<br />

with 10 6 spores/ml C. acutatum and placed at 5, 10,<br />

15, 17.5, 20, 25 or 30°C in humid conditions for 7 days. Lesion<br />

diameter was then measured.<br />

Field inoculations. ‘Royal Gala’ apples in two orchards in each of<br />

three major apple growing regions in New Zealand, viz. Waikato,<br />

Hawke’s Bay and Nelson, were wound‐inoculated monthly with<br />

10 6 spores/ml C. acutatum beginning in November 2005 until<br />

harvest in February 2006. Lesion diameter was then measured.<br />

Lesion diameter (mm)<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

November<br />

December<br />

January<br />

February<br />

5 10 15 20 25 30<br />

Temperature o (C)<br />

Figure 1. Mean lesion diameter (mm) of detached apple fruit inoculated<br />

at harvest with 10 6 spores/ml Colletotrichum acutatum. A Boltzmann<br />

plot was fitted.<br />

RESULTS<br />

Detached ‘Royal Gala’ apples were susceptible to infection by C.<br />

acutatum when a temperature of c. 15°C was exceeded,<br />

regardless of maturity (Fig. 1). A wetness period of 72 hours was<br />

required for infections without wounding (results not shown). In<br />

the field ‘Royal Gala’ apples were infected by C. acutatum after a<br />

temperature of 15.2°C was exceeded (Fig. 2).<br />

mean number of infected fruit (max. = 5)<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

Y = 3.1 + (0.02 - 3.1)/(1+exp ((x-15.2/0.01))<br />

R 2 = 73%<br />

12 13 14 15 16 17 18 19 20 21<br />

mean daily temperature ( o C)<br />

Figure 2. The mean number of infected apple fruit out of five inoculated<br />

with 10 6 spores/ml of Colletotrichum acutatum plotted against the mean<br />

daily temperature for 72 hours after inoculation. A Boltzmann plot was<br />

fitted. The x0 value (inflection point) = 15.2°C.<br />

DISCUSSION<br />

In New Zealand mean daily temperatures of 15°C are exceeded<br />

during December, January and February. Effective control<br />

without residues may be able to be achieved by reducing<br />

inoculum early in the season before temperatures are above<br />

mean daily termperatures of c. 15°C, as was achieved for control<br />

of B. dothidea (2). A biological control agent or a benign<br />

chemical such as calcium chloride (3) could be used to protect<br />

the fruit from infection during summer when temperature is not<br />

limiting infections.<br />

ACKNOWLEDGEMENTS<br />

This work was funded by MAF Sustainable Farming Fund, Pipfruit<br />

New Zealand, Waikato Fruitgrowers’ Association and Nelson<br />

Group 8.<br />

REFERENCES<br />

1. Peres NA, Timmer LW, Adaskaveg JE, Correll JC (2005) Lifestyles of<br />

Colletotrichum acutatum. <strong>Plant</strong> Disease 89, 784–96.<br />

2. Everett KR, Timudo‐Torrevilla OE, Taylor JT and Yu J (2007)<br />

Fungicide timing for control of summer rots of apples. New Zealand<br />

<strong>Plant</strong> Protection 60, 15–20.<br />

3. Boyd‐Wilson KSH and Walter M (2007) Effect of nutrients on the<br />

biocontrol activity of yeasts on apple pathogens. IN: Conference<br />

Handbook, 16th Biennial Conference of <strong>Australasian</strong> <strong>Plant</strong><br />

<strong>Pathology</strong> <strong>Society</strong>. P. 43.<br />

Session 2C—Epidemiology<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 45


Session 2C—Epidemiology<br />

Epidemiology of walnut blight, caused by Xanthomonas arboricola pv. juglandis, in<br />

Tasmania, Australia<br />

M.D. Lang A , K.J. Evans{ XE "Evans, K.J." } B and S.J. Pethybridge C<br />

A<br />

Tasmanian Institute of Agricultural Research (TIAR), University of Tasmania (UTAS), P.O. Box 3523, Burnie, 7320, TAS, Australia<br />

B<br />

TIAR, UTAS, 13 St Johns Ave, New Town, 7008, TAS, Australia<br />

C<br />

Botanical Resources Australia—Agricultural Services Pty Ltd., 44–46 Industrial Drive, Ulverstone, 7315, TAS, Australia<br />

INTRODUCTION<br />

Walnut blight is an important bacterial disease of walnuts worldwide,<br />

and can cause the premature drop of almost all fruit in<br />

some Australian orchards (1). In glasshouse trials, only five<br />

minutes of continuous wetness on the fruit surface is necessary<br />

for infection by Xanthomonas arboricola pv. juglandis (2), and<br />

extensive wetness periods may explain polycyclic disease<br />

epidemics in California (3). Objectives of this study were to<br />

characterise the temporal progression of disease incidence and<br />

severity in walnut fruit in Tasmania, Australia, and the<br />

relationship between disease severity and yield of marketable<br />

nuts.<br />

Table 1. Linear models of the temporal progression of walnut blight<br />

incidence in Vina and Franquette in Tasmania.<br />

Year Model R 2 A Intercept Slope Day 50% B<br />

Vina<br />

2004-05 Monomolecular 0.9677 -0.3203 0.0121 84<br />

2005-06 Logistic 0.9922 -10.0603 0.1926 52<br />

2006-07 Monomolecular 0.9886 -0.0655 0.0018 -<br />

Franquette<br />

2005-06 Gompertz 0.9926 -3.5383 0.0771 51<br />

2006-07 Monomolecular 0.9946 -0.1298 0.0042 -<br />

A Back transformed R 2 ; B Days (predicted) to 50% disease incidence.<br />

MATERIALS AND METHODS<br />

Randomised, complete block trials were conducted on single<br />

tree plots (n = 6) of Vina and Franquette over three and two<br />

growing seasons, respectively, in northern Tasmania. On nontreated<br />

trees, disease incidence and severity was assessed on<br />

the same 100 fruits from fruit set to harvest, or until a fruit fell.<br />

Severity was estimated visually as the perc ent area of the fruit<br />

covered with blight symptoms. Incidence was derived from the<br />

severity data as the percentage of fruit with symptoms. Daily<br />

cumulative disease incidence was analyzed as a function of time<br />

with linear, monomolecular, exponential, logistic and Gompertz<br />

models. Crop yield in non‐treated plots was defined as the<br />

percentage of the 100 fruit present at fruit set that produced<br />

marketable nuts at harvest.<br />

RESULTS AND DISCUSSION<br />

The time of disease onset was similar for each epidemic whereas<br />

the rate of disease increase was highest in 2005–06 with nearly<br />

100% incidence within 80 and 120 days of bud‐burst in Vina and<br />

Franquette respectively (Fig. 1). Temporal progression of disease<br />

incidence was described well by the monomolecular model in<br />

2004–05 and 2006–07, implying monocyclic disease epidemics<br />

(Table 1). In contrast, the logistic and Gompertz models<br />

described putative polycyclic disease in Vina and Franquette,<br />

respectively, in the relatively wet season of 2005–06. Disease<br />

severity of fruits at half full‐size diameter accounted for 74% of<br />

the variation in crop yield (Fig. 2). No marketable nuts were<br />

predicted when the mean blight severity on fruit was 10%. In<br />

contrast, 87% of fruits were predicted to produce marketable<br />

nuts when the mean blight severity was 2%.<br />

Incidence (%)<br />

Vina<br />

Franquette<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

20 40 60 80 100 120 140 160 20 40 60 80 100 120 140 160<br />

Days from bud-burst<br />

2004-05<br />

2005-06<br />

2006-07<br />

Yield (%)<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

y = -11.019x + 109.03<br />

R 2 = 0.74<br />

0.0 2.0 4.0 6.0 8.0 10.0<br />

Severity (%)<br />

Figure 2. Yield per plot of Vina nuts at harvest, expressed as a<br />

percentage of 100 fruit at fruit set, and disease severity observed on<br />

fruits at half full‐size diameter.<br />

DISCUSSION<br />

In Tasmania, epidemics of walnut blight appear to be either<br />

monocyclic or polycyclic, with polycyclic epidemics leading to all<br />

fruits developing disease symptoms. Preventing the onset and<br />

rate of increase of disease incidence and severity of fruits prior<br />

to fruits attaining half full‐size diameter appears critical to<br />

reducing crop loss. An empirical, weather‐based model for<br />

timing applications of copper is being developed.<br />

ACKNOWLEDGEMENTS<br />

This project was supported by the Australian Government<br />

through Horticulture Australia Limited in partnership with<br />

Webster Walnuts, and was managed by Agronico P/L., 175<br />

Allport St. East, Leith, TAS, 7315 www.agronico.com.au<br />

REFERENCES<br />

1. Lang, M.D., Hills, J.L. and Evans, K.J. (2006). Preliminary studies<br />

towards managing walnut blight in Tasmania. Acta Hort. 705: 451–<br />

456.<br />

2. Miller, P.W. and Bollen, W.B. (1946). Walnut bacteriosis and its<br />

control: Technical Bulletin No. 9, United States Department of<br />

Agriculture, Oregon State College, Corvallis, USA.<br />

3. Adaskaveg, J.E., Forster, H., Dieguez‐Uribeondo, J., Thompson, D.,<br />

Adams, C.J., Buchner, R. and Olsen, B. (2000). Epidemiology and<br />

management of walnut blight. See<br />

http://walnutresearch.ucdavis.edu/2000/2000_329.pdf<br />

Figure 1. Temporal progression of disease incidence observed on Vina<br />

and Franquette fruits in Tasmania.<br />

46 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


INTRODUCTION<br />

Sugarcane smut—disease development and mechanism of resistance<br />

S.A. Bhuiyan{ XE "Bhuiyan, S.A." } A , B.J. Croft B , R.C. James A , G. Bade A , and M.C. Cox A<br />

A BSES Limited, Private Bag 4, Bundaberg DC, QLD 4670, Queensland<br />

B BSES Limited, 90 Old Cove Road, Woodford, QLD 4514, Queensland<br />

Sugarcane smut caused by Ustilago scitaminea is an important<br />

disease worldwide. This disease can be effectively managed by<br />

replacing susceptible cultivars with resistant cultivars. The<br />

current smut rating system relies on artificial inoculation and a<br />

short plant and ratoon crop Ratings are based on % infected<br />

plants, not taking into account the severity of the disease.<br />

Infection of smut occurs through germinating or dormant buds<br />

of standing stalks or through germinating buds in the soil (1).<br />

Available literature suggests that there are two mechanisms of<br />

resistance to smut: i) bud scale or external resistance; and ii)<br />

internal resistance (2). Bud scale resistance is believed to be a<br />

combination of physical and chemical barriers to infection, and<br />

internal resistance is governed by interactions within plant<br />

tissue. It is also suggested that buds become increasingly<br />

resistant with age.<br />

The objectives of the study were to: i) monitor development of<br />

sugarcane smut; ii) compare disease incidence and severity; and<br />

iii) determine mechanisms of disease resistance of important<br />

sugarcane cultivars grown in Australia.<br />

MATERIALS AND METHODS<br />

Disease development. Twenty‐nine commercial cultivars were<br />

collected from various regions of Queensland. They were cut<br />

into one‐eye‐setts, inoculated by dipping in a spore suspension<br />

(10 6 spores/mL) and planted in September 2008 in 3 replications<br />

using a randomised complete block design. Disease incidence<br />

and severity were measured each month after planting.<br />

disease severity rating<br />

100<br />

80<br />

60<br />

40<br />

20<br />

y = 0.794x - 0.7452<br />

R 2 = 0.9199<br />

P


Session 2D—Disease management<br />

Dissemination of biological and chemical fungicides by bees onto Rubus and Ribes<br />

flowers<br />

M. Walter{ XE "Walter, M." } 1 , F.O. Obanor 2 , B. Donovan 3 , K.S.H. Boyd‐Wilson 1 , M. Neal 1 , H.J. Siefkes‐Boer 1 and G.I. Langford 1<br />

1 The New Zealand Institute for <strong>Plant</strong> and Food Research Limited, Private Bag 4704, Christchurch Mail Centre, Christchurch 8140, New<br />

Zealand<br />

2 CSIRO, <strong>Plant</strong> Industry, Queensland Bioscience Precinct, 306 Carmody Rd, St Lucia, QLD 4067, Australia<br />

3 DSIR, Private Bag 4704, Christchurch Mail Centre, Christchurch 8140, New Zealand<br />

INTRODUCTION<br />

Honey and bumble bees have shown to be capable of<br />

disseminating biological fungicides (BF) to achieve inoculation of<br />

berry flowers, such as strawberry, blueberry and blueberries.<br />

This in turn has caused a reduction in flower‐borne diseases. The<br />

aim of our work was to study the efficacy of a BF (year Y1) and a<br />

chemical fungicide (CF) (Y2) disseminated by honey bees and<br />

bumble bees (Y2 only) onto boysenberry (Rubus hybrid) and<br />

blackcurrant (Ribes nigrum) flowers.<br />

MATERIALS AND METHODS<br />

• BF: Sentinel® (Trichoderma atroviride)<br />

• CF: Switch® (fludioxonil+cyprodinil) finely ground and<br />

mixed with fluorescent dye (1:1, v/v)<br />

• Honey bees (Apis mellifera) and bumble bees (Bombus<br />

terrestris) at approx. 25000 and 250 bees/hive, respectively<br />

were used.<br />

• All experiments were conducted on commercial<br />

boysenberry and blackcurrant fields, with 2–4 replicate<br />

grower sites per crop and product.<br />

• The study was conducted separately for BF and CF over two<br />

flowering seasons. In Y1, honey bee and BF dissemination<br />

was examined. In Y2, honey and bumble bee CF<br />

disseminations were monitored, including CF residues in<br />

honey.<br />

In boysenberry, 2% and 1.3% of the flowers were visited by<br />

honey bees and bumble bees respectively, carrying<br />

Switch®+dye. The number of total honey bees (46) and bumble<br />

bees (0.7) observed was 0.5 honey and 0.08 bumble bees<br />

foraging on one boysenberry plant, with a daily average of 70–80<br />

open flowers/plant available. Bumble bee hive activity was,<br />

however, very low, with 0.05 bumble bees leaving the hive/15<br />

minutes. Approximately 5% of honey bees exposed to CF+dye<br />

vectored the fungicide to 2–4% of flowers. The product mixture<br />

probably was only available during approximately 15% of the<br />

actual foraging time for those bees. We can conclude that honey<br />

bees can vector biological and chemical fungicides to<br />

boysenberry flowers, however, in the level of disease control is<br />

not yet established. The role of bumble bees in boysenberry<br />

gardens is inconclusive, as the hive activities were very low.<br />

Switch® residues in honey from hives fitted with fungicide<br />

dispensers collected immediately after flowering were up to 5<br />

mg/kg active ingredient. No fungicide residues were measured in<br />

honey samples from adjacent hives, but without CF dispensers.<br />

ACKNOWLEDGEMENTS<br />

The work was funded by The New Zealand Boysenberry Council<br />

Ltd, Blackcurrant New Zealand Ltd and the Sustainable Farming<br />

Fund (SFF) of the Ministry of Agriculture and Forestry (MAF)<br />

Grant 06/007.<br />

RESULTS AND DISCUSSION<br />

Y1: Application efficacy (% flowers inoculated) of bee‐applied<br />

Sentinel® was similar to spray‐applied BF. Irrespective of<br />

application method in boysenberry, 60–80% of flowers were<br />

inoculated; in blackcurrant, 40–50%. This resulted in >160<br />

Trichoderma colony forming units or cfu/boysenberry ovary and<br />

3–4 cfu/blackcurrant style. Bee application could be maximised<br />

by either equipping all hives with dispensing units and/or by<br />

increasing the release rate from three per week to daily releases.<br />

The BF, bee‐ or spray‐applied, did not improve disease control.<br />

Therefore in Y2, a CF was selected.<br />

Y2: CF or dye recovery on blackcurrant flowers was in the order<br />

of 5–14% for honey bees and 5–9% for bumble bees. The<br />

average number insects foraging per 10‐m row was 5 honey bees<br />

and 0.1 bumble bees. The average number of bees exiting the<br />

experimental hives during 1 and 5 min of observation was 300<br />

honey bees and 3.4 bumble bees, respectively. Bumble bees<br />

visited similar numbers of flowers as the honey bees, although<br />

there were 50 times more honey bees active in the crop than<br />

bumble bees. In addition, there were only 5 bumble bee hives in<br />

the blackcurrant field (>5 ha) compared with 14 honey bee hives<br />

for the same area (only two hives were equipped with the<br />

dispensers).<br />

48 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Current studies on divergence and management of pepper yellow leaf curl disease in<br />

Indonesia<br />

INTRODUCTION<br />

Sri Hendrastuti Hidayat{ XE "Hidayat, S.H." } 1 , Sri Sulandari 2 , and Sriani Sujiprihati 1<br />

1 Bogor Agricultural University, Campus Darmaga, Bogor 16680, Indonesia<br />

2 Gadjah Mada University, Yogyakarta 55281, Indonesia<br />

Whitefly‐transmitted geminiviruses (WTGs) has been reported to<br />

infect several crops in Indonesia including tobacco, tomato, chilli<br />

pepper, ageratum, and cucumber. Infection of WTGs in chilli<br />

pepper causes severe crop damage and becoming a major threat<br />

since early 2000. The most unique symptoms associated with the<br />

virus infection involved yellowing and leaf curling, therefore it<br />

was known as pepper yellow leaf curl (PYLC) disease. Biological<br />

and molecular characterisation of the causal agent reveals that<br />

several WTGs are associated with the disease. Disease spread<br />

was very fast due to activity of its insect vector, Bemisia tabaci,<br />

which grows very prominently in the tropic climate. Therefore,<br />

disease control is becoming very difficult. Breeding program for<br />

WTGs resistance varieties is one of major activities in regard to<br />

disease control strategy in Indonesia since commercial cultivars<br />

carrying resistance to the diseases have not yet been released.<br />

Evaluation of chilli pepper genotypes showed that some<br />

germplasms are very promising for development of cultivars<br />

with resistance or tolerance to the disease.<br />

MATERIALS AND METHODS<br />

Analysis of Genetic Diversity. Pepper plant showing typical<br />

symptoms of PYLCV infection were collected from several chilli<br />

pepper production areas in Indonesia. Extraction of total DNA<br />

and PCR amplification was done according to procedure<br />

explained previously (1, 2). Sequence data obtained following<br />

nucleotide sequencing of the PCR product was analysed using<br />

ClustalW program version 1.83 EMBL‐EBI.<br />

Evaluation of 11 commercial cultivars and 27 genotypes of chilli<br />

pepper showed that the symptoms were developed within 2 to 3<br />

weeks after inoculation, although some genotypes required<br />

longer incubation period. Disease incidence was varied among<br />

different genotypes, i.e. in the range of 12 up to 100%. Selection<br />

of potential genotypes was proceeded for further breeding<br />

activity in order to develop resistant varieties for PYLCV.<br />

ACKNOWLEDGEMENTS<br />

This research was supported in part by ACIAR—AVRDC Chilli<br />

Integrated Disease Management Project.<br />

REFERENCES<br />

1. Doyle JJ, Doyle JJ (1999) Isolation of plant DNA from fresh tissue.<br />

Focus 12, 13–15.<br />

2. Rojas MR, Gilbertson RL. Russel DR, Maxwell DP (1993) Use of<br />

degenerate primers in the polymerase chain reaction to detect<br />

whitefly‐transmitted geminiviruses. <strong>Plant</strong> Disease 77, 340–347.<br />

3. Aidawati N, Hidayat SH, Suseno R, Sosromarsono s (2002)<br />

Transmission of an Indonesian isolate of tobacco leaf curl virus by<br />

Bemisia tabaci Genn. (Hemiptera:Aleyrodidae). <strong>Plant</strong> Pathol J 18,<br />

231–236.<br />

4. Ikegami M, Morinaga T, Miura K (1988) Potential gene product of<br />

bean golden mosaic virus have higher sequence homologies to<br />

those of tomato golden mosaic virus than those of cassava laten<br />

virus. Virus Genes 1,191–203.<br />

Session 2D—Disease management<br />

Evaluation of Chilli Pepper Genotypes. Inoculation of PYLCV by<br />

B. tabaci was conducted as explained previously (3). Response of<br />

different chilli pepper genotypes was classified into three groups<br />

i.e. resistant, moderately resistant, and susceptible based on<br />

symptoms expression and disease incidence.<br />

RESULTS AND DISCUSSION<br />

Identity of geminivirus infecting chilli pepper in Indonesia was<br />

determined based on their hairpin loop structure and repetitive<br />

sequence found in the common region. These hairpin loop<br />

structure was found in all geminivirus sequences so far (4).<br />

Variability in the structure as well as the length of hairpin loop<br />

region was observed among PYLCV isolates. This may indicate<br />

the possible genetic diversity among WTGs infecting chilli pepper<br />

in Indonesia. Phylogenetic analyses involving 32 sequences<br />

showed that PYLCV isolates can be differentiated into several<br />

clusters. Interestingly, they are all quite different from WTGs<br />

infecting tomato in Indonesia but more closely related to tomato<br />

yellow leaf curl virus from Thailand.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 49


Session 2D—Disease management<br />

INTRODUCTION<br />

Fungicide resistance in cucurbit powdery mildew<br />

G. MacManus, C. Akem{ XE "Akem, C." }, K. Stockdale, D. Lakhesar, E. Jovicich and P. Boccalatte<br />

Horticulture and Forestry Science, Queensland Primary Industries and Fisheries, P.O. Box 15, Ayr, Qld 4807<br />

Powdery mildew, caused by Podosphaera xanthii, (Castagne) is a<br />

major constraint to commercial cucurbit production in Australia<br />

and worldwide. Management of this disease has relied primarily<br />

on the use of foliar fungicides sprays. The development of strains<br />

of the pathogen with resistance to systemic fungicides is<br />

becoming increasingly widespread with the excessive use of<br />

these fungicides. Resistance problems were first reported in<br />

Queensland in the late 1980s (1) and since then there has been<br />

no resistance monitoring program.<br />

The aim of this study was to determine if resistance had<br />

developed to the four systemic fungicides (Amistar, Bayfidan,<br />

Nimrod and Spinflo) registered in Australia for the control of<br />

powdery mildew in cucurbit crops in the Burdekin region of<br />

north Queensland.<br />

in the main production regions. Promoting integrated crop<br />

management strategies is vital. These include spraying only<br />

when needed, using resistant varieties and using fungicide<br />

alternatives as substitutes for protectant fungicides, as well as<br />

destroying old crop residues in finished strips.<br />

% Leaf Area Infected<br />

80<br />

60<br />

40<br />

20<br />

0<br />

Amistar<br />

Bayfidan<br />

Clare (Watermelon)<br />

Control<br />

Nimrod<br />

Treatment<br />

Spinflo<br />

Rating 1<br />

Rating 2<br />

MATERIALS AND METHODS<br />

In 2008, 21‐day old seedlings of a powdery mildew susceptible<br />

zucchini variety (Congo, SPS) were used in bioassays to test for<br />

resistance against current registered fungicides. <strong>Plant</strong>s with good<br />

vigour were sprayed with the four systemic fungicides at half,<br />

full and double the recommended label rates of application and<br />

water as controls, 24 h prior to overnight field exposure in<br />

various cucurbit crops at seven different locations in the<br />

Burdekin production region. Actively growing apical shoots were<br />

removed from each plant leaving two cotyledons and three to<br />

four true leaves.<br />

% Leaf Area Infected<br />

80<br />

60<br />

40<br />

20<br />

0<br />

Amistar<br />

Guthalungra (Rockmelon)<br />

Bayfidan<br />

Control<br />

Nimrod<br />

Treatment<br />

Spinflo<br />

Rating 1<br />

Rating 2<br />

The exposed treated seedlings were randomly placed on<br />

benches in a glasshouse where night temperatures averaged<br />

about 20°C and day temperatures 30°C. Three lower true leaves<br />

of each seedling, which served as replications for each plant<br />

were rated for disease severity 11 and 15 days after field<br />

exposure. Disease severity (% leaf area infected) was estimated<br />

to the nearest 5%. The data collected was analysed using<br />

Genstat 11 to determine treatment differences.<br />

RESULTS<br />

Powdery mildew infection was first noticed on the watersprayed<br />

control seedlings 7 days after field exposure. Disease<br />

severity at the second rating was always higher than the first (Fig<br />

1; A‐C; based on the recommended label rate for each of the<br />

fungicides). There was low disease severity (≤15% on the<br />

controls) at four of the locations with no significant treatment<br />

differences for the first and second ratings. At the other three<br />

locations; Clare, Rocky Ponds and Guthalungra, disease severity<br />

on the controls was quite high (~60%). All the fungicide<br />

treatments were effective against the disease, except at Rocky<br />

Ponds and Guthalungra where they were not significantly<br />

different from the controls.<br />

DISCUSSION<br />

The results from Rocky Ponds and Guthalungra clearly show that<br />

there is a fungicide resistance problem in some areas in the<br />

Burdekin region. This is a major cause for concern. Similar results<br />

were recorded on seedlings exposed to isolates from the<br />

Bundaberg region of Central Queensland.<br />

% Leaf Area Infected<br />

60<br />

40<br />

20<br />

0<br />

Amistar<br />

Bayfidan<br />

Rocky Ponds (Honeydew)<br />

Control<br />

Nimrod<br />

Treatment<br />

Spinflo<br />

Rating 1<br />

Rating 2<br />

Figures 1. A‐C: Effect of systemic fungicides on powdery mildew disease<br />

severity in cucurbit crops in the Burdekin region of north Queensland.<br />

ACKNOWLEGEMENTS<br />

Funding for this work was provided by Horticulture Australia<br />

Limited (HAL) for which we are grateful.<br />

REFERENCE<br />

1. O’Brien, R.G., Vawdrey, L.L. and Glass, R.J. (1988). Fungicide<br />

resistance in cucurbit powdery mildew (Sphaerotheca fuliginea)<br />

and its effect on field control. Australian Journal of Experimental<br />

Agriculture 28, 417–423.<br />

These results reinforce the need for monitoring of isolate<br />

sensitivities to the main systemic fungicides on a seasonal basis<br />

50 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Population genetic analyses of plant pathogens: new challenges and opportunities<br />

C.C. Linde{ XE "Linde, C.C." }<br />

Botany and Zoology, Research School of Biology, College of Medicine, Biology and Environment, Bldg. 116, Daley Rd, Australian National<br />

University, Canberra, ACT 0200, Australia<br />

INTRODUCTION<br />

The study of population genetics attempts to investigate<br />

evolutionary forces such as mutation, migration, genetic drift,<br />

selection and recombination, and how gene frequencies change<br />

in populations to shape their genetic structure. These<br />

evolutionary forces and the interaction amongst them are<br />

particularly important in plant pathogens where, combined with<br />

the pathogen’s life history characteristics, they determine the<br />

evolutionary potential. The population genetics of plant<br />

pathogens has been investigated for at least 30 years. Early<br />

studies on population genetics of plant pathogens concentrated<br />

on the effect sexual reproduction has on levels of genetic<br />

diversity in populations (Burdon and Roelfs, 1985a, b) and what<br />

impact that had on disease control. Similar studies have<br />

continued with investigations of pathogen capacities to rapidly<br />

adapt to new environments such as developing resistance<br />

against a fungicide or overcoming a resistance gene in the plant<br />

host (McDonald and Linde, 2002).<br />

Although the questions we ask in the population genetics of<br />

plant pathogens has not changed significantly, advances in DNA<br />

sequencing and analytical approaches have significantly<br />

improved the accuracy of parameter estimates. In particular,<br />

coalescent based approaches are a powerful extension of<br />

classical population genetics because it is a collection of<br />

mathematical models that can accommodate biological<br />

phenomena as reflected in molecular data. The emphasis in<br />

coalescent thinking is to view populations backwards in time,<br />

using the divergence observable in a population to estimate the<br />

time to a most recent common ancestor. This ancestor is the<br />

point where gene genealogies `coalesce’, in a single biological<br />

organism.<br />

The barley scald pathogen, Rhynchosporium secalis, will be used<br />

as an example to illustrate the importance of some of these<br />

evolutionary forces and how coalescent based methods<br />

significantly improved our understanding of the pathogens’<br />

biology.<br />

MATERIALS AND METHODS<br />

Populations of R. secalis were characterised with 14<br />

microsatellite loci (Linde et al., 2009) and several sequence loci<br />

(Zaffarano et al., 2009). Several population genetic parameters<br />

were investigated, including migration among populations. This<br />

was investigated with a coalescent method in the program IM<br />

(Hey and Nielsen, 2004) and results were compared to estimates<br />

derived from traditional F ST estimates (Weir and Cockerham,<br />

1984).<br />

pathogen populations are constantly influenced by the host<br />

populations or human‐mediated migration.<br />

With coalescent methods, the direction of migration is obtained.<br />

This means the major source and sink populations for migration<br />

can be determined which is useful in determining breaches of<br />

quarantine or major migration routes due to eg prevailing wind<br />

currents. In R. secalis, unusually high migration rates in both<br />

directions between Australia and South Africa and Australia and<br />

New Zealand cause particular concern for disease management.<br />

A comparison of the results revealed that migration estimates<br />

based on coalescent analyses were frequently non‐symmetric,<br />

meaning one population contributed significantly more migrants<br />

than the other. This contributed to migration estimates derived<br />

from F st being over‐ or under‐estimated. Furthermore, F st derived<br />

migration estimates were usually underestimated when the<br />

migration was high, and/or when population sample sizes were<br />

low.<br />

Coalescent analyses provided population genetic parameter<br />

estimates that are more reliable, are less dependent on<br />

population sizes being stable and are affected less by<br />

populations with small sample sizes. Improved analyses and<br />

their usefulness in plant pathology are discussed.<br />

REFERENCES<br />

1. Burdon, J.J., Roelfs, A.P., 1985a. Isozyme and virulence variation in<br />

asexually reproducing populations of Puccinia graminis and<br />

Puccinia recondita on wheat. Phytopathology 75, 907–913.<br />

2. Burdon, J.J., Roelfs, A.P., 1985b. The effect of sexual and asexual<br />

reproduction on the isozyme structure of populations of Puccinia<br />

graminis. Phytopathology 75, 1068–1073.<br />

3. McDonald, B.A., Linde, C., 2002. Pathogen population genetics,<br />

evolutionary potential, and durable resistance. Annu. Rev.<br />

Phytopathol. 40, 349–379.<br />

4. Linde, C.C., Zala, M., McDonald, B.A., 2009. Molecular evidence for<br />

recent founder populations and human‐mediated migration in the<br />

barley scald pathogen Rhynchosporium secalis. Molecular<br />

Phylogenetics and Evolution 51, 454–464.<br />

5. Zaffarano, P.L., McDonald, B.A., Linde, C.C., 2009.<br />

Phylogeographical analyses reveal global migration patterns of the<br />

barley scald pathogen Rhynchosporium secalis. Molecular Ecology,<br />

279–293.<br />

6. Hey, J., Nielsen, R., 2004. Multilocus methods for estimation<br />

population sizes, migration rates and divergence times, with<br />

application to the divergence of Drosophila pseudoobscura and D.<br />

persimilis. Genetics 167, 747–760.<br />

7. Weir, B.S., Cockerham, C.C., 1984. Estimating F‐statistics for the<br />

analysis of population structure. Evolution 38, 1358–1370.<br />

Keynote speaker<br />

RESULTS AND DISCUSSION<br />

The results of this comparison revealed that coalescent based<br />

approaches offer several advantages over other analytical<br />

methods to estimate parameters such as migration and genetic<br />

drift. Traditional measures of the translation of F ST into gene<br />

flow assume that subpopulations have the same size, population<br />

sizes are constant, or that there are infinitely many populations,<br />

and that migration rates are all symmetric. Due to these<br />

underlying assumptions, migration estimates derived from F ST<br />

are almost always flawed and incorrect estimates are achieved<br />

when these assumptions are not met. This is often the case since<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 51


Session 3A—Population genetics<br />

Genetic diversity of Botryosphaeria parva (Neofusicoccum parvum) in New Zealand<br />

vineyards<br />

INTRODUCTION<br />

J. Baskarathevan{ XE "Baskarathevan, J." }, M.V. Jaspers, E.E. Jones and H.J. Ridgway<br />

Faculty of Agriculture and Life Sciences, Lincoln University, PO Box 84, Lincoln 7647, New Zealand<br />

Species of the Botryosphaeriaceae cause disease on numerous<br />

woody and non‐woody hosts. Worldwide several species of<br />

Botryosphaeria have been reported to cause various symptoms<br />

on grapevine including decline and die‐back. In a survey carried<br />

out in 2007/08, six Botryosphaeria species were isolated from<br />

symptomatic New Zealand grapevines. Among these<br />

Botryosphaeria species, Neofusicoccum parvum was identified as<br />

the most prevalent species (1). The aim of this study was to<br />

analyse the genetic diversity of the N. parvum isolates collected<br />

from the New Zealand vineyards and to compare these with<br />

international isolates.<br />

MATERIALS AND METHODS<br />

Fungal isolates and DNA extraction. New Zealand isolates of N.<br />

parvum (49), collected from six grape growing regions as well as<br />

three Californian and four Australian isolates were used in this<br />

study. The identity of the New Zealand isolates was confirmed by<br />

using a published PCR‐RFLP (ARDRA) method (2). Genomic DNA<br />

was extracted from mycelium using the PUREGENE ® genomic<br />

DNA isolation kit and concentration was adjusted to 20 ng /µl for<br />

UP‐PCR.<br />

UP‐PCR procedure. For genetic variation analysis of N. parvum<br />

isolates, 11 UP‐PCR primers (3) were tested, of which five<br />

(AA2M2, Fok1, L15, 0.3‐1 and 3‐2) were chosen for further use,<br />

based on total number of bands, number of polymorphic bands<br />

and the band distribution. UP‐PCR amplification products were<br />

separated by electrophoresis on a 1% agarose gel for 3 h at 100<br />

V.<br />

Genetic variation analysis. Bands were counted if they were<br />

strong and reproducible. A binomial matrix was produced as<br />

follows: if a band was present it was indicated by a “1” and if<br />

absent by a “0”. The binomial matrix thus generated was used<br />

for phylogenetic analysis using PAUP version 4.0b10. A<br />

neighbour joining cladogram was generated to show the<br />

relationships between the genotypes produced for all 56<br />

isolates.<br />

Pathogenicity test. From each general branch of the N. parvum<br />

neighbour joining tree, two isolates were selected (15 isolates in<br />

total) for in‐vitro green shoot assays. Shoots of similar age and<br />

diameter, 25–30 cm long, were cut from glasshouse grown<br />

Sauvignon Blanc plants. They were prick‐wounded and<br />

inoculated with mycelium colonised agar plugs from 3 day old<br />

cultures, wrapped onto the wounds with parafilm. Inoculated<br />

shoots were arranged randomly in a growth chamber with five<br />

replicates for each isolate. Pathogenicity of isolates was assessed<br />

as the external lesion lengths, measured at 7 d post inoculation.<br />

The data were analysed using ANOVA (GenStat 11) Means were<br />

separated by Fisher’s protected least significance difference<br />

(LSD) test.<br />

RESULTS<br />

The 61 informative bands produced with the five UP‐PCR<br />

primers were used to compile a neighbour joining tree. This tree<br />

revealed a high degree genetic variability in the N. parvum<br />

populations studied. Genetic variability was higher for intervineyard<br />

rather than intra‐vineyard populations. Only one of the<br />

four Australian isolates grouped together with New Zealand<br />

isolates. All of the Californian isolates were grouped into a<br />

branch with three New Zealand isolates collected from a single<br />

region. The N. parvum isolates selected from different branches<br />

of the neighbour joining tree differed significantly (P


Anthracnose disease of chili pepper—genetic diversity, pathogenicity and breeding<br />

for resistance<br />

P.W.J. Taylor{ XE "Taylor, P.W.J." } 1 , O. Mongkolporn 2,3 , P. Montri 3,4 , P. Mahasuk 3 , N. Khumpeng 2,5 , T. Supakaew 3 , A. Auyong 1 , N.<br />

Ranathunge 1 , R. Ford 1<br />

1 Centre for <strong>Plant</strong> Health/BioMarka, School of Land and Environment, The University of Melbourne, Victoria, Australia 3010<br />

2 Department of Horticulture, Kasetsart University, Kamphaeng Saen Campus, Nakhon Pathom 73140, Thailand<br />

3 Center for Agricultural Biotechnology, Kasetsart University, Kamphaeng Saen Campus, Nakhon Pathom 73140, Thailand<br />

4 Current address: Royal Project Foundation, Chiang Mai, Thailand<br />

5 Current address: Syngenta Crop Protection Limited, Khon Kaen, Thailand<br />

INTRODUCTION<br />

Anthracnose disease of chili pepper (Capsicum annuum) is<br />

caused by a complex of Colletotrichum species with C. capsici, C.<br />

acutatum, C. gloeosporioides, being the most severe pathogens<br />

in SE Asia and Australia (1). Elucidation of the disease cycle for C.<br />

capsici indicated that quiescent leaf infection was an important<br />

source of inoculum thus necessitating more efficient use of<br />

fungicides to prevent fruit infection.<br />

MATERIALS AND METHODS<br />

Genetic Diversity: The genetic structure of populations of C.<br />

capsici collected from Australia, India, Sri Lanka and Thailand<br />

were analysed using loci‐specific microsatellite markers (2) The<br />

number of effective alleles and the genetic diversity was<br />

determined using genalex 6 software.<br />

Pathogenicity: Differential reactions based on qualitative host<br />

reactions (infection vs no infection) on mature green and ripe<br />

chili fruit of 10 genotypes from four cultivated Capsicum<br />

species—C. annuum, C. baccatum, C. chinense and C. frutescens<br />

were investigated after being inoculated with 33 isolates of C.<br />

capsici, C. gloeosporioides and C. acutatum from Thailand.<br />

Bioassay and host reaction was recorded as described in Montri<br />

et al. (2009).<br />

Breeding for Resistance: Resistance to anthracnose caused by<br />

Colletotrichum capsici and C. acutatum was investigated in C.<br />

chinense PBC932 and Capsicum baccatum PBC80 and PBC1422.<br />

Mature green and ripe fruit were inoculated with 13 isolates of<br />

the two Colletotrichum species. Bioassay and host reaction was<br />

recorded as described in Montri et al. (2009).<br />

RESULTS AND DISCUSSION<br />

Genetic Diversity: Screening of 117 isolates against 27 STMS<br />

markers revealed 92 haplotypes and a total of 148 alleles<br />

present across all the populations. A highly significant population<br />

differentiation (0.154) was found among the populations. The<br />

Australian population was relatively homogeneous with low level<br />

of gene diversity and high population differentiation suggesting<br />

their recent origin (probably caused by genetic bottlenecks) in<br />

comparison with highly diverse Indian, Sri Lankan and Thai<br />

isolates. The overall high gene flow and diversity indicated that<br />

C. capsici had a high adaptive potential to overcome control<br />

measures such as host resistance and fungicides.<br />

capsici and an Agrobacterium‐mediated fungal transformation<br />

system developed for assessing function of these genes.<br />

Breeding for Resistance: The resistant C. chinense genotype<br />

PBC932 showed a strong hypersensitive response to infection by<br />

C. capsici pathotypes. Resistance was found to be controlled by<br />

three recessive genes at specific growth stages. This resistance is<br />

being incorporated into commercial varieties through marker<br />

assisted selection. ie co1 at mature green fruit, co2 at ripe red<br />

fruit, and co3 at seedling stages of plant growth. Linkage analysis<br />

suggested that co1 and co2 were linked (recombination<br />

frequency 0.25), and that the co3 was not linked to the fruit<br />

resistances. Resistance at mature green fruit stage in C.<br />

baccatum to C. acutatum was found to be controlled by a single<br />

recessive gene co4 and at ripe fruit stage by a single dominant<br />

gene Co5. Linkage analysis between the two genes showed the<br />

genes to be independent. Markers to these genes will be<br />

developed for use in Marker Assisted Selection to enable the<br />

development of highly resistant chili varieties.<br />

REFERENCES<br />

1. Montri P, Taylor PWJ, Mongkolporn O (2009) Pathotypes of<br />

Colletotrichum capsici, the Causal Agent of Chili Anthracnose, in<br />

Thailand. <strong>Plant</strong> Disease 93: 17–20.<br />

2. Ranathunge NP, Ford R, Taylor PWJ (2009). Development and<br />

optimization of sequence‐tagged microsatellite site markers to<br />

detect genetic diversity within Colletotrichum capsici, a causal<br />

agent of chilli pepper anthracnose disease. Molecular Ecology<br />

Resources DOI: 10.1111/j.1755‐0998.2009.02608.x<br />

3. Mahasuk P, Khumpeng N, Wasee S, Taylor PWJ, Mongkolporn, O<br />

(2009). Inheritance of resistance to anthracnose (Colletotrichum<br />

capsici) at seedling, and fruiting stages in chili pepper (Capsicum<br />

spp.). <strong>Plant</strong> Breeding. (In press).<br />

4. Mahasuk P, Taylor PWJ and Mongkolporn O (2009). Identification<br />

of two new genes conferring resistance to Colletotrichum acutatum<br />

in Capsicum baccatum L. Phytopathology. (In press).<br />

Session 3A—Population genetics<br />

Pathogenicity: Differential reactions on mature green and ripe<br />

chili fruit of 10 genotypes of cultivated Capsicum spp identified<br />

5, 11 and 3 pathotypes of C. capsici, C. gloeosporioides and C.<br />

acutatum respectively. This will have profound effect on chili<br />

breeding programs where novel sources of resistance genes<br />

from related species are being incorporated into commercial C.<br />

annuum varieties. Putative PR genes have been identified<br />

through transcriptional analysis from a virulent pathotype of C.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 53


Session 3A—Population genetics<br />

INTRODUCTION<br />

The diversity of Colletotrichum infecting lychee in Australia<br />

J.M. Anderson{ XE "Anderson, J.M." } A,B , L.M. Coates A , E.K. Dann A and E.A.B. Aitken B<br />

A Queensland Primary Industries and Fisheries, 80 Meiers Rd, Indooroopilly, 4068, Qld<br />

B University of Queensland, John Hines Building, St Lucia, 4072, Qld<br />

Pepper spot caused by Colletotrichum gloeosporioides is a<br />

relatively new disease of lychee (Litchi chinensis Sonn.) in<br />

Australia. While the disease only causes superficial damage to<br />

the skin of the fruit it does result in lower financial returns to<br />

lychee growers (1). Unlike anthracnose of lychee, symptoms of<br />

pepper spot develop prior to harvest with out an apparent<br />

quiescent period for C. gloeosporioides. Host stress and<br />

environment have been suggested to contribute to pepper spot<br />

development; however it is not clear if the genotype of C.<br />

gloeosporioides affects pepper spot development. In this study<br />

we compared isolates of Colletotrichum spp. from anthracnose<br />

and pepper spot lesions on lychee to determine if a new<br />

genotype of Colletotrichum sp. is responsible for pepper spot of<br />

lychee.<br />

MATERIALS AND METHODS<br />

Collection and characterisation of isolates. One‐hundred and<br />

fifty isolates of C. gloeosporioides were collected from pepper<br />

spot and anthracnose lesions of lychee from three orchards in<br />

eastern Australia. Isolates derived from single germinated<br />

conidia were characterised on the basis of morphology (conidia<br />

shape and size, production of teleomorph) and molecular<br />

fingerprint using arbitrary‐primed PCR (ap‐PCR) using two<br />

primers.<br />

Field pathogenicity testing. Thirteen isolates representing the<br />

main genotypes identified using ap‐PCR were selected for field<br />

pathogenicity testing. An isolate of C. gloeosporioides<br />

(BRIP28734) from mango was included in the pathogenicity<br />

testing. Lychee fruit were inoculated two weeks prior to harvest<br />

and were assessed for the development of pepper spot at<br />

harvest and for the development of anthracnose after 10 days of<br />

storage at 20°C.<br />

Field pathogenicity testing. Only isolates ALPS11, ALAN11 and<br />

RLPS24 caused significant levels of pepper spot (Figure 1), all of<br />

these isolates were from the closely related group identified<br />

using ap‐PCR which contained the majority of isolates. The other<br />

isolates from this group ALPS15, ALAN12 and RLAN11 did not<br />

cause pepper spot but did cause high levels of anthracnose after<br />

storage. None of the isolates in the large closely related group<br />

produced the teleomorph in culture, unlike the isolates of C.<br />

gloeosporioides from other genotypes of which 60% produced<br />

the teleomorph.<br />

Fruit inoculated with the mango isolate (BRIP 28734) had levels<br />

of anthracnose similar to the uninoculated control. Of the lychee<br />

isolates, GLPS12 caused only low levels of anthracnose which<br />

were not significantly different to those of the mango isolate.<br />

GLPS12 was from the genotype where the majority of the<br />

isolates produced the teleomorph. It is possible that this group<br />

of isolates is not specific to lychee.<br />

On the basis of ap‐PCR and morphological studies there were no<br />

apparent differences found between pepper spot and<br />

anthracnose isolates. However, when assessed together, the<br />

pathogenicity testing and molecular work suggest that there may<br />

be a group of isolates more likely to cause pepper spot. To<br />

confirm the host specificity of the isolates cross pathogenicity<br />

testing is being conducted with a range of hosts.<br />

RESULTS AND DISCUSSION<br />

Collection and characterisation of isolates. Of the 150 isolates in<br />

the collection, nine were C. acutatum, isolated from anthracnose<br />

lesions. The rest of the isolates were C. gloeosporioides. Analysis<br />

of conidial measurements did not indicate a distinct subpopulation<br />

within C. gloeosporioides responsible for the<br />

development of pepper spot. The production of the<br />

teleomorphic stage was not limited to isolates from one<br />

symptom type or location.<br />

The ap‐PCR analysis did not differentiate the pepper spot<br />

isolates from the anthracnose isolates. The majority of the<br />

isolates (73%) grouped together with 75% similarity to each<br />

other. Within this group were isolates from both pepper spot<br />

and anthracnose as well as isolates from all three locations.<br />

Most other genotypes identified had only one or very few<br />

isolates, except for a genotype which contained 15 isolates from<br />

the same location. Of the latter, 14 isolates produced the<br />

teleomorph under laboratory conditions. Both pepper spot and<br />

anthracnose isolates were in this group.<br />

Figure 1. Preharvest pepper spot (solid columns) and postharvest<br />

anthracnose (white columns) development on lychee fruit inoculated 2<br />

weeks prior to harvest. Error bars indicate standard error.<br />

ACKNOWLEDGEMENTS<br />

JA is recipient of a DAFF Science and Innovation Award<br />

REFERENCES<br />

1. Cooke AW and Coates LM (2002) Pepper spot: a preharvest disease<br />

of lychee caused by Colletotrichum gloeosporioides. <strong>Australasian</strong><br />

<strong>Plant</strong> <strong>Pathology</strong> 31:303–304.<br />

54 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


INTRODUCTION<br />

Variation in Phytophthora palmivora on cocoa in Papua New Guinea<br />

J.Y. Saul Maora{ XE "Maora, J.Y.S." } 1 , D.I. Guest 2 and E.C.Y. Liew 3<br />

1 PNG Cocoa Coconut Institute, PO Box 1846 Rabaul, ENBP, Papua New Guinea<br />

2 Faculty of Agriculture, Food and Natural Resources, The University if Sydney, NSW 2006, Australia<br />

3 Royal Botanical Gardens Sydney, Botanic Gardens Trust, Mrs Macquaries Rd, Sydney, NSW 2000, Australia<br />

Phytophthora palmivora is the major cause of cocoa disease in<br />

PNG, although P. arecae, P. megakarya, P. nicotianae, and P.<br />

citrophthora have also been suggested as pathogens. The<br />

success of disease control strategies depends on a thorough<br />

knowledge of the pathogen and its population biology.<br />

Furthermore, all cocoa planting material is bred in ENB and<br />

distributed throughout the country, without any comprehensive<br />

Session 3A—Population genetics<br />

This study tested the hypotheses that: 1. P. palmivora is the sole<br />

Phytophthora sp. causing disease on cocoa in PNG; and 2. there<br />

is variation between P. palmivora populations from different<br />

cocoa growing locations in PNG.<br />

MATERIALS AND METHODS<br />

ESP<br />

Karkar<br />

Madang<br />

Sampling locations in PNG<br />

ENB<br />

Bougainville<br />

Figure 1. Sampling locations in PNG<br />

Diseased pods were sampled hierarchical from 5 locations (Fig<br />

1); 8 farms/location, 8 diseased pods/farm. Isolates were grown<br />

on carrot agar in the dark at 25°C for 4 days and growth rates<br />

taken simultaneously with colony morphology. For sporangial<br />

morphology, isolates were grown as described for 10–14 days<br />

and sporangial length, breadth and pedicel length of fifty<br />

sporangia per isolate measured and length/breadth ratio<br />

calculated. Sporangiophore branching and caducity were also<br />

recorded. Mating type was determined using Duncan’s media<br />

(3). Genetic variation was studied using Randomly Amplified<br />

Microsatellite (RAMS) analysis. Genomic DNA was extracted and<br />

amplified by PCR and PCR products separated by agarose gel<br />

electrophoresis. Loci with clear bands only were scored.<br />

RESULTS<br />

Colony morphology of the isolates were stellate striate.<br />

Sporangiophore branching was simple sympodium and sporangia<br />

were caduceus. Mean Sporangia length, breadth, pedicel length<br />

and length:breadth ration were 50.1µm, 27.2µm, 4.7µm and 1.8<br />

respectively. Growth rates were not correlated with locations.<br />

Sporangia length, breadth and length:breadth ratio were<br />

different depending on locations. Both mating types A1 and A2<br />

are present in PNG. Genetic analysis revealed seven clonal<br />

groups (Figure 2).<br />

Figure 2. UPGMA dendogram based on RAMS analysis, representatives<br />

of 263 isolates of P. palmivora from PNG including references, P. capsici<br />

(Caps).<br />

DISCUSSION<br />

Isolates displayed striate/stellate colony morphology which is<br />

typical of P. palmivora (2). Sporangiophore branching, caducity<br />

and sporangial dimensions agree with descriptions of P.<br />

palmivora (4). Sporangia pedicel length was intermediate<br />

(~5µm), typical of P. palmivora (1). RAMS analysis clearly<br />

separated P. capsici from the PNG isolates and revealed that the<br />

population in PNG was generally clonal with emerging subpopulations<br />

in Bougainville and Madang. Both mating types A1<br />

and A2 are present in Madang posing a high risk of variation. P.<br />

palmivora is the sole Phytophthora sp causing disease on cocoa<br />

in PNG. The presence of a single species means concentrating<br />

cocoa breeding in one location is acceptable; however mixed<br />

cocoa cultivars should be deployed and integrated disease<br />

management promoted. Strict quarantine should be imposed,<br />

especially in Madang where both mating types are present.<br />

ACKNOWLEDGEMENTS<br />

The scholarship provided by AusAid to undertake this study is<br />

highly appreciated. The assistance of CCIPNG staff for isolate<br />

collection and assistance provided by Dr Rose Daniel at The<br />

Sydney University is acknowledged.<br />

REFERENCES<br />

1. Al‐Hedaithy SSA, Tsao PH (1979b) Sporangium pedicel length in<br />

Phytophthora species and consideration of it’s uniformity in<br />

determining sporangium caducity. Mycological Research 72, 1–13.<br />

2. Brasier CM, Griffin MJ (1979) Taxonomy of Phytophthora palmivora<br />

on cocoa. Transactions British Mycological <strong>Society</strong> 72, 111–143.<br />

3. Duncan JM (1988) A colour reaction associated with formation of<br />

oospores by Phytophthora spp. Transactions British Mycological<br />

<strong>Society</strong> 90, 366–337.<br />

4. Erwin DC and Riberiro OK (1996) Phytophthora diseases worldwide<br />

APS Press: Minnesota, USA.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 55


Session 3B—Modelling and crop loss assessment<br />

Spore traps for early warning of smut infestations in Australian sugarcane crops<br />

INTRODUCTION<br />

R.C. Magarey{ XE "Magarey, R.C." } A , G. Bade B , K.S. Braithwaite C , AK.J. Lonie A and B.J. Croft D<br />

A BSES Limited, PO Box 566, Tully, 4854, Queensland<br />

B BSES Limited, PO Box 651, Bundaberg, 4670, Queensland<br />

C BSES Limited, PO Box 86, Indooroopilly, 4068, Queensland<br />

D BSES Limited, 207 Old Cove Road, Woodford, 4514, Queensland<br />

Sugarcane is Australia’s second most important agricultural<br />

export crop on a monetary basis, being grown on over<br />

400,000ha of land to produce >30m tonnes of harvest product<br />

and >3m tonnes of crystal sugar. There have been many climatic,<br />

market and disease threats to the industry and recently<br />

sugarcane smut (Ustilago scitaminea) has posed a major threat.<br />

In 2006 the disease was found for the first time in the major<br />

eastern seaboard sugarcane production area. Previous research<br />

showed that the majority of the Australian commercial cultivars<br />

were highly susceptible to the disease; rapid spread and<br />

escalation of the disease therefore threatened crop yields and<br />

profitability of the Australian sugarcane industry. As the disease<br />

is difficult to find in mature crops, and as early warning of the<br />

disease would provide the greatest opportunity for farmers to<br />

transition to more resistant cultivars, it was decided to use<br />

atmospheric spore trapping as an early warning tool. This paper<br />

describes how spore trapping has provided early warning of the<br />

presence of smut to the Australian sugarcane industry.<br />

MATERIALS AND METHODS<br />

Smut outbreak. Smut was found for the first time in the Childers<br />

region in June 2006, then shortly afterwards in the Mackay<br />

production area in November 2006 and in the Herbert River<br />

district in December 2006. Early warning research was principally<br />

applied therefore to other production areas: these included<br />

northern Queensland (Tully north), the Burdekin Irrigation Area,<br />

specific parts of central and southern Queensland and northern<br />

New South Wales.<br />

Spore traps. Burkard ‘spore and pollen samplers’ were sourced<br />

from the UK manufacturer; traps incorporate an air intake at 10l<br />

/ minute, a revolving internal drum that rotates once every<br />

seven days, and an attached sticky tape made from clear plastic<br />

coated with petroleum jelly.<br />

Detection of smut spores. Initial diagnosis of spore tapes was by<br />

light microscopy. There were several major issues with this<br />

method: i. it was very difficult to categorically state that a spore<br />

on a spore trap tape was U. scitaminea and not that of another<br />

smut species, ii. the diagnosis was time consuming, causing<br />

there to be significant delays between a spore trapping event<br />

and supply of the results, iii. operator fatigue was a very real<br />

issue. For this reason, a specific molecular test for U. scitaminea<br />

was developed and used to assay spore trap tapes. The<br />

advantages were the generation of a specific, faster result,<br />

though outcomes were then qualitative rather than quantitative<br />

(plus or minus smut only)—no quantification of the airborne<br />

inoculum was possible using this assay.<br />

Spore trapping program. Fifteen spore traps were purchased<br />

soon after the initial smut detection in June 2006; trapping<br />

began in the nominated areas in late 2006. Sites were selected<br />

across the relevant district with care taken to minimise tape<br />

contamination from dust associated with vehicular traffic on<br />

unsealed roads and tracks. Traps were operated for 1–2 days at<br />

each trap site before movement to another location within the<br />

district. Records of weather conditions, site GPS details and crop<br />

details allowed sites to be characterised for later interpretation<br />

of results. Mapinfo (version 8) software was used to provide GIS<br />

information on smut spore detections.<br />

RESULTS<br />

Early warning. Many detections of sugarcane smut spores within<br />

districts and regions were made before disease symptoms were<br />

found. In the Burdekin Irrigation Area, over 40% of trap sites<br />

returned positive spore trap detections in July 2007. Although<br />

careful crop inspections were undertaken at this time, and over<br />

the next 18 months, disease symptoms were not detected until<br />

October 2008. Similar observations were made in a number of<br />

districts (spore detections before crop symptoms); these are<br />

listed in Table 1.<br />

Table 1. Spore trap detections of U. scitaminea in Australian sugarcane<br />

production areas compared to when crop symptoms were first<br />

identified.<br />

District<br />

Spores detected<br />

Crop symptoms<br />

identified<br />

Mossman July 2007 December 2008<br />

Tableland July 2008 September 2008<br />

Mulgrave July 2008 September 2008<br />

Burdekin April 2007 October 2008<br />

Proserpine July 2007 December 2007<br />

Maryborough March 2007 January 2008<br />

DISCUSSION<br />

The smut spore trapping program successfully provided early<br />

warning of the disease in Queensland and New South Wales<br />

cane‐fields. 18 months pre‐emptive warning was provided in the<br />

Burdekin and Mossman areas of northern Queensland. Smut<br />

spores have been found in northern NSW but after 18 months,<br />

no crop symptoms have been seen in this region. Early warning<br />

has provided farmers with the opportunity to implement smut<br />

management plans much earlier than otherwise would have<br />

occurred. As sugarcane is a semi‐perennial crop, it is not possible<br />

to change cultivars on a whole farm basis within one year; in fact<br />

cultivar rotations usually take 4–5 years to complete—so early<br />

warning of a disease threat is very important in the transition to<br />

resistant varieties. This type of work has not been undertaken in<br />

other sugarcane‐producing countries. The work reported here<br />

illustrates the potential for early warning of sugarcane smut<br />

using commercial spore trap technology.<br />

ACKNOWLEDGEMENTS<br />

We acknowledge the assistance of BSES Limited extension staff<br />

and Productivity Service personnel.<br />

56 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Software‐assisted gap estimation (SAGE) for measuring grapevine leaf canopy density<br />

G.N. Hill{ XE "Hill, G.N." } 1 , R.M. Beresford 1 , P.N. Wood 2 , K.S. Kim 1 and P.J Wright 3<br />

1 The New Zealand Institute for <strong>Plant</strong> and Food Research Ltd. (<strong>Plant</strong> and Food Research), Private Bag 92169, Auckland 1025, New Zealand<br />

2 <strong>Plant</strong> and Food Research, Private Bag 1401, Havelock North 4157, New Zealand<br />

3 <strong>Plant</strong> and Food Research, Pukekohe Research Station, Cronin Rd, RD 1, Pukekohe 2676, New Zealand<br />

INTRODUCTION<br />

Botrytis bunch rot (botrytis), caused by Botrytis cinerea, is a<br />

major disease of wine grapes. Reducing vine canopy density<br />

through leaf plucking can reduce harvest disease severity.<br />

Determining the level of leaf plucking required to achieve useful<br />

botrytis control requires accurate measurements of canopy<br />

density. Existing methods, such as point quadrat (PQ) analysis<br />

(1), can be labour‐intensive, subjective, damaging, too<br />

complicated and/or impractical for many researchers or for<br />

routine use by vineyard managers.<br />

Software‐assisted gap estimation (SAGE) gives a measurement of<br />

the percentage of gap in the vine canopy. This study investigated<br />

the relationship between SAGE and PQ analysis, and their<br />

relationship to botrytis severity, to establish whether or not<br />

SAGE is a satisfactory alternative to PQ analysis for vine canopy<br />

density estimation.<br />

MATERIALS AND METHODS<br />

Vineyard Trials. Two trials were conducted during the 2008–<br />

2009 growing season on Sauvignon blanc vines in New Zealand,<br />

one on a commercial vineyard in Hawke’s Bay and the other at<br />

the <strong>Plant</strong> and Food Research vineyard in Pukekohe. Treatments<br />

were imposed to produce a range of canopy densities. No<br />

botryticides were used in the trials. Canopy density was<br />

measured by both SAGE and PQ analysis at pre‐bunch closure<br />

(PBC), veraison and harvest. An additional measurement was<br />

taken at Pukekohe between PBC and veraison.<br />

SAGE. A blue tarpaulin is suspended behind the vine row and is<br />

then photographed. The image is analysed using software that<br />

calculates the ratio of the area of tarpaulin to the area of leaves<br />

in the canopy, termed ‘gap’.<br />

PQ Analysis. PQ analysis was carried out as described by Smart<br />

& Robinson in Sunlight into Wine (1). Leaf layer number (LLN)<br />

was compared with gap.<br />

Botrytis severity. Percentage severity of botrytis was visually<br />

estimated on 25 randomly selected bunches from each vine.<br />

RESULTS<br />

Gap vs LLN. Gap was found to be highly correlated with LLN<br />

(Figure 1). Fitted regression lines for the two site were not<br />

identical. Linear regression analysis testing the hypothesis<br />

relating to differences in slope and intercept of the two sites<br />

found that the intercepts were significantly different (P


Session 3B—Modelling and crop loss assessment<br />

Evaluation of the efficacy of Brassica spot models for control of white blister in<br />

Chinese cabbage<br />

D.P.F. Auer{ XE "Auer, D.P.F." } A , E.J. Minchinton A , R. Kennedy B , J.E. Petkowski A , R.F. Faggian A , R.J. Holmes A and F.M. Thomson A<br />

A DPI Victoria—Knoxfield Centre, Private Bag 15 Ferntree Gully DC, 3156, Victoria<br />

B<br />

Warrick, HRI, Wellesbourne, Warrickshire, CV359EF, United Kingdom<br />

INTRODUCTION<br />

White blister caused by Albugo candida is a major disease of<br />

Chinese cabbage, Brassica rapa. A. candida causes blisters on<br />

abbatial leaf surfaces, necessitating their removal from heads at<br />

harvest, which increases production costs. Brassica spot , a<br />

disease predictive model was developed for white blister on<br />

broccoli, B. oleracea (1). This abstract reports on evaluation of<br />

two version of the model, the infection model (old) and the<br />

latent incubation period model (new) in Chinese cabbage against<br />

weekly spray programs for control of white blister.<br />

MATERIALS AND METHODS<br />

Trial design. Chinese cabbage variety Matilda was direct seeded<br />

at 3 rows per bed on 27/11/2008 and harvested on 19/1/2009 at<br />

a commercial market garden in Devon Meadows, Victoria. The<br />

trial was laid out in 6 blocks each divided into 6 plots with<br />

unequal replication using the randomised procedure in GenStat.<br />

Each plot contained approximately 60 plants. The four<br />

treatments applied to plots were: (i) control (unsprayed); (ii)<br />

weekly sprays of Tribase Blue (copper sulphate tribase) and Li‐<br />

700 (soyal phospholypids and propionic acid); (iii) Amistar<br />

(azoxystrobin) sprayed according to the old version of the<br />

Brassica spot model and (iv) Amistar sprayed according to the<br />

new version of the Brassica spot model. The weekly program<br />

received five sprays starting at week 3, the old and new models<br />

predicted one spray each at weeks 3 and 4, respectively.<br />

Weather station and Brassica spot model. A ModelT weather<br />

station (Western Electronic Design) was placed in the crop and<br />

recorded average leaf wetness, temperature, relative humidity<br />

and total rainfall at 30 min. intervals. The model used this data<br />

to predict appearance of symptoms in the crop.<br />

Trial analysis. Disease incidence per plot was recorded as the<br />

number of plants out of 20, showing white blister symptoms. A<br />

Generalised Linear Mixed Model was used to analyse the data.<br />

Severity was scored as the number of the outer 4 free‐standing<br />

leaves that were showing white blister and these data were<br />

analysed using REML (residual maximum likelihood). Due to<br />

100% incidence of white blister on the unsprayed plots, these<br />

data were excluded from that analysis.<br />

RESULTS<br />

White blister first appeared in the crop four weeks after sowing<br />

in all treatments. At the harvest assessment, leaves appeared to<br />

be infected with white blister from oldest to youngest. No white<br />

blister symptoms were observed on the leaves covering the<br />

head. Conditions favoured white blister consistently throughout<br />

the trial, based on the Brassica spot model (Fig 1). Amistar<br />

applied according to the new Brassica spot model was the only<br />

treatment to significantly reduce incidence and severity of white<br />

blister on Chinese cabbage (Table 1).<br />

Figure 1. Brassica spot predictions for the old (top) and new (bottom)<br />

versions of the model for prediction for white blister in the Chinese<br />

cabbage trial. Old Brassica spot model: black bars = high disease<br />

pressure, grey bars = moderate disease pressure and white bars = low<br />

disease pressure. New Brassica <br />

spot model: when the 5% line (top)<br />

crosses the dashed index bar a spray is predicted (arrow).<br />

Table 1. Effect of timing sprays with the two versions of the Brassica spot <br />

disease predictive models to control white blister on Chinese cabbage at<br />

harvest.<br />

Treatment<br />

Mean incidence Mean severity on outer 4<br />

on plants (%) 1 leaves (scale 0-4) 1<br />

Control (unsprayed) 100 2.33a<br />

Weekly 95.59a 2.24a<br />

Old Brassica spot model 94.48a 2.17a<br />

New Brassica spot model 72.90b 1.17b<br />

1 Different letters against the predicted means indicates that they are significantly<br />

different (p=0.008)<br />

DISCUSSION<br />

Disease freedom on the outer 4 leaves of the harvested head is a<br />

critical commercial quality factor, but none of the treatments<br />

achieved this. Amistar applied according to the new Brassica spot <br />

model was the only treatment giving significant control of white<br />

blister on these outer leaves. The new model recommended a<br />

single spray 14 days before harvest, based on disease<br />

progression data from the crop inspections and environmental<br />

data, but spraying 14 d prior to harvest may co‐incidentally be<br />

the best phenological timing to protect the 4 unfolding leaves.<br />

Using the model is time consuming because it involves crop<br />

inspections. Further work is suggested to compare spraying<br />

according to new Brassica spot model against a single spray of<br />

fungicide 14 d before harvest.<br />

ACKNOWLEDGEMENTS<br />

The authors thank HAL, AusVeg, the State Government of<br />

Victoria and the Federal Government for financial support and<br />

the growers for providing field trial sites.<br />

REFERENCES<br />

1. Kennedy, R. and T. Gilles (2003). Brassica spot a forecasting system<br />

for foliar disease of vegetable brassicas. 8th ICPP Christchurch, New<br />

Zealand, 2–7 February 2003, Christchurch New Zealand.<br />

58 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


evaluating an infection model of prune rust to improve the management of disease<br />

for almond and prune growers<br />

INTRODUCTION<br />

P.A. Magarey{ XE "Magarey, P.A." } A , T.J. Wicks B , N.A. Learhinan A and A. Horsfield C<br />

A South Australian Research and Development Institute, PO Box 411, Loxton, 5333, SA<br />

B SARDI, GPO Box 1671, Adelaide, 5064 SA<br />

C FarmOz, PO Box 1932, Milton BC QLD 4064<br />

Trials of an almond rust infection model could lead to financial<br />

savings by effectively controlling the disease with minimum<br />

chemical use. The threat of infection events for almond and<br />

prune rust (Tranzschelia discolour) in orchards is responsible for<br />

many fungicides being applied needlessly. We aim to identify the<br />

conditions that favour the pathogen to 1) more effectively<br />

predict disease and thus more accurately time spray<br />

applications, and 2) address key issues such as: the rising cost of<br />

fungicides and fuel, and the industry’s move to improve the<br />

carbon footprint of almond orchards. We plan to assess the<br />

potential for a disease advisory service for almond and prune<br />

growers similar to CropWatch as used in the South Australian<br />

grape industry.<br />

MATERIALS AND METHODS<br />

For the past two years, Model T MetStations®—two at Loxton in<br />

the Riverland and one on the Adelaide Plains ‐have monitored<br />

orchard micro‐climate while we have monitored disease<br />

progress. An infection model developed for prune rust (1) has<br />

been adapted for use in almonds and installed in the<br />

MetStations® as disease predictors. The MetStation® software<br />

used the model to process leaf wetness and temperature data<br />

and predict infections. We compared these with field<br />

observations to evaluate the model for accuracy. The foliage was<br />

monitored usually every 8 days (range 2–13 days).<br />

RESULTS<br />

We present data for 2008/09 at Loxton. In that season, rain<br />

events of >1.5 mm were rare: x2 in September; x1 October; x3<br />

November; x4 December and none in January‐March. Consistent<br />

with the dry conditions, new occurrence of rust (light infection<br />

only) occurred on only four occasions (Table 1).<br />

For example, on 2 November, a 10.7 mm rain induced 15 hours<br />

leaf wetness while temperatures ranged from 24.8°C–13.8°C.<br />

The model predicted an Infection Score (InfSc) of 5026 for this<br />

event. Since this was below the previously set threshold of 6,500<br />

for significant disease (data not shown), a light infection was<br />

expected. Previous experiments had measured incubation<br />

periods of between 17–21 days, so we anticipated a little rust<br />

would be first seen in the vicinity of 19 November. This matched<br />

well with observations of a few rust pustules on 17 November<br />

which increased in number by 24th.<br />

Table 1 shows the model outputs for this and three other events<br />

in which infection was observed. Similar or better accuracy was<br />

achieved on each of those occasions but an apparent failure<br />

occurred on two others, when no rust developed.<br />

In the period January‐March 2009, there were many days with<br />

no leaf wetness. The prototype model correctly predicted ‘no<br />

disease’ for these events.<br />

DISCUSSION<br />

The rarity of the rain events made it possible to decipher when<br />

infection actually occurred. One of the ‘failures’ occurred on 28<br />

November with a 12‐hour leaf wetness period for which the<br />

model predicted InfSc 3920, a light infection, comparable to that<br />

noted on 12 December, but none was seen. Analyses of the data<br />

for this and an additional predicted light infection event on 5<br />

December showed that relative humidity (RH) was low for much<br />

of the associated leaf wetness periods. This raised the question:<br />

would the prototype infection model be improved by adding a<br />

RH factor?<br />

Further review of the model is planned in subsequent seasons<br />

but evaluation to‐date suggests incorporation of RH as a factor<br />

would increase the model’s sensitivity in distinguishing potential<br />

light infections caused by leaf wetness alone, from conditions<br />

when RH is also high and more favourable for infection. For<br />

instance, a simple arbitrary multiplier could be included as<br />

follows:<br />

RH 00.0 – 89.9%, InfSc = InfSc x 1.0 (ie no change)<br />

RH 90.0 – 97.9%, InfSc = InfSc x 1.2; and<br />

RH ≥ 98%, InfSc = InfSc x 1.5.<br />

Conclusions The potential for success in adapting the Model T<br />

MetStation as a rust disease predictor to‐date has led to<br />

optimism in achieving project objectives. The current data have<br />

advanced the prospects of the prototype infection model as a<br />

useful tool for almond (and prune) growers to manage disease.<br />

The model has shown capacity to provide advice as to when<br />

sprays can be confidently withheld and when they are needed in<br />

rust control programs.<br />

ACKNOWLEDGEMENTS<br />

We appreciate the assistance of Ben Brown, the Almond Board<br />

of Australia, and Horticulture Australian Ltd in facilitating this<br />

project.<br />

REFERENCES<br />

1. Kable, PF et al (1991) A computer‐based system for the<br />

management of the rust disease of French prunes. EPPO Bulletin<br />

Volume 21(3): 573–580.<br />

Session 3B—Modelling and crop loss assessment<br />

Table 1. Predictions of Infection Events and Disease Scores Compared to Observations of the Almond Rust Fungus Tranzschelia discolor. Site 1: Loxton<br />

Research Centre, Loxton, SA. 2008/09<br />

Date Time Score, Severity and Date of Predicted Disease Orchard Disease Observations<br />

2 Nov 16:00 5026 Light infection From 19 Nov Between 17–24 Nov. Few pustules<br />

13 Nov 23:40 5338 Light infection From 30 Nov Between 26 Nov – 2 Dec. Few pustules<br />

28 Nov 0130 3920 Light infection From 15 Dec None seen<br />

5 Dec 0250 3528 Light infection From 22 Dec None seen<br />

12 Dec 12:30 3808 V. light infection From 29 Dec On 29 Dec. Few pustules<br />

18 Dec 01:30 4088 Light infection From 4 Jan Between 29 Dec and 6 Jan. Few pustules<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 59


Session 3C—Disease management<br />

Management of white blister on vegetable brassicas with irrigation and varieties<br />

INTRODUCTION<br />

E.J. Minchinton{ XE "Minchinton, E.J." } A , D.P.F. Auer A , J.E. Petkowski A , R.F. Faggian A , V. Galea B and F. Thomson A<br />

A BRD, DPI Victoria—Knoxfield Centre, Private Bag 15 Ferntree Gully DC, 3156, Victoria<br />

B School of Agronomy and Horticulture, The University of Queensland, Gatton Campus, 4334, Queensland<br />

White blister caused by the oomycete Albugo candida, has been<br />

the main foliage disease on red radish for at least 30 years. In<br />

the summer of 2001/2002, it caused up to 100% crop losses in<br />

broccoli and cauliflower in Victoria. A. candida infecting radish is<br />

generally classified as race 1 and A. candida infecting broccoli is<br />

classified as race 9.<br />

This abstract reports on systematic surveys undertaken<br />

seasonally during 2002 to identify if irrigation practices impacted<br />

upon the level of disease in commercial red radish crops, and<br />

whether similar IPM practices could be used to manage white<br />

blister on broccoli crops.<br />

MATERIALS AND METHODS<br />

Systematic surveys. Red radishes of all ages were assessed<br />

seasonally during 2002 by a ‘two‐stage sampling method’ (Nam<br />

Ky Nguyen, pers. comm.). This method was used to determine<br />

the number of beds to be selected for assessment of white<br />

blister in each half of a bay (section between over‐head sprinkler<br />

lines). The beds were then randomly selected. A bed of radish<br />

usually consists of 6 rows. In each of the selected beds, two<br />

sections each of 20 cm in length were randomly selected to<br />

assess for the incidence of white blister. During summer,<br />

autumn, winter and spring 26, 23, 25 and 27 crops were<br />

surveyed, respectively. Only growers who irrigated crops at the<br />

same times were included in the analysis. Due to the binary<br />

nature of the data, logistic regression was used to analyse the<br />

results.<br />

Irrigation field trial for broccoli. The trial, located at Dairy Road,<br />

Werribee, Victoria, was originally designed as a general split‐plot<br />

design. Each of 3 blocks contained 2 replicates. Each of the 6<br />

replicates consisted of two whole plots to which the two<br />

irrigation times of early morning (4.00 am) or evening (8.00 pm)<br />

were randomly allocated. The 8 treatment combinations of<br />

variety (consisting of the resistant variety ‘Tyson’ from Syngenta,<br />

or the susceptible variety ‘Ironman’ from Seminis) and four spray<br />

regimes (not reported here) were randomly allocated to 8<br />

subplots within each of the whole plots.<br />

Table 1. Effect of irrigation time on the incidence of white blister on red<br />

radish<br />

Incidence of white blister in 2002 (%)<br />

Time of irrigation Summer Autumn Winter Spring<br />

(Feb and Mar) (May) (Jul and Aug) (Oct and Nov)<br />

Evening (8-12pm) 7.9a 1 23.0a 4.8a 12.5a<br />

Late morning (9-12am) 0.8b 12.1b 0.4c 4.9b<br />

Early morning (approx.6.00 0.8b 3.4c 1.7b 1.5c<br />

1 Within each column (season), different letters against the means indicates that<br />

they are significantly different at the 5% level.<br />

Irrigation trial for broccoli. Irrigating the broccoli crop in the<br />

evening approximately doubled the incidence of white blister<br />

compared with early morning irrigation for both varieties (Table<br />

2). Tyson which is a resistant variety showed few symptoms of<br />

white blister.<br />

Table 2. Effect of irrigation time on broccoli varieties susceptible and<br />

tolerant to white blister<br />

Time of irrigation<br />

Mean incidence<br />

(%) of white<br />

Mean incidence (%) of<br />

white blister<br />

blister var. Tyson var. Ironman<br />

Evening (8.00 pm) 2.9 a 0.57 14.37<br />

Early morning (4.00 am) 1.3 b 0.25 6.8<br />

The means are significantly different (p=0.005).<br />

DISCUSSION<br />

Early morning irrigation (4.00 am or 6.00 am) can be a useful IPM<br />

tool to reduce the risk of white blister in red radish and broccoli<br />

crops. Growing a resistant variety may further reduce the risk of<br />

white blister.<br />

ACKNOWLEDGEMENTS<br />

The authors thank HAL, AusVeg, the State Government of<br />

Victoria and the Federal Government for financial support and<br />

the growers for providing field trial sites.<br />

Seedlings, aged 8 weeks, were planted 2 rows per bed on 23 July<br />

2008. Subplot dimensions of beds were 8 m long and contained<br />

approximately 52 plants. The middle 20 plants per subplot were<br />

assessed for the incidence (presence or absence) of white blister<br />

on broccoli heads at harvest. A generalised linear mixed model<br />

(GLMM) was fitted to the data.<br />

RESULTS<br />

Systematic surveys of red radish. White blister on red radish<br />

was most prevalent during autumn and spring (Table 1). Red<br />

radish crops irrigated in the evening (8.00pm–12.00pm) had<br />

significantly higher incidence of white blister compared with<br />

other times of irrigation. Crops that were irrigated in the early<br />

morning (approximately 6.00 am) showed lower levels of white<br />

blister in 3 of the 4 seasons surveyed (Table 1). Data collected on<br />

age of crop, cultivar and regional differences are not reported<br />

here.<br />

60 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Alternative screening methods for sugarcane smut using natural infection and tissue<br />

staining<br />

INTRODUCTION<br />

S.A. Bhuiyan{ XE "Bhuiyan, S.A." } A , V. Barden A , R.C. James A , G. Bade A , B.J. Croft B and M.C. Cox A<br />

A<br />

BSES Limited, Private Bag 4, Bundaberg DC, QLD 4670, Queensland<br />

B<br />

BSES Limited 90 Old Cove Road, Woodford, QLD 4514, Queensland<br />

Sugarcane smut caused by Ustilago scitaminea was first<br />

detected in the Ord River Irrigation Area, WA in 1998, and in<br />

Queensland in 2006 (1). BSES had rated cultivars for resistance<br />

to smut in Indonesia and WA prior to the arrival of the disease in<br />

Queensland but commenced a program to screen sugarcane<br />

clones for resistance in Queensland when it clear the disease<br />

could not be eradicated. The current screening method involves<br />

inoculation of sugarcane setts by dipping into smut spore<br />

suspension. Although this method is internationally accepted, it<br />

has some drawbacks: i) it does not replicate natural infection;<br />

and ii) it is relatively time consuming and expensive.<br />

The objectives of this research were: i) obtain data on sugarcane<br />

cultivar reaction to natural infection and compare with existing<br />

standard ratings; and ii) develop a histological method to screen<br />

for smut resistance.<br />

MATERIALS AND METHODS<br />

Natural infection. Nine cultivars with a range of smut ratings<br />

were planted at the Bundaberg smut research farm, using<br />

randomised complete block design with 10 replicates in<br />

September 2007. Rows of smut susceptible cultivar Q205 A<br />

inoculated with smut were planted between the rows of the test<br />

plots. The trial was inspected for smut twice in 2008 in the plant<br />

crop, and monthly from February 2009 in the first ratoon (regrowth)<br />

crop. The number of infected plants and total number<br />

plants in each plot were counted. Data were log‐transformed<br />

(log 10 (smut%+1)) for regression analysis to determine the<br />

relationship between standard smut rating and smut incidence<br />

for natural infection.<br />

there is potential to reduce the time of screening by discarding<br />

susceptible cultivars early in the screening program.<br />

Table 1. Incidence (%) of smut infected plants in natural<br />

infection trial in Bundaberg, assessed in March 2009<br />

Cultivar Standard rating Mean smut % *<br />

Q151 1 1.43 (1.43)<br />

Q232 A 3 2.00 (2.0)<br />

Q190 A 4 9.71 (5.5)<br />

QS97‐2067 4 6.43 (5.0)<br />

Q135 5 12.20 (4.6)<br />

QS94‐91 6 23.02 (6.5)<br />

Q188 A 7 22.08 (5.6)<br />

Q138 8 21.78 (3.3)<br />

Q205 A 9 93.17 (3.3)<br />

* values in parenthesis are ±standard error of means<br />

Log(smut%+1)<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

y = 0.1836x + 0.1479<br />

R 2 = 0.9<br />

P1 mm sections of bud were cut, and stained with trypan blue<br />

(Figure 2), and observed under a light microscope, and<br />

subsequently photographed.<br />

RESULTS AND DISCUSSION<br />

Natural infection. Except Q205 A , no smut symptoms were<br />

observed in the plant crop in 2008. High incidence of smut was<br />

observed on susceptible cultivars (rating 6–9) in the first ratoon<br />

crop compared with resistant cultivars (Table 1). The regression<br />

results suggest that there was a highly significant (P


Session 3C—Disease management<br />

Interruption of cool chain and strawberry fruit rot by leak‐causing fungi Rhizopus<br />

species<br />

M. Walter{ XE "Walter, M." }, H.J. Siefkes‐Boer and K.S.H. Boyd‐Wilson<br />

The New Zealand Institute for <strong>Plant</strong> and Food Research Limited, Canterbury Research Centre, PO Box 5, Lincoln 7640, New Zealand<br />

INTRODUCTION<br />

The causal agents of strawberry leak are several different species<br />

grouped in the fungus‐like class of zygomycetes (kingdom:<br />

Chromista), which are fast growing organisms 1 . Zygomycetes<br />

have been associated with strawberry fruit leak, mostly from the<br />

Rhizopus genus, but similar disease symptoms are also<br />

associated with Mucor infections 1 . The recent detection of<br />

several cold‐tolerant leak isolates 2 has prompted research into<br />

cool chain management for NZ strawberry. The work described<br />

here shows how breaks in the cool chain increase leak rots and<br />

therefore dramatically decrease the shelf‐life of strawberry fruit.<br />

MATERIALS AND METHODS<br />

Six leak isolates (Rhizopus stolonifer) were used to inoculate<br />

(mycelium) potato dextrose agar (PDA) and sliced strawberry<br />

fruit (5 mm). Cool chain incubation (4ºC) was interrupted at<br />

staggered intervals (daily) by exposure to room temperature<br />

(20°C) for 2 h, resulting in continuous cool chain incubation or 1–<br />

6 interruptions. Fungal growth was assessed daily during the ‘2 h<br />

at room temperature cycle’. There were 2 replicates/isolate and<br />

substrate. The experiment was repeated.<br />

An additional whole fruit experiment was conducted. Eight fruit<br />

per tray (with 10 individual compartments) were inoculated with<br />

a mycelial tuft from one of the six isolates and two fruit served<br />

as non‐inoculated controls (injury only). There were two<br />

replicate trays per isolate. Incubation and interruptions were as<br />

above. Rot was assessed employing a fruit score, where 0=no<br />

symptoms; 1=small sunken lesion; 2=large sunken lesion;<br />

3=sunken lesion with juice leaking; 4=fruit covered in mycelia. At<br />

completion of the experiment all fruit were left for 2 days at<br />

20ºC to check for delayed onset of disease symptoms.<br />

RESULTS<br />

Mycelial growth and fruit score for continuous incubation at<br />

20ºC and 4ºC (with 1–6 interruptions) are shown in Figures 1.<br />

Growth and disease symptoms increased during incubation at<br />

both temperature regimes. At the 4ºC incubation, the number of<br />

interruptions (or hours exposed to 20ºC) significantly increased<br />

(P


Enhancing Papua New Guinea smallholder cocoa production through greater adoption<br />

of integrated pest and disease management<br />

INTRODUCTION<br />

Yak Namaliu{ XE "Namaliu, Y." } A , John Konam A,B , Josephine Saul A , Rosalie Daniel C and David Guest C<br />

A PNG Cocoa and Coconut Institute, PMB 10 Rabaul, PNG<br />

B Secretariat of the Pacific Community, Suva, FIJI<br />

C Faculty of Agriculture, Food and Natural Resources, The University of Sydney AUSTRALIA<br />

More than 80% of Papua New Guinea’s annual cocoa production<br />

is produced by 150,000 smallholder farming families. The current<br />

average yield of 300 kg dry beans/ha reflects poor management<br />

and high losses to Phytophthora pod rot and canker<br />

(Phytophthora palmivora) and Vascular Streak Dieback<br />

(Oncobasidium theobromae). Since 2007 cocoa pod borer (CPB)<br />

has also been recognised as a serious pest after the eradication<br />

program in Gazelle district of ENB failed. Apart from improved<br />

cocoa genotypes, technology adoption is poor and over 95% of<br />

108 farmers surveyed had no knowledge of cocoa disease and<br />

pest management, leading observers like Frank Jarrett (1) to ask<br />

‘‘how do farmers (in PNG) find out about innovations and just<br />

what sources of information are important?”<br />

We developed interventions that are synchronised with the<br />

cocoa cropping cycle and the resources available to farmers and<br />

link stakeholders through active participation. Four Integrated<br />

Pest and Disease Management (IPDM) options have been piloted<br />

in 3 different provinces.<br />

MATERIALS AND METHODS<br />

The IPDM strategy was designed with inputs programmed in<br />

relation to peak flowering, cherelle setting and peak ripening<br />

and also in relation to the pest and disease cycle so that the<br />

IPDM inputs are applied when the pest and disease are at their<br />

weakest point of their cycle (2). Four IPDM packages including a<br />

conventional management option were tested in East New<br />

Britain (ENB), Bougainville and Madang (Table 1).<br />

Table 1. IDPM options developed for PNG cocoa farmers<br />

Option IPDM Input Activities<br />

1 Low Current practice (minimal)<br />

2 Medium Sanitation, weekly harvests, cocoa and<br />

shade tree pruning, weed management<br />

3 High Option 2 + canker treatment, fertiliser<br />

and manures<br />

4 Very high Option 3 + insect management<br />

The piloting of options involved three village communities in<br />

Bougainville, Madang and ENB. At each site 12 to 15 farmer<br />

families were selected and the four options were established<br />

through Participatory Action Research (PAR). In the start up<br />

workshop in November 2005, selected and interested farmers<br />

participated in presentations and discussions. Some smallholder<br />

communities were provided with information on improved<br />

cocoa management from trials at the CCI research station.<br />

Selected farmers, their families and extension officers were<br />

trained and were engaged in observing and analysing the<br />

condition of trees in each option before applying treatments for<br />

each management package. Each selected family treated one<br />

tree in each option and the trees were given the name of the<br />

farmer involved. The four options were applied in each farmers<br />

block, and fully replicated by the 12 participating farmers. Thus<br />

there were 36 farmers trained in each province. Farmers visited<br />

their trees every month, carried out observations and took<br />

notes. The managed cocoa tree themselves have became the<br />

farmer’s principle educator. Baseline surveys were carried out at<br />

the beginning and a second followup survey using the same<br />

questionnaire was conducted in 2008 to determine if farmers<br />

had changed their management of cocoa.<br />

RESULTS AND DISCUSSION<br />

We aim to transform the industry from the current 90% low<br />

input to 50% medium input farms. Over 108 PAR trials have been<br />

established and more than 1500 farmers trained. The uptake of<br />

options is close to 100%, and over 80% of farmers prefer the<br />

higher input options. Yield increases of more than 100% have<br />

been reported and PNG smallholders are investing in cocoa for<br />

the first time. National production has increased from 42,000T in<br />

2005 to 56,000T in 2008. Farmers in CPB‐infested areas in ENBP<br />

report increased yields following the implementation of IPDM<br />

despite the impact of CPB.<br />

In the delivery of IPDM, direct transfer of the technology and<br />

establishment of research programs with farmers via PAR<br />

provides a unique opportunity to increase adoption of research<br />

results. Through this approach smallholders were trained to<br />

record inputs and production data, and are able to understand<br />

plant health management and develop improved cocoa<br />

management and production.<br />

Establishing demonstration plots and conducting field days has<br />

increased the profile of research and extension agencies, which<br />

are now much more engaged with the day‐to‐day problems<br />

faced by the farmers. The feedback from farmers has in turn<br />

improved the capacity of supporting researchers at CCI to focus<br />

their research on industry needs.<br />

The work highlights the importance of packaging research results<br />

into recommendations to improve technology adoption.<br />

ACKNOWLEDGMENTS<br />

This work was supported by ACIAR (PHT/2003/015).<br />

REFERENCES<br />

1. Jarrett FG. 1985. Innovation in Papua New Guinea Agriculture.<br />

Discussion Paper 23. Institute of National Affairs, Port Moresby.<br />

2. Konam JK et al. 2008. Integrated Pest and Disease Management for<br />

Sustainable Cocoa Production. Monograph 131, ACIAR, Canberra.<br />

Session 3C—Disease management<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 63


Keynote address<br />

INTRODUCTION<br />

Molecular cytology of Phytophthora‐plant interactions<br />

Adrienne R. Hardham{ XE "Hardham, A.R." }<br />

<strong>Plant</strong> Cell Biology Group, Research School of Biology, The Australian National University, Canberra, 2601 ACT<br />

Many of the more than 60 species of Phytophthora are<br />

aggressive plant pathogens that cause extensive losses in<br />

agricultural crops, horticultural plants and natural ecosystems.<br />

Some Phytophthora species have narrow host ranges; others<br />

have extremely broad host ranges. P. cinnamomi, for example, is<br />

now known to infect over 3,500 plant species, many of them<br />

native to Australia.<br />

motifs. Cysts germinate from a pre‐determined site that has<br />

been oriented towards the plant surface. Germ tubes penetrate<br />

either directly through the outer periclinal wall or along an<br />

anticlinal wall. Phytophthora species contain large multigene<br />

families encoding cell wall degrading enzymes whose secretion<br />

facilitates penetration and colonisation of host tissues. Nutrients<br />

are acquired from living or dead plant cells, allowing host<br />

colonisation and pathogen reproduction within 2–3 days.<br />

Phytophthora and other members of the class Oomycetes form<br />

fungus‐like hyphae and conidia‐like asexual sporangia, but they<br />

are not fungi. The Oomycetes group with a range of other<br />

protists such as diatoms, coloured algae and malarial parasites<br />

within the Stramenopiles, an assemblage whose taxon‐defining<br />

characteristics include possession of tubular hairs on their<br />

flagella.<br />

Species of Phytophthora produce motile, biflagellate zoospores<br />

that play a key role in the initiation of plant disease. Zoospores<br />

target suitable infection sites where they encyst and attach.<br />

Cysts soon germinate and attempt to invade the underlying<br />

plant tissues. Some Phytophthora species are hemibiotrophs and<br />

initially establish a stable relationship with living host cells,<br />

obtaining nutrients through the development of haustoria within<br />

infected cells. The majority are necrotrophs that feed on dead or<br />

dying cells. Like fungi, Phytophthora and other Oomycetes<br />

secrete effector proteins that are required for pathogenicity.<br />

Some effectors, such as cell wall degrading enzymes, function in<br />

the plant apoplast but others are transported across the plant<br />

plasma membrane into the host cytoplasm from where, in<br />

susceptible plants, they orchestrate metabolic changes that<br />

favour pathogen growth. In resistant plants, recognition of the<br />

invading pathogen induces a rapid defence response that<br />

inhibits disease development. In this presentation, I will review<br />

our current understanding of cellular and molecular aspects of<br />

the interactions between plants and Phytophthora pathogens. In<br />

so doing, I will highlight how modern molecular cytology is<br />

revolutionising our ability to elucidate the roles of selected<br />

proteins and cell components in Phytophthora pathogenicity and<br />

plant defence.<br />

MATERIALS AND METHODS<br />

Early studies of the interactions between plants and species of<br />

Phytophthora used light and electron microscopy to describe the<br />

major features of disease development and the plant defence<br />

response. More recently, these traditional approaches have<br />

been extended by advanced light and electron microscopy<br />

techniques that use a variety of methods, such as<br />

immunocytochemical labelling, GFP‐tagging and confocal<br />

microscopy, to mark and visualise a range of plant and pathogen<br />

molecules and cell components. This molecular cytology not only<br />

facilitates identification of specific cell structures in fixed and<br />

sectioned material but it can also do so in living cells.<br />

<strong>Plant</strong> defence. <strong>Plant</strong>s react rapidly to attempted infection by<br />

Phytophthora. Some of the earliest responses observed include<br />

an increase in cytoplasmic Ca 2+ concentration, cytoplasmic<br />

aggregation, formation of wall appositions beneath the invading<br />

hyphae and synthesis of reactive oxygen species, pathogenesisrelated<br />

proteins and phytoalexins. Immunocytochemistry and<br />

GFP‐tagging have revealed that cytoplasmic aggregation is<br />

accompanied by dramatic and dynamic reorganisation of actin<br />

and microtubular cytoskeletons, the endoplasmic reticulum (ER),<br />

Golgi bodies (GA) and peroxisomes. Actin, ER, GA and<br />

peroxisomes become focused on the infection site and are likely<br />

to be responsible for secretion of toxins and formation of wall<br />

appositions that inhibit hyphal penetration of the plant cell wall.<br />

Recent studies of GFP‐tagged Arabidopsis plants indicate that<br />

the rapid plant cell response may be triggered by detection of<br />

the pressure exerted by the invading pathogen hypha.<br />

Figure 1. Attack and defence. A. CryoScanning electron micrograph of a<br />

Phytophthora spore that has penetrated the plant surface along the<br />

anticlinal wall between adjacent epidermal cells. Events leading up to<br />

this stage include zoospore chemotaxis to a suitable infection site,<br />

polarised cyst germination and secretion of cell wall degrading enzymes<br />

from the hyphal tip. B. Confocal microscopy of transgenic Arabidopsis<br />

plants expressing GFP‐tagged hTalin to visualise actin microfilament<br />

arrays in living plant epidermal cells responding to attack. Aggregations<br />

of filamentous actin form directly underneath sites of attempted<br />

penetration.<br />

The genomes of four Phytophthora species have now been<br />

sequenced with others in the pipeline. Extensive transcriptome<br />

data are also available for P. infestans and P. nicotianae.<br />

Together with data from a number of host plants, this sequence<br />

information provides an invaluable resource for molecular<br />

cytology studies of protein and organelle function during<br />

Phytophthora‐plant interactions.<br />

RESULTS AND DISCUSSION<br />

Phytophthora pathogenicity. Phytophthora zoospores are<br />

chemotactically and electrotactically attracted to specific regions<br />

on the plant surface that are favourable infection sites.<br />

Zoospores encyst and attach to the plant through rapid secretion<br />

of adhesive material that includes high molecular weight<br />

proteins containing multiple thrombo‐spondin type1 repeat<br />

64 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Gene expression changes during host‐pathogen interaction between Arabidopsis<br />

thaliana and Plasmodiophora brassicae<br />

INTRODUCTION<br />

A. Agarwal{ XE "Agarwal, A." } A,D , V. Kaul B , R. Faggian C and D.M. Cahill D<br />

A Department of Primary Industries, Private Bag 15, Ferntree Gully DC, 3156, Victoria<br />

B School of Botany, The University of Melbourne, Parkville, 3010, Victoria<br />

C Department of Primary Industries, 32 Lincoln Square, North Carlton, 3052, Victoria<br />

D School of Life and Environmental Sciences, Deakin University, Geelong, 3217, Victoria<br />

Plasmodiophora brassicae is a biotrophic obligate plant<br />

pathogen that causes developmental changes in susceptible host<br />

cells, leading to the development of root galls (clubroot) and<br />

stunted growth. It is one of the most devastating diseases of<br />

vegetable brassicas worldwide and significantly reduces crop<br />

yield. In Australia, clubroot is managed using a combination of<br />

integrated control methods and recently introduced resistant<br />

varieties. Breeding resistant cultivars is difficult because of the<br />

genetic variation in the field populations of the pathogen. A<br />

host‐pathogen interaction between Arabidopsis thaliana<br />

ecotype Col‐0 and Plasmodiophora brassicae (Australian field<br />

population) was examined at the cellular and molecular levels to<br />

gain a better understanding of the pathogenic mechanisms. This<br />

study reports on the gene expression study (microarray analysis)<br />

conducted for this compatible host‐pathogen interaction during<br />

the key developmental stages of the disease prior to ten days<br />

after inoculation.<br />

MATERIALS AND METHODS<br />

A modified sand‐liquid culture method was developed to grow<br />

test‐plants such that observation of the primary life‐cycle stages<br />

of P. brassicae within Arabidopsis (ecotype Col‐0) roots was<br />

possible at very early time points. A real‐time quantitative PCR<br />

(qPCR) assay was also developed to quantify P. brassicae DNA in<br />

roots from day 1 onwards and up to 23 days after inoculation.<br />

Microarray analysis was conducted at 4, 7 and 10 days after<br />

inoculation using the 22K Arabidopsis ATH1 microarray chip.<br />

Gene expression at greater than a 1.5‐fold increase or decrease<br />

at a 95% confidence level relative to controls was considered<br />

biologically significant in this study. Data analysis was carried out<br />

using the AVADIS software package and selected genes were<br />

validated using real‐time reverse transcriptase quantitative PCR<br />

(RT‐qPCR).<br />

RESULTS AND DISCUSSION<br />

According to the microscopic study, pathogen attachment and<br />

penetration occurred from day 4 onwards and root galls were<br />

fully developed within 28 days. QPCR confirmed that P. brassicae<br />

DNA was detectable in infected Arabidopsis roots from day 4<br />

onwards. The amount of amplified pathogen DNA increased by<br />

day 23 as the disease progressed equating to the amount of<br />

pathogen DNA amplified from 10 8 resting spores.<br />

Microarray analysis conducted at these early time points (days 4,<br />

7 and 10) demonstrated significant changes in gene expression.<br />

At 4 days after inoculation (dai), 147 genes were differentially<br />

up‐ or down‐regulated relative to control plants compared to 27<br />

genes at 7 dai and 37 genes at 10 dai. All genes were categorised<br />

into their functional groups and metabolic pathways. At day 4,<br />

when the pathogen had attached to the root hair and had<br />

possibly commenced penetration, differential expression of<br />

several genes known to be important for pathogen recognition<br />

and signal transduction in resistant interactions, such as the<br />

WRKY transcription factor, TIR‐NBS‐LRR and leucine‐rich repeat<br />

protein, were induced. Genes involved in cell growth, jasmonic<br />

acid biosynthesis and lipid biosynthesis were also induced.<br />

However, genes involved in the biosynthesis of lignin,<br />

phenylpropanoids (salicylic acid), ethylene, cytokinin, reactive<br />

oxygen species and pathogenesis related proteins (chitinase)<br />

were repressed (Table 1).<br />

Table 1. Summary of up- and down-regulated genes expressed in<br />

Arabidopsis roots 4 days after inoculation with P. brassicae<br />

Up-regulated genes<br />

Down-regulated genes<br />

Signalling<br />

Oxidative burst/stress<br />

WRKY transcription factor<br />

leucine-rich repeat protein<br />

MAPKK<br />

CDPK<br />

TIR-NBS-LRR<br />

Cell growth and modification<br />

expansin<br />

xyloglucan:xyloglucosyl transferase<br />

pectinesterase<br />

hydroxyproline-rich glycoprotein<br />

arabinogalactan<br />

Jasmonic acid biosynthesis<br />

Lipid metabolism<br />

peroxidase<br />

glutathione S-transferase<br />

NADP oxidoreductase<br />

Lignin biosynthesis<br />

Salicylic acid biosynthesis<br />

Ethylene biosynthesis<br />

Cytokinin biosynthesis<br />

PR proteins<br />

Genes differentially expressed were fewer in number at the 7<br />

and 10 day time points, which is the time when the pathogen<br />

has established within the roots and has developed into primary<br />

plasmodia and zoosporangia, without any morphological<br />

changes in the host. Four candidate genes expressed at day 4<br />

(WRKY transcription factor, lipoxygenase, phytoalexin‐deficient 4<br />

protein and TIR‐NBS‐LRR) were confirmed by RT‐qPCR.<br />

In conclusion, the microarray study identified changes in gene<br />

expression among many host‐plant genes that are known to<br />

have important roles during plant‐pathogen interactions that<br />

may be amenable to manipulation to increase disease<br />

resistance. Microscopic observations of host roots during the<br />

infection process showed correlations between gene expression<br />

and pathogen life‐cycle stage. The most important time point, in<br />

terms of gene‐expression changes in the plant was 4 days after<br />

inoculation. Suppression of specific gene activity and/or<br />

functional groups of genes in the host may lead to susceptibility<br />

in this host‐pathogen interaction.<br />

ACKNOWLEDGEMENTS<br />

This work has been funded by DPI Victoria and Horticulture<br />

Australia Limited (HAL) using the vegetable levy and matched<br />

funds from the Australian Government. We would also like to<br />

thank DPI, Victoria for providing access to facilities.<br />

Session 4A—<strong>Plant</strong> pathogen interactions<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 65


Session 4A—<strong>Plant</strong> pathogen interactions<br />

Hairpin RNA derived from viral NIa gene confers immunity to wheat streak mosaic<br />

virus infection in transgenic wheat plants<br />

INTRODUCTION<br />

Muhammad Fahim{ XE "Fahim, M." } 1,2 , Ligia Ayala‐Navarrete 1 , Anthony A. Millar 2 and Philip J. Larkin 1<br />

1 CSIRO <strong>Plant</strong> Industry, GPO Box 1600, Canberra, ACT 2601, Australia<br />

2 Australian National University, Canberra ACT 2601, Australia<br />

Wheat streak mosaic virus (WSMV) has recently been identified<br />

from wheat and other cereals in Australia. The difficulties in<br />

finding adequate natural resistance in bread wheat and durum<br />

wheat prompted us to develop transgenic resistance based on<br />

induced siRNA mechanisms. We are reporting a successful<br />

strategy to develop resistance in Wheat against WSMV.<br />

MATERIALS AND METHODS<br />

A hairpinRNA construct was designed derived from Nuclear<br />

Inclusion ‘a’ gene of WSMV. BobWhite26 was stably cotransformed<br />

with two separate plasmids: one containing a<br />

hairpin with WSMV sequences and the other one with the nptII<br />

selectable marker.<br />

RESULTS<br />

Using biolistics we obtained a transformation efficiency of 3.5%.<br />

When progeny T 1 individuals were assayed against WSMV ten<br />

out of 16 tested families showed extreme resistance in<br />

transgenic segregants. The resistance in transgenic T 1 segregants<br />

was classified as immunity by four criteria: no disease symptoms<br />

were produced; ELISA readings were as in uninoculated plants;<br />

viral sequence could not be amplified from sap; the same saps<br />

failed to give infections in susceptible plants when used in testinoculation<br />

experiments. In one of four transgenic families<br />

examined in greater detail, the resistance segregated in a simple<br />

Mendelian ratio along with the transgene (Fig 1); the T 0 parent<br />

(hpWS2b) of this family had a single transgene insert by<br />

Southern hybridisation. Also in the T 1 family of hpWS2b the<br />

antibiotic resistance gene nptII, introduced on a separate<br />

plasmid by co‐bombardment, segregated independently of the<br />

hairpin transgene.<br />

DISCUSSION<br />

This paper reports for the first time engineered RNAi mediated<br />

immunity in wheat against WSMV using a hairpin RNA derived<br />

from Nuclear inclusion protein a (NIa) protease gene. WSMV is<br />

arguably the third most important virus of wheat behind BYDV<br />

and CYDV (barley and cereal yellow dwarf viruses). Marker‐free<br />

transgenics have advantages for regulatory approval and public<br />

acceptance (1, 2). Our study indicates that marker‐free WSMV<br />

immune plants can be readily produced using hairpinRNA genes,<br />

biolistics and co‐bombardment.<br />

ACKNOWLEDGEMENTS<br />

We acknowledge AusAID for the studentship support of MF,<br />

Peter Waterhouse for invaluable advice on hairpin RNA design,<br />

Terese Richardson and Anna Mechanicos for technical support.<br />

REFERENCES<br />

1. Miki B, Abdeen A, Manabe Y and MacDonald P (2009) Selectable<br />

marker genes and unintended changes to the plant transcriptome.<br />

<strong>Plant</strong> Biotechnology Journal 7:211–218.<br />

2. Miki B and McHugh S (2004) Selectable marker genes in transgenic<br />

plants: applications, alternatives and biosafety. Journal of<br />

Biotechnology 107:193–232.<br />

ELISA ratio (I/H ± SEM)<br />

Syptoms Severity<br />

<strong>Plant</strong> Height<br />

15<br />

10<br />

5<br />

0<br />

4<br />

3<br />

2<br />

1<br />

0<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

HpWS2b-1<br />

HpWS2b-2<br />

HpWS2b-3<br />

HpWS2b-4<br />

HpWS2b-5<br />

HpWS2b-6<br />

HpWS2b-7<br />

HpWS2b-8<br />

HpWS2b-9<br />

HpWS2b-10<br />

HpWS2b-11<br />

HpWS2b-12<br />

HpWS2b-13<br />

HpWS2b-14<br />

HpWS2b-15<br />

HpWS2b-16<br />

HpWS2b-17<br />

HpWS2b-18<br />

HpWS2b-19<br />

HpWS2b-20<br />

HpWS2b-21<br />

HpWS2b-22<br />

HpWS2b-23<br />

HpWS2b-24<br />

HpWS2b-25<br />

HpWS2b-26<br />

HpWS2b-27<br />

HpWS2b-28<br />

HpWS2b-29<br />

HpWS2b-30<br />

HpWS2b-31<br />

HpWS2b-32<br />

HpWS2b-33<br />

HpWS2b-34<br />

Figure 1. WSMV inoculation of hpWS2b T 1 transgenic family. Shown are<br />

the ELISA ratio at 14 dpi, symptom severity and plant height at booting<br />

stage. The asterisk shows which plants amplified both ends of the hairpin<br />

transgene. This family showed simple mendelian inheritance of the<br />

transgene cosegregating with the immunity.<br />

66 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Characterising inositol signalling pathways in Phytophthora spp. for future<br />

development of selective antibiotics<br />

D. Phillips{ XE "Phillips, D." } 1 , D. Cahill 2 and P.L. Beech 1<br />

1 Centre for Cellular and Molecular Biology, Deakin University, Burwood Vic 3125<br />

2 Deakin University, Waurn Ponds Vic 3216<br />

Phytophthora is probably best known for causing the Irish potato<br />

famine of the 1840s, but this plant pathogen is not just an issue<br />

of history. Recent estimates show P.sojae to cause $1–2 billion in<br />

soy bean and $400 million in tomato crop losses p.a., and that<br />

P.infestans costs ~$290 million p.a. in management and losses of<br />

potato crops in the USA (e.g. 1). In Australia, P. cinnamomi is one<br />

of the greatest risks to our terrestrial ecosystems: it destroys<br />

numerous non‐arid habitats, having a wide host range of up to<br />

2500 plant species (2), and is thus considered a key threat under<br />

the Commonwealth Environmental Protection and Biodiversity<br />

Conservation Act 1999. Development of a control for<br />

Phytophthora spp. is critical to future sustainable agriculture and<br />

will be invaluable in maintaining numerous ecosystems in<br />

Australia and abroad. This project aims to directly target<br />

Phytophthora in the same manner used to develop the antiinfluenza<br />

drug, Relenza⎢(3) by identifying and solving the<br />

structure of a protein(s) unique to and critical for the survival of<br />

Phytophthora, we aim to provide the information required for<br />

future development of customised antibiotics.<br />

Only a handful of biological components are so critical to life that<br />

they are conserved throughout all organisms. For example,<br />

ribosomes are required for protein synthesis and as such are<br />

found in every living cell. So too, several signaling cascade<br />

enzymes appear to have been conserved across life;<br />

phospholipase C (PLC) is one such enzyme conserved from<br />

bacteria to humans and, just as the loss of ribosomes would be<br />

fatal, the loss of PLC would be catastrophic to the cell. How is it<br />

then, that plant pathogens of the genus Phytophthora do not<br />

have any recognizable PLC (4)? Phospholipase C is a transient<br />

membrane protein that, upon GTP activation, hydrolyses the<br />

phospholipid phosphoinositide bisphosphate (PIP 2 ) into the<br />

secondary messengers (1,4,5) inositol triphosphate (IP 3 ) and<br />

diacylglycerol (DAG), which inter alia activate protein kinase<br />

pathways, phospholipase D pathways and mediate rapid calcium<br />

release from the endoplasmic reticulum (5).<br />

We propose that PLC has been replaced by an alternative<br />

protein we call AltPLC. Existence of such an alternate protein<br />

would not only represent significant insight into the evolution of<br />

Phytophthora, but may indeed represent an ideal target for anti‐<br />

Phytophthora antibiotics.<br />

We have approached the problem of identifying the AltPLC from<br />

three directions, utilising structural bioinformatics, differential<br />

proteomics, and biochemical analysis.<br />

METHODS<br />

Structural bioinformatics: We have identified all proteins within<br />

the P. sojae genome which bind to PIP 2 using Hidden Markov<br />

Models to search for patterns which convey the structure of<br />

Plekstrin homology domains –a structure known to specifically<br />

bind to PIP 2. This data set was then scrutinised by a number of<br />

domain‐architecture mapping and structural prediction<br />

algorithms.<br />

lipid‐regulating proteins to the membrane fragment. After<br />

washing the membrane, elution is achieved by removing Ca 2+<br />

and allowing dissociation. Analysis of these fractions was<br />

performed by MS/MS and in vitro hydrolysis reactions.<br />

Biochemical analysis: IP 3 was isolated using the method of Lorke<br />

et al. (2004)(6) and analyzed by MDD‐HPLC (7).<br />

RESULTS<br />

Using MDD‐HPLC we have shown that P. cinnamomi does<br />

produce IP 3 endogenously and DAG in vitro by hydrolysis<br />

reactions with differentially isolated transient membrane protein<br />

fractions. This evidence supports our hypothesis of an<br />

alternative PLC in the Phytophthora genus. Using our<br />

combinatorial bioinformatics approach we have uncovered a<br />

single protein with all necessary structural components to<br />

perform PIP 2 hydrolysis. Furthermore, this protein is conserved<br />

among P. sojae P. ramourum and P. infestans and, as<br />

hypothesised, is unique to the Phytophthora genus. We have<br />

cloned and continue to isolate, recombinant AltPLC and its<br />

activator RAS protein for functional and structural analysis.<br />

although final correlation between PIP 2 hydrolysis and our<br />

putative AltPLC protein has yet to be achieved. Beyond the<br />

obvious development of novel control methods, identification of<br />

a phospholipase C protein of independent evolutionary origin is<br />

a unique and significant discovery that may ultimately aid in<br />

elucidating / refining the evolutionary origins of Phytophthora,<br />

and give us an insight into the process of independent<br />

convergence events in general terms.<br />

REFERENCES<br />

1. Tyler,B.M., Henkart,M. (2005) Genome information from plant<br />

destroyers could save trees, beans and chocolate. National science<br />

foundation<br />

press.<br />

http://www.nsf.gov/news/news_summ.jsp?cntn_id=107973&org=<br />

NSF&from=news<br />

2. Hardham, A. (2005) Phytophthora cinnamomi. Molecular <strong>Plant</strong><br />

<strong>Pathology</strong> 6: 589–604<br />

3. Coleman,P,M., Hoyne,P,A., Lawrence,M,C. (1983) Sequence and<br />

structure alignment of paramyxovirus hemagglutininneuraminidase<br />

with influenza virus neuraminidase. Journal of<br />

Virology 67: 2972–2980<br />

4. Tyler BM. et al. (2006) Phytophthora genome sequences uncover<br />

evolutionary origins and mechanisms of pathogenesis. Science 313:<br />

1261–1266.<br />

5. Durrell,J. Sodd,M,A. Friedel,R.O. (1968) Acetylcholine stimulation of<br />

phosphodiesteratic cleavage of the guinea pig brain<br />

phosphoinositides. Journal of Life Science 7: 363–368<br />

6. Lorke DE. Gustke H, Mayr GW (2004) An Optimized Fixation and<br />

Extraction Technique for High Resolution of Inositol Phosphate<br />

Signals in Rodent Brain Neurochemical Research, Vol. 29, No. 10,<br />

pp. 1887–1896<br />

7. Mayr GW (1988) A novel metal‐dye detection system permits<br />

picomolar‐range h.p.l.c. analysis of inositol polyphosphates from<br />

non‐radioactively labelled cell or tissue specimens Biochem. J.<br />

(1988) 254, 585–591<br />

Session 4A—<strong>Plant</strong> pathogen interactions<br />

Differential proteomics: we have developed a method of<br />

isolating transient membrane proteins. This involves cracking the<br />

cells under high calcium and low temperature conditions, to bind<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 67


Session 4A—<strong>Plant</strong> pathogen interactions<br />

INTRODUCTION<br />

Systemic acquired resistance—a new addition to the IPM clubroot toolbox?<br />

A. Agarwal A , E.C. Donald{ XE "Donald, E.C." } A , R. Faggian B , D.M. Cahill C , D. Lovelock C and I.J. Porter A<br />

A<br />

Department of Primary Industries Victoria, Private Bag 15, Ferntree Gully Delivery Centre, 3156, Victoria<br />

B<br />

Department of Primary Industries Victoria, 32 Lincoln Square Nth, Carlton, 3053, Victoria<br />

C Deakin University, School of Life and Environmental Sciences, Waurn Ponds Campus, Geelong, 3217, Victoria<br />

Clubroot caused by Plasmodiophora brassicae affects the<br />

Brassicaceae family of plants causing root galling, stunting and<br />

wilting of many important vegetable crops. There has been no<br />

‘silver bullet’ solution to clubroot but a number of ‘tools’ are<br />

available to manage the disease. Integrated use of these ‘tools’,<br />

including detection of P. brassicae and prediction of yield loss<br />

due to clubroot, identification and elimination of hygiene risks<br />

together with in‐field cultural methods, use of resistant varieties,<br />

manipulation of soil pH, calcium and boron amendment and<br />

strategic use of pesticides has been extremely effective in<br />

vegetable production systems (1).<br />

Figure 1. Arabidopsis plants 50 days after inoculation with P. brassicae.<br />

<strong>Plant</strong>s on the right have been pretreated with salicylic acid to induce<br />

SAR.<br />

Microarray analysis conducted at the early time points during<br />

the infection process of P. brassicae in Arabidopsis (4, 7 and 10<br />

days after inoculation) identified a number of genes and<br />

pathways that may regulate disease expression in Arabidopsis<br />

(2). Manipulation of the salicylic acid (SA) signalling pathway may<br />

induce systemic acquired resistance (SAR), a state of heightened<br />

defensive capacity in plant species. This paper describes<br />

preliminary experiments to study the effect of SA as an inducer<br />

of SAR in Arabidopsis and broccoli, and assess the potential for<br />

SAR to be incorporated into the IPM ‘toolbox’ for clubroot<br />

management.<br />

MATERIALS AND METHODS<br />

A proof of concept study was conducted using Arabidopsis.<br />

Roots were treated with 0.5 mM SA for 1 minute and then<br />

inoculated with P. brassicae resting spores 4 hours after<br />

treatment. <strong>Plant</strong>s were assessed for disease expression 50 days<br />

after inoculation.<br />

A broader range of SA dip rates (1–10 mM) and contact times<br />

were evaluated in order to induce SAR in broccoli. <strong>Plant</strong>s were<br />

inoculated with a spore suspension of P. brassicae 24 hours after<br />

treatment and assessed for disease expression 6 weeks after<br />

inoculation. A real‐time reverse transcriptase quantitative PCR<br />

(RT‐qPCR) assay was developed to determine the expression of<br />

the chitinase gene in broccoli roots and leaves. Biochemical<br />

methods are also being developed to confirm SAR induction.<br />

RESULTS AND DISCUSSION<br />

Clubroot disease was strongly suppressed in salicylic acid treated<br />

Arabidopsis plants (Fig 1). Fifty days post‐inoculation SA treated<br />

plants had a much lower disease index and infection rate (DI=20,<br />

IR=50%) compared to untreated plants (DI=81.5, IR=100%).<br />

A 15 min root dip in 1 mM SA 24 hours before inoculation was<br />

the most effective method of SAR induction in broccoli. This<br />

treatment consistently increased expression of the chitinase<br />

gene by between 2.3 and 5.5 fold in roots and leaves confirming<br />

a systemic response. At concentrations in excess of 1 mM SA,<br />

changes in the expression of the chitinase gene were less<br />

consistent. Frequently these higher concentrations of SA caused<br />

a decrease in the expression of the chitinase gene. At the higher<br />

rates SA might not be translocated or it may alter the physiology<br />

of the plant. A similar result (ie. increased control only at the<br />

lowest rate 1 mM) was obtained from disease expression studies<br />

using broccoli (Fig 2). SA was phytotoxic to plants at 10 mM.<br />

Figure 2. Symptoms of clubroot on roots of broccoli plants 6 weeks after<br />

inoculation which occurred 24 hrs after treatment with SA (clockwise<br />

from top left 0, 1, 2.5 and 10 mM SA). Each image shows the range of<br />

symptoms in each treatment group.<br />

This is the first evidence that SAR induction may be a useful<br />

addition to the IPM clubroot ‘toolbox’. Work is ongoing to<br />

further optimise rates and timing of application of SA, to identify<br />

and evaluate other inducers and to extend the work to other<br />

pathogens.<br />

ACKNOWLEDGEMENTS<br />

This work has been funded by DPI Victoria and Horticulture<br />

Australia Limited (HAL) using the vegetable levy and matched<br />

funds from the Australian Government.<br />

REFERENCES<br />

1. Donald C, Porter I (2009) Integrated control of clubroot. Journal of<br />

<strong>Plant</strong> Growth Regulation (In Press). doi: 10.1007/s00344‐009‐9094–<br />

7<br />

2. Agarwal A (2009) Interactions of Plasmodiophora brassicae with<br />

Arabidopsis thaliana. PhD thesis, Deakin University.<br />

68 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Prevalence and pathogenicity of Botryosphaeria lutea isolated from grapevine<br />

nursery materials in New Zealand<br />

INTRODUCTION<br />

R.G. Billones{ XE "Billones, R.G." }, E.E. Jones, H.J. Ridgway and M.V. Jaspers<br />

Lincoln University, PO Box 84, Lincoln 7647, Canterbury, New Zealand<br />

Botryosphaeria species are considered important pathogens of<br />

grapevines worldwide; they are associated with dieback in<br />

mature vines and decline of young vines. They have been<br />

isolated from scion and rootstock canes in France and Spain (1,<br />

2). A 2008 survey of nine grapevine nurseries around New<br />

Zealand showed that 23% of the canes and grafted plants<br />

collected were infected with Botryosphaeria spp. From the<br />

isolates recovered, 59% were identified as B. lutea (unpublished<br />

data), making it the most important of the Botryosphaeria<br />

species to threaten New Zealand vineyards.<br />

This paper reports the distribution and prevalence of B. lutea in<br />

plant materials from different grapevine nurseries in New<br />

Zealand, as well as the variation between nurseries in<br />

pathogenicity of their isolates.<br />

MATERIALS AND METHODS<br />

Isolation and Identification of Botryosphaeria spp. from<br />

nursery plant materials. <strong>Plant</strong> materials comprising 5–15<br />

samples of each tissue type [apparently healthy grafted plants,<br />

failed grafted plants (or Grade 2 plants), scion and rootstock<br />

cuttings of different varieties] were collected from 9 grapevine<br />

nurseries from different climatic zones in New Zealand.<br />

Isolations were made from the surface‐sterilised plant samples,<br />

with 0.5 cm pieces cut from different parts of each sample<br />

placed onto potato dextrose agar with 0.5 g/L streptomycin<br />

sulphate (PDAS). Plates were incubated for 72 h and<br />

Botryosphaeria‐like colonies were subcultured onto prune<br />

extract agar (PEA) plates to induce sporulation. Isolates were<br />

later identified by conidial characteristics and by molecular<br />

methods.<br />

Pathogenicity Tests of B. lutea. Mycelium plugs from 4 day old<br />

B. lutea PDA cultures were inoculated onto the wounds created<br />

when the shoot tips were cut from rooted one‐year‐old<br />

Sauvignon blanc canes. Four rooted cuttings were used per<br />

isolate and control plants were inoculated with sterile agar. The<br />

inoculated plants were kept in the greenhouse for 28 days and<br />

then the bark peeled off so that the cane lesions could be<br />

measured. Re‐isolation onto PDAS was done with 1 cm sections<br />

of the canes that were cut from 0 to 5 cm beyond each lesion.<br />

The plates were incubated in the dark for 72 h at room<br />

temperature and assessed for characteristic growth of B. lutea.<br />

RESULTS<br />

Botryosphaeria lutea Prevalence and Distribution. B. lutea was<br />

isolated from seven of the nine nurseries sampled (Table 1).<br />

However, one of the two nurseries that were negative for B.<br />

lutea submitted only part of the sample types requested. The<br />

presence of B. lutea in different grapevine nurseries was<br />

statistically significant using the Pearson Chi‐square test<br />

(P


Session 4B—Disease surveys<br />

Infection and disease progression of Neofusicoccum luteum in grapevine plants<br />

INTRODUCTION<br />

N.T. Amponsah{ XE "Amponsah, N.T." }, E.E. Jones, H.J. Ridgway and M.V. Jaspers<br />

Department of Ecology, Faculty of Agriculture and Life Sciences, PO Box 84, Lincoln University, New Zealand<br />

Species of the Botryosphaeriaceae are major pathogens of<br />

grapevines worldwide that are frequently found to be associated<br />

with trunk and cane dieback, internal necrotic stem tissues and<br />

bud mortality (3). An accurate monitoring of disease progression<br />

is therefore important to evaluate disease susceptibility in<br />

grapevine plants. Although some fungi have evolved a variety of<br />

morphogenic strategies to enter plants using infection structures<br />

such as appressoria (2), the invasion of grapevine tissues by<br />

species of the Botryosphaeriaceae mostly occurs through<br />

wounds made by pruning or other injuries (unpublished data).<br />

Initial pathogenicity experiments indicated that shoots of the<br />

major grapevine cultivars grown in New Zealand were equally<br />

susceptible to infection, with Neofusicoccum luteum being the<br />

most prevalent and pathogenic species. The aim of this research<br />

was to investigate infection processes and the progression of N.<br />

luteum infection on grapevine leaves and shoots.<br />

MATERIALS AND METHODS<br />

Conidia of N. luteum were obtained by inducing their production<br />

on infected, green grapevine shoots (1). The conidial<br />

suspensions used for all inoculations were adjusted to a final<br />

concentration of 10 4 conidia/mL. The grapevine plants used were<br />

18 months old potted Pinot noir growing in a shade house.<br />

Wounding on leaves and shoots was done by scraping the<br />

surface layer of a tiny section (2–5 mm 2 ) with a sterile scalpel,<br />

after which a drop of the conidial suspension was applied onto<br />

both wounded and non wounded sections of attached or<br />

detached leaves and shoots. Controls were inoculated with<br />

sterile distilled water. There were six replicates for each type of<br />

inoculated site.<br />

The detached leaf and stem tissues were observed by SEM 24 h<br />

after inoculation. For the attached tissues, plants were grown for<br />

4 months and watered daily. Scanning electron microscopy<br />

(SEM) observation or pathogen re‐isolation was done on the<br />

leaves 24 to 72 h after inoculation and on the shoots at monthly<br />

intervals for 4 months. Pathogen re‐isolation from the shoots<br />

tissues were taken at 1 cm intervals below and above the<br />

inoculation point.<br />

RESULTS<br />

No infection occurred in any non wounded shoots or leaves<br />

(attached or detached); no fungal cultures characteristic of N.<br />

luteum were isolated from these tissues. The SEM observations<br />

of inoculation sites on the unwounded attached leaves revealed<br />

no conidia on the surfaces at 24 h after inoculation. However, on<br />

the surfaces of the attached (wounded) shoots and leaves, the<br />

conidia were observed to have germinated, with the germ tubes<br />

penetrating into the wounded tissue at 24 h after inoculations<br />

(Fig. 1A). On the detached (wounded) shoots and leaves, there<br />

were networks of mycelium at 24 h after the inoculations. Reisolation<br />

from these tissues yielded 100% N. luteum.<br />

Further SEM of the pathogen at 3 months after inoculation onto<br />

the wounded shoots showed long threads of mycelium growing<br />

in through the vessels (Fig 1B). Detection of pathogen movement<br />

by monthly re‐isolation at 1 cm intervals below and above the<br />

inoculation points showed pathogen progression increased with<br />

time, being greater in the upward than the downward direction<br />

(Fig. 2).<br />

A<br />

Figure 1. Development of N. luteum on wounded grapevine tissue (A)<br />

arrow shows conidia germinating on an attached leaf surface after 24 h,<br />

(B) mycelium growing through shoot xylem vessel 3 months after<br />

inoculations.<br />

Distance moved by<br />

pathogen (mm)<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

-10<br />

-20<br />

Upwards<br />

Downwards<br />

Jan-08 Feb-08 Mar-08 Apr-08<br />

Dates on which pathogen was assessed<br />

Figure 2. Recovery of N. luteum from tissue above and below wounded<br />

inoculation sites of 18 months old Pinot noir grapevines.<br />

DISCUSSION<br />

No infection by N. luteum conidia was seen on non wounded<br />

tissue and no conidia were observed at 24 h after the<br />

inoculations on non wounded (attached) leaf surfaces. This<br />

suggests that the conidia could not attach to the surfaces and<br />

were lost from them. The faster development of germinating<br />

conidia on the attached shoots and leaf surfaces (wounded) than<br />

on the detached tissues at 24 hr after infection could be due to<br />

active inhibitory plant‐pathogen interactions. This may also<br />

explain why plants under water stress had enhanced<br />

susceptibility to infection. In stems of 18 months old vines,<br />

movement of the pathogen was faster in the upward than<br />

downward direction, possibly because it followed xylem flow,<br />

SEM observations showed pathogen presence in xylem vessels,<br />

which may allow for the non‐symptomatic progression<br />

previously observed (unpublished data)<br />

ACKNOWLEDGEMENTS<br />

The authors gratefully acknowledge funding from the New<br />

Zealand Winegrowers Inc.<br />

REFERENCES<br />

1. Amponsah N.T, Jones E.E, Ridgway, H.J, and Jaspers MV (2008).<br />

Production of Botryosphaeria species conidia using grapevine green<br />

shoots. New Zealand <strong>Plant</strong> Protection, 61, 301–305<br />

2. Egan M.J and Talbot N.J (2008). Genomes, free radicals and plant<br />

cell invasion: recent developments in plant pathogenic fungi.<br />

Current Opinion in <strong>Plant</strong> Biology, 11(4), 367–372.<br />

3. Phillips A. J. L. (1998). Botryosphaeria dothidea and other fungi<br />

associated with excoriose and dieback of grapevines in Portugal.<br />

Journal of Phytopathology 146, 327–332.<br />

B<br />

70 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Carbohydrate stress increases susceptibility of grapevines to Cylindrocarpon black<br />

foot disease<br />

INTRODUCTION<br />

D.S. Dore{ XE "Dore, D.S." }, E.E. Jones, H.J. Ridgway and M.V. Jaspers<br />

Agriculture and Life Sciences Faculty, Lincoln University 7647, Canterbury, New Zealand<br />

Stress can predispose woody plants to pathogen infection,<br />

increasing both disease incidence and severity (1). Defoliation by<br />

means of leaf plucking can stress a plant by decreasing the<br />

availability and concentration of photosynthate, compromising<br />

its resistance to other biotic and abiotic stresses (2). This<br />

experiment investigated the effect of carbohydrate stress, as<br />

induced by leaf plucking of grapevines, on the disease severity<br />

and incidence of black foot disease, a serious threat to vineyards<br />

around the world. This disease is caused by Cylindrocarpon<br />

species including C. destructans, C. macrodidymum, and C.<br />

liriodendri (3).<br />

MATERIALS AND METHODS<br />

<strong>Plant</strong>s (1 year old) of Sauvignon Blanc scion wood grafted to<br />

rootstocks 101–14 or Schwarzmann (ten per treatment), were<br />

grown in a 50/50 mix of vineyard soil and potting mix. Pots were<br />

laid out in a completely randomised design in a greenhouse in<br />

September 2006. Leaves were plucked (Nov 2006) from scion<br />

shoots above the fourth node, to induce three different levels of<br />

carbohydrate stress, being none (level 0; no removal), moderate<br />

(level 1; every third leaf removed) and high (level 2; every third<br />

leaf was left). Plucking treatments were performed three times<br />

at three weeks apart on new shoot growth and then the root<br />

systems of all the vines were wounded by slicing down into the<br />

soil, four times around each plant. For each stress level, half the<br />

plants were inoculated with 50 mL (per plant) of a mixed (three<br />

C. destructans isolates) conidial suspension (10 6 /mL) poured<br />

over the soil surface followed by 50 mL of tap water. The<br />

remaining plants were treated with 100 mL of tap water.<br />

varieties, 101–14 (19.2 g) and Schwarzmann (16.8 g), which<br />

responded differently to stress (P=0.062), being 17.9 g and 13.2<br />

g, respectively in the highly stressed treatment. There was no<br />

significant effect of carbohydrate stress (P=0.259) or inoculation<br />

(P=0.885) on shoot dry weight<br />

Table 1. Effects of the three carbohydrate stress treatments (analysis a),<br />

and two stress treatments (0+1 compared with 2; analysis b) on C.<br />

destructans disease severity (% infected wood pieces) of grapevines.<br />

Data are combined averages for inoculated and uninoculated plants for<br />

two rootstock varieties.<br />

Stress level *<br />

0:none 7.15<br />

1:moderate 8.15<br />

Disease<br />

severity Stress level **<br />

Disease<br />

severity<br />

0 + 1 7.60<br />

2:high 19.65 2 19.65<br />

P value P=0.138 P=0.043<br />

*Analysis a<br />

**Analysis b<br />

DISCUSSION<br />

This study showed that carbohydrate stress caused by leaf<br />

plucking significantly increased the severity of black foot disease<br />

and decreased root dry weight. There was also some indication,<br />

although not significant, that Schwarzmann was more affected<br />

by carbohydrate stress than 101–14. These results are relevant<br />

to the industry since canopy thinning is a regular practice in the<br />

vineyard. Hunter et al. (4) reported that no negative effects were<br />

thought to be associated with the practice; however, in light of<br />

these results care should be taken with deciding the intensity of<br />

canopy thinning in areas at risk to Cylindrocarpon infection.<br />

Session 4B—Disease surveys<br />

<strong>Plant</strong>s were grown for a further six months prior to assessment.<br />

Root and shoot dry weights were recorded and isolations made<br />

by plating sections of surface sterilised trunk tissue onto potato<br />

dextrose agar. Plates were incubated at 20˚C for 7 d and<br />

assessed for the presence of C. destructans colonies (3).<br />

RESULTS<br />

Cylindrocarpon destructans disease severity and incidence was<br />

similar for both Schwarzmann (8.4% and 29.3%, respectively)<br />

and 101–14 rootstocks (14.9% and 31.0%, respectively).<br />

Although not significant, the highest disease severity was seen<br />

with high carbohydrate stress compared with moderate or no<br />

stress (Table 1). However, when data for the moderate and no<br />

stress treatments were combined (because the effects were<br />

similar), the disease severity was significantly higher for the<br />

highly stressed plants (P=0.043). Stress did not influence disease<br />

incidence (P=0.551). Infection also occurred in the un‐inoculated<br />

plants, due to the soil being infested by Cylindrocarpon spp., but<br />

disease severity was higher in the plants inoculated with C.<br />

destructans (40.5%) than those that were not (19.5%).<br />

ACKNOWLEDGEMENTS<br />

New Zealand Winegrowers and Lincoln University for funding<br />

this project.<br />

REFERENCES<br />

1. Schoeneweiss DF (1981) The role of environmental stress in<br />

diseases of woody plants. <strong>Plant</strong> Disease 65, 308–314<br />

2. Marcais B, Breda N (2006) Role of an opportunistic pathogen in the<br />

decline of stressed oak trees. Journal of Ecology 94, 1214–1223<br />

3. Halleen F, Schroers HJ, Groenwald JZ, Crous PW (2004) Novel<br />

species of Cylindrocarpon (Neonectria) and Campylocarpon gen.<br />

nov. associated with black foot disease of grapevines (Vitis spp.).<br />

Studies in Mycology 50, 431–455<br />

4. Hunter JJ, Ruffner HP, Volschenk CG, Le Roux DJ (1995) Partial<br />

defoliation of Vitis vinifera L. cv. Carbernet Sauvignon/99 Richter:<br />

Effect on root growth, canopy efficiency, grape composition and<br />

wine quality. American Journal of Enology and Viticulture 46, 306–<br />

314.<br />

Root dry weights were similar in inoculated and uninoculated<br />

plants, but were significantly lower for highly stressed plants<br />

(15.6 g) than both the moderately stressed (19.9 g; P=0.000) and<br />

unstressed plants (18.5 g; P=0.003). An interaction between<br />

inoculation and stress (P=0.031) showed that inoculated and<br />

highly stressed plants had the lowest root dry weight (14.9 g).<br />

Root dry weights differed (P=0.0001) between the rootstock<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 71


Session 4B—Disease surveys<br />

Botryosphaeria spp. associated with bunch rot of grapevines in south‐eastern<br />

Australia<br />

N. Wunderlich{ XE "Wunderlich, N." } A , G.J. Ash A , C.C. Steel A , H. Raman B and S. Savocchia A<br />

A National Wine and Grape Industry Centre, School of Agricultural and Wine Sciences, Charles Sturt University, Locked Bag 588, Wagga<br />

Wagga, 2678, NSW<br />

B New South Wales Department of Primary Industries, Private Mail Bag, Wagga Wagga, 2650, NSW<br />

INTRODUCTION<br />

Species of Botryosphaeria are common wood pathogens of<br />

grapevines and are responsible for the disease known as ‘Bot<br />

canker’ (1). Recently some species have also been implicated in<br />

bunch rot of Vitis vinifera in Australia (2). While pathogenicity<br />

tests have been conducted on grapevine wood for several<br />

species, it is unknown which of these infect bunches and how<br />

they enter the berry.<br />

MATERIALS AND METHODS<br />

<strong>Plant</strong> survey. Between 2007 and 2009 two vineyards in the<br />

lower Hunter Valley, NSW, were sampled for species of<br />

Botryosphaeria. Samples were collected from 200 each of<br />

Chardonnay and Shiraz grapevines symptomatic of Bot canker at<br />

different phenological stages: dormant buds (B), flowers (F), peasized<br />

berries (P) and berries at harvest (H). Samples were also<br />

collected from the margin of healthy and discoloured internal<br />

wood (W) from the trunks of each plant.<br />

Fungal isolation and identification. Samples were surface<br />

sterilised, placed onto Potato Dextrose Agar (PDA) amended<br />

with Streptomycin Sulfate and incubated at 25ºC in the dark.<br />

Fungal cultures characteristic of Botryosphaeria spp. were subcultured<br />

onto PDA and/or triple autoclaved pine needles on 1%<br />

water agar maintained under near UV light (12hr light/dark) to<br />

encourage sporulation. Botryosphaeria spp. were identified<br />

according to spore morphology and sequencing of the rDNA<br />

internal transcribed spacer region.<br />

Pathogenicity tests on berries. Disease‐free, surface sterilised<br />

Shiraz and Chardonnay berries at harvest were inoculated with<br />

10 μL spore suspensions from 19 isolates belonging to the<br />

species listed in table 1. Conidial suspensions at concentrations<br />

of 10 6 and 10 4 spores/mL were used in trial 1 and 2, respectively.<br />

Control berries were inoculated with 10 μL of sterile distilled<br />

water. The berries were incubated in 24 well plates at 27ºC in<br />

the dark for 15 days. A constant relative humidity was<br />

maintained by the addition of 20 mL of sterile water to each<br />

plate. Disease incidence (%) and severity (1–10) was recorded for<br />

each treatment. A disease index (DI) was calculated for each<br />

treatment replicate at each time of recording:<br />

DI = ∑diseased berries ÷ ∑total berries X ∑disease severity<br />

scores ÷ sum max disease severity scores.<br />

mutila and Lasiodiplodia theobromae. Abundance and origin of<br />

each species are listed in Table 1.<br />

Table 1. Botryosphaeria spp. isolated from grapevine tissue.<br />

Species<br />

Origin tissue<br />

B F P H W<br />

D. seriata 83 2 3 6 27<br />

B. dothidea 15 0 0 3 7<br />

N. parvum 7 1 0 2 4<br />

N. luteum 3 0 0 2 2<br />

D. viticola 3 0 0 0 2<br />

D. mutila 0 0 0 0 4<br />

L. theobromae 0 0 0 0 1<br />

Pathogenicity tests All isolates produced bunch rot symptoms<br />

on berries including the formation of mycelia and pycnidia,<br />

darkening of the berry skin, oozing and berry collapse. Disease<br />

indices for each replicate treatment varied significantly from<br />

control berries. Variation in the rate of increase of the disease<br />

index was detected within and between species. Inoculation of<br />

canes showed lesion development and pycnidia formation on<br />

the cane surface. There were significant variations in lesion<br />

lengths between species. In both berry and cane pathogenicity<br />

tests, virulence appeared to be independent from the origin of<br />

the isolate.<br />

DISCUSSION<br />

Results suggest that Botryosphaeria spp. have the potential to<br />

contribute to grapevine bunch rots and infect grapevine canes.<br />

Botryosphaeria spp. appear to be non‐tissue specific in their<br />

pathogenicity toward grapevine. Trials are in progress to<br />

establish if these results can be reproduced under field<br />

conditions and whether the infection of buds or flowers by<br />

Botryosphaeria results in bunch rot at harvest.<br />

ACKNOWLEDGEMENTS<br />

The authors would like to thank Chris Haywood for assistance<br />

with field sampling and the growers for access to their vineyards.<br />

This work was supported by a National Wine and Grape Industry<br />

Centre (NWGIC) Scholarship and through the Wine Growing<br />

Futures program, a joint initiative of the Grape and Wine<br />

Research and Development Corporation and the NWGIC.<br />

Pathogenicity tests on canes. Detached one year old canes were<br />

inoculated with 4 mm diameter mycelium plugs of 14<br />

Botryosphaeria isolates, previously tested on berries, by<br />

inserting the plugs into the wood. Canes were incubated on<br />

moist filter paper in Petri dishes at 27ºC in the dark. After 15<br />

days, lesion lengths were measured for each isolate.<br />

RESULTS<br />

Survey and fungal identification. To date, a collection of 177<br />

isolates of Botryosphaeria spp. has been established. The species<br />

isolated included Diplodia seriata, Botryosphaeria dothidea,<br />

Neofusicoccum parvum, N. luteum, Dothiorella viticola, Diplodia<br />

REFERENCES<br />

1. Savocchia S, Steel CC, Stodart, BJ, Somers A (2007). Pathogenicity<br />

of Botryosphaeria species isolated from declining grapevines in<br />

sub‐tropical regions of eastern Australia. Vitis 46, 27–32.<br />

2. Taylor AS (2007) ‘Scoping study on the non‐Botrytis bunch rots that<br />

occur in Western Australia’. Grape and Wine Research<br />

Development Corporation Final Report RT 05/05‐2.<br />

72 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Honey bees—do they aid the dispersal of Alternaria radicina in carrot seed crops?<br />

INTRODUCTION<br />

R.S. Trivedi{ XE "Trivedi, R.S." } 1 , J.G. Hampton 1 , M.V. Jaspers 2 , J.M. Townshend 3 and H.J. Ridgway 2<br />

1 Bio‐Protection Research Centre, PO Box 84, Lincoln University, Canterbury, New Zealand<br />

2 Department of Ecology, PO Box 84, Lincoln University, Canterbury, New Zealand<br />

3 Midlands Seed Ltd, PO Box 65, Ashburton, Canterbury, New Zealand<br />

A. radicina is a seed and soil‐borne pathogen of carrot (Daucus<br />

carotae) crops now common in Canterbury, New Zealand. A.<br />

radicina causes black rot disease, but can also reduce seed<br />

germination and seedling establishment. For carrot seed<br />

production, honey bees (Apis mellifera) are introduced into the<br />

crop to improve pollination. During this process, if bees visit A.<br />

radicina infected umbels and pathogen spores adhere to them,<br />

they may disperse the pathogen within the crop.<br />

MATERIALS AND METHODS<br />

To test this hypothesis, both live and dead bees were collected<br />

from around beehives that had been placed in three seed crops<br />

in the Ashburton region, Canterbury, NZ. Fifty bees were<br />

selected at random, placed in 100 ml sterile distilled water in a<br />

250 ml bottle and shaken using a Wrist action shaker (Griffin) at<br />

maximum speed (1000 rpm) for 15 min to dislodge any spores<br />

on the cadavers. The suspension was filtered using sterile<br />

Whatman paper No. 105 and the filtrate centrifuged in 50 ml<br />

tube at 4000 rpm for 5 min. The supernatant was discarded, the<br />

pellet reconstituted in 100 µl water and plated on A. radicina<br />

semi selective agar and incubated at 27°C. After 7 d fungi<br />

growing on these plates were isolated into pure culture. The<br />

isolated fungi were identified by morphological, cultural and<br />

sporulation characteristics. To confirm those cultures<br />

preliminarily identified as A. radicina, the fungal DNA was<br />

obtained using Puregene® DNA purification kit protocol. The<br />

universal primers ITS 4 and ITS 5 were used to amplify a portion<br />

of the rDNA. For each PCR reaction 1 µl of DNA (10 ng/μl) was<br />

mixed with 24 µl of PCR mixture containing the manufacturer’s<br />

buffer, 200 µM of each dNTP, 1.5 mM MgCl 2 , 5 pmol of each<br />

primer, 1.25 U Faststart Taq (Roche). A negative control without<br />

DNA template was also included. Amplification was done using a<br />

Mastercycler® Gradient (Eppendorf, USA) using the following<br />

thermal cycles: initial denaturation at 94°C for 3 min then 35<br />

cycles of: denaturation at 94°C for 2 min, annealing at 57°C for<br />

30 s and extension at 72°C for 1 min for 35 cycles following final<br />

extension at 72°C for 10 min and hold at 4°C. The PCR product<br />

was separated by electrophoresis on a 1% agarose gel in 1X TAE<br />

buffer and visualised using ethidium bromide dye under UV light<br />

(Versadoc Imaging Systems Model‐3000; Bio‐Rad, USA). The PCR<br />

product was sequenced in 3130 xl genetic analyser (ABI Prism,<br />

Applied Biosystems) and the obtained sequence of the rDNA was<br />

then compared with sequences present on GenBank<br />

(http://www.ncbi.nlm.nih.gov/) using a Blast search to confirm<br />

its identity.<br />

RESULTS<br />

In addition to A. radicina, a few other fungal colonies were also<br />

grown on selective agar medium. Honey bees from different<br />

carrot fields carried different amounts of inoculum of A. radicina<br />

on their bodies. The average numbers of colonies recovered<br />

from 50 bees were ~166 and spore per bee ratio was 3.3:1. The<br />

electrophoresis resulted in a single band of ~600 bp on the<br />

agarose gel (Fig 1). The DNA sequence obtained from this band<br />

was confirmed using Blast as A. radicina. The sequence had<br />

maximum similarity (100%) with accession numbers AY154704.1,<br />

DQ394073.1 and EU807870.1. The sequence from the honey bee<br />

(597 bp) was deposited in GenBank as accession number<br />

FJ958190.<br />

2000 bp‐<br />

1000 bp‐<br />

650 bp‐<br />

Figure 1. Agarose gel electrophoresis of the amplified rDNA region of A.<br />

radicina colonies recovered from honey bee bodies collected from three<br />

infected fields. Lane M contains 1 kb plus DNA ladder (Invitrogen); lane<br />

1, negative control; lane 2, infected site A; lane 3 infected site B; lane 4,<br />

infected site C.<br />

DISCUSSION<br />

The detection of A. radicina conidia (38 x 19 µm in size) on the<br />

body of honey bees suggests that they have the potential to play<br />

a role in the spread of the disease within and among the carrot<br />

seed crops. This research supports that of many past researchers<br />

that concluded some other insects and mollusc, like flea beetles<br />

(1) and slugs (2) in cabbage and, pollen beetles and seed pod<br />

weevil (3) in oil rape were able to play an important role in<br />

dispersal of Alternaria spp. Future work is now focused on<br />

quantification of A. radicina through real time PCR and the<br />

identification of other fungal species carried on the body of<br />

honey bees.<br />

ACKNOWLEDGEMENTS<br />

The Foundation for Research Science and Technology and<br />

Midlands Seed Ltd funded this research.<br />

REFERENCES<br />

M 1 2 3 4<br />

1. Dillard, H.R., Cobb, A.C. & Lamboy, J.S. 1998. Transmission of<br />

Alternaria brassicicola to cabbage by flea beetles (Phyllotreta<br />

cruciferea). <strong>Plant</strong> Disease 82: 153–157.<br />

2. Hasan, S. & Vago, C. 1966. Transmission of Alternaria brassicicola<br />

by slugs. <strong>Plant</strong> Disease Reporter. 50:764–767.<br />

3. Quak, F. 1956. De biologie en de bestrijdingsmogelikheden van de<br />

veroorzakers van spikkelziekte (Alternaria sp.) in Koolzaad<br />

(Brassicae napus) In Alternaria brassicicola and Xanthomonas<br />

Campestris pv. Campestris in organic seed production of Brassicae:<br />

Epidemiology and seed infection. <strong>Plant</strong> Research International,<br />

Wageningen.<br />

Session 4C—Epidemiology<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 73


Session 4C—Epidemiology<br />

Translating research into the field: meta‐analysis of field pea blackspot severity and<br />

yield loss to extend model application for disease management in Western Australia<br />

INTRODUCTION<br />

M.U. Salam{ XE "Salam, M.U." } A , J. Galloway A , W.J. MacLeod B , M. Seymour A , I. Pritchard A , M.J. Barbetti A,B and T. Maling A<br />

A<br />

Department of Agriculture and Food Western Australia, Locked Bag 4, Bentley Delivery Centre WA 6983, Australia<br />

B School of <strong>Plant</strong> Biology, The University of Western Australia, 35 Stirling Highway, Crawley, WA 6009, Australia<br />

Blackspot or ascochyta blight, caused predominantly by<br />

Mycosphaerella pinodes, is the most destructive foliar pathogen<br />

of field peas and causes considerable yield loss. The amount of<br />

yield loss is mainly driven by primary infection, i.e. spread of<br />

wind‐borne ascospores from infected pea stubble of previous<br />

seasons’ crops. In this paper, we present a meta‐analysis to<br />

show observed disease severity and associated yield loss in<br />

Western Australia (WA), explore the feasibility of chemical<br />

control, and describe the development and application of a<br />

weather‐based model to manage the disease.<br />

MATERIALS AND METHODS<br />

The meta‐analysis, using different forms of regressions, was<br />

done using data from 14 experiments conducted at 13 sites over<br />

8 seasons in WA. The model, “Blackspot Manager” was<br />

developed using daily weather data to predict the timing of<br />

onset, and progression, of ascospore maturity (Salam et al.,<br />

2006). The model was tested with independent field data from<br />

agricultural regions in WA. A system was developed to provide<br />

model output to the agribusiness community and farmers via<br />

internet (http://www.agric.wa.gov.au/cropdiseases).<br />

RESULTS<br />

Potential blackspot severity decreases as the season progresses<br />

and can be quantified as a function of time of sowing (results not<br />

shown). Under the “No control” scenario, the potential disease<br />

severity remains above rating 2 until early July (Fig. 1); however,<br />

it varies between regions (data not shown). Application of an infurrow<br />

fungicide cannot decrease the severity below rating 2<br />

before mid‐June; and fortnightly sprays not before mid‐May.<br />

growing season. Disease severity rating does not exceed 2 until<br />

crop has been exposed to more than predicted 40% of seasonal<br />

ascospores (Fig. 3).<br />

Potential yield reduction (%)<br />

60<br />

45<br />

30<br />

15<br />

0<br />

0 1 2 3 4 5<br />

Disease severity<br />

Figure 2. Relationship between blackspot disease severity ratings and<br />

potential yield reduction in field pea.<br />

Blackspot severity<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

0 10 20 30 40 50 60 70 80 90 100<br />

Predicted exposed ascospore-load<br />

(% of seasonal total)<br />

No control<br />

In-furrow fingicide<br />

Fortnightly spray<br />

No control (fitted)<br />

In-furrow fingicide (fitted)<br />

Fortnightly spray (fitted)<br />

Figure 3. Relationship between predicted exposures to ascospores from<br />

infected stubble and disease severity rating at flowering on standing field<br />

pea crops.<br />

Disease severity<br />

5<br />

4<br />

3<br />

2<br />

1<br />

CONCLUSION<br />

This research shows that chemical control for blackspot<br />

management is unlikely to be economical. Consequently,<br />

growers must rely of appropriate sowing dates. The outcome of<br />

this work is a model that predicts the temporal ascospore‐load<br />

and allows for the determination a sowing date that minimises<br />

potential yield loss.<br />

0<br />

27-Apr 14-May 31-May 17-Jun 4-Jul 21-Jul<br />

Time of sowing<br />

Figure 1. Fitted curves of potential severity of blackspot disease on field<br />

pea across WA in response to no control, in‐furrow fungicide or<br />

fortnightly fungicide spray.<br />

Potential yield reduction does not occur before disease severity<br />

rating 1 (Fig. 2). At disease severity rating 2, it reaches around<br />

15%, whereas, the maximum rating of 5 corresponded to a<br />

potential yield reduction of around 55%.<br />

ACKNOWLEDGEMENTS<br />

We thank the Australian Grains Research and Development<br />

Corporation (GRDC) the Department of Agriculture and Food<br />

Western Australia for supporting this research.<br />

REFERENCES<br />

1. Salam, M.U., Galloway, J., MacLeod, W.J. and Diggle, A. (2006)<br />

Development and use of computer models for managing ascochyta<br />

diseases in pulses in Western Australia. 1st International ascochyta<br />

workshop on grain legumes. 2–6 July, 2006, Le Tronchet, France.<br />

The Blackspot Manager uses the pre‐season temperature and<br />

rainfall to forecast the onset and progression of ascospore<br />

release from infected field pea stubble prior to and during the<br />

74 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Development of a model to predict spread of exotic wind and rain borne fungal pests<br />

INTRODUCTION<br />

S.A. Coventry A , J.A. Davidson B , M. Salam{ XE "Salam, M." } C and E.S. Scott A<br />

CRC for National <strong>Plant</strong> Biosecurity, FECCA House, 4 Phipps Close, Deakin, 2600, ACT<br />

A School of Agriculture, Food and Wine, The University of Adelaide, PMB1, Glen Osmond, 5064, South Australia<br />

B South Australian Research and Development Institute, GPO Box 397, Adelaide, 5001, South Australia<br />

C Department of Agriculture and Food Western Australia, PO Box 483, Northam, 6401, Western Australia<br />

Exotic fungal plant pathogens pose a threat to Australian<br />

agriculture. Some of the most devastating fungal pathogens are<br />

transported by rain splash, wind dispersal or a combination of<br />

both. The transport of fungal pathogens via rain and wind makes<br />

containment and eradication difficult. Ascochyta rabiei, causal<br />

agent of ascochyta blight of chickpea, is a wind/rain borne<br />

pathogen already present in Australia. A. rabiei, therefore,<br />

provides a suitable pathogen for modelling the potential spread<br />

of an exotic fungal pathogen dispersed by wind and rain.<br />

MATERIALS AND METHODS<br />

Field trials and laboratory studies were conducted to examine<br />

key environmental factors influencing the short distance (rain<br />

splashed) and long distance (wind borne) distribution of spores.<br />

Field trials. were undertaken at Kingsford Research Station,<br />

53 km north‐east of Adelaide in 2007 (S 34.54521, E 138. 78117),<br />

and at Turretfield Research Station, approximately 60 km northeast<br />

of Adelaide in 2008 (S 34.54760, E 138.82225). Plots (11 x<br />

11 m) at each site were planted with 3 cultivars of chickpea;<br />

Howzat (moderately susceptible), Almaz (moderately resistant)<br />

and Genesis 090 (resistant) (Figure 1). Infested stubble was<br />

placed at the centre of each plot and disease spread was<br />

recorded weekly. The percentage of plants infected in 1 m 2<br />

quadrats and the number of growing points (number of main<br />

stems and branches) for three plants of each cultivar were<br />

recorded over time. Weather data were collected via an<br />

automated weather observation system at the Roseworthy,<br />

within 13 km of the field sites.<br />

Almaz<br />

(mr)<br />

Genesis<br />

Howzat<br />

(ms)<br />

parameters. The adjusted parameters produced a model output<br />

in Mathematica that best fit field disease observations.<br />

RESULTS<br />

The model. For (a) number of growing points (Figure 2) collected<br />

from the 2007 field experiments and (b) spore dispersal in wind<br />

and rain tunnel experiments were used as parameter inputs for<br />

the model. <strong>Plant</strong> infection data from the 2007 field experiments<br />

were compared and calibrated with simulations run in the<br />

model. Data for disease incidence and severity in the 2008 field<br />

trial were then used to validate the model.<br />

Growing points (m -2 )<br />

6000<br />

5000<br />

4000<br />

3000<br />

2000<br />

1000<br />

0<br />

Gp (model)<br />

Gp (observation-Howzat)<br />

0 500 1000 1500 2000<br />

Growing degree-days<br />

Figure 2. Growing points (Gp, main stems and branches) influenced by<br />

degree days as predicted by the model and compared to field<br />

observations in chickpea cv. Howzat at Kingsford 2007.<br />

DISCUSSION<br />

The outcome of this work is a model calibrated with<br />

experimental data and tested with field observations for<br />

accuracy. Weather data are entered into the model and<br />

pathogen spread is predicted by graphical output showing<br />

disease occurrence in the field. The number of conidia produced<br />

per lesion on each of the cultivars will be estimated, to add more<br />

information to the model. When fully developed, the model will<br />

provide a basis for predictive models for exotic plant pathogens.<br />

It will also facilitate improved management of disease through<br />

forecasting, and more precise application of fungicide and timing<br />

of crop sowing.<br />

Session 4C—Epidemiology<br />

Figure 1. Aerial photograph of 11 x 11‐m chickpea plots at Turretfield in<br />

2008. Top, Almaz with moderate disease incidence; middle, Genesis 090<br />

with no disease; bottom, Howzat with severe disease.<br />

Laboratory experiments. were conducted to investigate the<br />

effect of wind speed (m/s), rain splash (ml/m) and a combination<br />

of the two factors on the dispersal of conidia in a purpose‐built<br />

wind and rain tunnel.<br />

Model development. A model for determining the spread of rain<br />

and wind‐borne pathogens was developed for ascochyta blight<br />

based on the spatiotemporal model for simulating the spread of<br />

anthracnose in lupin fields (1). The data collected from the field<br />

trials and laboratory experiments were entered as the model<br />

ACKNOWLEDGEMENTS<br />

We thank the CRC for National <strong>Plant</strong> Biosecurity, the University<br />

of Adelaide, South Australian Research and Development<br />

Institute, and Department of Agriculture and Food Western<br />

Australia for supporting this research.<br />

REFERENCES<br />

1. Diggle, A., Salam, M., Thomas, G., Yang, H., O’Connell, M. &<br />

Sweetingham, M. (2002) AnthracnoseTracer: A Spatiotemporal<br />

Model for Simulating the Spread of Anthracnose in a Lupin Field.<br />

Phytopathology, 92, 1110–1121.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 75


Session 4C—Epidemiology<br />

Psyllid transmission of huanglongbing from naturally infected Shogun mandarin to<br />

orange jasmine<br />

INTRODUCTION<br />

R. Sdoodee{ XE "Sdoodee, R." } A , W. Teeragulpisut A and P. Tippeng A<br />

A<br />

Department of Pest Management, Faculty of Natural Resources, Prince of Songkla University, Hat Yai, Thailand 90112<br />

Huanglongbing (HLB), previously known as citrus greening is a<br />

devastating disease of citrus. It affects all citrus cultivars and<br />

causes rapid decline of trees. HLB is caused by a phloem limited<br />

fastidious bacterium, Candidatus Liberibacter a gram negative<br />

bacterium belonging to alpha proteabacteria (1). At least 3<br />

species of Candidatus Liberibacter have been reported to be<br />

associated with HLB, Ca. L. africanus, Ca. L. asiaticus, and Ca. L.<br />

americanus. The two Ca. Liberibacter species, africanus and<br />

asiaticus, are transmitted by the psyllid vectors Trioza erytreae<br />

(Del Guercio) in Africa and Diaphorina citri (Kuwayama) in Asia,<br />

respectively.<br />

Candidatus Liberibacter can infect nearly all citrus species,<br />

cultivars and hybrids. Later, orange jasmine, Murraya paniculata<br />

(L.) Jack and Chinese box orange, Severnia buxifolia (Poiret)<br />

Men., have been reported to harbor HLB pathogen (2, 3).<br />

This paper reported transmission of Ca. L. asiaticus, Asian form<br />

HLB, by its psyllid vector, D. citri from naturally infected Shogun<br />

mandarin, Citrus reticulata Blanco, to orange jasmine (M.<br />

paniculata).<br />

MATERIALS AND METHODS<br />

Insect vector. Adults of disease free D. citri were caged on<br />

naturally infected Shogun mandarins grown at Prince of Songkla<br />

university experimental plot for one month. The vectors were<br />

then used in transmission test.<br />

Transmission test. The insect vectors were released to feed on<br />

healthy M. paniculata seedlings (test plants) using 0, 1 and 15<br />

insects per test plant. Healthy test plants were also placed under<br />

infected Shogun mandarin trees to obtain natural transmission.<br />

Detection of HLB pathogen. Total DNA of M. paniculata plant<br />

was extracted from 0.1–0.5 g of leaf midribs using CTAB method<br />

(3). PCR detection of Ca. Liberibacter asiaticus, HLB pathogen,<br />

was carried out with primers specific to 16S rRNA gene of Asian<br />

HLB (1).<br />

RESULTS AND DISCUSSION<br />

D. citri successfully transmitted HLB pathogen from the naturally<br />

infected shogun mandarin to M. paniculata even by single insect<br />

vector (Table 1). Transmission occurred within 7 and 9 weeks<br />

after released the insect vector on M. paniculata test plant using<br />

1 and 15 insects, respectively. For natural transmission (NT,<br />

placing test plants under diseased mandarin tree), test plant<br />

became infected within 28 weeks. Single insect transmission rate<br />

was as high as by 15 insects (50%) but it was slightly lower than<br />

the natural transmission (75%). All of HLB infected M. paniculata<br />

by insect vector showed typical symptom (Fig 1) resembling HLB<br />

infected citrus (3). HLB pathogen as tested by PCR remained in<br />

infected M. paniculata with typical symptom more than 30<br />

weeks which was contrast to a previous report (4).<br />

Table 1. Transmission of HLB pathogen (Ca. L.asiaticus) by D. citri insect<br />

vector.<br />

No. of insect 1<br />

No. of transmission<br />

/total test plant 2<br />

Transmission rate<br />

(%)<br />

0 0/4 0<br />

1 2/4 50<br />

15 2/4 50<br />

NT 3/4 75<br />

1 Number of D. citri insect vector feeding on one test plant (healthy Murraya<br />

paniculata), NT (natural transmission), healthy test plant placed under infected<br />

shogun mandarin.<br />

2 Transmission determined by HLB symptom on test plant and followed by PCR<br />

amplification of HLB DNA.<br />

A<br />

Figure 1. Infected Murraya paniculata showing typical HLB symptoms<br />

(interveinal chlorosis and blochy mottle leaves)<br />

A by single Diaphorina citri insect vector<br />

B by 15 insect vectors<br />

C by natural transmission<br />

ACKNOWLEDGEMENTS<br />

We wish to thank Thailand Research Fund (TRF) and Nakajima<br />

Peace Foundation for financial support and also to Department<br />

of Pest Management, Faculty of Natural Resources, Prince of<br />

Songkla University, for providing laboratory facilities.<br />

REFERENCE<br />

1. Jagoueix S, Bove JM, Garnier M (1996) PCR detection of two<br />

Liberibacter associated with greening disease of citrus. Molecular<br />

and Cellular Probes 10, 43–50.<br />

2. Hung TH, Wu ML, Su HJ (2000) Identification of the alternative host<br />

of the fastidious bacteria causing citrus greening disease. Journal of<br />

Phytopathology 148, 321–326.<br />

3. Sdoodee R, Tothaum P, Jumpaduang A (2008) PCR detection of<br />

Candidatus Liberibacter asiaticus from Murraya paniculata and<br />

psyllid vector in Thailand. Journal of <strong>Plant</strong> <strong>Pathology</strong> 90 (2,<br />

supplement) S2.450.<br />

4. Li T, Ke C, (2002) Detection of the bearing rate of Liberibacter<br />

asiaticum, in citrus psylla and its host plant Murraya paniculata by<br />

nested PCR. Acta Phytophylacica Sinica 29, 31–35.<br />

B<br />

C<br />

76 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Transmission of ‘Candidatus Phytoplasma australiense’ to Cordyline and Coprosma<br />

INTRODUCTION<br />

R.E. Beever{ XE "Beever, R.E." } A , M.T. Andersen B and C.J. Winks A<br />

A Landcare Research, Private Bag 92170, Auckland 1142, New Zealand<br />

B <strong>Plant</strong> and Food Research, Private Bag 92169, Auckland 1142, New Zealand<br />

Phytoplasmas are specialised plant pathogenic bacteria (class<br />

Mollicutes) that live in the phloem tissue of host plants and in<br />

the tissues of phloem‐feeding insects that transmit them. The<br />

phytoplasma “Candidatus Phytoplasma australiense” is<br />

associated in New Zealand with diseases of three naturally<br />

occurring host species, New Zealand flax, (Phormium tenax,<br />

family Phormiaceae), New Zealand cabbage tree (Cordyline<br />

australis, family Laxmanniaceae), and karamū (Coprosma<br />

robusta, family Rubiaceae), as well as one cultivated host,<br />

strawberry (Fragaria x ananassa, family Rosaceae). Previous<br />

work (4, 5) has established that “Ca. P. australiense” can be<br />

transmitted from Phormium to Phormium by the New Zealand<br />

flax planthopper Zeoliarus (Oliarus) atkinsoni (family Cixiidae). In<br />

this study, we have identified that a second species in this genus,<br />

Zeoliarus oppositus, acts as a vector between two of the<br />

naturally occurring hosts.<br />

Table 1. Phytoplasma detection in tissue samples from symptomatic<br />

cabbage trees.<br />

<strong>Plant</strong><br />

Days after<br />

insects<br />

caged<br />

Shoot<br />

apex<br />

Leaf<br />

bases<br />

Ca7 89 + not +<br />

tested<br />

Ca2 131 + + +<br />

Ca1 164 ‐ + ‐<br />

Ca9 196 + + ‐<br />

Rhizome<br />

apex<br />

DISCUSSION<br />

We conclude that Z. oppositus is a vector of “Ca. P. australiense”,<br />

capable of transmitting this phytoplasma from Coprosma to both<br />

Cordyline and Coprosma.<br />

Session 4D—Prokaryotic pathogens<br />

MATERIALS AND METHODS<br />

Small saplings of Cordyline australis and Coprosma robusta were<br />

grown from seed. Z. oppositus adults were collected from the<br />

wild by ‘beating’ symptomatic and non‐symptomatic plants of C.<br />

robusta. Transmission experiments were set up by caging 10<br />

individuals of Z. oppositus onto seedlings of the test species (10<br />

replicates).<br />

DNA was extracted from various plant tissues and whole insects.<br />

Phytoplasma presence in insects and plant samples was<br />

examined by one stage and nested PCR using the “universal”<br />

phytoplasma 16S primers (P1+P7, R16F2 + R16R2) (1).<br />

RESULTS<br />

Insect donors. Zeoliarus oppositus adults collected from the<br />

donor population in early summer were examined for the<br />

presence of phytoplasma using one stage PCR. A total of 4 were<br />

positive from 33 individuals examined, indicating an infection<br />

rate of c. 12%. Other adults were caged onto Cordyline and<br />

Coprosma seedlings. Over 60% were still alive on removal at 3–4<br />

weeks.<br />

Cabbage tree. Four of the ten plants exposed to Z. oppositus<br />

developed symptoms within one year. Symptoms closely<br />

resembled those of ‘initially‐affected’ tufts of diseased trees<br />

observed in the field (1). At least one of the tissue samples from<br />

all 4 symptomatic plants were positive using one stage PCR<br />

(Table 1).<br />

Karamū. After one year, 8 of the 10 Coprosma plants exposed to<br />

Z. oppositus showed leaf reddening of older leaves and one of<br />

these also showed dieback of the main shoot, symptoms<br />

consistent with phytoplasma infection in this host (3). Two<br />

plants that had been exposed to Z. opposites tested positive for<br />

phytoplasma.<br />

Both species of Zeoliarus are endemic to New Zealand. The<br />

ecology of Z. oppositus is consistent with the proposal that it is a<br />

vector for “Ca. P. australiense”. It is a very common species<br />

where it is found in natural and modified habitats throughout<br />

the country. Given the polyphagous nature of Z. oppositus, it is<br />

apparent that it may transmit the phytoplasma to other plants.<br />

ACKNOWLEDGEMENTS<br />

This work was funded by New Zealand’s Foundation for<br />

Research, Science and Technology and Ministry of Research,<br />

Science and Technology. We thank Lia Liefting and John Charles<br />

for useful discussions.<br />

REFERENCES<br />

1. Andersen MT, Beever RE, Sutherland PW, Forster RLS (2001)<br />

Association of ‘Candidatus Phytoplasma australiense’ with sudden<br />

decline of cabbage tree in New Zealand. <strong>Plant</strong> Disease 85, 462–469.<br />

2. Beever RE, Forster RLS, Rees‐George J, Robertson GI, Wood GA,<br />

Winks CJ (1996) Sudden decline of cabbage tree (Cordyline<br />

australis): search for the cause. New Zealand Journal of Ecology 20,<br />

53–68.<br />

3. Beever RE, Wood GA, Andersen MT, Pennycook SR, Sutherland PW,<br />

Forster RLS (2004) “Candidatus Phytoplasma australiense” in<br />

Coprosma robusta in New Zealand. New Zealand Journal of Botany<br />

42, 663–675<br />

4. Cumber RA (1953) Investigations into yellow‐leaf disease of<br />

Phormium. IV. Experimental induction of yellow‐leaf condition on<br />

Phormium tenax Forst. by the insect vector Oliarus atkinsoni Myers<br />

(Hem. Cixiidae). New Zealand Journal of Science and Technology<br />

34A (Supplement 1), 31–40.<br />

5. Liefting LW, Beever RE, Winks CJ, Pearson MN, Forster RLS (1997)<br />

<strong>Plant</strong>hopper transmission of Phormium yellow leaf phytoplasma.<br />

<strong>Australasian</strong> <strong>Plant</strong> <strong>Pathology</strong> 26, 148–154.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 77


Session 4D—Prokaryotic pathogens<br />

Australian grapevine yellows phytoplasma found in symptomless shoot tips after a<br />

heat wave in South Australia<br />

P.A. Magarey{ XE "Magarey, P.A." } A , N. Habili B , J.A. Altmann C and J.D. Prosser D<br />

A South Australia Research and Development Institute, PO Box 411, Loxton, 5333, SA, B The University of Adelaide, PMB1, Glen Osmond,<br />

SA, 5064<br />

C Fruit Doctors, Balfour Ogilvy Avenue, Loxton, South Australia, 5333, D 115 Nildottie‐Bakara Road, Nildottie, South Australia, 5238<br />

INTRODUCTION<br />

In 2008/09, extremely high incidences of Australian Grapevine<br />

Yellows (AGY) in cv. Riesling vineyards at Nildottie, near Loxton,<br />

South Australia, suggested incursion of a new yellows pathogen.<br />

To investigate, in March 2009 we surveyed vineyards, tested<br />

material via PCR and examined the role of a period of very high<br />

temperature on symptom expression. The findings challenged<br />

our previous theory on the pathogenesis of AGY.<br />

MATERIALS AND METHODS<br />

Vineyard Surveys. In Nildottie vineyards (Table 1), 50<br />

vines/block, ≥ 3 blocks/transect and ≥ 3 transects /vineyard were<br />

scored for typical AGY viz. yellowed, downward curled leaves,<br />

unlignified shoots and shrivelled bunches.<br />

PCR‐Tests. Following a heat wave in Jan‐Feb 2009 with 13<br />

consecutive days ≥ 38°C (5 ≥ 42°C), ~5 shoots with typical AGY<br />

symptoms were selected from each of several cultivars (Table 2)<br />

for replicated two‐step PCR analysis with both positive and<br />

negative controls to detect AGY phytoplasma (AGYp) as per (1,2).<br />

Samples were taken from: 1) mature, yellowed leaves; and 2)<br />

symptomless new growth on shoot‐tips.<br />

respectively) were warmer compared with March averages<br />

(28.2°C/11.8°C).<br />

Table 2. PCR‐tests of AGY‐affected shoot material and new tip‐growth in<br />

several cultivars after a heat wave, Nildottie, South Australia, March<br />

2009.<br />

Vineyard 1 #<br />

Shoots<br />

with AGYp<br />

PCR‐Tests for AGYp<br />

#<br />

Shoots<br />

Tested<br />

#<br />

Shoot Tips<br />

with AGYp<br />

Riesling 1 4 5 4 4<br />

Riesling 2 5 7 4 6<br />

Riesling 3 4 6 1 5<br />

Sangiovese 1 1 ‐ ‐<br />

Shiraz 2 1 1 ‐ ‐<br />

Sauv. Blanc 0 1 ‐ ‐<br />

1<br />

The same vineyards as in Table 1. 2 Weakly positive for AGYp.<br />

#<br />

Shoot Tips<br />

Tested<br />

Elsewhere in 2008/09, surveys also showed higher severity but<br />

little new incidence (data not shown). 2) Incidence: The higher<br />

incidence at Nildottie suggests an increased activity of AGYp<br />

vector(s) but raises the possibility of an incursion of a more<br />

infective vector(s).<br />

RESULTS AND DISCUSSION<br />

Vineyard Surveys. Previous levels of AGY in the Riesling<br />

vineyards at Nildottie had averaged 3–5% vines but, in 2008/09,<br />

the incidence was extreme and unprecedented in Australia<br />

(Table 1). Adjacent, usually symptomless red cultivars were also<br />

diseased.<br />

Table 1. Incidence of AGY in vineyards of several cultivars, Nildottie,<br />

South Australia, March 2009.<br />

Scores of AGY Incidence<br />

Vineyard<br />

#<br />

Vines<br />

%<br />

AGY<br />

Vineyard<br />

#<br />

Vines<br />

%<br />

AGY<br />

Riesling 1 562 99.8% Chardonnay 1 292 1.7%<br />

Riesling 2 366 94.8% Chardonnay 2 146 Nil<br />

Riesling 3 388 88.9% Shiraz 1 150 3.3%<br />

Sangiovese 1 150 19.3% Sauv. Blanc 150 Nil<br />

1<br />

From conservative assessment—actual incidence was likely higher.<br />

The highest incidence we had previously seen on any cultivar<br />

was 86% (on Riesling in Renmark, SA, in 1978/79). At Nildottie,<br />

the severity of disease was also extreme. Whereas usually only<br />

3–5 shoots/vine are affected by AGY, in 2008/09, most shoots<br />

were diseased and in Riesling Vineyards 1 and 2, crop loss was<br />

complete.<br />

PCR‐Tests. AGYp were found in a high proportion of samples<br />

(Table 2). This was consistent with previous tests (data not<br />

shown) and counteracted the idea that incursion of a new<br />

pathogen caused the extreme disease.<br />

Why were levels of AGY so high in 2008/09? Two possibilities<br />

are: 1) Severity: Seasonal increases in temperatures from<br />

autumn to early spring in 2008 may have increased the<br />

multiplication of overwintering AGYp, raising their titre and so<br />

the severity of symptoms seen in 2008/09. In March 2008 at<br />

Loxton, mean max./min. temperature (T) (at 32.7°C/12.5°C<br />

Why the observed remission of symptoms? New growth of AGY<br />

affected shoots 10–14 days after heat waves has been seen<br />

many times (data not shown). Symptomless new shoot growth<br />

produced after the 13‐day heat wave of Jan‐Feb 2009 was PCR<br />

+ve for AGYp (Table 2), suggesting that not all AGYp were killed<br />

or denatured by the heat. Hot water treatments (e.g. 45 minutes<br />

at 50°C) delivering ~40°C‐hrs/treatment at 50°C or ~200°C‐hrs at<br />

45°C, are used in Europe to reduce transmission rates of FD and<br />

other yellows pathogens (3). The 2009 heat wave delivered a<br />

lesser though similar heat treatment to the vines at Nildottie. A<br />

46°C max T delivered ~180°C‐hrs at ≥45°C and a 45°C max T<br />

delivered ~230°C‐hrs at ≥43°C. We had supposed that these<br />

temperatures were lethal to AGYp and thus triggered the<br />

observed new shoot growth but this theory is now questioned.<br />

AGY‐affected shoots show signs of disturbed phloem cell<br />

function, hormone imbalance and deposition of callose in sieve<br />

cells (data not shown) but these were insufficient to prevent<br />

significant reactivation of shoot growth soon after the heat<br />

wave. If AGY symptoms were the result of that damage, the swift<br />

‘remission’ of AGYp‐infected and severely diseased shoots raises<br />

conjecture as to the cause of AGY symptoms.<br />

ACKNOWLEDGEMENTS<br />

Dr David Cartwright, Primary Industries and Resources SA,<br />

provided some financial assistance for the PCR‐tests.<br />

REFERENCES<br />

1. Davis R., et al, 1997. “Candidatus Phytoplasma australiense”, a new<br />

phytoplasma taxon associated with Australian grapevine yellows.<br />

International Journal of Systematic Bacteriology 47: 262–269.<br />

2. Kirkpatrick, B.C. et al, 1994. Phylogenetic relationship of plant<br />

pathogenic MLOs established by 16/23S rDNA spacer sequences.<br />

IOM Let. 3: 228–229.<br />

3. Caudwell A., et al. 1997. Flavescence dorée (FD) elimination from<br />

dormant wood of grapevines by hot‐water treatment. Australian<br />

Journal of Grape and Wine Research 3, 21–25.<br />

78 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Association of phytoplasmas with papaya crown yellows disease—a new disease of<br />

papaya in Northern Mindanao, Philippines<br />

R.G. Billones{ XE "Billones, R.G." } A , M.P. Natural B , C.M. Vera Cruz C , E.Y. Ardales B , C. Streten D , L. Tran‐Nguyen D and K.S. Gibb D<br />

A Del Monte Phils, Inc, Bukidnon, Philippines<br />

B Crop Protection Cluster, University of the Philippines‐Los Baños, Philippines<br />

C International Rice Research Institute, Los Baños, Philippines<br />

D Faculty of Education, Health and Science, Charles Darwin University, NT, 0909 Australia<br />

INTRODUCTION<br />

“Papaya crown yellows” (PCY) disease of unknown etiology was<br />

first observed in Northern Mindanao, Philippines in 2001. PCY<br />

was thought to be associated with phytoplasma because its<br />

symptoms are similar to phytoplasma diseases of papaya in<br />

Australia. The occurrence of PCY poses a potential threat to the<br />

papaya industry in Northern Mindanao as well as the whole<br />

country. Thus, this research aimed to investigate the etiology of<br />

the PCY disease and to determine its identity using sensitive<br />

molecular and genetic techniques.<br />

1650 bp<br />

1000 bp<br />

100 bp<br />

DNA marker<br />

Symptomless<br />

Natumolan 1<br />

Natumolan 2<br />

Natumolan 3<br />

Natumolan 4<br />

Malitbog 1<br />

1 2 3 4 5 6 7 8 9 10 11 12 13<br />

Malitbog 2<br />

Malitbog 3<br />

Malitbog 4<br />

Ca. P. Aus<br />

TBB strain<br />

SDW<br />

880 bp<br />

Session 4D—Prokaryotic pathogens<br />

MATERIALS AND METHODS<br />

Sample Collection. Seventy symptomatic and 33 symptomless<br />

papaya leaf samples were collected at 8 different locations in<br />

Northern Mindanao in April‐June 2004. An additional 12<br />

symptomless samples were collected at Los Baños, Laguna<br />

province in July 2004 where PCY disease was not observed.<br />

Detection of Phytoplasmas from Papaya DNA using Polymerase<br />

Chain Reaction (PCR). DNA was isolated from papaya leaf<br />

samples using phytoplasma enrichment nucleic acid extraction<br />

method by Dellaporta (1). Universal primer pair fu5/rU3 (2) of<br />

the 16S rRNA gene was used for general detection of<br />

phytoplasmas in papaya DNA. The group‐specific primer pair Tuf<br />

f40 and Tuf r1150 (Streten, unpublished) that amplifies 1110 bp<br />

of the tuf gene was used to detect papaya dieback phytoplasma<br />

Candidatus P. australiense in papaya DNA. Ca. P. aurantifolia<br />

(TBB strain) and Ca. P. australiense were used as positive<br />

controls while healthy papaya DNA and sterile distilled water<br />

(SDW) were used as negative controls.<br />

Restriction Fragmnent Length Polymorphism (RFLP) Analysis.<br />

PCR products were digested using 1U of TaqI restriction enzyme<br />

(Biolabs, Australia) and were incubated overnight at 65°C. DNA<br />

fragments were separated by electrophoresis in an 8%<br />

polyacrylamide gel and visualised by staining with 5% ethidium<br />

bromide. Ca. P. aurantifolia (TBB and SPLL‐V4 strains) and Ca. P.<br />

australiense were used as references.<br />

RESULTS<br />

Detection of phytoplasmas from papaya using PCR. Of 70<br />

symptomatic samples collected in Northern Mindanao, 8<br />

samples (4 from Natumolan area and 4 from Malitbog area)<br />

were positive to phytoplasmas using fU5/rU3 universal primers<br />

(Figure 1) but none were positive using group‐specific primer Tuf<br />

f40/r1150. All symptomless samples were negative to PCR.<br />

RFLP Analysis. RFLP analyses of the 16S rRNA gene using TaqI<br />

restriction digest enzyme indicated that the phytoplasmas<br />

associated with PCY disease were identical to Ca. Phytoplasma<br />

aurantifolia, the phytoplasma associated with papaya yellow<br />

crinkle and papaya mosaic in Australia and not Ca. P.<br />

australiense causing papaya dieback.<br />

Figure 1. Lane 1, 1 kb plus DNA ladder; 2, symptomless papaya; 3–6,<br />

diseased papaya from Natumolan area; 7–10, diseased papaya from<br />

Malitbog area; 11, Ca. P. australiense (Ca. P. aus); 12, Ca. P. aurantifolia‐<br />

TBB strain; 13, sterile distilled water<br />

650 bp<br />

500 bp<br />

400 bp<br />

300 bp<br />

200 bp<br />

100 bp<br />

DNA ladder<br />

Malitbog 1<br />

Malitbog 2<br />

Malitbog 3<br />

Malitbog 4<br />

Natumolan 1<br />

1 2 3 4 5 6 7 8 9 10 11 12<br />

Natumolan 2<br />

Figure 2. Digestion of fU5/rU3 PCR products with TaqI restriction<br />

enzyme. Lane 1, 1 kb plus DNA ladder, 2–5, diseased papaya from<br />

Malitbog area; 6–9, diseased papaya from Natumolan area, 10, Ca. P.<br />

aurantifolia‐TBB strain; lane 11, Ca. P. aurantifolia‐SPLL‐V4 strain, lane<br />

12, Ca. P. australiense.<br />

DISCUSSION.<br />

This is the first report of a phytoplasma association in diseased<br />

papaya in the Philippines. The identification of the associated<br />

pathogen for PCY is an important step to be able to establish<br />

control measures for the disease.<br />

ACKNOWLEDGEMENT<br />

Del Monte Philippines, Inc. for the research fund.<br />

REFERENCES<br />

1. Dellaporta SL, Wood J, Hicks JB, 1983. A plant DNA<br />

minipreparation: Version II. <strong>Plant</strong> Molecular Biology Reporter 1, 19–<br />

21.<br />

2. Lorenz K‐H, Schneider B, Ahrens U, Seemueller E, 1995. Detection<br />

of the apple proliferation and pear decline phytoplasmas by PCR<br />

amplification of ribosomal and nonribosomal DNA. Phytopathology<br />

85, 771–776.<br />

Natumolan 3<br />

Natumolan 4<br />

TBB strain<br />

SPLL-V4<br />

Ca. P. aus<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 79


Session 4D—Prokaryotic pathogens<br />

Phytoplasma diseases in citrus orchards of Pakistan<br />

Shazia Mannan{ XE "Mannan, S." } A , Shahid Nadeem Chohan B,C , Muhammad Ibrahim A , Obaid Aftab B , Raheel Qamar B , M. Kausar Nawaz<br />

Shah B , Iftikhar Ahmad D , Paul Holford C , G. Andrew C. Beattie C<br />

A Department of Biosciences, COMSATS Institute of Information Technology, Sahiwal, 520‐B Civil Lines, Sahiwal, Pakistan<br />

B Department of Biosciences, COMSATS Institute of Information Technology, Chak Shahzad Campus, Islamabad, Pakistan<br />

C Centre for <strong>Plant</strong> and Food Science, University of Western Sydney, Locked Bag 1797, Penrith South DC, NSW 1797, Australia<br />

D National Agricultural Research Centre, Pakistan Agricultural Research Council, Park Road, Islamabad, Pakistan<br />

INTRODUCTION<br />

Citrus fruits are one of the major export commodities of Pakistan<br />

and are grown in an area of 160,000 ha with production of 1.5<br />

MMT annually (1). Most of this citrus is grown in the province of<br />

Punjab. Citrus species and hybrids [Sapindales: Rutaceae] are<br />

affected by a number of destructive diseases and plants infected<br />

with ‘Candidatus Phytoplasma asteris’ have been detected in<br />

orchards close to Islamabad (4). However, the extent of infection<br />

is not known and the vector(s) has not been identified.<br />

Therefore, a study is under way to detect the presence of<br />

phytoplasmas in citrus from orchards in the districts of Sahiwal,<br />

Pakpattan and Multan, some 500 km west of Islamabad in the<br />

province of Punjab, as well as to identify possible alternative<br />

hosts and the vector(s).<br />

MATERIALS AND METHODS<br />

Leaves showing phytoplasma‐like symptoms were collected from<br />

plants including sweet and blood oranges, grapefruit (C. ×<br />

aurantium L.), mandarins (C. reticulata Blanco), lemons (C. ×<br />

limon (L.) Osbeck) and limes (C. × aurantiifolia (Christm.)<br />

Swingle) [Sapindales: Rutaceae]. In order to identify possible<br />

alternative hosts, weeds including couch grass Cynodon dactylon<br />

(L.) Pers. and wild oat Avena fatua L., [Poales: Gramineae], field<br />

bindweed Convolvulus arvensis L. [Solanales: Convolvulaceae],<br />

and fat‐hen Chenopodium album L. [Caryophyllales:<br />

Chenopodiaceae] were collected. Potential vectors, including<br />

Asiatic citrus psylla Diaphorina citri Kuwayama [Hemiptera:<br />

Psyllidae], and leafhoppers (possibly Balclutha punctata<br />

(Fabricius) and Empoasca decipiens Paoli [Hemiptera:<br />

Cicadellidae]) were collected, most from around plants showing<br />

phytoplasma‐like disease symptoms.<br />

DNA was extracted from the petioles and midribs of samples<br />

using the method of Doyle and Doyle (3). Both single and nested<br />

PCR were used to amplify phytoplasma DNA sequences. Single<br />

PCR used the O‐MLO primers of Doyle and Doyle (3). The P1/P7<br />

primer pair of Deng and Hiruki (2) and Schneider et al (6) was<br />

used in conjunction with primers R16F2n/R16R2 and<br />

R16mF2/R16mR1 (5) for nested PCR. DNA of ‘Candidatus<br />

Phytoplasma aurantifolia’, obtained from Central Science<br />

Laboratory, UK, was used as a positive control during<br />

amplifications.<br />

Figure 1. Typical PCR amplification products. Lanes 1–3, single PCR; lanes<br />

4–7, nested PCR; lanes 1, 2 and 4–6, infected sweet orange; lanes 3 and<br />

7, control (‘Ca. Phytoplasma aurantifolia’); lane 8 molecular weight<br />

markers<br />

ACKNOWLEDGEMENT<br />

We would like to acknowledge the Higher Education Commission<br />

of Pakistan for their financial support of this study.<br />

REFERENCES<br />

1. Anon (undated) All about—Citrus.<br />

http://www.pakissan.com/english/allabout/orchards/citrus/index.s<br />

html (accessed 30/4/2009)<br />

2. Deng S, Hiruki D (1991) Amplification of 16S rRNA genes from<br />

cultureable and nonculturable mollicutes. J. Microbiol. Methods 14,<br />

53–61.<br />

3. Doyle JJ, Doyle JL (1990) Isolation of plant DNA from fresh tissue.<br />

Focus 12, 13–15.<br />

4. Fahmeed, F, Arocha Rosete Y, Acosta Perez K, Boa E, Lucas J (2009)<br />

First report of ‘Candidatus Phytoplasma asteris’ (Group 16SrI)<br />

infecting fruits and vegetables in Islamabad, Pakistan. J.<br />

Phytopathology doi: 10.1111/j.1439‐0434.2009.01549.x<br />

5. Gundersen DE, Lee I‐M (1996) Ultrasensitive detection of<br />

phytoplasmas by nested‐PCR assays using two universal primer<br />

pairs. Phytopathologia Mediterranea 35: 144–151.<br />

6. Schneider B, Seemuller E, Smart CD, Kirkpatrick BC (1995)<br />

Phylogenetic classification of plant pathogenic mycoplasma‐like<br />

organisms or phytoplasmas. In ‘Molecular and diagnostic<br />

procedures in mycoplasmology, Vol. 1’.(Eds S Razin, JG Tully) pp.<br />

369–380. (Academic Press: San Diego)<br />

RESULTS<br />

Single PCR amplified a 558 bp sequence from the phytoplasma<br />

16S rRNA gene of from DNA extracted from infected plants and<br />

nested PCR amplified a 1.2 kb fragment confirming infection<br />

with a phytoplasma (Fig. 1). The amplicons will be sequenced to<br />

determine which group the phytoplasma belongs to. To date, a<br />

total of 20 samples of sweet orange have been tested from the<br />

Sahiwal district, 15 from farmer orchards and 5 from the<br />

Horticulture Research Center at Sahiwal: 6 samples from<br />

farmers’ orchards and 3 from the Center were found to be<br />

infected. Screening of the insects as well as the weeds collected<br />

from the orchards is in progress and the survey is currently being<br />

extended to include the rest of Punjab Province to ascertain the<br />

incidence and prevalence of disease caused by phytoplasmas.<br />

80 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Mechanisms modulating fungal attack in postharvest pathogen interactions and their<br />

modulation for improved disease control<br />

Dov Prusky<br />

Department of Postharvest Science of Fresh Produce, Agricultural Research Organization, Bet Dagan 50250, Israel<br />

Email: dovprusk@agri.gov.il<br />

Keynote address<br />

As biotrophs, insidious fungal infections of postharvest<br />

pathogens remain quiescent during fruit growth while at a<br />

particular phase during ripening and senescence the pathogens<br />

transform to necrotrophs causing typical decay symptoms.<br />

Exposure of unripe hosts to pathogens (hemi‐biotroph or<br />

necrotrophs), initiates defensive signal‐transduction cascades<br />

that limit fungal growth and development. Exposure to the same<br />

pathogens during ripening and storage activates a substantially<br />

different signalling cascade which facilitates fungal colonisation.<br />

This presentation will focus on modulation of postharvest hostpathogen<br />

interactions by pH and the consequences of these<br />

changes. Host pH can be raised or lowered in response to host<br />

signals, including alkalisation by ammonification of the host<br />

tissue as observed in Colletotrichum and Alternaria, or<br />

acidification by secretion of organic acids as observed in<br />

Penicillium and Botrytis. These changes sensitise the host and<br />

activate transcription and secretion of fungal hydrolases that<br />

promote maceration of the host tissue. This sensitisation is<br />

further enhanced at various stages by accumulation of fungal<br />

ROS that can further weaken host tissue and amplifies fungal<br />

development. Several particular examples of coordinated<br />

responses which follow this scheme in Colletotrichum and<br />

Penicillium will be described, followed by discussion of the<br />

means to exploit these mechanisms for establishment of new<br />

approaches for postharvest disease control.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 81


Session 5A—<strong>Plant</strong> pathogen interactions<br />

ABA‐dependant signalling of PR genes and potential involvement in the defence of<br />

lentil to Ascochyta lentis<br />

INTRODUCTION<br />

B.M. Mustafa, D.T.H. Tan, P.W.J. Taylor and R. Ford{ XE "Ford, R." }<br />

BioMarka, Melbourne School of Land and Environment, The University of Melbourne, 3010, Victoria<br />

Pathogenesis‐related (PR) proteins are an important component<br />

of inducible defence mechanisms in plants. Their accumulation is<br />

triggered by pathogen attack or by abiotic stress (1). Recently, a<br />

PR‐4 and a PR‐10a gene were differentially transcribed in<br />

response to the important fungal pathogen, Ascochyta lentis,<br />

between the resistant (ILL7537) and susceptible (ILL6002) lentil<br />

genotypes (2).<br />

This paper outlines the cloning and expression analyses of the<br />

PR‐4 and PR‐10a genes from ILL7537 in response to exogenous<br />

treatments of the global signalling molecules abscisic acid (ABA),<br />

the immediate ethylene precursor aminocyclopropane<br />

carboxylic acid (ACC), methyl jasmonate (MeJA) and salicylic acid<br />

(SA); known to control plant defence pathways (3). This will give<br />

an insight into the regulatory mechanism controlling specific PR<br />

gene expression in lentil and if broadly applicable to defence,<br />

these genes may be targeted for future resistance breeding<br />

strategies.<br />

MATERIALS AND METHODS<br />

SMART RACE cDNA amplification (Clontech, USA) enabled fulllength<br />

cDNA cloning of the lentil PR‐4 and PR‐10a cDNAs. For<br />

expression studies, 14‐day‐old seedlings were sprayed with a<br />

100 µM solution of ABA, ACC, MeJA or SA. Seedlings were also<br />

inoculated with a 10 5 A. lentis spore suspension. Control plants<br />

were sprayed with sterile water. Bulk seedling foliage (leaf and<br />

shoot) from five plants was harvested at 6, 24, 48, and 96 hours<br />

post treatment (hpt) using the RNeasy <strong>Plant</strong> Mini Kit (Qiagen,<br />

USA). The bioassay was repeated with another set of<br />

independently grown seedlings. cDNA synthesis was carried out<br />

by reverse‐transcribing 1.5 µg of each RNA sample using an oligo<br />

dT18 primer (Roche, Germany) and the Omniscript RT kit<br />

(Qiagen, USA). Triplicate qPCR reactions were performed on<br />

each hpt cDNA sample. All PCR products were subjected to<br />

melting curve analysis and the comparative C t method (ΔΔC t )<br />

was used to calculate the relative fold changes of gene<br />

expression.<br />

RESULTS<br />

The full length PR‐10a cDNA was 783 bp long with an ORF<br />

encoding a peptide of 156 amino acids with an N‐terminal<br />

methionine and a C‐terminal leucine. The full length PR‐4 cDNA<br />

was 636 bp long, with 39 bases of 5` untranslated and 157 bases<br />

of 3` untranslated sequence, and a poly(A) tail.<br />

PR‐4 and PR‐10a gene expression was up‐regulated in lentil by<br />

ABA at all hpt with PR‐10a being more highly expressed than PR‐<br />

4. Neither gene was up‐regulated by the other signalling<br />

molecules (ACC, MeJA and SA; Fig 1). Thus we proposed that the<br />

signalling pathway for both of these genes is ABA‐dependent<br />

and JA‐independent.<br />

Figure 1. Relative fold changes in transcript levels of PR‐4 and PR‐10a in<br />

14‐day‐old ILL7537 seedlings at 6, 24, 48 and 96 hpt following ABA, ACC,<br />

MeJA, SA or Ascochyta lentis treatment.<br />

DISCUSSION<br />

Since, both PR‐4 and PR‐10a were also up‐regulated in response<br />

to the important fungal pathogen Ascochyta lentis (2), we<br />

propose that ABA may play a pivotal role in the signal<br />

transduction of defence responses against A. lentis in lentil. This<br />

is in agreement with other host/pathogen interaction studies (4).<br />

However, further functional validation of ABA modulation of<br />

defence responses to such pathogenic stimuli is required to<br />

facilitate the identification and characterisation of key genes<br />

involved in the ABA‐dependent signalling pathway in lentil.<br />

REFERENCES<br />

1. van Loon LC, Rep M, Pieterse CM (2006) Significance of inducible<br />

defense‐related proteins in infected plants. Annual Review of<br />

Phytopathology 44, 135–162.<br />

2. Mustafa BM (2008) Functional mechanisms of Aschochyta blight<br />

resistance in lentil. PhD Thesis. The University of Melbourne,<br />

Australia.<br />

3. Bostock RM (2005) Signal Crosstalk and Induced Resistance:<br />

Straddling the Line Between Cost and Benefit. Annual Review of<br />

Phytopathology, 43, 545–580.<br />

4. Kaliff M, Staal J, Myrenas M, Dixelius C (2007) ABA is required for<br />

Leptosphaeria maculans resistance via ABI1‐ and ABI4‐dependent<br />

signaling. Molecular <strong>Plant</strong>‐Microbe Interactions 20, 335–345.<br />

82 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Fundamental components of resistance to Phytophthora cinnamomi: using model<br />

system approaches<br />

INTRODUCTION<br />

J.A. Cullum, T.K. Gunning, J.E. Rookes and D.M. Cahill{ XE "Cahill, D.M." }<br />

School of Life and Environmental Sciences, Deakin University, Geelong, 3217, Victoria<br />

Compared with the number of plant species known to be<br />

susceptible to Phytophthora cinnamomi, few are known to be<br />

resistant (1). Understanding how plants are able to resist this<br />

pathogen will enable strategies to be developed to enhance<br />

individual species survival and to restore structure and<br />

biodiversity to the ecosystems under threat. Natural resistance<br />

to pathogens can be characterised at a number of levels ranging<br />

from specific gene involvement to whole plant responses.<br />

The aims of this project are to determine at the cellular,<br />

biochemical and molecular levels what constitutes resistance of<br />

native plants to P. cinnamomi by identifying the pathways within<br />

plants that regulate resistance and then exploring the potential<br />

to manipulate them. Our research is centred on further defining<br />

and understanding natural resistance in native Australian plant<br />

species. Most studies on resistance against pathogens have been<br />

undertaken on aerial plant parts (ie stems and leaves) thus there<br />

is a requirement for a good, fundamental understanding of rootbased<br />

resistance. Here we have used a model system approach<br />

to characterise the fundamental components of resistance<br />

against P. cinnamomi.<br />

MATERIALS AND METHODS<br />

Arabidopsis model. The interaction between Arabidopsis<br />

thaliana ecotype Columbia‐0 and P. cinnamomi was examined in<br />

detail (2). Leaf and root analysis and examination of responses in<br />

signal transduction pathway mutants were carried out.<br />

Zea model. Gene expression analysis of defence‐related genes<br />

was conducted following inoculation of Zea mays root tissue<br />

over several time points ranging from 0–120 hours. In addition,<br />

regulation of defence‐related genes in response to the defence<br />

hormones jasmonic acid, salicylic acid and ethylene was<br />

examined to determine which signalling pathways may operate<br />

in this pathosystem.<br />

Lupin model L. angustifolius was grown in a soil‐free plant<br />

growth system (3) and roots inoculated with zoospores. This well<br />

recognised, highly susceptible interaction is typified by the<br />

development of dark lesions in the root that extend through the<br />

vascular system causing root decay. We developed a nontargeted<br />

method to extract, separate and identify metabolites<br />

produced following inoculation.<br />

RESULTS<br />

Arabidopsis thaliana. P. cinnamomi was found to induce active<br />

defence responses in Arabidopsis. Tissue specific differences in<br />

levels of infection and defence responses induced were found<br />

between inoculated root and leaf tissue. Molecular analysis of<br />

defence‐related genes also showed differential induction<br />

between root and leaf tissue. It is likely that the<br />

resistance/tolerance that Arabidopsis displays against P.<br />

cinnamomi is provided by a multi‐faceted defence response.<br />

maize specific microarray has identified genetic pathways that<br />

regulate defence to P. cinnamomi.<br />

Lupinus angustifolius. The metabolic profiles (in the form of<br />

HPLC chromatograms) from both un‐inoculated (healthy) and<br />

inoculated (diseased) plant root tissue were compared. The data<br />

strongly suggests differences in metabolic activity. Multiple<br />

experimental repeats were performed and were statistically<br />

analysed using principal components analysis which confirmed<br />

that the profiles were statistically different. The variables loading<br />

identified the peaks/metabolites that caused the data to<br />

separate into distinct groups. The separated extracts were then<br />

passed through a Time of Flight Mass Spectrometer (TOF<br />

MS/MS) to determine identity/chemical structure of<br />

metabolites.<br />

DISCUSSION<br />

We have examined the potential for using model plant systems<br />

for the analysis of the interaction of P. cinnamomi with roots.<br />

This research has provided a number of important outcomes<br />

that will alter the way in which we approach disease caused by<br />

P. cinnamomi in Australian native plants. If we know what the<br />

mechanisms of resistance are then there is the potential to<br />

manipulate them and to devise new methods for control that<br />

may include induction of specific resistance mechanisms in<br />

susceptible species and development of markers for resistance<br />

(for example, anatomical, morphological, molecular‐based).<br />

Additionally, we are developing screening techniques that will<br />

enable concentration of effort on those species that are the<br />

most susceptible, vulnerable and rare. Resistant plants can also<br />

be used to restore vegetation structure on P. cinnamomiaffected<br />

sites where the pathogen may still be present.<br />

ACKNOWLEDGEMENTS<br />

We thank the Australian Government Department of<br />

Environment, Heritage, Water and the Arts for funding. Dr Xavier<br />

Conlan and Professors Neil Barnett and Mike Adams (Deakin<br />

University and RMIT University) have provided valuable advice<br />

on chemical analyses.<br />

REFERENCES<br />

1. Cahill DM, Rookes JE, Wilson BA, Gibson L, McDougall K (2008)<br />

Phytophthora cinnamomi and Australia’s biodiversity: impacts,<br />

predictions and progress towards control. Australian Journal of<br />

Botany 56, 279–310.<br />

2. Rookes JE, Wright M, Cahill DM (2008) Elucidation of defence<br />

responses and signalling pathways induced in Arabidopsis thaliana<br />

following challenge with Phytophthora cinnamomi. Physiological<br />

and Molecular <strong>Plant</strong> <strong>Pathology</strong>, 72, 151–161.<br />

3. Gunning TK and Cahill DM (2009) A soil‐free plant growth system to<br />

facilitate analysis of plant pathogen interactions in roots. Journal of<br />

Phytopathology, In press.<br />

Session 5A—<strong>Plant</strong> pathogen interactions<br />

Zea mays. A resistant monocot model system (Zea mays) for<br />

gene expression analysis of defence pathways was optimised. A<br />

suite of maize defence‐related genes was analysed in response<br />

to infection with P. cinnamomi. Genome‐wide analysis using a<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 83


Session 5A—<strong>Plant</strong> pathogen interactions<br />

Genes involved in hypersensitive cell death responses during Fusarium crown rot<br />

infection in wheat<br />

Jill E. Petrisko{ XE "Petrisko, J.E." } 1 , Mark W. Sutherland 1 , Juliet M. Windes 2<br />

1 Centre for Systems Biology, University of Southern Queensland, Toowoomba, Queensland, 4350, Australia<br />

2 University of Idaho, <strong>Plant</strong>, Soils, and Entomological Sciences, University Place, 1776 Science Center Drive, Suite 205, Idaho Falls, Idaho<br />

INTRODUCTION<br />

Hypersensitive plant cell death is activated by the accumulation<br />

of hydrogen peroxide and nitric oxide (1), and is strictly<br />

controlled by several genes including cysteine proteases,<br />

hydrogen peroxide and superoxide scavengers, and cell death<br />

regulators (2). In contrast to biotrophic fungal pathogens,<br />

necrotrophic pathogens like Fusarium pseudograminearum and<br />

F. culmorum that cause Fusarium crown rot infections, benefit<br />

from plant cell death by utilising dying plant tissue to facilitate<br />

their spread throughout the plant (3).<br />

83402<br />

RESULTS AND DISCUSSION<br />

Table 1. Gene transcript levels expressed during infection with<br />

F. culmorum.<br />

Genes 2–49 Puseas<br />

Cathepsin B 1.58 2.09<br />

Mlo‐like protein 2.26 ‐1.05<br />

Catalase ‐2.19 ‐4.40<br />

Manganese SOD<br />

superoxide dismutase **<br />

40.69 1<br />

MATERIALS AND METHODS<br />

Seedling Germination.‐Seeds of the Fusarium crown rot<br />

susceptible wheat cultivar Puseas and partially resistant wheat<br />

line 2–49 were sterilised in 5% NaOCl for 1 hour and were then<br />

germinated in the dark on petri dishes containing 2% water agar.<br />

Seedling Inoculation.‐Seedlings were inoculated with a single<br />

spore of F. culmorum or F. pseudograminearum on a 2% water<br />

agar block using an adapted procedure of Mergoum et al. (4) and<br />

harvested 10 days post‐inoculation.<br />

Microarray analysis of F. culmorum infection. RNA was<br />

extracted from non‐inoculated and F. culmorum inoculated<br />

seedlings of 2–49 and Puseas and hybridised to Affymetrix®<br />

wheat gene chips. Gene transcripts in the inoculated treatments<br />

determined to be significantly induced or repressed two‐fold<br />

over the non‐inoculated treatments were analysed using the<br />

GeneSpring GX_7_3 program (Agilent).<br />

Deoxynivalenol (DON) Application.‐10 ul of 10 mg/ml<br />

deoxynivalenol was applied to a block of 2% water agar attached<br />

to growing seedlings of 2–49 and Puseas and was taken up by<br />

the seedling for 24 hours.<br />

Staining for cell death was visualised in some of the seedlings by<br />

applying a second agar block containing 10 ul of 0.1% Evans blue<br />

dye below the block containing DON and allowing the stain to be<br />

taken up with the DON for 24 hours.<br />

RNA extraction, cDNA, and real‐time PCR.‐RNA from F.<br />

pseudograminearum inoculated or DON applied seedlings was<br />

extracted using the <strong>Plant</strong> RNA Purification Reagent protocol<br />

(Invitrogen). cDNA was produced using gene specific primers in a<br />

reverse transcriptase reaction. cDNA transcripts were assayed<br />

using real‐time quantitative PCR using SYBR green in the Rotor‐<br />

Gene 6000 thermocycler.<br />

Genes involved in the hypersensitive cell death response during<br />

F. culmorum infection (Table 1) were identified using a<br />

microarray analysis with the Affymetrix® wheat chip. Cathepsin<br />

B, a plant cysteine protease, was induced in the susceptible<br />

cultivar Puseas during F. culmorum infection. Catalase, an<br />

enzyme preventing hydrogen peroxide accumulation, was<br />

repressed in Puseas. A Mlo‐like protein (cell death regulator) and<br />

manganese superoxide dismutase were up‐regulated in the<br />

resistant wheat line 2–49. These genes are under current<br />

investigation during infection studies with F.<br />

pseudograminearum and DON application to determine what<br />

role they have in the response of these cultivars to infection.<br />

Infection with F. pseudograminearum spores and DON has been<br />

shown to elicit hydrogen peroxide formation and plant cell death<br />

as well induce genes involved in defence responses in wheat (5).<br />

It is not known whether hypersensitive cell death or avoidance<br />

of hypersensitive cell death during infection with Fusarium<br />

species plays a role in the susceptibility or resistance of wheat<br />

cultivars to Fusarium crown infection. Further investigation of<br />

these genes during the infection process with F.<br />

pseudograminearum and DON is needed in order to determine if<br />

different levels influence hypersensitive cell death and the role<br />

they have in either enhancing susceptibility or resistance in<br />

wheat during Fusarium crown infection.<br />

REFERENCES<br />

1. Delledonne M, Zeier J, Marocco A, Lamb C. (2001) Signal<br />

interactions between nitric oxide and reactive oxygen<br />

intermediates in the plant hypersensitive disease resistance<br />

response. Proc of the Nat Acad of Sci 98, 13454–59.<br />

2. Gilroy EM, Hein I, van der Horn R, Boevink PC, Venter E, McLellan<br />

H, Kaffarnik F, Hrubikova K, Shaw J, Holeva M, Lopez EC, Borras‐<br />

Hidlago O, Pritchard L, Loake, GJ, Lacomme C, Birch PR. (2007)<br />

Involvement of cathepsin B in the plant resistance hypersensitive<br />

response. <strong>Plant</strong> J 52, 1–13.<br />

3. Govrin EM, Levine A (2000) The hypersensitive response facilitates<br />

plant infection by the necrotrophic pathogen Botrytis cinerea. Curr<br />

Biol 10, 751–57.<br />

4. Mergoum M, Hill JP, Quick J (1998). Evaluation of resistance of<br />

winter wheat to Fusarium accuminatum by inoculation of seedling<br />

roots with single, germinated macroconidia. <strong>Plant</strong> Dis 82, 300–2.<br />

5. Desmond OJ, Manners JM, Stephens AE, Maclean DJ, Schenk PM,<br />

Gardiner DM, Munn AL, Kazan K (2008) The Fusarium mycotoxin<br />

deoxynivalenol elicits hydrogen peroxide production, programmed<br />

cell death, and defence responses in wheat. Mol <strong>Plant</strong> Pathol 9,<br />

435–45<br />

84 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Fishing for Phytophthora across Western Australia’s water bodies<br />

D. Hüberli{ XE "Hüberli, D." } A,B , T.I. Burgess A and G.E.St.J. Hardy A<br />

A Centre for Phytophthora Science and Management, School of Biological Sciences and Biotechnology, Murdoch University, Perth, 6150,<br />

WA<br />

B Present Address: <strong>Plant</strong> <strong>Pathology</strong>, Department of Agriculture and Food, 3 Baron‐Hay Court, South Perth, 6151, WA<br />

INTRODUCTION<br />

Most Phytophthora surveys in native ecosystems in Australia<br />

have focused exclusively on isolations from samples of soil and<br />

symptomatic plant tissue including the extensive vegetation<br />

health surveys conducted in Western Australia (WA) (1). The<br />

outbreak of P. ramorum in California and Europe, where early<br />

detection of an infested area was important to the success of<br />

containment and eradication efforts, has popularised the stream<br />

surveys in native ecosystems. In Australia, it has recently been<br />

used to detect Phytophthora spp. in Victoria resulting in several<br />

species being isolated from four streams which varied according<br />

to the winter and summer sampling season (2). In our study, the<br />

baiting technique was used to survey a wide range of WA’s<br />

waterways for Phytophthora spp. during October to early<br />

December 2008.<br />

(VHS16108) and P. inundata. That P. inundata is apparently<br />

widespread in the south coast and some wheatbelt regions is of<br />

concern given that it has been associated with dying native<br />

vegetation in WA’s southwest (3). Currently, little is known<br />

about P.sp.12 and P.sp.13, but our study clearly shows that they<br />

are widespread across many regions of WA.<br />

P.sp.3 and P.sp.11 were not common (Table 1); P.sp.3 has<br />

previously been isolated from dying Eucalyptus marginata,<br />

Banksia spp. and Pinus radiata (1). P.sp.8, P.sp.11, P.sp.12 and<br />

P.sp.13 fit within a strongly supported clade with P.sp.3, and (1)<br />

proposed that these would be common in water bodies. Our<br />

sampling during October to December 2008 shows that this<br />

hypothesis is true with the exception of P.sp.3 and P.sp.11.<br />

Table 1. Phytophthora species found in water bodies in nine regions of<br />

Western Australia during the October to December 2008 sampling (see<br />

Figure 1). Reg = Region, na = not applicable (too few samplings)<br />

Reg. Phytophthora spp. Most common spp.<br />

1 P. inundata, P.sp.11, P.sp.13 P.sp.13<br />

2 P. inundata, P.sp.11, P.sp.13 P. inundata<br />

3 P. inundata, P.sp.13 P. inundata<br />

4 P.sp.3, P.sp.12, P.sp.13 P.sp.12, P.sp.13<br />

5 P.sp.12 na<br />

6 P. hydropathica, P. inundata, P.sp.3,<br />

P.sp.8, P.sp.11, P.sp.12, P.sp.13<br />

P.sp.12<br />

7 P.sp.8, P.sp.11 P.sp.8<br />

8 P. inundata, P.sp.3, P.sp.8, P.sp.11,<br />

P.sp.13<br />

9 P. parvispora, P. hydropathica na<br />

P. inundata, P.sp.8<br />

Session 5B—Disease surveys<br />

Figure 1. Water bodies in nine regions of Western Australia sampled for<br />

Phytophthora species in October to December 2008. Some sites were<br />

sampled on four occasions. Region 6 is Perth.<br />

MATERIALS AND METHODS<br />

Waterways (streams, lakes, ponds and estuaries) in WA were<br />

sampled up to four times during October to early December<br />

2008. Sites were selected across 50 regional locations from<br />

Kununurra (northern site) to Esperance (southern site), and 37<br />

locations in the Perth suburbs (Figure 1, Table 1). Bait bags<br />

containing leaves of Banksia attenuata, Pittosporum sp., Hakea<br />

sp. and Quercus sp., and lupin seedlings were sent via overnight<br />

post to 16 volunteers who deployed, and retrieved and returned<br />

baits after ~10 days in the water. Leaves were plated onto<br />

NARPH agar plates, a medium selective for Phytophthora, and<br />

incubated in darkness for up to 2 weeks at 20°C during which<br />

plates were periodically checked for Phytophthora colonies.<br />

Colonies were isolated into pure culture and grouped into<br />

morpho‐types. One representative morpho‐type from each site<br />

per sampling was identified using the sequence of the ITS region<br />

of the rDNA, conducting a BLAST search on Genbank and a<br />

phylogenetic analysis (1).<br />

RESULTS AND DISCUSSION<br />

A total of eight Phytophthora species were isolated during the<br />

October to December 2008 survey of water bodies in nine<br />

regions of WA (Table 1). Species yet undescribed were assigned<br />

taxa numbers as described in (1). The most frequently isolated<br />

species in the southwest were P.sp.12 (VHS5185), P.sp.13<br />

P. hydropathica frequently isolated from irrigation water, was<br />

only isolated on two occasions; once in Perth and once in the<br />

north of WA from an irrigation channel (Table 1). Additionally, P.<br />

cinnamomi var. parvispora was only recovered once during the<br />

sampling, also from the irrigation channel in the north. Our<br />

results in relation to the Victorian study will be discussed.<br />

ACKNOWLEDGEMENTS<br />

We thank all the volunteers who deployed and returned the<br />

baits and WWF for funding support. Additional assistance was<br />

provided by C. Fletcher, T. Paap, D. White and N. Williams. More<br />

information at www.f4p.murdoch.edu.au.<br />

REFERENCES<br />

1. Burgess TI, Webster JL, Ciampini JA, White D, Hardy GEStJ, Stukely<br />

MJC (2009) Re‐evaluation of Phytophthora species isolated during<br />

30 years of vegetation health surveys in Western Australia using<br />

molecular techniques. <strong>Plant</strong> Disease 93, 215–223.<br />

2. Smith BW, Smith IW, Cunnington J, Jones RH (2007) An evaluation<br />

of stream monitoring techniques for surveys for Phytophthora<br />

species in Victoria, Australia. In ‘Fourth Meeting of IUFRO Working<br />

Party 7.02.09 on Phytophthoras in Forests and Natural Ecosystems,<br />

26–31 August.’ Monterey, California, US. (Poster)<br />

3. Stukely MJC, Webster JL, Ciampini JA, Dunstan WA, Hardy GEStJ,<br />

Woodman GJ, Davison EM, Tay FCS (2007) Phytophthora inundata<br />

from native vegetation in Western Australia. <strong>Australasian</strong> <strong>Plant</strong><br />

<strong>Pathology</strong> 36, 606–608.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 85


Session 5B—Disease surveys<br />

Incidence of fungi isolated from grape trunks in New Zealand vineyards<br />

D.C. Mundy{ XE "Mundy, D.C." } A , S.G. Casonato B and M. A. Manning B<br />

A<br />

The New Zealand Institute for <strong>Plant</strong> and Food Research Limited, Marlborough Wine Research Centre, PO Box 845, Blenheim, New<br />

Zealand<br />

B<br />

The New Zealand Institute for <strong>Plant</strong> and Food Research Limited, Mt Albert Research Centre, Private Bag 92169, Auckland Mail Centre,<br />

INTRODUCTION<br />

With the growth of the New Zealand wine industry in size and<br />

geographic distribution, the number of observations of grape<br />

vine trunk disease symptoms has increased. Within the industry,<br />

identification of fungi present in trunks of unthrifty vines has<br />

been a low priority. Previous studies have often focused on a<br />

single genus (1) or regional record of disease (2).<br />

In order to reduce the impact of trunk diseases in New Zealand<br />

vineyards, it is important first to establish which fungi are<br />

present. This survey is a preliminary investigation of the range<br />

and incidence of fungi isolated from trunk wood across a<br />

number of vineyard sites in New Zealand.<br />

New Zealand<br />

Table 1. Recoveries of fungal genera regarded as wood pathogens from<br />

a survey of thirty‐seven vineyard blocks. The survey of incidence of<br />

fungi was conducted in the North and South Islands of New Zealand<br />

during 2007 and 2008.<br />

Genus<br />

North Is.<br />

n=24<br />

South Is.<br />

n=13<br />

Botryosphaeria 20 5<br />

Phaeomoniella 19 3<br />

Phaeoacremonium 3 0<br />

Phomopsis 8 3<br />

Cylindrocarpon 2 3<br />

Eutypa 12 8<br />

MATERIALS AND METHODS<br />

Field survey. A survey was conducted on vines from 37 vineyard<br />

blocks in the North and South Islands of New Zealand. Core<br />

samples were taken from the trunks of five vines at each site,<br />

using a MATTSON N° 4333 forestry corer. This device removed a<br />

5‐mm core approximately 80 cm up the trunk, passing directly<br />

through the wood until the bark was ruptured on the far side.<br />

The corer was cleaned between samples with 70% ethanol to<br />

prevent cross contamination. The entire core sample was<br />

transferred in a sterile tube to the laboratory.<br />

Isolations. Each core was surface sterilised for 30 sec in 70%<br />

ethanol, 2 min in 3.5% w/v sodium hypochlorite and 30 sec in<br />

70% ethanol. Samples were cut into 5–10 mm pieces and placed<br />

on potato dextrose agar (PDA; DIFCO) amended with 100 µg/mL<br />

streptomycin sulphate and 100 µg/mL Penicillin G potassium<br />

salts and incubated at 20°C with lights (12 h photoperiod).<br />

Identification. After one week, fungi were identified by<br />

morphological features and confirmed by amplifying the internal<br />

transcribed spacer regions of the rDNA using the polymerase<br />

chain reaction (PCR) primers ITS1‐F and ITS‐4. PCR products were<br />

sequenced using the BigDye Terminator V. 3.1 cycle sequencing<br />

kit (Applied Biosystems, UK). The resultant sequences were<br />

characterised by Basic Local Alignment Search Tool analysis from<br />

the most closely related sequences on GenBank. Fungal<br />

morphological characteristics were re‐examined after a month<br />

to confirm identification further and to allow time for slower<br />

growing fungi to be isolated and identified. Not all fungi were<br />

identified to the species level so results are given as a summary<br />

by genera.<br />

RESULTS<br />

The most commonly isolated fungi were species of the genera<br />

Botryosphaeria, Phaeomoniella and Eutypa at multiple sites<br />

(Table 1). Less commonly, Phaeoacremonium spp. and<br />

Cylindrocarpon spp. isolates were also found. The same genera<br />

were not isolated from all 37 sites sampled. Phaeoacremonium<br />

sp. were found only in Hawke’s Bay, although the other fungal<br />

isolates were not confined to a single region. Other fungi<br />

isolated during the survey included species of Acremonium,<br />

Alternaria, Cadophora, Cladosporium, Epicoccum, Gliocladium,<br />

Mucor, Penicillium, Phoma, Trichoderma, Ulocladium and<br />

Xylaria.<br />

DISCUSSION<br />

The incidence of species of Botryosphaeria, Phaeomoniella and<br />

Eutypa in many of the vineyard blocks surveyed suggests that<br />

these fungi should be the focus of more detailed research to<br />

establish their importance to the industry. The isolation of a<br />

fungus from a block does not prove that it is the organism<br />

responsible for the disease symptoms. If multiple fungi are<br />

present, the incidence does not allow the determination of<br />

which fungi are the most likely to be causing symptoms at that<br />

site.<br />

As the sample size in each vineyard block was limited, we cannot<br />

be certain that failure to detect a particular species indicates<br />

that these fungi were not present. For example additional<br />

isolations will be needed to determine if Phaeoacremonium is<br />

restricted to Hawke’s Bay only.<br />

ACKNOWLEDGEMENTS<br />

Funding for this project was provided by the Ministry of<br />

Agriculture and Forestry Sustainable Farming Fund and the<br />

Marlborough Wine Research Centre Trust. We would also like to<br />

thank all industry representatives who allowed us to sample<br />

vines.<br />

REFERENCES<br />

1. Amponsah NT, Jones EE, Ridgway HJ, Jaspers MV 2008. Production<br />

of Botryosphaeria species conidia using grapevine green shoots.<br />

New Zealand <strong>Plant</strong> Protection 61: 301–305.<br />

2. Mundy DC, Manning MA 2006. Initial investigation of grapevine<br />

trunk health in Marlborough, New Zealand. 5th International<br />

Workshop on Grapevine Trunk Diseases. Department of <strong>Plant</strong><br />

<strong>Pathology</strong> University of California, Davis, CA.<br />

86 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Isolation and characterisation of strains of Pseudomonas syringae from waterways of<br />

the central North Island of New Zealand<br />

J.L. Vanneste{ XE "Vanneste, J.L." } A , D.A. Cornish A , J. Yu A , and C.E. Morris B<br />

A<br />

The New Zealand Institute for <strong>Plant</strong> and Food Research Limited, Ruakura Research Centre, Private Bag 3123, Hamilton, New Zealand<br />

B<br />

INRA, UR 407 Pathologie Végétale, F‐84140 Montfavet, France<br />

INTRODUCTION<br />

Epiphytotics of plant bacterial diseases can occur where<br />

previously no or little inoculum was thought to be present. The<br />

source of the primary inoculum of these epiphytotics is not<br />

always easy to determine, especially for pathogenic bacteria,<br />

which are good epiphytes, such as Pseudomonas syringae, one<br />

of the most economically important bacterial plant pathogens.<br />

This project aimed to determine whether rivers from the Central<br />

North Island of New Zealand could constitute a reservoir for P.<br />

syringae. P. syringae is a complex group of strains which can be<br />

grouped in about 50 different pathovars (strains which have the<br />

same pathogenicity and the same host range) or in nine<br />

genomospecies (strains which belong to the same species based<br />

on DNA/DNA homology) (1). In this study we isolated some<br />

strains of P. syringae and tried to determine to which<br />

genomospecies or which pathovar they belong based on<br />

polymerase chain reaction (PCR) experiments.<br />

METHODS AND RESULTS<br />

Collection of water samples and isolation of bacteria. Water<br />

samples were collected from the Waikato River (Hamilton) and<br />

from the Whakapapanui Stream (Tongariro National Park). The<br />

isolation of bacteria was carried out as described previously (4).<br />

Ten strains from Whakapapanui and five strains from the<br />

Waikato River, which showed all the characteristics of strains of<br />

P. syringae: ability to produce a fluorescent pigment on a<br />

modified King’s B medium, ability to cause a hypersensitive<br />

reaction when infiltrated into tobacco plant, absence of a<br />

cytochrome c oxidase and inability to utilise arginine, were<br />

retained for further characterisation.<br />

Characterisation by Polymerase Chain Reaction (PCR). All PCR<br />

experiments were carried out on an Eppendorf Mastercycler®<br />

Gradient using 20 ng of total DNA per reaction. The final reaction<br />

volume was 30 μl including 10 μM of each primers and 1 unit of<br />

i‐Taq from INtRON Biotechnology Inc. The primers and the<br />

programs were those published earlier (e.g. 2). For each<br />

experiment, a negative control, in which the DNA solution was<br />

replaced by water, and a positive control, in which the DNA was<br />

that of a strain we knew would give a positive reaction, were<br />

used. Of the 15 strains analysed, five gave a positive reaction<br />

with primers specific to strains of genomospecies 1, which is<br />

represented by P. syringae pv. syringae. None gave a positive<br />

response with primers specific for strains of genomospecies 2,<br />

which is represented by P. syringae pv. phaseolicola and P.<br />

syringae pv morsprunorum. None gave a positive response when<br />

PCR protocols specific for P. syringae pv. papulans, P. syringae pv<br />

tagetis, P. syringae pv helianthi or P. syringae pv actinidiae were<br />

used.<br />

water cycle, as proposed by Morris et al. (3). In this scenario, rain<br />

and melt water containing cells of P. syringae feed streams and<br />

rivers that bring those cells of P. syringae in contact with wild<br />

and cultivated plants. The subsequent multiplication of these<br />

bacteria as pathogens or epiphytes provides a huge inoculum,<br />

part of which might form aerosols that can be taken up by<br />

clouds. The ability of some of these bacteria to induce ice<br />

nucleation might help the formation of rain and/or snow and<br />

explain the presence of these bacteria in rain and snow.<br />

Although strains of P. syringae have been found in a river and a<br />

stream fed by melt water, to complete the cycle we need to<br />

demonstrate that those same strains are also found on or in<br />

plants as epiphytes or as pathogens. The characterisation of the<br />

strains of P. syringae isolated from New Zealand waterways<br />

would allow this demonstration. The characterisation is<br />

continuing, with more strains being analysed including strains<br />

being isolated from different waterways, and with more<br />

techniques being utilised including molecular techniques<br />

different from PCR.<br />

ACKNOWLEDGEMENTS<br />

We thank The New Zealand Institute for <strong>Plant</strong> and Food<br />

Research Limited for their support.<br />

REFERENCES<br />

1. Gardan L, Shafik H, Belouin S, Broch R, Grimont F, Grimont PDA<br />

1999. DNA relatedness among the pathovars of Pseudomonas<br />

syringae and description of Pseudomonas tremae sp. nov. and<br />

Pseudomonas cannabina sp. nov. (ex Sutic and Dowson 1959).<br />

International Journal of Systematic Bacteriology. 49: 469–478.<br />

2. Kerkoud M, Manceau C, Paulin J‐P 2002. Rapid diagnosis of<br />

Pseudomonas syringae pv. papulans, the causal agent of blister<br />

spot of apple, by polymerase chain reaction using specifically<br />

designed hrpL gene primer Phytopathology 92: 1077–1083.<br />

3. Morris CE, Sands DC, Vinatzer BA, Glaux C, Guilbaud C, Buffiere A,<br />

Yan S, Dominguez H Thompson BM (2008) The life history of the<br />

plant pathogen Pseudomonas syringae is linked to the water cycle.<br />

The International <strong>Society</strong> for Microbial Ecology Journal 1, 1–14.<br />

4. Vanneste JL, Cornish DA, Yu J, Boyd RJ, Morris CE (2008) Isolation of<br />

copper and streptomycin resistant phytopathogenic Pseudomonas<br />

syringae from lakes and rivers in the Central North Island of New<br />

Zealand. New Zealand <strong>Plant</strong> Protection 61, 80–85.<br />

Session 5B—Disease surveys<br />

DISCUSSION<br />

Strains of P. syringae were isolated from two different water<br />

systems: the Waikato River, a complex system which includes<br />

lakes and goes through some cultivated and non cultivated<br />

lands, and the Whakapapanui Stream, which is fed by melt water<br />

from Mount Ruapehu and does not cross cultivated lands. These<br />

results and similar ones presented earlier (3) support the<br />

hypothesis that the life history of P. syringae is linked to the<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 87


Session 5C—Chemical control<br />

INTRODUCTION<br />

Evaluation of fungicides to manage brassica stem canker<br />

B.H. Hall, L. Deland{ XE "Deland, L." }, B. Rawnsley, T. Barlow, C. Hitch and T.J. Wicks<br />

South Australian Research and Development Institute, GPO Box 397, Adelaide, 5001, South Australia<br />

Leptosphaeria maculans and Rhizoctonia solani AG 2.1, 2.2 and 4<br />

are the dominant soil borne fungal pathogens causing Brassica<br />

stem canker (1). Brassica stem canker causes stem rot, resulting<br />

in plant collapse before harvest or stem breakage during<br />

harvest. Greenhouse trials were undertaken to evaluate preplanting<br />

and post‐planting fungicide drenches for the control of<br />

R. solani AG 2.1, 2.2 and 4 and L. maculans.<br />

MATERIALS AND METHODS<br />

R. solani. Coco peat potting mix was inoculated with either AG<br />

2.1, 2.2 or 4 by mixing in a 1L slurry of 12 macerated plates of 5–<br />

7 day old cultures grown on potato dextrose agar (PDA). The<br />

inoculated mix was placed into MK12 pots and incubated in the<br />

greenhouse at 22°C for 7 days to allow the Rhizoctonia to<br />

establish in the soil.<br />

Six‐week‐old susceptible cauliflower speedlings (cv. Chaser) (2)<br />

were immersed in a fungicide mix (Table 1) for 5 mins to ensure<br />

the fungicide had permeated the soil and root matrix before<br />

planting into the inoculated soil mix. Ten replicate plants were<br />

used per treatment, with a water drench used as the control<br />

treatment.<br />

<strong>Plant</strong>s were maintained in a greenhouse at 22°C and assessed<br />

weekly for stem canker using a disease severity rating scale of 0–<br />

100 where; 0 = healthy, 20 = superficial staining, 40 = canker<br />

girdling ½ stem, 60 = canker girdling full stem, 80 = severe canker<br />

(wilt) and 100 = plant death. The presence of R. solani in the pots<br />

was confirmed 4–6 weeks after planting by baiting with<br />

toothpicks, whereby toothpicks were placed in the pots for 24<br />

hrs, washed and incubated on PDA (3).<br />

L. maculans. Four‐‐week‐old cauliflower seedlings cv. Chaser<br />

were planted into MK12 pots with coco peat. Six replicate plants<br />

were treated with 30 ml each of fungicide drench (Table 2)<br />

applied to the soil surface either two days before or two days<br />

after mycelial plug inoculation (2). <strong>Plant</strong>s were maintained and<br />

assessed similarly to the R. solani trial.<br />

RESULTS AND DISCUSSION<br />

R. solani. All controls were infected with R. solani, AG 2.1 being<br />

the most virulent and AG 2.2 the least virulent, with 100% and<br />

70% of untreated plants infected respectively. None of the<br />

fungicides effectively controlled the disease; however Amistar,<br />

Cabrio, Maxim, Sumisclex and Jockey did reduce the severity<br />

(Table 1). The different AG groups responded differently to the<br />

fungicides, for example Rizolex at 0.4ml/L did not suppress AG<br />

2.1 or 4, but was effective against AG 2.2. R. solani was detected<br />

in soil from all treatments (data not presented).<br />

L. maculans. The disease developed slowly, with symptoms not<br />

showing on many plants until 10 weeks after inoculation. All<br />

control plants were infected, and none of the treatments were<br />

effective when applied after inoculation (Table 2). Maxim, Cabrio<br />

and Rovral provided some suppression of disease when applied<br />

before inoculation, and complete control was achieved with a<br />

pre‐ plant drench of the higher rate of Amistar.<br />

Table 1. Mean per cent severity of stem canker symptoms on plants 8<br />

weeks after being drenched with a fungicide prior to planting into R.<br />

solani inoculated soil.<br />

Treatment Rate /L AG2.1 AG2.2 AG4<br />

Untreated 74 24 46<br />

Sumisclex 500 0.75ml 18 12 18<br />

Rizolex liquid 0.2ml 78 14 42<br />

Rizolex liquid 0.4ml 62 2 82<br />

Jockey 1ml 10 6 26<br />

Terrachlor 2g 52 16 24<br />

Rovral Aquaflo 0.5ml 36 6 36<br />

Rovral Aquaflo 1ml 30 6 34<br />

Amistar 0.5ml 20 4 10<br />

Amistar 1ml 16 10 14<br />

Cabrio 0.4ml 16 12 22<br />

Maxim 0.4ml 16 6 10<br />

L.S.D (P=0.05) 14.5 12.7 17.0<br />

Table 2. Per cent incidence (inc) and severity (sev) of stem canker<br />

symptoms on plants drenched with fungicides and inoculated with L.<br />

maculans before or after treatment.<br />

Pre inoc.<br />

Post inoc. drench<br />

drench<br />

Treatment Rate /L Inc Sev Inc Sev<br />

Untreated 100 70 100 27<br />

Sumisclex 500 0.75ml 100 40 100 67<br />

Rizolex liquid 0.2ml 67 17 83 27<br />

Rizolex liquid 0.4ml 83 30 100 27<br />

Jockey 1ml 83 27 100 24<br />

Terrachlor 2g 100 20 100 30<br />

Rovral Aquaflo 0.5ml 83 17 100 44<br />

Rovral Aquaflo 1ml 17 3 100 23<br />

Amistar 0.5ml 50 14 67 24<br />

Amistar 1ml 0 0 67 20<br />

Cabrio 0.4ml 33 7 100 27<br />

Maxim 0.4ml 33 7 67 20<br />

LSD (P=0.05) ‐ 16.2 ‐ 16.5<br />

CONCLUSION<br />

No fungicide treatments were effective in controlling stem<br />

canker; however suppression was achieved with a pre planting<br />

drench of either Amistar, Cabrio or Maxim.<br />

ACKNOWLEDGEMENTS<br />

This project was facilitated by HAL in partnership with AUSVEG<br />

and was funded by the National Vegetable Levy. The Australian<br />

Government provides matched funding for all HAL’s R&D<br />

activities.<br />

REFERENCES<br />

1. Hitch C.J. et al (2006). Identifying the cause of brassica stem<br />

canker. 4th <strong>Australasian</strong> Soilborne Diseases Symposium, NZ, 2006.<br />

2. Hall, B. et al (2009). Varietal resistance of cauliflower cultivars to<br />

soilborne diseases Rhizoctonia solani and Leptosphaeria maculans.<br />

5th <strong>Australasian</strong> Soilborne Diseases Symposium, NSW, 2009.<br />

3. Paulitz TC & Schroeder KL (2005) A new method for the<br />

Quantification of Rhizoctonia solani and R. oryzae from soil. <strong>Plant</strong><br />

Disease 89, 767–772.<br />

88 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Evaluation of spray programs for powdery mildew management in greenhouse<br />

cucumbers<br />

INTRODUCTION<br />

K.L Ferguson{ XE "Ferguson, K.L." }, B.H. Hall and T.J. Wicks<br />

South Australian Research and Development Institute, GPO Box 397, Adelaide SA 5001<br />

Powdery mildew (Podosphaera xanthii) is a serious disease of<br />

greenhouse cucumbers worldwide causing defoliation and<br />

premature senescence of crops. The disease is mainly managed<br />

with fungicides and management is becoming increasingly<br />

difficult due to fungicide resistance and the lack of registered<br />

products for greenhouse use. Trials were conducted to examine<br />

spray intervals for conventional fungicides and to determine<br />

whether ‘soft’ products could be effectively incorporated into<br />

spray programs for powdery mildew.<br />

MATERIALS AND METHODS<br />

Spray program trials were conducted on cucumber plants in a<br />

research greenhouse. In the first trial various fungicide programs<br />

were compared at spray intervals of 7 or 14 days (Table 1). The<br />

second evaluated programs incorporating ‘soft’ products that<br />

were effective in preliminary screening trials (Table 2). Two week<br />

old seedlings were exposed to infected plants (Day 0) and<br />

exposure continued for the entire trial. Sprays commenced at<br />

day 14, with 10 plants per treatment. Severity of powdery<br />

mildew was assessed on leaves on day 0 and then every 7 days<br />

on sprayed leaves. Relative Area Under Disease Progression<br />

Curves (RAUDPC) (1) for each spray program were analysed with<br />

ANOVA (Statistix 8).<br />

RESULTS AND DISCUSSION<br />

All 7 day interval programs reduced the severity of powdery<br />

mildew significantly more than the 14 day programs. Although<br />

the program with Bayfidan ® and Cabrio ® at 14 day intervals had<br />

significantly more disease than the 7 day programs, the disease<br />

level may still be commercially acceptable (Figure 1; Table 1).<br />

Table 2. Spray schedules and Relative Area Under Disease Progress<br />

Curves (RAUDPC) for trial examining ‘soft’ products in conventional spray<br />

programs.<br />

Day 14 Day 28 Day 42 Day 56 RAUDPC<br />

A R A R 4.26 a<br />

M M Bi Bi 5.15 ab<br />

E A E A 6.57 ab<br />

M M B B 6.92 b<br />

B A B A 7.39 bc<br />

B R B R 9.74 c<br />

Untreated control<br />

21.98 d<br />

A=Amistar ® 250SC (250g/L azoxystrobin) 0.8mL/L; M=Morestan ® (250g/kg<br />

oxythioquinox) 0.4g/L; B=Bayfidan ® 250EC (250g/L triadimenol) 0.4mL/L;<br />

Bi=BioCover ® (840g/L petroleum oil) 10mL/L; E=Ecocarb ® (940g/kg potassium<br />

bicarbonate) 4g/L +Synertrol ® HortiOil (905g/L emulsifiable botanical oil) 2.5mL/L;<br />

R=Rezist ® 1.5mL/L+Sett Enhanced 2mL/L ® . Means with the same letter are not<br />

significantly different (P


Session 5C—Chemical control<br />

The incidence of copper tolerant bacteria In Australian pome and stone fruit orchards<br />

INTRODUCTION<br />

S.C. Gouk{ XE "Gouk, S.C." }. A , E. Mace B and EA. Byrne B<br />

A Department of Primary Industries, Private Bag 15, Ferntree Gully Business Centre, 3156, Victoria<br />

B Department of Primary Industries, Private Bag 1, Ferguson Road, Tatura, 3616, Victoria<br />

Australian pome and stone fruit production relies heavily on<br />

copper‐based fungicides for control of bacterial diseases<br />

including bacterial blast and bacterial canker (Pseudomonas<br />

syringae). Copper fungicides have had a long history of use for<br />

control of fungal diseases and general clean up during the<br />

dormant period. The impact of their long term application on<br />

development of copper tolerant bacterial populations in fruit<br />

orchard is not known. The authors had initiated collection and<br />

screening of copper tolerant bacteria from Australian pome and<br />

stonefruit fruit orchards since 2007. This study presents an<br />

analysis of the copper tolerant (CuT) bacterial populations<br />

detected in the 2008–2009 growing season.<br />

MATERIALS AND METHODS<br />

Pome and stone fruit blossoms were collected during the spring<br />

of 2008, from major fruit growing regions in Victoria, New South<br />

Wales, Queensland, Tasmania and South Australia. Up to four<br />

samples consisting of ten blossoms, one from each tree, were<br />

collected from monitored orchard block. Each sample was<br />

washed in 5 ml sterile distilled water and a 0.1 ml aliquot was<br />

spread on Pseudomonas Fluorescent Agar (PFA, Difco). Pure<br />

cultures were obtained by repeated streaking and dilution<br />

plating. The purified bacteria were characterised based on<br />

cultural and biochemical properties: namely, production of<br />

fluorescent pigments, hypersensitivity reaction on tobacco, and<br />

oxidase, levan and arginine reactions (1,2). Whilst additional<br />

tests are being conducted to further identify the bacteria, for the<br />

purpose of this report, they were categorised as fluorescent<br />

Pseudomonas spp., yellow bacteria, and non‐fluorescent, nonyellow<br />

(NFY) bacteria.<br />

The sensitivity of bacteria to copper ions was tested using a ten<br />

fold dilution of a slightly turbid bacterial suspension, yielding<br />

10 6 –10 8 colonly forming units, determined by dilution plating on<br />

PFA. Four 10 ul drops of each bacterial suspension were added<br />

to casitone‐yeast‐extract medium (3) amended with 0, 0.16,<br />

0.32, 0.48 and 0.64 millimolar (mM) of copper sulphate (Cu).<br />

Formation of bacterial colony was assessed after incubation at<br />

27±1°C for 3 days.<br />

RESULTS<br />

A total of 315 isolates comprising 155, 77 and 83 Pseudomonas,<br />

spp., yellow bacteria, and NFY bacteria respectively were tested<br />

for sensitivity to Cu ions (Table 1).<br />

Table 1. The incidence of copper tolerant bacteria in pome and stone<br />

fruit orchards in the 2008–2009 season.<br />

Pseudomonas<br />

spp. Yellow bacteria NFY bacteria<br />

Source Tot CuT Tot CuT Tot CuT<br />

Apple and pear 68 44 36 5 51 16<br />

Stone fruit 87 65 41 6 32 17<br />

Total 155 106 77 11 83 33<br />

NFY—Non‐fluorescent, non‐yellow bacteria.<br />

Tot—Total number of bacterial isolates tested.<br />

CuT—Number of copper tolerant bacterial isolates.<br />

DISCUSSION<br />

The findings indicate detection of CuT bacteria in pome and<br />

stone fruit orchards for the first time in Australia. The detection<br />

of high proportions of CuT bacteria suggests the potential risk of<br />

selection pressure associated with application of copper sprays.<br />

The implication of this finding on the effectiveness of copperbased<br />

sprays, which is the only chemical treatment available for<br />

the control of bacterial diseases, needs further research. Further<br />

evaluation of the impact of copper sprays on CuT bacterial<br />

populations is required to determine the continued efficacy of<br />

copper‐based bactericides.<br />

ACKNOWLEDGEMENTS<br />

Funding from DPI, Victoria; Horticulture Australia Limited, Apple<br />

and Pear Australia Limited, Canned Fruit Industry Council of<br />

Australia and Fruit Grower Victoria. Collaborators: A. Steinhauser<br />

(FGT, Tasmania), S. Hetherington, A. Moorey (DPI NSW), P.<br />

James (PIRSA, SA), L. Haselgrove (QDPI, Queensland).<br />

REFERENCES<br />

1. Lelliot RA, Billing E, Hayward, AC (1966) A determinative scheme<br />

for the fluorescent plant pathogenic pseudomonads. J. Applied<br />

Bacteriol. 29, 470–489.<br />

2. Schaad NW (2001) Initial identification of common bacteria. In<br />

Laboratory guide for identification of plant pathogenic bacteria.<br />

(Eds NW Schaad, JB Jones, W Chun) pp. 1–16. (APS Press:<br />

Minnesota) pp. 373.<br />

3. Andersen GL et al. (1991) Occurrence and properties of copper<br />

tolerant strains of Pseudomonas syringae isolated from fruit trees<br />

in California. Phytopathology 81, 648–656.<br />

The number of isolates that were able to form confluent colony<br />

at 0.32 mM Cu (CuT), the threshold concentration for tolerance<br />

to copper ions (4), are presented in Table 1.<br />

Overall, 47.6% of the isolates tested were CuT. Twenty‐seven<br />

CuT isolates were potentially plant pathogenic based on their<br />

ability to induce hypersensitivity (HR+) response in tobacco. Of<br />

155 isolates from apple and pear hosts, 40.0% were CuT, but<br />

only 10 isolates i.e. 6.5%, were both CuT and HR+. Six<br />

Pseudomonas spp. from pome fruit were CuT and HR+, but<br />

fourteen CuT and HR+ isolates were obtained from stone fruit<br />

hosts. Whilst more than half (68.4%) of the Pseudomonas spp.<br />

from both pome and stone fruits were CuT, only 14.3 and 39.8%<br />

of yellow and NFY bacteria respectively were CuT.<br />

90 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Lessons from the tropics—the unfolding mystery of vascular‐streak dieback of cocoa,<br />

the importance of genetic diversity, horizontal resistance, and the plight of farmers<br />

P.J. Keane{ XE "Keane, P.J." }<br />

Department of Botany, La Trobe University, Bundoora Victoria 3086, Australia<br />

Cocoa in Papua New Guinea and South‐East Asia has been<br />

devastated by a mysterious dieback disease at least since the<br />

rapid expansion of planting in the 1950s. Investigations into the<br />

nature of this disease, its control and the ongoing mystery<br />

surrounding it will be described as an example of unusual<br />

biology likely to be common in the relatively unexplored biology<br />

of the wet tropics. The incredible genetic diversity of plant<br />

communities and its importance especially in the wet tropics will<br />

be introduced in relation to disease control in food crops and<br />

selection of resistance to vascular‐streak dieback of cocoa. The<br />

nature of the durable resistance of cocoa to vascular‐streak<br />

dieback will be discussed as a tribute to J.E. van der Plank.<br />

Recent changes in the nature of the disease found during a<br />

current ACIAR project in Sulawesi will be described as an<br />

example of the uncertainty of human knowledge in the face of<br />

ever‐changing biology. Finally, the plight of cocoa farmers facing<br />

serious pest and disease problems in the region will be discussed<br />

as an example of the poor situation of farmers throughout the<br />

world.<br />

McAlpine lecture<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 91


Keynote address<br />

Translating research into the field: how it started, how it is practised and how we<br />

carry out grape powdery mildew research<br />

Robert Seem{ XE "Seem, R." }<br />

Department of <strong>Plant</strong> <strong>Pathology</strong> and <strong>Plant</strong>‐Microbe Biology, Cornell University, Geneva, New York 14456 USA<br />

INTRODUCTION<br />

To understand research translation, it is helpful to understand<br />

our past efforts to extend knowledge, especially agricultural<br />

knowledge, to those who need and benefit from that<br />

information. The traditional term to define this process is<br />

Extension. Extension as it exists today in the United States has a<br />

rich history of nearly 150 years. We will examine the<br />

development of the Cooperative Extension Service in the US,<br />

especially at the time (mid 19th century) when educational and<br />

research institutions were evolving dramatically (1). In that era<br />

of change, colleges began to be more egalitarian and outward<br />

looking (extension). Extension has survived and grown over the<br />

years, but some of the mission fervour had faded. A renewed<br />

commitment to extension is now embodied within the latest<br />

buzzword of the 21st century: engagement. We will examine<br />

how extension and engagement influence research using the<br />

example of grapevine powdery mildew. We will consider what<br />

drives the process and how we can sustain it. Some final<br />

reflections look towards the future and thoughts on the value of<br />

translating research to the field.<br />

EXTENSION’S CREATION AND TRANSITION TO ENGAGEMENT<br />

The Genesis of Extension It was during the American Civil War<br />

that an innovative approach was taken to permit a broadened<br />

reach of higher education. The Morrill Act of 1862 established a<br />

federal partnership with states to establish centres of learning<br />

for agriculture and the mechanic arts. The act provided tracts of<br />

federal land for states to use or sell in order to creat “landgrant”<br />

colleges. Shortly thereafter (1865) Cornell University was<br />

established as New York State’s land‐grant college. Almost thirty<br />

years later (1894) Cornell’s first professor of horticulture, Liberty<br />

Hyde Bailey, implemented the nation’s first extension program,<br />

“a plain, earnest, and continuous effort to meet the needs of the<br />

people on their own farms and in the localities.”<br />

Extension in the 21st Century Fast forward to 2009, and the<br />

modern Extension Service remains one of the very unique<br />

features of American agriculture. However, it has moved beyond<br />

farmers and addresses many aspects of rural and urban life. In<br />

an analysis of the future of the land‐grant universities, the<br />

Kellogg Commission exhorted these institutions to become<br />

“engaged universities” (2). By engagement they refer to<br />

institutions that have redesigned their teaching, research, and<br />

extension and service functions to become even more<br />

sympathetically and productively involved with their<br />

communities, however community may be defined.”<br />

MEETING THE CHALLENGE<br />

So how does this new engagement play out for a university<br />

researcher who has no official extension responsibility? For<br />

those of us who carry out mission‐oriented research we are<br />

always asking two questions: What are the challenges facing the<br />

industry? How does my research address those challenges? In<br />

the context of plant pathology, our first goal is to carry out good<br />

science; but our science is always focused on solving problems as<br />

we expand our knowledge.<br />

GRAPEVINE POWDERY MILDEW CHALLENGES WE HAVE<br />

ADDRESSED<br />

Over the past 30 years with the assistance of colleagues and<br />

students we have examined, and in some cases, redefined the<br />

role and impact of Erysiphe necator on grape, always driven by<br />

industry needs (e.g., 3‐5):<br />

• The importance of early‐season disease management<br />

• The role of cleistothecia in epidemics<br />

• Environment, ascospore release, and infection<br />

• Early season disease spread within vineyards<br />

• Effect of cold nights on mildew development<br />

• Ontogenic resistance in grape berries<br />

• Heterogeneity of berry development and susceptibility<br />

• Diffuse mildew infection on berries and wine quality<br />

• Signals that turn sporulation on and off<br />

• Biological control of powdery mildew by Tydeid mites<br />

CONCLUSIONS<br />

Our research on the biology and epidemiology of grape powdery<br />

mildew has had a major impact on grape production in New York<br />

and has influenced research and extension programs throughout<br />

the US and beyond. The program is not substantially different<br />

from many other mission‐oriented research programs, but for<br />

those who seek to translate research to the field, this is what we<br />

have learned:<br />

• Be grounded – understand and appreciate agriculture<br />

• Observe; look at the big picture, but mind the little stuff<br />

• Know your patients and your pathogens<br />

• Become engaged; know your stakeholders and deal with<br />

them as peers<br />

• Write proposals that are clear succinct, and interesting<br />

• Collaborate, locally and professionally<br />

• Train the next generation; get them excited about<br />

translational research<br />

• Have fun while doing all this!<br />

ACKNOWLEDGEMENTS<br />

For their vital collaborations, I thank David Gadoury and the labs<br />

of Wayne Wilcox, Greg Loeb, Thomas Henick‐Kling, Lance Cadle‐<br />

Davidson, and Ian Dry.<br />

REFERENCES<br />

1. Boyer, EL (1990) Scholarship reconsidered: priorities of the<br />

professoriate. (Carnegie Foundation for the Advancement of<br />

Teaching: Princeton, NJ) 147pp.<br />

2. Kellogg Commission on the Future of State and Land‐Grant<br />

Universities (1999) Returning to our roots: The engaged institution,<br />

Report 3. (National Association of State Universities and Land‐<br />

Grant Colleges: Washington DC) 57 pp.<br />

3. English‐Loeb, G, Norton, AP, Gadoury, D, Seem, RC, and Wilcox, WF<br />

(2007) Biological control of grape powdery mildew using<br />

mycophagous mites. <strong>Plant</strong> Dis. 91, 421‐429.<br />

4. Gadoury, DM, Seem, RC, Ficke, A, and Wilcox, WF (2003) Ontogenic<br />

resistance to powdery mildew in grape berries. Phytopathology 93,<br />

547‐555.<br />

5. Gadoury, DM, Seem, RC, Wilcox, WF, Henick‐Kling, T, Conterno, L,<br />

Day, A, and Ficke, A (2007) Effects of diffuse colonization of grape<br />

berries by Uncinula necator on bunch rots, berry microflora, and<br />

juice and wine quality. Phytopathology 97, 1356‐1365.<br />

92 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Stem rust race Ug99: international perspectives and implications for Australia<br />

INTRODUCTION<br />

R.F. Park B , H.S. Bariana B and C.R. Wellings{ XE "Wellings, C.R." } A,B<br />

The University of Sydney, <strong>Plant</strong> Breeding Institute, PMB 11, Camden, 2570, NSW<br />

B seconded to The University of Sydney, <strong>Plant</strong> Breeding Institute, PMB 11, Camden, NSW 2570<br />

Stem rust of wheat, caused by Puccinia graminis f. sp. tritici<br />

(Pgt), is one of the most feared plant diseases. It was particularly<br />

problematic in early efforts to grow wheat in Australia, being<br />

described by McAlpine (1) as “positively injurious”. Reports of<br />

losses in Australia include £400,000 in 1903, £2 million in 1916,<br />

£7 million in 1947, and $200 to 300 million in 1973 (2).<br />

Concerted efforts to control the disease with genetic resistance<br />

began with the release of cultivar Eureka in 1938, and have led<br />

to a decline in the incidence of the disease and in the frequency<br />

of epidemics.<br />

The detection of stem rust race “Ug99” in Uganda in 1999 has<br />

had significant implications for the control of stem rust.<br />

Referring to “Ug99”, Nobel Laureate Dr Norman Borlaug, stated<br />

“The prospect of a stem‐rust epidemic in wheat in Africa, Asia,<br />

and the Americas is real and must be stopped before it causes<br />

untold damage and human suffering”<br />

(http://www.sciencenews.org/articles/20050924/food.asp).<br />

PATHOGENIC VARIABILITY IN AUSTRALIA<br />

Annual pathogenicity surveys of Pgt conducted at the University<br />

of Sydney since 1919 have provided a sound basis for rust<br />

resistance breeding efforts. These surveys have revealed 3<br />

incursions of 4 exotic Pgt isolates, all of which had significant<br />

impacts on wheat production, highlighting the importance of<br />

current exotic threats such as Pgt race “Ug99”. The surveys have<br />

also shown clear evidence of the importance of pathogen<br />

aggressiveness, of single step mutation and of somatic<br />

hybridisation in overall pathogen population structure (2).<br />

STEM RUST RACE Ug99<br />

Since its first detection in Uganda in 1999, “Ug99” has been<br />

detected in Kenya, Ethiopia, Sudan and Yemen, and in 2007 was<br />

detected in Iran (3). It carries virulence matching many<br />

resistance genes in hexaploid wheat: genes rendered ineffective<br />

are Sr5, Sr6, Sr7b, Sr8a, Sr8b, Sr9a, Sr9b, Sr9d, Sr9e, Sr9g, Sr11,<br />

Sr15, Sr17, Sr21, Sr30, Sr31, and Sr38; those that remain<br />

effective are: Sr7a, Sr13, Sr22, Sr24, Sr25, Sr26, Sr27, Sr28, Sr29,<br />

Sr32, Sr33, Sr35, Sr36, Sr37, Sr39, Sr40 and Sr44. Of particular<br />

concern is virulence for gene Sr31, one of the most widely<br />

deployed stem rust resistance genes, which remained effective<br />

until the detection of “Ug99”. Since its first detection, two<br />

presumed mutational derivatives with virulence for Sr24 (4) and<br />

for Sr36 (5) have been detected in Kenya (4), and a race with<br />

identical virulence but lacking virulence for Sr31 has been<br />

detected in South Africa (6).<br />

“Ug99” has had a large impact on a wide range of international<br />

wheat germplasm. In response, the Borlaug Global Rust Initiative<br />

(BGRI) was launched in Kenya in 2005 and acknowledged the<br />

threat of “Ug99” to stable wheat production in eastern Africa,<br />

and also recognised the threat posed to many other parts of the<br />

world. More recently, an international project, “Durable Rust<br />

Resistance in Wheat” was funded by the Bill and Melinda Gates<br />

Foundation and is managed by Cornell University. This project<br />

has a range of activities that include pre‐breeding, rust pathogen<br />

surveillance and cultivar deployment.<br />

IMPLICATIONS FOR AUSTRALIA<br />

The Australian Cereal Rust Control Program (ACRCP) provides<br />

support to all groups engaged in cereal breeding in Australia,<br />

and undertakes research on the pathology and genetics of rust<br />

diseases. This strategy has led to a robust understanding of the<br />

resistance genes deployed in Australia, and a resulting ability to<br />

predict response of Australian germplasm to new rust races such<br />

as “Ug99”. These predictions have been refined by field testing<br />

germplasm in Kenya with the assistance of the Kenyan<br />

Agricultural Research Institute from 2005–07. Because Sr31 has<br />

not been used widely in Australia, the greatest impact of “Ug99”<br />

on germplasm to date has been due to virulence for Sr30,<br />

combined virulence for Sr38 with other genes, and more<br />

recently, virulence for Sr24 and Sr36. While virulences for Sr30,<br />

Sr36 and Sr38 have been detected in Australia, virulence for Sr24<br />

has not. The genes Sr2, Sr12, Sr13, Sr22 and Sr26, effective<br />

against “Ug99” and derivatives, are important contributors to<br />

the resistance present in current germplasm.<br />

CONCLUDING COMMENTS<br />

Although “Ug99” may never reach our shores, it must be<br />

regarded as a serious exotic rust threat. Whilst acknowledging<br />

this, it is equally important not to forget other exotic rust<br />

threats, including races of endemic cereal rust pathogens, plus<br />

the cereal rust diseases that do not occur here (stripe rust of<br />

barley, Puccinia striiformis f. sp. hordei; leaf rust of durum<br />

wheat, Puccinia sp. Group II Type A; crown rust of barley, P.<br />

coronata var. hordei.<br />

ACKNOWLEDGEMENTS<br />

The Australian Cereal Rust Control Program is supported<br />

financially by the Grains Research and Development<br />

Corporation, Australia.<br />

REFERENCES<br />

1. McAlpine D (1906) ‘The Rusts of Australia. Their Structure, Nature,<br />

and Classification’. (Department of Agriculture, Victoria:<br />

Melbourne).<br />

2. Park RF (2007) Stem rust of wheat in Australia. Australian Journal of<br />

Agricultural Research 58, 558–566.<br />

3. Narari K, Mafi M, Yayhaoui A, Singh RP, Park RF (2009). Detection<br />

of wheat stem rust (Puccinia graminis f. sp. tritici) race TTKS (Ug99)<br />

in Iran. <strong>Plant</strong> Disease 93, 317.<br />

4. Jin Y, Szabo LJ, Pretorius ZA, et al. (2008). Detection of virulence to<br />

resistance gene Sr24 within race TTKS of Puccinia graminis f. sp.<br />

tritici. <strong>Plant</strong> Disease 92, 923–926.<br />

5. Jin Y, Szabo LJ, Rouse MN et al. (2009) Detection of virulence to<br />

resistance gene Sr36 within the TTKS race lineage of Puccinia<br />

graminis f. sp. tritici. <strong>Plant</strong> Disease 93, 367–370.<br />

6. Visser B, Herselman L, Pretorius ZA (2009) Genetic comparison of<br />

Ug99 with selected South African races of Puccinia graminis f. sp.<br />

tritici. Molecular <strong>Plant</strong> <strong>Pathology</strong> 10, 213–222.<br />

Session 6A—Cereal pathology 1<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 93


Session 6A—Cereal pathology 1<br />

Mitigating crop losses due to stripe rust in Australia: integrating pathogen population<br />

dynamics with research and extension programs<br />

INTRODUCTION<br />

C.R. Wellings{ XE "Wellings, C.R." } A,B and K.R. Kandel B<br />

A NSW Department Primary Industries, seconded to<br />

B The University of Sydney, <strong>Plant</strong> Breeding Institute, PMB 11, Camden, NSW 2570<br />

Wheat stripe rust (caused by Puccinia striiformis f. sp. tritici, Pst)<br />

was first recorded in Australia in 1979 and became endemic to<br />

the eastern Australian wheat zone causing serious losses in the<br />

mid eighties (1). Concerted pathology and breeding R&D<br />

combined with industry adoption of resistant varieties resulted<br />

in minimal losses for nearly 20 years. The first report of stripe<br />

rust in Western Australia in 2002 was the result of a foreign<br />

pathotype incursion (2). This aggressive pathotype widened its<br />

distribution in following years to encompass the entire<br />

Australian wheat production zone, and caused serious losses<br />

including increased annual fungicide expenditure ranging from<br />

$AUD40–90million (1).<br />

The stripe epidemic in eastern Australia in 2008 was the most<br />

intensive in the 30 year history of the disease in Australia. The<br />

dynamics of host resistance and pathogen variability gave rise to<br />

a situation that required a close connection between extension<br />

and research staff in order to maximise the available resources<br />

of host resistance and fungicide availability. This paper presents<br />

details of the epidemic development and the interplay of variety<br />

resistance and pathogen population dynamics during the 2008<br />

season.<br />

MATERIALS AND METHODS<br />

Rust samples collected and forwarded to PBI by co‐operators<br />

(advisors, farmers, researchers) were assessed for pathotype<br />

determination using described methods (3). Results were<br />

immediately reported to co‐operators by email. The relationship<br />

between pathotype and the resistance genes present in<br />

commercial wheats provided a basis for predicting expected<br />

disease response.<br />

RESULTS AND DISCUSSION<br />

Samples received from various regions of Australia are illustrated<br />

in Figure 1. The epidemic began early from presumed green<br />

bridge survival sites, developed slowly in winter, and became<br />

explosive in spring; the epidemic was largely confined to NSW<br />

and Queensland.<br />

Sample Number<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

Queensland<br />

n NSW<br />

s NSW<br />

Victoria<br />

South Australia<br />

Western Australia<br />

Table 1. Features and frequency of the four Pst pathotypes detected in<br />

Australia in 2008. (R=resistant; S=susceptible)<br />

Resistance Gene (Yr)<br />

First 2008 Response<br />

Pathotype<br />

Report n=830 17 J 27<br />

‘WA’ 2002 20% R R R<br />

‘WA Yr17’ 2006 12% S R R<br />

‘Jackie’ 2007 55% R S R<br />

‘Jackie Yr27’ 2008


INTRODUCTION<br />

Impact of sowing date on crown rot losses<br />

S. Simpfendorfer{ XE "Simpfendorfer, S." }<br />

NSW Department of Primary Industries, 4 Marsden Park Rd, Tamworth, 2340, NSW<br />

Crown rot caused by the fungus Fusarium pseudograminearum<br />

(Fp) is a major constraint to winter cereal production in the<br />

northern cropping region especially under no‐till farming<br />

systems (3). Yield loss from crown rot interacts heavily with<br />

moisture stress during grain‐fill. One way to manipulate this<br />

interaction is through sowing time. Only two studies have ever<br />

examined this interaction under natural field infections which<br />

both found that earlier sowing increased the incidence of crown<br />

rot (1,2). An issue with these previous studies is that they are<br />

unable to differentiate seasonal interactions from the direct<br />

crown rot effects. An inoculated versus uninoculated<br />

experimental design, as suggested in (2), was adopted in this<br />

study to allow the direct effects of crown rot to be determined<br />

on yield and quality across three sowing dates.<br />

MATERIALS AND METHODS<br />

Three bread wheat varieties (EGA Gregory, Strzelecki and EGA<br />

Wylie) were used with the first two being longer season and<br />

rated as being more susceptible to crown rot and EGA Wylie a<br />

main season variety which has the best resistance rating. Plots of<br />

each variety were either uninoculated or inoculated with<br />

sterilised durum grain colonised by Fp at a rate of 2g/m of row.<br />

Plots of each treatment were then sown on three different dates<br />

at Tamworth in 2008 being: 1st sowing = 21st May, 2nd sowing =<br />

10th June and 3rd sowing = 27th June. There were four<br />

replicates of each treatment which were blocked for sowing time<br />

with treatments randomised within each block. Hand samples<br />

were removed from each plot at physiological maturity to obtain<br />

pathology measures while yield and quality were obtained from<br />

samples collected using a small plot harvester.<br />

RESULTS<br />

Good rainfall occurred at Tamworth late in the season during<br />

grain‐fill which prevented the formation of whiteheads in all<br />

treatments. There was no significant variety x inoculum or<br />

variety x sowing time x inoculum effect on yield given this good<br />

finish to the season. Sowing time had a significant impact on<br />

final grain yield in all three varieties with the average percentage<br />

yield reduction between the 1st and 2nd sowing for the three<br />

varieties being ‐9% and between the 1st and 3rd sowing ‐22.6%.<br />

Crown rot had less of an impact on yield at each sowing date<br />

causing ‐4.1% yield loss at 1st sowing, ‐3.4% 2nd sowing and ‐<br />

6.7% at 3rd sowing date.<br />

Percentage screenings were also significantly affected by sowing<br />

time (1st to 2nd sowing date +1.0%; 1st to 3rd sowing date<br />

+3.5%). Crown rot also had a direct effect of increasing<br />

screenings at each sowing date with a trend towards increased<br />

negative impacts with delayed sowing (1st +0.7%, 2nd +1.5% and<br />

3rd +1.7%).<br />

There was no difference between the three varieties in the levels<br />

of infection initiated by Fp at any sowing date i.e. longer season<br />

varieties did not have greater numbers of plants infected<br />

irrespective of sowing time. In plots where no additional Fp<br />

inoculum was added (background infections) there was no<br />

difference in the percentage of plants infected at harvest<br />

between the three sowing times. When Fp inoculum was added,<br />

all sowing times resulted in around 80% of plants or greater<br />

being infected at harvest with the 2nd sowing time being<br />

significantly higher at 94% than the other two sowing times.<br />

However, with both inoculum levels it was obvious that early<br />

sowing did not result in increased numbers of infected plants at<br />

harvest.<br />

Although early or delayed sowing time did not impact on the<br />

percentage of plants ultimately infected by Fp, it did appear to<br />

influence disease expression as measured by the extent of basal<br />

browning (i.e. crown rot severity). Delaying sowing time<br />

significantly increased disease severity across the three sowing<br />

dates at both inoculum levels.<br />

DISCUSSION<br />

Sowing time and hence length of exposure to infection over the<br />

season did not result in different levels of plants being infected<br />

by Fp at harvest. The 2008 season was very conducive to<br />

infection with good soil moisture for much of the year. Certainly<br />

longer season varieties and earlier sowing did not increase<br />

susceptibility to infection.<br />

Earlier sowing increased yield and reduced screenings<br />

irrespective of crown rot infection. The actual % yield loss to<br />

crown rot did not vary greatly between sowing times with each<br />

of the varieties. There was an indication that crown rot resulted<br />

in increased screenings with later sowings. The 2008 season was<br />

not overly conducive to yield and quality loss from crown rot but<br />

differences were still evident. It would be interesting to repeat<br />

this experiment in a season with a tougher finish. In theory,<br />

bringing grain‐fill forward even 1–2 weeks may have a<br />

considerable impact on disease expression by limiting moisture<br />

and evaporative stress.<br />

The major effect on yield and quality comes from the sowing<br />

time itself. Later sowing decreases yield potential and grain size<br />

and increases screenings. Adding crown rot into the picture on<br />

top of this further exacerbates these losses thus increasing the<br />

probability of downgrading. The % yield and quality losses<br />

attributable to crown rot were pretty consistent across the three<br />

sowing dates. If anything they got slightly worse with the later<br />

sowings. Hence, sowing earlier in the window, if soil moisture<br />

allows, maximises the genetic yield potential, grain size and<br />

limits screenings in a variety. This provides buffering from any<br />

detrimental effect that crown rot infection may then have.<br />

ACKNOWLEDGEMENTS<br />

Partial funding for this research was provided by the Grains<br />

Research and Development Corporation.<br />

REFERENCES<br />

1. Purss GS (1971) Effect of sowing time on the incidence of crown rot<br />

(Gibberella zeae) in wheat. Australian Journal of Experimental<br />

Agriculture and Animal Husbandry 11, 85–89.<br />

2. Klein TA, Burgess LW, Ellison FW (1989) The incidence of crown rot<br />

in wheat, barley and triticale when sown on two dates. Australian<br />

Journal of Experimental Agriculture 29, 559–563.<br />

3. Burgess LW, Backhouse D, Summerell BA, Swan LJ (2001) Crown rot<br />

of wheat. In: Fusarium. (Ed BA Summerell, JF Leslie, D Backhouse,<br />

WL Bryden, LW Burgess) APS Press. pp 271–294.<br />

Session 6A—Cereal pathology 1<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 95


Session 6A—Cereal pathology 1<br />

Symptom development and pathogen spread in wheat genotypes with varying levels<br />

of crown rot resistance<br />

C.D. Malligan{ XE "Malligan, C.D." } A,B , M.W. Sutherland A and G.B. Wildermuth C<br />

A<br />

University of Southern Queensland, West St, Toowoomba, 4350, QLD<br />

B Department of Employment, Economic Development and Innovation, Queensland Primary Industries and Fisheries, Leslie Research<br />

Centre, Toowoomba, 4350, QLD<br />

C Retired from Department of Employment, Economic Development and Innovation, Queensland Primary Industries and Fisheries, Leslie<br />

Research Centre, Toowoomba, 4350, QLD<br />

INTRODUCTION<br />

Crown rot, caused by Fusarium pseudograminearum (Fpg), is an<br />

important soilborne disease of winter cereals. Complete<br />

resistance has yet to be reported in any wheat genotypes and<br />

hence is an ongoing issue for Australian wheat growers. In order<br />

to understand the nature of the partial resistance identified to<br />

crown rot we have examined the patterns of disease and<br />

pathogen spread in both susceptible and partially resistant<br />

tissues. Field trials were designed to study disease symptom<br />

development and localisation of Fpg hyphae in the bread wheat<br />

varieties Puseas, Vasco and Sunco, and the line 2–49.<br />

MATERIALS AND METHODS<br />

Inoculated field trials were conducted, using a randomised block<br />

design. Inoculum was placed in a band lying above the seed at<br />

sowing. Five plants from three replicates were harvested at<br />

approximately fortnightly intervals throughout the growing<br />

season. Leaf sheaths and internodes of the 1st 5 tillers were<br />

rated for disease using a scale from 0 to 4 as described in<br />

Wildermuth & McNamara (1), where 0 = no lesions evident and<br />

4 = >75% of tissue lesioned. Following disease rating, each tissue<br />

piece from two replicates was surface sterilised and plated out<br />

on Czapek Dox agar. Plates were checked daily for 5 days after<br />

plating. Sites of colony emergence were marked with ink on the<br />

abaxial plate surface.<br />

Disease rating data were analysed untransformed and the<br />

isolation counts were square‐root transformed prior to analysis.<br />

Restricted Maximum Likelihood (REML) Variance Components<br />

Analysis was used to determine significance of the fixed factors<br />

harvest, genotype, tiller and the corresponding two and three<br />

way interactions. To determine where individual means were<br />

significantly different 95% confidence intervals of error were<br />

calculated for each analysed plant part.<br />

RESULTS AND DISCUSSION<br />

Differences in moisture conditions between the two field trials<br />

resulted in differences in overall plant development, extent of<br />

Fpg colonisation and symptom expression. The results of the<br />

second trial conducted under higher moisture conditions will be<br />

presented here.<br />

Disease symptoms developed and Fpg was isolated from plant<br />

parts of all tillers of all genotypes. Statistically significant<br />

differences between genotypes were not expressed in the<br />

disease rating or isolation of Fpg from leaf sheath tissue in field<br />

trials even at the seedling stage (data not shown). Significant<br />

differences were seen between partially resistant and<br />

susceptible wheat genotypes in both disease rating and<br />

isolations from internode tissues and this could be detected<br />

soon after stem extension commenced (Figures 1 and 2).<br />

for rating field material for crown rot screening. At later harvests<br />

differences between genotypes were clearly expressed in higher<br />

internodes and at maturity lesions had developed as high as the<br />

2nd internode in 2–49, 4th in Sunco, and the 5th internode in<br />

Puseas and Vasco. At maturity Fpg was consistently recovered<br />

from the 4th internode in 2–49 and the 5th in all other tested<br />

genotypes, indicating a delay in symptom expression in the<br />

infected 2–49 tissues.<br />

Mean Disease Rating<br />

4<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

Mean Disease Rating of Internode 1<br />

(Tillers combined) +/- 95% confidence interval of error<br />

Booting Anthesis Milk<br />

Development<br />

Dough<br />

Development<br />

Puseas<br />

Vasco<br />

Sunco 2-49<br />

Ripening<br />

Figure 1. Mean disease rating of internode 1 at 13 (booting) to 22 weeks<br />

after planting (WAP) (ripening). n = 75 tillers<br />

Mean Number of Isolations<br />

3.00<br />

2.50<br />

2.00<br />

1.50<br />

1.00<br />

0.50<br />

0.00<br />

Square Root of Isolations from Internode 1<br />

(Tillers combined) +/- 95% confidence interval of error<br />

Booting Anthesis Milk<br />

Development<br />

Dough<br />

Development<br />

Puseas<br />

Vasco<br />

Sunco 2-49<br />

Ripening<br />

Figure 2. Mean isolations from internode 1 at 13 (booting) to 22 WAP<br />

(ripening). n = 50 tillers<br />

CONCLUSIONS<br />

Resistance was expressed as a slowing down of colonisation of<br />

plant parts in 2–49 and to a lesser extent in Sunco when<br />

compared to the susceptible genotype Puseas. Both colonisation<br />

and disease symptoms are initially slowed in young tissues of<br />

partially resistant genotypes however at later harvest times<br />

these same tissues may be as infected and symptomatic as the<br />

tissues of susceptible genotypes.<br />

ACKNOWLEDGEMENTS<br />

Financial support provided by GRDC PhD scholarship to CDP.<br />

Large differences in symptom expression were seen between<br />

genotypes in internode 1 around anthesis but not at maturity.<br />

This is an important observation as maturity is a favoured time<br />

REFERENCES<br />

1. Wildermuth, G. B. and McNamara, R. B. (1994). Testing wheat<br />

seedlings for resistance to crown rot caused by Fusarium<br />

graminearum Group 1. <strong>Plant</strong> Disease 78: 949–953.<br />

96 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Development of an eradication strategy for exotic grapevine pathogens<br />

M.R. Sosnowski{ XE "Sosnowski, M.R." } A,D , R.W. Emmett B,D , W.F. Wilcox C,D and T.J. Wicks A,D<br />

A South Australian Research and Development Institute, GPO Box 397, Adelaide, South Australia 5001<br />

B Department of Primary Industries Victoria, PO Box 905, Mildura, Victoria 3502<br />

C Department of <strong>Plant</strong> <strong>Pathology</strong>, Cornell University, New York State Agricultural Experiment Station, PO Box 462 Geneva, NY 14456 USA<br />

D Cooperative Research Centre for National <strong>Plant</strong> Biosecurity, LPO Box 5012, Bruce ACT 2617<br />

INTRODUCTION<br />

Eradication of exotic grapevine diseases can incur significant<br />

costs to growers and the industry using current strategies which<br />

include complete removal of affected and suspected vines.<br />

Alternative strategies need to be developed which optimise<br />

efficiency of the eradication process and minimise the economic<br />

cost of returning the crop to its previous quality and production<br />

levels (1). The endemic disease of grapevine, black spot (Elsinoe<br />

ampelina), was used as a model to develop a drastic pruning<br />

eradication strategy for the exotic disease black rot (Guignardia<br />

bidwellii). These pathogens have similar biology and<br />

epidemiology and as surface pathogens, inhabit fruit, leaves and<br />

shoots of grapevines producing similar looking lesions on these<br />

parts of the vine (2, 3).<br />

MATERIALS AND METHODS<br />

In 2006, a trial was established in the Sunraysia district of<br />

Victoria to develop and assess a drastic pruning protocol. Using a<br />

randomised block design, the trial comprised four table grape<br />

cultivars (Red Globe, Christmas Rose, Blush Seedless and Fantasy<br />

Seedless) as blocks. Plots consisted of three vines with standard<br />

two‐bud spur pruning. Vines in each plot were either drastically<br />

pruned (as described below) or left as controls. Spacing between<br />

plots within rows was at least 7.3 m and between rows was 10.5<br />

m.<br />

Vines were inoculated in spring 2007 by spraying a suspension of<br />

E. ampelina conidia on new shoots with 2–4 unfolded leaves.<br />

Inoculations were conducted at three different times to cater for<br />

differences in phenology between the cultivars. Shoots were<br />

covered with polyethylene bags overnight to provide high<br />

humidity to promote spore germination and infection.<br />

emerging leaves on potted grapevines (cv. Thompson Seedless).<br />

The leaves were incubated overnight as described above. The<br />

vines were assessed for symptoms 12 days later.<br />

RESULTS<br />

Assessment of the vines in December 2007 showed that 5–12<br />

inoculated shoots on each vine had black spot leaf lesions and<br />

stem cankers. This indicated that the vines had sufficient<br />

infection to simulate an exotic pathogen incursion for<br />

eradication.<br />

In December 2008, following the eradication, symptoms were<br />

recorded on all control vines and on 4 of 36 treated vines. On<br />

treated vines, each symptomatic shoot grew from the trunk<br />

within 20 cm of the ground. The bioassay indicated that<br />

symptoms were most likely caused by inoculum produced from<br />

vine debris remaining on the vineyard floor directly beneath low<br />

shoots.<br />

Assessment of the sentinel vines revealed that there was no<br />

spread of disease between plots or from external sources.<br />

DISCUSSION<br />

As a result of the assessment following eradication, the protocol<br />

has been modified to include removal of lower shoots when<br />

regrowth occurs on vine trunks and the use of straw mulch on<br />

the vineyard floor. The revised protocol will be applied in the<br />

second year of the eradication trial in Australia and the<br />

assessment in December 2009 will determine if the eradication<br />

was successful. Validation of the protocol for eradicating black<br />

rot has been initiated in an infected vineyard in New York USA,<br />

where the disease is endemic.<br />

Session 6B—Quarantine and exotic pathogens<br />

In July 2008, vines were drastically pruned as follows. Vines were<br />

cut at the crown using a chainsaw and excised material from<br />

above the crown was removed and placed in an excavated area<br />

about 25 m from the trial plots. The vineyard floor around the<br />

treated vines was raked and the debris was placed in the<br />

excavated area to be burnt and buried. Soil between vines was<br />

disc cultivated to bury any remaining debris. Trunks of the<br />

treated vines were drenched with lime sulphur using a back pack<br />

sprayer.<br />

Canopy misters were used in the spring as new shoots emerged<br />

to provide conditions conducive for disease development. Vines<br />

were assessed for recurrence of symptoms in December 2008.<br />

Healthy sentinel vines in pots were placed strategically within<br />

and around the trial site during spring and early summer to<br />

detect any movement of the pathogen between plots or from<br />

external sources. After periods of 2–3 weeks, the potted vines<br />

were placed in a glasshouse, drip irrigated, incubated for 4<br />

weeks at 22–28°C and inspected for symptoms.<br />

A bioassay was conducted to determine if vine debris in the soil<br />

below vines was a source of inoculum for emerging shoots. Soil<br />

from the base of treated trunks was collected and organic debris<br />

was separated using a sieve. The debris was soaked in water<br />

overnight and the water was decanted and sprayed over<br />

This research has potential to save the Australian wine industry<br />

over $18 million in lost production and vineyard reestablishment<br />

if there is an exotic disease incursion (R. Mcleod,<br />

unpublished data).<br />

ACKNOWLEDGEMENTS<br />

We would like to thank K. Clarke (DPI Vic), A. Loschiavo and D.<br />

Sosnowski (SARDI) for technical assistance and the CRC for<br />

National <strong>Plant</strong> Biosecurity for funding this research.<br />

REFERENCES<br />

1. Sosnowski MR, Fletcher JD, Daly AM, Rodoni BC and Viljanen‐<br />

Rollinson SLH (2009) Techniques for the treatment, removal and<br />

disposal of host material during programmes for plant pathogen<br />

eradication. <strong>Plant</strong> <strong>Pathology</strong> Early view online DOI: 10.1111/j.1365‐<br />

3059.2009.02042.x.<br />

2. Magarey RD, Coffey BE and Emmett RW (1993) Anthracnose of<br />

grapevines, a review. <strong>Plant</strong> Protection Quarterly 8, 106–110.<br />

3. Wilcox W (2003) Grapes: Black rot (Guignardia bidwelli (Ellis) Viala<br />

and Ravaz.). Cornell Cooperative Extension Disease Identification<br />

Sheet No. 102GFSG‐D4, Cornell University.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 97


Session 6B—Quarantine and exotic pathogens<br />

Green grassy shoot disease of sugarcane, a major disease in Nghe An Province,<br />

Vietnam<br />

INTRODUCTION<br />

R.C. Magarey{ XE "Magarey, R.C." } A , L.W. Burgess B , P.J. Nielsen C and Ngo Van Tu C<br />

A BSES Limited, PO Box 566, Tully, 4854, Queensland<br />

B Faculty of Agriculture, Food and Natural Resources, University of Sydney, 2006, New South Wales<br />

C NAT&L Sugar Factory, Quy Hop, Nghe An Province, Vietnam<br />

Sugarcane is a major crop in many South East Asian countries<br />

and provides an important cash crop as a rotation in a farming<br />

system often consisting of crops such as rice, corn, peanuts and<br />

watermelon. In contrast to first world countries, cropping areas<br />

in Vietnam are small with up to 24,000 farmers supplying<br />

sugarcane from 1ha plots to the local factory. This compares to<br />

around 250 farmers supplying each Australian sugar factory. In<br />

the mid‐1990s, a new sugarcane disease called green grassy<br />

shoot disease (GGSD) was identified in Thailand. Characterised<br />

by the production of many small grassy tillers, and caused by a<br />

phytoplasma, the disease had severe consequences on crop<br />

yields. Several other diseases in neighbouring countries are also<br />

caused by phytoplasmas; these include white leaf disease (WLD)<br />

and grassy shoot disease (GSD). In 2006, symptoms of GGSD<br />

were identified in the NAT&L factory area, Quy Hop, Nghe An<br />

Province, Vietnam. This paper briefly describes GGSD symptoms<br />

and the current epidemic occurring in Vietnam.<br />

GGSD<br />

Symptoms. The disease is characterised by the production of<br />

many small green grassy tillers. These first appear at the base of<br />

mature sugarcane stools late in the cropping period; in this crop,<br />

yields are not unduly affected. Being a semi‐perennial crop,<br />

second and third annual harvests (first and second ratoon crops)<br />

are made from the same planting. The following ratoon crops<br />

arising from an infested crop suffer very serious yield effects.<br />

Healthy ratoon shoots are replaced by profuse green, grassy<br />

shoots that lead to complete crop failure. Harvest yields often<br />

progress from 80 tonnes biomass per ha in a largely disease‐free<br />

plant crop to 15 tonnes / ha in the first ratoon crop; second<br />

ratoon crops in susceptible cultivars often fail altogether. In<br />

contrast to GSD and WLD, there is no chlorosis in leaves of GGSD<br />

affected sugarcane.<br />

Transmission. As a vegetatively propagated crop, infected<br />

planting material leads to diseased crops; the supply of diseasefree<br />

seed‐cane is essential for limiting disease spread. There are<br />

no recorded vectors for GGSD but circumstantial evidence, such<br />

as speed of spread, suggests a vector is likely to be associated<br />

with disease transmission.<br />

Control. The most important control measures for GGSD are the<br />

termination of heavily diseased crops, the planting of new crops<br />

with disease‐free planting material and the choice of the most<br />

resistant cultivars—though there are few resistant cultivars<br />

currently available in Vietnam. Further importation of<br />

germplasm into Vietnam is needed to select suitably‐resistant<br />

cultivars. Research has shown that immersion of infested<br />

planting material in water maintained at 50C for 3 hours (HWT)<br />

leads to the elimination of the disease in >85% of the axillary<br />

buds. The selection of the cleanest planting material for HWT<br />

provides the best opportunity for producing disease‐free nursery<br />

cane.<br />

NAT&L sugar factory, Nghe An Province. The disease has been<br />

widely detected in the two most widely planted cultivars MY55‐<br />

14 and ROC 10; ROC 10 is more susceptible than MY55‐14. The<br />

disease quickly expanded beyond the initial finding with severe<br />

GGSD observed in >6,000ha of crops in early 2009; lighter<br />

infection has been widely observed across the sugar factory<br />

area. The sugar factory has pro‐actively addressed the problem<br />

with incentives paid to farmers to eliminate badly diseased<br />

crops. Concurrently an intense extension program has been run<br />

by the factory in the local communes; over 175 commune<br />

meetings were staged from January to May 2009. In late Aprilearly<br />

May 2009, there has been an expanded program, with<br />

further funding, focused on the elimination of infested crops in<br />

an attempt to further reduce disease spread.<br />

DISCUSSION<br />

The extent of the disease in the Quy Hop sugar factory area, the<br />

speed of spread and the effect on yield all suggest that GGSD is a<br />

very significant threat to sugarcane crop production in Vietnam.<br />

Not enough is known about the disease, including the nature of<br />

possible vectors, the resistance of cultivars to the disease, and<br />

potential replacement canes, and the distribution of the disease<br />

in Vietnam. There is a suspicion that GGSD also occurs in other<br />

Provinces of Vietnam, but at lower severity levels. Further<br />

research is needed, not only with GGSD but also to develop<br />

reliable diagnostic tools for GGSD, GSD and WLD. Findings of<br />

white leaves associated with diseased cane crops suggest that<br />

GSD and / or WLD may also be present in Vietnam. It is<br />

important that the status of the various pathogens is known to<br />

ensure appropriate control measures are applied.<br />

Figure 1. Symptoms of GGSD in sugarcane crop (cultivar MY55‐14) in<br />

Nghe An Province, Vietnam. Note the small green grassy tillers in the<br />

midst of normal ratoon shoots.<br />

Causal agent. Research undertaken in Thailand suggests that a<br />

phytoplasma is the causal agent of GGSD.<br />

ACKNOWLEDGEMENTS<br />

We acknowledge the assistance provided by NAT&L factory staff<br />

in gathering information on this disease.<br />

98 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Molecular detection of Mycosphaerella fijiensis in the leaf trash of ‘Cavendish’ banana<br />

S.G. Casonato{ XE "Casonato, S.G." } A , J. Henderson B and R.A. Fullerton A<br />

A The New Zealand Institute for <strong>Plant</strong> and Food Research Limited, Private Bag 92169, Auckland Mail Centre 1142, New Zealand<br />

B Tree <strong>Pathology</strong> Centre, 80 Meiers Road, Indooroopilly, Queensland, 4068, Australia<br />

INTRODUCTION<br />

Black Sigatoka (black leaf streak) caused by, Mycosphaerella<br />

fijiensis Morelet (anamorph Paracercospora fijiensis (Morelet)<br />

Deighton), the most destructive foliar pathogen of bananas<br />

globally. The disease is present in commercial plantations in<br />

Africa, Asia and Central and South America, where extensive<br />

fungicide applications are required for its control. The potential<br />

for M. fijiensis to be carried into countries free of the disease in<br />

leaf trash carried in commercial consignments is unknown. This<br />

study was undertaken to determine whether M. fijiensis could<br />

be detected in leaf trash in cartons of bananas imported from<br />

the Philippines to New Zealand.<br />

MATERIALS AND METHODS<br />

Samples of leaf tissue and banana skin were collected from<br />

cartons of a commercial consignment of bananas imported to<br />

New Zealand from the Philippines in December 2005. The<br />

samples were stored at ‐20ºC until assayed in July 2006.<br />

DNA extraction. DNA was extracted from 11 of the samples<br />

supplied (Table 1) using a QIAGEN DNeasy ® <strong>Plant</strong> Mini Kit.<br />

Insufficient sample of S40 (particulate leaf material) was present<br />

for extraction. Samples of M. fijiensis (748 ex banana leaf,<br />

Tongatapu, Tonga) and M. musicola (yellow Sigatoka) (Mf589<br />

[Cultures are held in the culture collection maintained by Dr R.A.<br />

Fullerton at <strong>Plant</strong> and Food Research, Mt Albert Research Centre,<br />

Auckland.] ex banana leaf South Johnston, Queensland) were<br />

used as control samples and had DNA extracted from mycelium<br />

growing on a potato dextrose agar plate. DNA was quantified<br />

using a NanoDrop spectrophotometer. DNA extracts were kept<br />

at ‐20°C.<br />

of known identity (Mf 748). This was shown to be 0.243 fg/µL (1<br />

fg = 10 ‐15 g) of pure M. fijiensis DNA. When assaying the<br />

extractions from leaf trash, a PCR product of approximately 1050<br />

bp indicated a positive amplification of M. fijiensis (Figure 1).<br />

Where a positive result was achieved, the PCR reaction was<br />

repeated a minimum of three times. A control of healthy,<br />

uninfected banana leaf was also included in reactions.<br />

Sequencing. Direct sequencing was carried out to confirm the<br />

identity of the amplified products. PCR products were purified<br />

using a QIAGEN MinElute PCR Purification Kit. The PCR product<br />

was fluorescently labelled using a BigDye ® Terminator v3.1 Cycle<br />

Sequencing Kit (Applied Biosystems, Foster City, USA). Each 10<br />

µL reaction contained approximately 25 ng/µL of PCR product, 2<br />

µM primer (ITS‐1 or ITS‐4), 2 µL terminator‐ready reaction mix<br />

and the volume made up with sterile Milli‐Q water. Sequences<br />

obtained were compared with those in the database GenBank ®<br />

using the Basic Local Alignment Search Tool (BLAST), BLASTN.<br />

RESULTS AND DISCUSSION<br />

Four of the 11 tissue samples consistently yielded a PCR product<br />

using M. fijiensis specific primers MFFor and R635‐mod (Figure<br />

1). They were: S2 (floral), S31 (leaf material), S36 (particulate<br />

trash) and S56 (stem or petiole). Sequenced products were<br />

homologous (at least 99%) with M. fijiensis sequences lodged in<br />

GenBank. This study has shown that M. fijiensis was present in<br />

fragments of leaf trash found in cartons of banana fruit imported<br />

into New Zealand from the Philippines. The viability of the<br />

organism within the sample cannot be ascertained from these<br />

tests nor can the quantity of M. fijiensis be verified using these<br />

techniques.<br />

Session 6B—Quarantine and exotic pathogens<br />

Table 1. List of banana leaf, floral and trash samples from which DNA<br />

was extracted. DNA was also extracted from Mycosphaerella fijiensis and<br />

M. musicola cultures.<br />

Sample<br />

S2<br />

S7<br />

S31<br />

S32<br />

S36<br />

S39<br />

S56<br />

S113<br />

S155<br />

S348<br />

S351<br />

Mf 748<br />

Mf 589<br />

Banana leaf<br />

Type<br />

Floral<br />

Edge of leaf?<br />

Leaf material<br />

Leaf material<br />

Particulate trash?<br />

Leaf material<br />

Stem/petiole on fruit<br />

Fruit spots under trash<br />

Particulate trash<br />

Unknown<br />

Unknown<br />

Mycosphaerella fijiensis culture Fullerton<br />

Mycosphaerella musicola culture Fullerton<br />

Healthy glasshouse grown plant in NZ<br />

PCR protocol. Samples were initially amplified using primers<br />

MF137 and R635 (1). These primers were found to be not<br />

specific for M. fijiensis and alternative primers were sought,<br />

MFFor and R635‐mod (Henderson et al. unpublished). These<br />

primers are designed to amplify part of the internal transcribed<br />

spacer (ITS) regions between the 18S and the 28S rDNA subunits<br />

of M. fijiensis. Prior to amplifying the extracted DNA, the<br />

minimum detection limit of M. fijiensis for the PCR protocol<br />

being used was determined using DNA extracted from a culture<br />

—A—B—C—D—E—F—G—H—I—J—K—L—M—N—O—P—Q<br />

Figure 1. Amplified products of Mycosphaerella fijiensis using primers<br />

MFFor and R635‐mod. Reaction used 5 µL DNA per reaction. Lane A:100<br />

bp ladder; B:Banana leaf; C:S2: D:S7; E:S31; F:S32; G:S39; H:S36; I:S56;<br />

J:S113; K:S155; L:S348; M:S351; N:Mf589 yellow sigatoka; O:Mf748 black<br />

sigatoka; P:blank (negative control); Q:100 bp ladder. Double‐ended<br />

arrow indicates product of approximately 1050 bp. Deteriorating DNA<br />

lessened the band brightness for some sampels. Note: gel has been cut.<br />

REFERENCES<br />

1. Johanson, A. and Jeger, MJ. 1993. Use of PCR for detection of<br />

Mycosphaerella fijiensis and M. musicola, the causal agents of<br />

Sigatoka leaf spots in banana and plantain. Mycological Research<br />

97, 670–674.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 99


Session 6B—Quarantine and exotic pathogens<br />

INTRODUCTION<br />

Optimising responses to incursions of exotic plant pathogens<br />

Biological, spatial and economic data, linked through modelling,<br />

can assist in optimising responses to incursions of exotic plant<br />

pathogens. The approach allows predictions of the behaviour of<br />

linked biological and agronomic systems within defined bounds<br />

despite many uncertainties involved in individual parameters.<br />

Uncertainties are to be expected because each incursion of an<br />

exotic pest into a new environment is a novel situation for which<br />

there may be no precedents. The biological and agronomic<br />

parameters having the greatest impact can be identified, and the<br />

response designed to optimise the benefit:cost ratio.<br />

The value of this approach is shown in examples of two relatively<br />

recent incursions into Australia by exotic pathogenic nematodes;<br />

(a) Bursaphelenchus hunanensis, a relative of the Pine Wilt<br />

Nematode (1), (b) Potato Cyst Nematode (PCN) (2).<br />

MATERIALS AND METHODS<br />

The model initially simulated possible scenarios for the arrival,<br />

establishment, and expansion of the geographic range of a pest<br />

in the absence of biosecurity measures. The effects of various<br />

measures were then added and the results compared with the<br />

first run.<br />

The model was a stochastic simulation model using random<br />

number generators to simulate chance or random events.<br />

Probability distributions were used as parameters within an<br />

abstract model rather than point estimates, and a Monte Carlo<br />

algorithm used to sample from each of these distributions (3).<br />

Many parameters were used to estimate the ecological<br />

processes of establishment, spread, population growth and crop<br />

damage, together with their economic consequences in terms of<br />

crop yields, testing for disease, and control measures. Each<br />

parameter was given one of a number of statistical distributions<br />

with a defined mean or modal value, depending on the<br />

distribution chosen. In each of the 5,000 iterations of the model,<br />

one value was randomly sampled across the range of each<br />

distribution. The model used Markov chains to estimate<br />

transitional probabilities between time periods of 1 year. The<br />

model was run over 20 years and used a standard discount rate<br />

of 8% (a margin of 3% on top of a real risk free rate of 5%).<br />

RESULTS<br />

Impacts of both pests studied were large over the time period<br />

considered. Under most possible scenarios, annual impact rises<br />

steeply initially, followed by slower growth, before eventually<br />

declining (Fig. 1). Raw crop losses in the field were only a small<br />

proportion of the aggregate impact. Parameters had different<br />

effects and time courses on the aggregate impact of the pests.<br />

Potential rate of geographic expansion of the pest was<br />

important, but the cost of testing for the pest during its<br />

expansion was also important. This cost occurs soon after<br />

invasion; it can be largely independent of the actual expansion<br />

rate or range, but is affected by the accuracy and efficiency of<br />

the test. Efficacy of testing affects the impact of a pest on other<br />

crops occurring in the region. Cost of mitigation of the pest may<br />

be large, and the failure rate of control is an important cost.<br />

M. Hodda{ XE "Hodda, M." } and D.C. Cook<br />

CSIRO Entomology, GPO Box 1700, Canberra, 2601, ACT<br />

highly significant. With increasing distances from production to<br />

market, the chances of barriers to trade arising or loss of<br />

markets following arrival of a pest are increased. Disinfestation<br />

and certification costs were substantial in the long term.<br />

DISCUSSION<br />

Rapid initial rise in impact of pests makes early detection and<br />

action desirable, even when there is great uncertainty over the<br />

future behaviour of the pest. The substantial impact beyond lost<br />

crop production means that eradication or other control<br />

measures are often the best option. The problem is that cost of<br />

this strategy precedes any benefits. Benefits of control programs<br />

may be wider than the direct crop losses, so wider contributions<br />

to costs may be justified.<br />

The predicted decline in annual impacts may be largely related<br />

to discount rates and this requires further investigation since the<br />

real costs of many forms of pest control, eg chemicals, are<br />

increasing, along with environmental, social and regulatory<br />

costs.<br />

Expected impact (million $)<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 2 4 6 8 10 12 14 16 18 20<br />

Time (years)<br />

Figure 1. Simulated impact of PCN in Australia.<br />

mean<br />

standard error<br />

95% confidence limit<br />

REFERENCES<br />

1. Hodda M, Smith DI, Smith IW, Nambiar L, Pascoe I (2008) Incursion<br />

management in the face of multiple uncertainties: a case study of<br />

an unidentified nematode associated with dying pines near<br />

Melbourne, Australia. In ‘Pine Wilt Disease—a threat to forest<br />

ecosystems’. (Eds P Viera, Mota M) pp. ()<br />

2. Hodda M, Cook DC (in press) Economic impact from unrestricted<br />

spread of Potato Cyst Nematodes in Australia. Phytopathology 33,<br />

3. Cook DC, Thomas MB, Cunningham SA, Anderson, DL and DeBarro<br />

PJ 2007. Predicting the economic impact of an invasive species on<br />

an ecosystem service. Ecol Appl 17: 1832–1840<br />

Impacts related to trade in the crop, both in terms of quantity<br />

and value are highly uncertain, but under most scenarios are<br />

100 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


The influence of soil biotic factors on the ecology of Trichoderma biological control<br />

agents<br />

INTRODUCTION<br />

A. Stewart{ XE "Stewart, A." }, Y.W. Workneh and K.L. McLean<br />

Bio‐Protection Research Centre, PO Box 84, Lincoln University, Canterbury 7647, NZ<br />

The influence of environmental factors such as pH, moisture,<br />

temperature and other abiotic factors such as fertiliser and<br />

pesticide application on the establishment, proliferation and<br />

persistence of biocontrol agents in the field has been intensively<br />

investigated (1). However, there is little understanding of the<br />

nature of biotic influences which are likely to play an equally<br />

important role in determining the nature of the biocontrol<br />

outcome. This paper reports on a preliminary study that<br />

examines the effect of common soil microbes on two biocontrol<br />

agents, Trichoderma atroviride LU132 active against onion white<br />

rot and Trichoderma hamatum LU593 active against Sclerotinia<br />

lettuce rot (2).<br />

MATERIALS AND METHODS<br />

Soil microbes. Forty‐eight microbes representing 11 fungal<br />

genera (Acremonium, Alternaria, Aspergillus, Beauveria,<br />

Chaetomium, Cladosporium, Fusarium, Metarhizium,<br />

Paecilomyces, Penicilllium, Verticillium), seven bacterial genera<br />

(Agrobacterium, Azotobacter, Bacillus, Burkholderia,<br />

Flavobacterium Paenibacillus) and four actinomycete genera<br />

(Actinomyces, Arthrobacter, Rhodococcus, Streptomyces) were<br />

obtained from NZ culture collections (Lincoln University,<br />

Landcare Research, AgResearch).<br />

Dual culture assays. Test microbes were inoculated 3d prior to<br />

or simultaneously with the Trichoderma on 9 or 15cm diameter<br />

PDA plates. An inoculum plug of the test microbe was placed<br />

3cm apart from the Trichoderma in the centre of the plate.<br />

Colony interactions were monitored every 24h until Trichoderma<br />

colony growth stopped or was constrained by the edge of the<br />

plate. Trichoderma colony area (mm 2 ) was measured and<br />

percentage inhibition compared to the Trichoderma control<br />

calculated.<br />

Soil pot assays. Inocula of six test microbes were produced on<br />

rice grains and incorporated into Templeton silt loam soil in pots<br />

to give 10 6 cfu/g soil. Trichoderma was applied to the soil (10 6<br />

cfu/g soil) as a granular formulation (Agrimm Technologies Ltd).<br />

Pots were incubated at constant temperature and moisture for<br />

30d. At weekly intervals, soil samples were taken from three<br />

random spots in each pot and Trichoderma population counts<br />

(cfu/g soil) determined using soil dilution plating on Trichoderma<br />

selective medium.<br />

Statistical analyses. Data was analysed using one‐way ANOVA<br />

and treatment means compared using Fishers LSD.<br />

RESULTS<br />

Dual culture assays. Co‐culture on PDA revealed six fungi and<br />

one bacterium that significantly inhibited Trichoderma colony<br />

growth (Table 1). Greatest inhibition (>85%) occurred with<br />

Aspergillus niger and Paecilomyces lilacinus for T. atroviride and<br />

T. hamatum, respectively. In general, T. hamatum was less<br />

sensitive than T. atroviride to the test microbes, in particular to<br />

C. globosum.<br />

Soil pot assays. T. atroviride populations were significantly<br />

reduced in soil treated with Alternaria, Aspergillus, Metarhizium,<br />

Paecilomyces and Daldinia (Fig. 1) T. hamatum was less sensitive<br />

to the test microbes but the trend was similar (data not shown).<br />

Table 1. Percentage inhibition of Trichoderma colony growth after 10d<br />

dual culture with test microbes<br />

Test microbes T. atroviride T. hamatum<br />

Alt. alternata 94.7 a* 72.8 c<br />

P. lilacinus 84.6 b 85.1 a<br />

Asp. niger 96.5 a 82.2 ab<br />

C. globosum 87.2 b 44.5 d<br />

M. anisopliae 82.5 b 74.2 bc<br />

D. eschscholzii 56.4 c 51.2 d<br />

Agrobacterium sp 67.1 c 31.3 e<br />

* Values within columns followed by the same letter are not significantly different.<br />

CFU/gram soil<br />

1.40E+05<br />

1.20E+05<br />

1.00E+05<br />

8.00E+04<br />

6.00E+04<br />

4.00E+04<br />

2.00E+04<br />

0.00E+00<br />

A. alternata<br />

A. niger<br />

C. globosum<br />

M. anisopliae<br />

P. lilacinus<br />

D. eschscholzii<br />

Agrobacterium<br />

Control<br />

Figure 1. T. atroviride population (cfu/g soil) after 30d in soil treated with<br />

different test microbes.<br />

DISCUSSION<br />

Six fungi and one bacterium significantly inhibited Trichoderma<br />

colony growth on agar plates with differential sensitivity<br />

observed between the two Trichoderma strains. Preliminary<br />

studies suggest the inhibition is due to the production of<br />

antifungal metabolites by the test microbes. The high inhibition<br />

observed in culture was not reproduced in the soil assay where<br />

five of the seven test microbes reduced Trichoderma populations<br />

but only by ten‐fold. Since test microbe populations in the field<br />

are likely to be lower than those used here, the results likely<br />

overestimate the potential negative impact on Trichoderma<br />

biocontrol agents applied to soil. However, further work<br />

examining the effect of the test microbes in different soil types is<br />

needed since metabolite production by the test microbes may<br />

be influenced by soil abiotic factors.<br />

ACKNOWLEDGEMENTS<br />

Funding for this project was provided by the NZ Tertiary<br />

Education Commission.<br />

REFERENCES<br />

1. Kredics L, Antal Z, Manczinger L, Szekeres A, Kevei F, Nagy E (2003)<br />

Influence of environmental parameters on Trichoderma strains<br />

with biocontrol potential. Food Tech. and Biotechnol. 41, 37–42.<br />

2. Stewart A, McLean K L (2004) Optimising Trichoderma bioinoculants<br />

for integrated control of soilborne disease. Proc 3rd<br />

<strong>Australasian</strong> Soilborne Diseases Symposium 55–56.<br />

Session 6C—Alternatives to chemical control<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 101


Session 6C—Alternatives to chemical control<br />

Understanding Trichoderma bio‐inoculants in the root system of Pinus radiata<br />

P. Hohmann{ XE "Hohmann, P." } A , E.E. Jones B , R. Hill A , A. Stewart A<br />

A Bio‐Protection Research Centre, PO Box 84, Lincoln University, Lincoln 7647, New Zealand<br />

B Department of Ecology, Faculty of Agriculture and Life Sciences, PO Box 84, Lincoln University, Lincoln 7647, New Zealand<br />

INTRODUCTION<br />

The genus Trichoderma are beneficial soil‐borne fungi and a well<br />

known source of biological control agent active against a wide<br />

range of crop diseases, including those of pine trees (1). Several<br />

isolates of Trichoderma have been shown to improve<br />

establishment and reduce pathogen infection of Pinus radiata in<br />

the nursery and in forestry plantations (2). Three isolates of<br />

different Trichoderma species were selected for this study. T.<br />

hamatum (LU592) and T. harzianum (LU686), known to stimulate<br />

growth and improve establishment of P. radiata seedlings, and T.<br />

atroviride (LU132) which had no stimulatory activity. To enable<br />

more predictable and effective use of Trichoderma bioinoculants,<br />

their establishment and population dynamics was<br />

determined. In addition, the effect of each isolate on P. radiata<br />

seedling vitality and growth was assessed.<br />

MATERIALS AND METHODS<br />

Each Trichoderma isolate was applied either as a seed coat<br />

formulation (4 x 10 5 spores/seed; SC) or a spore‐suspension (5 x<br />

10 6 spores/pot; SA) sprayed directly after sowing the P. radiata<br />

seeds. P. radiata seeds were grown in root‐pruning containers<br />

for 7 months under conditions reflecting those used in the<br />

commercial PF Olsen nursery. Health and growth assessments<br />

included mortality rate, shoot height and shoot and root dry<br />

weight measurements. During the 7 month trial period,<br />

Trichoderma populations were enumerated in the bulk potting<br />

mix, rhizosphere, rhizoplane and endorhizosphere subsystems<br />

by dilution plating. At the 20 week assessment, recovered<br />

Trichoderma colonies were identified using morphological and<br />

molecular techniques to differentiate between introduced and<br />

indigenous species. A large‐scale experiment was set up at the<br />

PF Olsen nursery under commercial conditions to verify the<br />

results for LU592.<br />

RESULTS<br />

T. hamatum LU592 performed the best out of the three<br />

introduced isolates. Seedling mortality rate was reduced from<br />

5.2% for the control to 0.2% for LU592 and 0.4% for T.<br />

harzianum LU686 SC. LU592 and LU686 SC also increased shoot<br />

height by 17% and 11%, respectively. Results also indicated that<br />

T. atroviride LU132 increased the root/shoot ratio.<br />

Trichoderma populations of all SA treatments were significantly<br />

higher in the rhizosphere (by 2.1 to 3.3 times) compared with<br />

the control. Applied Trichoderma spp. could be differentiated<br />

from indigenous isolates by colony morphology and confirmed<br />

by molecular sequencing. Introduced Trichoderma isolates could<br />

be detected even though overall Trichoderma populations did<br />

not reveal significant differences to the control. In the<br />

rhizosphere, introduced isolates established with levels of ~20%<br />

for LU132 SA and LU592 SC. T. harzianum LU686 was not<br />

recovered from the rhizosphere after 20 weeks. T. hamatum<br />

LU592, when spray applied, was the only isolate clearly<br />

dominating all four subsystems bulk potting mix, rhizosphere,<br />

rhizoplane and endorhizosphere.<br />

Table 1. Increase (%) in P. radiata seedling growth parameters by T.<br />

hamatum LU592 compared with the control. All values significant at P ><br />

0.05 from the control.<br />

Growth factor Seed coat Spray Suspension<br />

Shoot height 7.4 9.5<br />

Dry weights ‐ roots 17 21<br />

‐ shoots 23 24<br />

Stem diameter 9.0 9.4<br />

number root tips 11 n.s.<br />

n.s. = not significant<br />

DISCUSSION<br />

Both T. hamatum LU592 and T. harzianum LU686 increased the<br />

growth of P. radiata seedlings, with the subsequent large‐scale<br />

experiment confirming the growth promotion effects of LU592.<br />

The spray application performed slightly better than the seed<br />

coat application. This reduction in seedling morality and increase<br />

in seedling growth represents a substantial economic benefit to<br />

the industry.<br />

The spray application method clearly promoted the<br />

establishment of the introduced isolates in the root system of P.<br />

radiata. T. harzianum LU686 was found to be an early<br />

rhizosphere coloniser (declining after 12 weeks). Strong<br />

rhizosphere competence was identified for T. hamatum LU592.<br />

The ability of Trichoderma, LU592 in particular, to establish in<br />

the rhizosphere and penetrate the roots is a crucial indicator of<br />

beneficial activity (1).<br />

LU592, being the most effective isolate at colonising all P.<br />

radiata root subsystems, was selected for more detailed<br />

ecological studies using a genetically marked strain. Future<br />

experiments will focus on the use of a fluorescently marked<br />

isolate of LU592 to verify rhizosphere competence, examine<br />

spatio‐temporal distribution within the rhizosphere and<br />

determine endophytic activity including interactions with<br />

ectomycorrhizae.<br />

ACKNOWLEDGEMENTS<br />

This study is part of the Ecosystem Bioprotection program<br />

LINX0304 funded by the NZ Foundation for Research Science and<br />

Technology (FRST). We would like to thank PF Olsen for nursery<br />

access.<br />

REFERENCES<br />

1. Harman, G.E., Howell, C.R., Viterbo, A., Chet, I. & Lorito, M. (2004).<br />

Trichoderma species—Opportunistic, avirulent plant symbionts.<br />

Nature Reviews Microbiology. 2(1): 43–56.<br />

2. Hood, I.A., Hill, R.A., Horner, I.J. (2006). Armillaria Root Disease in<br />

New Zealand Forests. A Review. Review document written for the<br />

Forest Biosecurity Research Council.<br />

T. hamatum LU592 as a seed coat and spray application<br />

significantly increased shoot height, shoot and root dry weight<br />

and stem diameter compared with the control in the large‐scale<br />

experiment (Table 1).<br />

102 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


A bioassay to screen Trichoderma isolates for their ability to promote root growth in<br />

willow<br />

INTRODUCTION<br />

M. Braithwaite{ XE "Braithwaite, M." }, R. Minchin, R.A. Hill, and A. Stewart<br />

Bio‐Protection Research Centre, PO Box 84, Lincoln University, Lincoln 7647, New Zealand<br />

Trichoderma species have been shown to increase the biomass<br />

of both root and shoots of a range of plants including willow (1).<br />

These positive benefits of biocontrol agents have been<br />

attributed to antibiotic production, parasitism or competition of<br />

pathogenic fungi, activation of the host defence response, and<br />

possibly the direct stimulation of plant growth.<br />

Cuttings of some plant species such as Pinus radiata are slow<br />

and difficult to root resulting in poor plant establishment. The<br />

Bio‐Protection Research Centre culture collection has a large<br />

number of Trichoderma isolates and a bioassay to rapidly screen<br />

a selection of these isolates for their ability to promote root<br />

growth and establishment of cuttings was investigated. Willow<br />

was chosen as a model system because of its ease and speed to<br />

produce roots, enabling rapid screening of potential isolates.<br />

This paper describes the development of this assay.<br />

MATERIALS AND METHODS<br />

Dormant cuttings of Salix x matsudana willow, approximately<br />

300 mm in length, were collected in July and August. Cuttings<br />

were given a two hour soak in tap water prior to storage in<br />

plastic bags at 4°C until required for treatment and planting.<br />

Representative isolates (65 in total) from a range of Trichoderma<br />

species from the Biocontrol Microbial Culture Collection (Bio‐<br />

Protection Research Centre, Lincoln) were grown on PDA (pH<br />

4.0) to produce inoculum (conidia). Conidia were harvested in<br />

reverse osmosis water, filtered through Mira cloth (22–25µm)<br />

and added to 0.5% methyl cellulose to produce an inoculum<br />

concentration of 1 x 10 6 mL ‐1 .<br />

The trial was separated into four experiments set up on different<br />

days. Each experiment included three control/standard<br />

treatments (methyl cellulose (0.5%), fulvic acid (0.3%) and<br />

Thiram (12 g L ‐1 )). Willow cuttings were dipped in each<br />

treatment (Trichoderma spore suspensions, controls) for 10<br />

minutes, except Thiram (4 minutes as per manufacturer’s<br />

recommendation). Treated cuttings were planted into individual<br />

planter bags (special Long PB ¾, 64 x 64 x 300 mm) filled with a<br />

pine bark, sand, pumice mix (50, 30, 20% respectively). The trial<br />

was laid out in a stratified random block design consisting of 4<br />

blocks and 40 replicates per treatment. <strong>Plant</strong>s were assessed for<br />

root development 35 days post planting with the following<br />

measurements recorded; cutting girth, cutting length, number of<br />

root initials and number of shoots. Statistical analysis was by<br />

ANOVA and included a range of parameters including root dry<br />

weight, root dry weight per unit volume cutting and root dry<br />

weight per root initial.<br />

RESULTS/DISCUSSION<br />

This bioassay proved successful for screening a large number of<br />

Trichoderma isolates for their ability to promote root growth on<br />

willow. In general, the majority of isolates had no effect on root<br />

promotion compared to the methyl cellulose control. A small<br />

number of isolates reduced root growth. However, three of the<br />

65 Trichoderma isolates screened significantly promoted root<br />

growth (up to 40% more total root dry weight) compared to the<br />

methyl cellulose control. These isolates performed well when<br />

other parameters were examined with increased root:shoot<br />

ratio and root weight per root initial. Figure 1 presents example<br />

data from experiment one demonstrating the ability of<br />

Trichoderma isolate T6 to increase all measured parameters.<br />

Root dry wt (g)<br />

Root wt / root initial (g)<br />

Root : shoot<br />

0.80<br />

0.60<br />

0.40<br />

0.20<br />

0.00<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

0.00<br />

1.00<br />

0.80<br />

0.60<br />

0.40<br />

0.20<br />

0.00<br />

Total root dry weight<br />

Root weight / root initial<br />

Root : shoot<br />

T 1 T 2 T 3 T 4 T 5 T 6 Methyl Fulvic Thiram<br />

cellulose acid<br />

*<br />

* *<br />

Figure 1. The effects of six Trichoderma isolates (T1‐T6) on root<br />

weight/root initial, root/shoot ratio, and total root dry weight in willow<br />

compared to the control treatments methyl cellulose, fulvic acid, and<br />

Thiram for experiment 1. 5% LSD bars are shown (* indicates treatments<br />

significantly better than the methyl cellulose control).<br />

These three isolates plus others which showed potential for<br />

promoting root growth will be further evaluated in additional<br />

experiments on other plant host cutting/seedling systems.<br />

ACKNOWLEDGEMENTS<br />

We wish to thank Candice Barclay, Bronwyn Braithwaite,<br />

Prashant Kumar Chohan, Emily Duerr, Stuart Larsen, Kirstin<br />

McLean, Mana Ohkura, for their technical assistance.<br />

REFERENCES<br />

1. Adams P, De‐Leij AM, Lynch JM (2007) Trichoderma harzianum Rifai<br />

1295–22 mediates growth promotion of crack willow (Salix fragisis)<br />

saplings in both clean and metal‐contaminated soil. Microbial<br />

Ecology 54, 306–313.<br />

*<br />

*<br />

Session 6C—Alternatives to chemical control<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 103


Session 6C—Alternatives to chemical control<br />

Biofumigation for reducing Cylindrocarpon spp. in New Zealand vineyard and nursery<br />

soils<br />

C.M. Bleach{ XE "Bleach, C.M." }, E.E. Jones and M.V. Jaspers<br />

A<br />

Ecology Department, Faculty of Agriculture and Life Sciences, Lincoln University, PO Box 84, Lincoln 7647, New Zealand<br />

INTRODUCTION<br />

The soil‐borne disease, Cylindrocarpon black foot disease (BF), is<br />

reported as the cause of decline of young grapevines in new and<br />

replanted vineyards world‐wide. Cylindrocarpon destructans,<br />

C. liriodendri and C. macrodidymum have been found to be<br />

equally associated with this disease in New Zealand vineyards<br />

and nurseries.<br />

Brassica species contain significant quantities of the<br />

thioglucoside compounds known as glucosinolates (GSLs). When<br />

GSLs are hydrolysed by the myrosinase enzyme present in<br />

Brassica tissue, volatile isothiocyanates (ITCs) are produced [1].<br />

ITCs are known to have broad biocidal activity in the suppression<br />

of pathogenic fungal species, nematodes, weeds, and certain<br />

insect species [2]. The principal suppressive effect of brassica<br />

amendments on soil‐borne diseases occurs at flowering,<br />

immediately after their maceration and incorporation into soil<br />

[3]. ITCs are also exuded from the roots of brassicas throughout<br />

their growth. The efficacy of Brassica spp. for control of<br />

Cylindrocarpon spp. in grapevines was tested in field<br />

experiments.<br />

MATERIALS AND METHODS<br />

A preliminary experiment used crops of mustard (Brassica<br />

juncea), rape (B. napus) and oats (Avena sativa). They were<br />

grown for 5 weeks in an infested site, where young vines<br />

previously grown had been infected with the above<br />

Cylindrocarpon spp. The brassica crops were cultivated into the<br />

soil and the area covered with polythene. After 2 weeks, callused<br />

rootstock cuttings (varieties 101–14 and 5C) were planted in the<br />

treated soil in a randomised split plot design (5 plots per<br />

treatment) according to standard nursery practices. The plants<br />

were grown for 9 months, harvested and infection assessed by<br />

isolation onto potato dextrose agar. Cylindrocarpon spp. were<br />

identified morphologically after 7–10 days incubation at 20ºC.<br />

The second experiment used 3 mustard treatments (Trt 1–3)<br />

• Trt 1. Mustard meal cultivated into the soil.<br />

• Trt 2. Grown once to flowering with cultivation.<br />

• Trt 3. Grown twice to flowering with cultivation each time.<br />

When Trt 3 was flowering (4 weeks), the entire field was<br />

inoculated with Cylindrocarpon spp. grown on wheat grains (3<br />

isolates of each species) and 2 days later the mustard plants<br />

were cultivated into the soil and the area covered with<br />

polythene. After 3 days the polythene was removed and mustard<br />

seed was sown for Trt 2 and Trt 3 and grown to flowering. At this<br />

time, the mustard meal (Trt 1) was broadcast and all 3<br />

treatments were incorporated into the soil. The area was<br />

covered with polythene for 2 days then callused cuttings of<br />

rootstocks 101–14 and 5C (20 per plot) were planted in a<br />

randomised split plot design (5 plots per treatment). The plants<br />

were grown for 10 months then infection assessed as above.<br />

Analysis of data was with a general linear model appropriate to<br />

the split‐plot randomised block design.<br />

RESULTS<br />

In the preliminary experiment (Fig 1a), although not significant<br />

(P=0.137), disease incidence (DI) was reduced by the mustard<br />

treatment in rootstocks 101–14 and 5C (11 and 43%,<br />

respectively). DI was decreased in rootstock 5C by the oats<br />

treatment but increased in 101–14 and increased in both<br />

rootstocks by the rape treatment. The mustard treatment was<br />

further tested in a second experiment (Fig 1b). Again, treatment<br />

effects were not significant for DI (P=0.359), which was reduced<br />

by Trt 1 and 3 in both rootstocks and in 5C by Trt 2. Overall, DI<br />

was reduced by the mustard treatments in rootstock 5C by more<br />

than 41% and in experiment two DI was reduced by Trt 1 and Trt<br />

3 in rootstock 101–14 by 30 and 18%, respectively.<br />

Mean disease incidence<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

Figure 1. Mean Cylindrocarpon incidence in grapevine varieties 101–14<br />

and 5C after cultivation of (a) brassica and oat crops and (b) mustard as<br />

crops and mustard meal.<br />

DISCUSSION<br />

The results indicate that biofumigation with mustard was<br />

effective for reducing infection by Cylindrocarpon spp.<br />

particularly in rootstock 5C, but less so in 101–14, which is more<br />

susceptible to BF disease. Overall, treatments 1 and 3 were able<br />

to reduce disease incidence by 36 and 27%, respectively. Despite<br />

the lack of statistical significance (which may be due to high<br />

variability), these findings suggest that mustard treatments may<br />

be a highly effective method for the control of BF disease.<br />

Biofumigation may well be a natural, effective, and economical<br />

way to eliminate pests and diseases in vineyard and nursery<br />

soils, to improve soil structure and soil organic matter, and to<br />

increase soil microbial activity. A key to improving the efficacy of<br />

biofumigation in the field lies in the development of application<br />

technologies that macerate and incorporate the biofumigant<br />

evenly in soils, in addition to incorporating it under optimal<br />

edaphic conditions for release of ITCs [3].<br />

ACKNOWLEDGEMENTS<br />

New Zealand Winegrowers, Corbans Viticulture and Technology<br />

NZ (TIF) for assistance and financial support.<br />

REFERENCES<br />

5C<br />

101-14<br />

Control<br />

Mustard<br />

Oats<br />

(a)<br />

Rape<br />

Treatment<br />

Control<br />

Mustard meal<br />

Mustard flower 1<br />

Mustard flower 2<br />

1. Kirkegaard, J.A. and M. Sarwar, Biofumigation potential of brassicas<br />

<strong>Plant</strong> and Soil. 1998. 201(Number 1): p. 71–89.<br />

2. Brown, P.D. and M.J. Morra, Control of soil‐borne plant pests using<br />

glucosinolate‐containing plants. Advances in Agronomy. 1997.<br />

61(167–231).<br />

3. Mattner, S.W., et al., Factors that impact on the ability of<br />

biofumigants to suppress fungal pathogens and weeds of<br />

strawberry. Crop Protection, 2008. 27(8): p. 1165–1173.<br />

<br />

(b)<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

104 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Crown rot of winter cereals: integrating molecular studies and germplasm<br />

improvement<br />

Mark W. Sutherland{ XE "Sutherland, M.W." } A , W.D. Bovill A,D , F.S. Eberhard A , A. Lehmensiek A , D. Herde B and S. Simpfendorfer C<br />

A Centre for Systems Biology, Faculty of Sciences, University of Southern Queensland, Toowoomba, QLD 4350<br />

B DEEDI, Primary Industries and Fisheries, Leslie Research Centre, Toowoomba, QLD 4350<br />

C NSW Department of Primary Industries, Marsden Park Road, Tamworth, NSW 2340<br />

D Current address: School of Food, Agriculture and Wine, University of Adelaide, Waite Campus, PMB1, Glen Osmond, SA 5064<br />

INTRODUCTION<br />

Crown rot of winter cereals is a major constraint on grain<br />

production across most growing regions in Australia, particularly<br />

where stubble retention is practiced to maintain soil structure<br />

and retain soil water. The predominant cause of this disease is<br />

infection with Fusarium pseudograminearum (Fpg), although in<br />

some southern areas Fusarium culmorum infections are also<br />

significant. These Fusarium species are able to grow<br />

saprophytically on stubble remnants over the summer and<br />

provide inoculum for crop infection in the following season.<br />

Losses due to crown rot are highest in seasons featuring a dry<br />

finish in which maturing plants experience water stress, with<br />

symptoms including basal stem browning and white heads<br />

bearing no grain.<br />

Control of this disease is challenging and is currently based on<br />

management practices centred on crop rotation strategies. At<br />

present, there are no resistant commercial varieties of bread<br />

wheat, durum or barley available for deployment. Durum wheats<br />

are particularly susceptible.<br />

We are currently undertaking a long‐term collaborative research<br />

program which aims to:<br />

• characterise known resistance sources<br />

• develop molecular markers for quantitative trait loci (QTL)<br />

to assist selection in breeding programs<br />

• transfer QTL for resistance from hexaploid (bread) to<br />

tetraploid (durum) wheats<br />

• pyramid resistance in bread wheats<br />

• understand the fundamental biology of this host/pathogen<br />

interaction.<br />

Central to the task is an integration of laboratory and field‐based<br />

investigations to ensure outcomes that not only advance our<br />

knowledge but also reduce yield losses and increase<br />

management options for primary producers.<br />

Here we report on recent successes in pyramiding sources of<br />

partial resistance and discuss progress in transferring resistance<br />

from hexaploid sources into a durum background.<br />

METHODOLOGY<br />

Two doubled haploid wheat populations produced from crosses<br />

of partially resistant parents, Sunco/2‐49 and 2‐49/W21MMT70<br />

were evaluated for resistance to crown rot using a standard<br />

seedling pot test inoculated with a mixture of aggressive Fpg<br />

isolates(1). Based on genetic maps constructed from SSR and<br />

DArT markers, QTL for resistance were then identified.<br />

RESULTS AND DISCUSSION<br />

To date, a wide selection of resistance sources have been<br />

partially characterised and quantitative trait loci (QTL) identified<br />

(2, 3, 4). Inoculated seedling and field trials indicate overlapping<br />

sets of loci that contribute at these different stages of<br />

development. These partial sources of resistance contain largely<br />

different sets of QTL which suggest that improved resistance<br />

may be obtained by gene pyramiding. Results from QTL analysis<br />

of the Sunco/2‐49 and 2‐49/W21MMT70 populations following<br />

seedling trials indicate that the more resistant lines inherited the<br />

major QTL from each parent and that a number of lines in the 2‐<br />

49/W21MMT70 population expressed significantly higher<br />

resistance than either of the parents. Hence pyramiding of<br />

independent resistance sources can produce significant<br />

improvements in the resistance of derived progeny towards<br />

crown rot.<br />

Marker analysis of hexaploid x tetraploid crosses shows that the<br />

bread wheat markers were readily transferred to the progeny.<br />

Furthermore even after several generations these markers<br />

remained linked to the resistance character and were<br />

independent of remnant D genome material. Crosses of durum<br />

wheats with the hexaploid resistance sources significantly<br />

reduced the disease severity in derived materials, demonstrating<br />

the potential of this approach for improving the resistance of<br />

durums to crown rot.<br />

ACKNOWLEDGEMENTS<br />

We acknowledge the contributions of Dr Graham Wildermuth<br />

and Dr Ray Hare to this work. This work was funded by the<br />

Grains Research and Development Corporation (GRDC).<br />

REFERENCES<br />

1 Wildermuth GB, McNamara RB (1994) Testing wheat seedlings for<br />

resistance to crown rot caused by Fusarium graminearum Group 1.<br />

<strong>Plant</strong> Disease 78: 949–953.<br />

2 Collard BCY, Grams RA, Bovill WD, Percy CD, Jolley R, Lehmensiek A,<br />

Wildermuth GB, Sutherland MW (2005) Development of molecular<br />

markers for crown rot resistance in wheat: mapping of QTLs for<br />

seedling resistance in a 2‐49 x Janz population. <strong>Plant</strong> Breeding 124:<br />

1–6.<br />

3 Collard BCY, Jolley R, Bovill WD, Grams RA, Wildermuth GB,<br />

Sutherland MW (2006) Confirmation of QTL mapping and marker<br />

validation for partial seedling resistance to crown rot in wheat line<br />

‘2‐49’. Aust J Agric Res 57: 967–973.<br />

4 Bovill WD, Ma W, Ritter K, Collard BCY, Davis M, Wildermuth GB,<br />

Sutherland MW (2006) Identification of novel QTL for resistance to<br />

crown rot in the doubled haploid wheat population ‘W21MMT70’ x<br />

‘Mendos’. <strong>Plant</strong> Breeding 125: 538–543.<br />

Session 7A—Cereal pathology 2<br />

Crosses between hexaploid wheat lines with partial resistance<br />

and a range of durum lines were obtained from Dr Ray Hare,<br />

NSW DPI. These materials were field grown near Tamworth NSW<br />

in Fpg infected plots through to the F7 generation and assessed<br />

for crown rot susceptibility each season.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 105


Session 7A—Cereal pathology 2<br />

INTRODUCTION<br />

Infection of wheat tissues by Fusarium pseudograminearum<br />

N.L. Knight{ XE "Knight, N.L." } A , A. Lehmensiek A , D.J Herde B , M.W. Sutherland A<br />

A Centre for Systems Biology, Faculty of Sciences, University of Southern Queensland, Toowoomba, 4350, QLD<br />

B DEEDI, Primary Industries and Fisheries, Leslie Research Centre, Toowoomba, 4350, QLD<br />

Crown rot of wheat, caused by Fusarium pseudograminearum<br />

(Fp), is a serious disease threat across the Australian wheat belt.<br />

Currently control of this disease relies on farming practices (e.g.<br />

crop rotation) and planting of less susceptible cultivars. Partial<br />

resistance has been identified in a small number of wheat lines,<br />

such as 2–49 and Sunco, but the mechanisms of resistance<br />

shown by these lines have not been identified.<br />

RESULTS AND DISCUSSION<br />

A strong relationship was observed between visual rating scores<br />

of wheat leaf sheaths and the normalised Fp DNA (Fig. 1). This<br />

demonstrates that the degree of visual discolouration of wheat<br />

leaf sheaths correlates with the quantity of Fp mycelium present<br />

in the tissue, validating visual rating systems of basal<br />

discolouration (2,3) as a relative estimation of tissue infection<br />

levels across a seedling trial.<br />

Partial resistance can be expressed in either the seedling or adult<br />

stage, depending on the genotype, with the majority of current<br />

screening methods being based on seedling scoring. Extensive<br />

seedling trial comparisons between susceptible and partially<br />

resistant host genotypes suggest a significantly slower spread of<br />

the fungus in the younger tissues of resistant individuals (1).<br />

The current project aims to assess growth of Fp during crown rot<br />

development across partially resistant and susceptible wheat<br />

lines in order to determine key elements in the progress of<br />

disease and when resistance mechanisms are induced. Current<br />

disease rating systems for seedlings rely heavily on browning of<br />

leaf sheaths and tiller bases (2). Our current investigations are<br />

centred on the relationship between the expression of these<br />

visible disease symptoms and the extent of fungal infection.<br />

These studies are also comparing the progress of fungal spread<br />

in both susceptible and partially resistant wheats. The increase<br />

in fungal load in each inoculated host genotype has been<br />

measured using a quantitative real time multiplex polymerase<br />

chain reaction (PCR) assay, allowing simultaneous detection of<br />

both pathogen and host DNA. Microscopy of infected wheat<br />

tissues is also in progress.<br />

MATERIALS AND METHODS<br />

Inoculation. Two week old seedlings were inoculated using a 10 6<br />

conidia per mL suspension (3). Seedling tissues (particularly leaf<br />

sheaths) were harvested at different time points after infection,<br />

with four host genotypes being compared (Table 1).<br />

Table 1. Genotypes tested and their resistance rating from field trials.<br />

DNA Extraction and Multiplex Quantitative PCR. DNA was<br />

extracted using the DNeasy Minikit (Qiagen). Primers and probes<br />

were designed from Genbank sequences with Primer3 software<br />

using the translation elongation factor (TEF)‐α sequence of Fp<br />

and the TEF‐G sequence of wheat. PCR results were normalised<br />

by expressing Fp DNA content relative to the host DNA.<br />

Normalised<br />

(Pathogen/Host DNA)<br />

0.05<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

0<br />

LS1<br />

LS2<br />

LS3<br />

R 2 = 0.78<br />

0 1 2 3 4<br />

Visual Rating (1-4)<br />

Figure 1. Comparison of levels of normalised Fp DNA and visual rating<br />

scores of leaf sheaths (LS) 1, 2 and 3 of the four host genotypes at 7 days<br />

after inoculation.<br />

Observation of Fp growth at increasing periods after inoculation<br />

has revealed significant differences in growth of Fp in seedlings<br />

of the four standard genotypes.<br />

Microscopic assessment of Fp growth has observed intra‐ and<br />

inter‐cellular growth associated with the leaf sheath epidermis,<br />

including trichomes and stomata. Current investigations are<br />

examining growth of mycelium in vascular tissues of expanded<br />

tillers.<br />

REFERENCES<br />

1. Percy C (2009) Crown rot (Fusarium pseudograminearum) symptom<br />

development and pathogen spread in wheat genotypes with<br />

varying disease resistance. PhD Thesis, University of Southern<br />

Queensland.<br />

2. Wildermuth GB, McNamara RB (1994) Testing wheat seedlings for<br />

resistance to crown rot caused by Fusarium graminearum Group 1.<br />

<strong>Plant</strong> Disease, 78, 949–953.<br />

3. Mitter V, Zhang MC, Liu CJ, Ghosh R, Ghosh M, Chakraborty S<br />

(2006) A high‐throughput glasshouse bioassay to detect crown rot<br />

resistance in wheat germplasm. <strong>Plant</strong> <strong>Pathology</strong>, 55, 433–441.<br />

4. Hückelhoven R, Kogel KH (1998) Tissue‐specific superoxide<br />

generation at interaction sites in resistant and susceptible nearisogenic<br />

barley lines attacked by the powdery mildew fungus<br />

(Erysiphe graminis f. sp. hordei). Molecular <strong>Plant</strong> Microbe<br />

Interactions, 11, 292–300.<br />

Microscopy. Fixation and clearing of tissues was performed as<br />

described in (4). Differential staining used safranin and<br />

solophenyl flavine dyes. <strong>View</strong>ing of tissue was performed using a<br />

fluorescence microscope (Nikon Eclipse) under the UV‐2A filter.<br />

106 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Monitoring sensitivity to strobilurin fungicides in Blumeria graminis on wheat and<br />

barley in Canterbury, New Zealand<br />

INTRODUCTION<br />

S.L.H. Viljanen‐Rollinson{ XE "Viljanen‐Rollinson, S.L.H." } A , M.V. Marroni A , R.C. Butler A and G. Stammler B<br />

A New Zealand Institute for <strong>Plant</strong> and Food Research Limited, Private Bag 4704, Christchurch, New Zealand<br />

B BASF Aktiengesellschaft, APR/FB‐LI470, D‐67117, Limburgerhof, Germany<br />

Strobilurin fungicides (Quinone ‘outside’ Inhibitors (QoI)) act at<br />

the Qo binding site of the cytochrome bc1 complex in target<br />

fungi (1). Because QoIs have a specific, single‐site mode of<br />

action, there is a greater risk of developing resistance to these<br />

types of fungicides than to multi‐site inhibitor fungicides. QoIs<br />

were first commercialised in 1996 and have since been widely<br />

used overseas and in New Zealand to control various plant<br />

diseases. QoI‐resistance occurred in Europe in 1998. QoI<br />

resistance has been encountered in 29 different pathogens to<br />

date (2). In most cases, QoI resistance is conferred by a single<br />

point mutation in the cytochrome b gene at the amino acid<br />

codon 143, causing glycine to be replaced with alanine (G143A).<br />

Sensitivity to QoIs in the pathogens causing powdery mildew of<br />

wheat (Blumeria graminis f. sp. tritici) and barley (B. graminis f.<br />

sp. hordei) was investigated in Canterbury, New Zealand cereal<br />

crops during two growing seasons (2004–05 and 2005–06).<br />

MATERIALS AND METHODS<br />

Three different surveys (spring 2004, autumn 2005 and spring<br />

2005) were carried out using two methods of spore collection. In<br />

the first method, isolates of powdery mildew were collected<br />

from infected plants, bulked up in a glasshouse and tested on<br />

detached leaves of barley and wheat either treated with<br />

kresoxim‐methyl or left untreated. Numbers of colonies formed<br />

on the leaves were counted. In the second method young trap<br />

seedlings of wheat (cv. Kotare) or barley (cv. Cask), either<br />

treated with kresoxim‐methyl or left untreated, were exposed to<br />

the air spora at a specific location for 5–7 days, then placed in a<br />

glasshouse until lesions appeared. At each site, 15 untreated and<br />

15 treated pots, with 10 plants per pot, were used. Lesions on<br />

untreated and treated leaves were counted, and colonies that<br />

grew on treated leaves were kept in isolation chambers for<br />

further analysis on detached leaves. There were a total of 33<br />

wheat and 28 barley sites over the two seasons. Data (only<br />

spring 2005 results shown) were analysed using a log‐linear<br />

model for count data with GenStat.<br />

Fifteen isolates of B. graminis f.sp. tritici collected from wheat<br />

plants at six sites and 12 isolates of B. graminis f.sp. hordei<br />

collected from barley plants at six sites in the three surveys were<br />

tested for presence of mutation G143A. These isolates had<br />

shown increased resistance to QoIs in the lab tests. DNA was<br />

extracted from the lesions and the cytochrome b ‐gene was<br />

amplified with PCR. For qualitative analysis the product was<br />

digested with restriction enzyme Sat1 to detect the presence of<br />

the gene G143A for resistance and for quantitative analysis the<br />

PCR‐product was analysed using pyrosequencing.<br />

RESULTS<br />

Spring 2004. Trap plant studies indicated that in nine out of ten<br />

sampled locations, wheat powdery mildew colonies were able to<br />

grow on fungicide‐treated plants. The results for barley powdery<br />

mildew showed that in seven out of nine locations, barley<br />

powdery mildew colonies were able to grow on fungicidetreated<br />

plants. No resistance was detected in any of the isolates<br />

collected from infected plants in the field.<br />

Autumn 2005. Survey results indicated no resistance to wheat<br />

powdery mildew, but resistant barley powdery mildew isolates<br />

were found at all five sites where trap plants were placed, and<br />

four out of six sites where isolates were collected from infected<br />

plants in the field.<br />

Spring 2005. Trap plant results indicated that resistant wheat<br />

powdery mildew isolates were found in seven out of eight trap<br />

plant sites although in one of these six sites, only one colony was<br />

found on all of the 150 plants. One isolate from these sites was<br />

found to contain the gene G143A. Barley trap plant results<br />

indicated that resistant powdery mildew colonies were found in<br />

all seven sites. Two isolates from these sites were found to<br />

contain the gene G143A.<br />

The mean number of colonies per leaf was greater (P < 0.05) on<br />

unsprayed plants than on sprayed plants at all sites except at<br />

one wheat site. At this site, the mean number of colonies was<br />

significantly greater (P = 0.007) on the fungicide‐treated leaves<br />

than on the untreated leaves. For wheat, mean numbers of<br />

colonies per leaf on unsprayed plants varied from below 1 to<br />

27.5. For barley, mean numbers of colonies per leaf varied from<br />

below 1 to more than ten per leaf.<br />

Four of the 15 B. graminis f. sp. tritici isolates and all 12 B.<br />

graminis f. sp. hordei isolates contained the gene G143A as<br />

indicated by PCR.<br />

DISCUSSION<br />

These results confirm that populations of B. graminis resistant to<br />

QoI fungicides are affecting wheat and barley crops in<br />

Canterbury. Isolates carrying the gene G143A express high<br />

(complete) resistance. As a consequence, continued applications<br />

of solo QoIs over time are very likely to result in severe loss in<br />

disease control. Further surveys of resistance to these fungicides<br />

have not been carried out since spring 2005. These should be<br />

carried out as soon as possible to monitor the situation so that<br />

growers can adopt appropriate fungicide resistance<br />

management strategies.<br />

ACKNOWLEDGEMENTS<br />

We thank Foundation for Arable Research for funding, and<br />

participating growers and Mr Grant Hagerty for assistance in<br />

these studies.<br />

REFERENCES<br />

1. Holomon D (2007) Are some diseases unlikely to develop QoI<br />

resistance? Pest Management Science 63, 217–218.<br />

2. Fungicide Resistance Action Committee (FRAC), Pathogens with<br />

field resistance towards QoI fungicides (Status Dec 2008),<br />

http://www.frac.info/<br />

Session 7A—Cereal pathology 2<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 107


Session 7A—Cereal pathology 2<br />

INTRODUCTION<br />

Cross inoculation of crown rot and Fusarium head blight isolates of wheat<br />

P.A.B. Davies{ XE "Davies, P.A.B." } A , L.W. Burgess B , R. Trethowan A , R. Tokachichu A , D. Guest B<br />

A<br />

<strong>Plant</strong> Breeding Institute, University of Sydney, PMB 11, Camden, NSW, 2750<br />

B<br />

Faculty of Agriculture, Food and Natural Resources, University of Sydney, NSW, 2006<br />

Fusarium head blight (FHB) and crown rot (CR) of wheat are two<br />

cereal diseases caused by Fusarium pathogens in Australia. Both<br />

Fusarium graminearum and F. pseudograminearum have been<br />

responsible for FHB epidemics (1, 2), while F.<br />

pseudograminearum is most commonly associated with CR.<br />

Recently, reports of F. graminearum causing CR have emerged<br />

from glasshouse seedling tests (3).<br />

Crown rot and FHB are linked through aetiology, pathogen<br />

biology and epidemiology. The presence of perithecia on stubble<br />

could potentially allow the infection of wheat seedlings following<br />

ascospore release. Similarly, lodging of crops during heavy rain<br />

would allow wheat heads to come in close proximity of wheat<br />

crowns or residue with sporodochia of F. pseudograminearum,<br />

allowing FHB infection through splashed macroconidia (1). This<br />

does suggest however that there is no pressure for pathogenic<br />

specialisation as an individual isolate must retain pathogenicity<br />

for both crown and head infections.<br />

RESULTS AND DISCUSSION<br />

Results of the CR assay display a significant affect of genotype<br />

for both F. graminearum and F. pseudograminearum (P


Twenty years of quarantine plant disease surveillance on the island of New Guinea:<br />

key discoveries for Australia and PNG<br />

R.I. Davis{ XE "Davis, R.I." } A , P. Kokoa B<br />

A Northern Australia Quarantine Strategy (NAQS), Australian Quarantine and Inspection Service (AQIS), PO Box 96, Cairns International<br />

Airport, Cairns 4870, Queensland<br />

B National Agricultural Quarantine and Inspection Authority (NAQIA), PO Box 741, Port Moresby, NCD, Papua New Guinea<br />

INTRODUCTION<br />

The island of New Guinea lies just 5 km from Australia’s northern<br />

border and has a history of plant pest incursions, presumably<br />

from the west. Since 1989, quarantine plant pathologists from<br />

Papua New Guinea (PNG) and the Australian Quarantine and<br />

Inspection Service (AQIS) have conducted regular joint surveys in<br />

PNG’s border regions. Since the mid 1990s, this search for exotic<br />

pathogens became annual. Less frequently over this period,<br />

Australian teams have also worked with Indonesian counterparts<br />

in the Indonesian province of Papua, which occupies the eastern<br />

half of the New Guinea land mass.<br />

MATERIALS AND METHODS<br />

Herbarium specimens and other kinds of plant tissue samples<br />

are collected, rendered quarantine secure, and returned under<br />

permit to various laboratories in Australia, PNG, and occasionally<br />

elsewhere, for diagnostic testing.<br />

RESULTS<br />

Major detections of quarantine significance are listed in Table 1.<br />

These are backed up by a list of finds of other important crop<br />

pathogens published in the last decade including first records on<br />

the island of New Guinea of Banana streak GF virus, Banana<br />

streak Mys virus, and Banana streak OL virus; first records in<br />

PNG of Citrus tristeza virus, Papaya ringspot virus –type W,<br />

Watermelon mosaic virus, and Zucchini yellow mosaic virus; first<br />

valid laboratory confirmation of Fiji disease virus, in sugarcane;<br />

and unpublished first records of Cucumber mosaic virus and teak<br />

rust caused by Olivea tectonae (P. Kokoa, NAQIA unpublished<br />

data, 2008).<br />

DISCUSSION<br />

These collaborative surveys have provided a rich source of<br />

quarantine incursion information for both Australia and PNG.<br />

Critical initial detections of major pathogen threats to<br />

horticultural, field and ornamental commodities of importance<br />

to both countries have been made (Table 1). Moreover, the<br />

ongoing and regular nature of this work in PNG facilitated close<br />

monitoring of spread, following incursion, of fusarium wilt of<br />

banana and huanglongbing (HLB) of citrus. Whilst later finds of<br />

Foc VCG0126 were quite distant from the original one, HLB has<br />

been effectively contained since its discovery in 2002.<br />

These surveys have also yielded critical information on the<br />

distribution of certain diseases endemic in PNG, but still of<br />

extreme quarantine concern to Australia. These include citrus<br />

canker, caused by Xanthomonas citri subsp. citri, and black<br />

Sigatoka disease of banana caused by Mycosphaerella fijiensis.<br />

Although citrus canker is widespread in the border regions, it is<br />

patchy in distribution: detected only in well separated larger<br />

communities. Black Sigatoka, in contrast, is ubiquitous, and over<br />

100 samples have been collected.<br />

In addition, certain key pathogens of regional quarantine<br />

concern, because of their recent history of spread in south east<br />

Asia, have not been detected on any AQIS surveys on the island<br />

of New Guinea. These include Banana bunchy top virus, cause of<br />

bunchy top disease of banana, and Cavendish competent strains<br />

of Guignardia musae, cause of freckle disease of banana. No<br />

convincing symptoms of bunchy top disease have ever been<br />

seen and all samples so far tested by enzyme linked<br />

immunosorbent assay were negative. Over 50 G. musae or<br />

Phyllosticta musarum (anamorph) herbarium specimens have<br />

been collected. However, all were from cooking bananas (ABB<br />

genome) or plantains (AAB genome). In the case of many of<br />

these collections, uninfected Cavendish sub group bananas were<br />

growing adjacent or nearby.<br />

The results summarised here underline the critical importance,<br />

to Australia and PNG, of effective quarantine vigilance. This can<br />

only be achieved with an ongoing program of regular<br />

surveillance.<br />

Session 7B—Quarantine and exotic pathogens<br />

Table 1. Highlights of general disease surveys on the island of New Guinea, 1989–2009<br />

Year Location A Host Pathogen Key discovery Published B<br />

1993 Papua,<br />

Indonesia<br />

Musa sp.<br />

Fusarium oxysporum f.sp.<br />

cubense (Foc) ‐VCG0126<br />

First record of fusarium wilt of banana on the<br />

island of New Guinea<br />

APP, 25<br />

1996 SP, PNG Musa sp. Foc –VCG 0126 First record of fusarium wilt of banana in PNG APP, 25<br />

1997 Papua,<br />

Indonesia<br />

1999 Papua,<br />

Indonesia<br />

1999 Papua,<br />

Indonesia<br />

1999 Papua,<br />

Indonesia<br />

1999 Papua,<br />

Indonesia<br />

Musa spp. Foc –VCG 01213/16 First record of ‘tropical race 4’ of Foc on the<br />

island of New Guinea<br />

Musa sp. Blood disease bacterium First record of blood disease of banana on the<br />

island of New Guinea<br />

Citrus spp.<br />

‘Candidatus Liberibacter<br />

asiaticus’<br />

First record of huanglongbing (greening disease)<br />

on the island of New Guinea<br />

Oryza sativa Rice tungro bacilliform virus First record of rice tungro disease on the island<br />

of New Guinea<br />

Arachis hypogea<br />

Bean common mosaic virus,<br />

peanut stripe strain<br />

Confirmation of widespread occurrence of<br />

peanut stripe disease in Papua<br />

APP, 29<br />

APP, 29<br />

APP, 29<br />

APP, 29<br />

APP, 31<br />

2002 SP, PNG Citrus spp. ‘Ca. L. asiaticus’ First record of HLB in PNG APP, 33<br />

2006 SP, PNG Heliconia sp. Puccinia heliconiae First record in PNG APDN, 3<br />

A SP: MP: Morobe Province, Sandaun Province, WP: Western Province<br />

B APP: <strong>Australasian</strong> <strong>Plant</strong> pathology, APDN: <strong>Australasian</strong> <strong>Plant</strong> Disease Notes, volume number is listed.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 109


Session 7B—Quarantine and exotic pathogens<br />

The importance of reporting suspect exotic or emergency plant pests to your State<br />

Department of Primary Industry<br />

INTRODUCTION<br />

S.A. Peterson{ XE "Peterson, S.A." } A , and F.J. Macbeth B<br />

A<br />

<strong>Plant</strong> Health Australia, 5/4 Phipps Close, Deakin, 2600, ACT<br />

B<br />

Australian Government Department of Agriculture, Fisheries and Forestry, GPO Box 858, Canberra, 2601, ACT<br />

The Emergency <strong>Plant</strong> Pest Response Deed (EPPRD) (1) is a formal<br />

legally binding agreement between <strong>Plant</strong> Health Australia (PHA),<br />

the Australian Government, all State and Territory Governments<br />

and plant industry signatories covering the management and<br />

funding of eradication responses to Emergency <strong>Plant</strong> Pest (EPP)<br />

Incidents. <strong>Plant</strong> Health Australia is the Custodian of the EPPRD<br />

and it became operative on October 26, 2005.<br />

The EPPRD replaces previous informal arrangements and<br />

provides a formal role for industry to participate and assume a<br />

greater responsibility in decision making in relation to EPP<br />

responses.<br />

DISCUSSION<br />

The EPPRD only operates for the eradication of EPPs. There are<br />

four criteria in the EPPRD for the definition of an EPP and the<br />

Pest only has to satisfy one of these to be considered an EPP.<br />

Briefly, these are:<br />

1. A new pest to Australia<br />

2. A different variation or strain of established pest<br />

3. A previously unknown pest<br />

4. A confined or contained pest<br />

The formal definitions can be found in Clause 1.1 Definitions of<br />

the EPPRD. There is a list of EPPs that have already met one of<br />

the definitions and have been Categorised in Schedule 13 of the<br />

EPPRD.<br />

For an eradication response to be agreed it must be both<br />

technically feasible and cost beneficial to eradicate the pest. As<br />

such, early reporting of suspect emergency plant pests is a<br />

critical step in the process. The longer it takes for a suspected<br />

EPP to be reported, the more time the pest has to become<br />

established and more wide spread. This increases the costs of<br />

containment, control and eradication measures, reduces the<br />

technical feasibility and therefore reduces the likelihood of<br />

success of eradication. Figure 1 demonstrates the differences in<br />

probability of successful eradication compared to time from<br />

detection to reporting.<br />

Government Signatories to the EPPRD undertake to provide<br />

formal notification within 24 hours of becoming aware of an<br />

incident and take all reasonable steps to ensure that persons<br />

within their jurisdiction are aware they need to advise that<br />

government within 24 hours of becoming aware of an incident<br />

so that the formal notification can be made. It is important that<br />

diagnosticians and researchers understand their responsibility,<br />

not only a moral obligation to protect Australian agriculture and<br />

horticulture but this legal obligation that now exists for<br />

jurisdictions and their personnel. Personnel of government<br />

agricultural agencies need to report a ‘reasonably held suspicion’<br />

of an exotic pest to their jurisdiction’s Chief <strong>Plant</strong> Health<br />

Manager directly or via the Exotic <strong>Plant</strong> Pest Hotline (1800 084<br />

881).<br />

It is also important for the integrity of the EPPRD that all Parties<br />

(government and industry) understand their rights and<br />

responsibilities and adhere to them. If Parties do not adhere, as<br />

closely as possible, to their responsibilities under the EPPRD, the<br />

work done to improve collaboration and trust between all<br />

Parties is eroded.<br />

Many significant plant pests are cryptic and not readily visible.<br />

Red Imported Fire Ant is estimated to have been present for as<br />

many as five years before detection, likewise European House<br />

Borer may have been present for as long as 50 years before<br />

detection.<br />

Both of these pests would have cost considerable less to<br />

eradicate if they had been detected and reported within the first<br />

few generations.<br />

Not only does the EPPRD provide an obligation to report a<br />

‘reasonably held suspicion’ but there is the potential for cost<br />

sharing of actions taken to be rejected if it is deemed that there<br />

has been a failure to report in a timely manner.<br />

ACKNOWLEDGEMENTS<br />

The authors would like to thank Dr Peter Caley of the Bureau of<br />

Rural Sciences, Department of Agriculture, Fisheries and Forestry<br />

for his assistance with generating the graphical data.<br />

REFERENCES<br />

1. <strong>Plant</strong> Health Australia (2005) Government and <strong>Plant</strong> Industry Cost<br />

Sharing Deed in respect of Emergency <strong>Plant</strong> Pest Responses (<strong>Plant</strong><br />

Health Australia; Canberra)<br />

Figure 1. Probability of successful eradication compared to reporting<br />

time lag following identification.<br />

110 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


The use of sentinel plantings in forest biosecurity; results from mixed eucalypt species<br />

trails in South‐East Asia and Australia<br />

INTRODUCTION<br />

T.I. Burgess{ XE "Burgess, T.I." } A , B. Dell A , G.E.StJ. Hardy A , P. Thu B<br />

A School of Biological Sciences and Biotechnology, Murdoch University, Murdoch, 6150, Australia<br />

B Forest Science Institute of Vietnam, Hanoi, Vietnam<br />

Many diseases of Eucalyptus species have emerged as pathogens<br />

in exotic plantations. Guava rust (Puccinia psidii), cryphonectria<br />

canker (Crysoporthe cubensis) coniotherium canker<br />

(Colletogloeopsis zuluensis) and Kirramyces leaf blight<br />

(Kirramyces destructans) are all serious pathogens that have not<br />

been found in native forests or in plantations in Australia<br />

(Burgess & Wingfield 2002; Cortinas et al. 2006; Glen et al. 2007;<br />

Wingfield et al. 2001). The susceptibility to these pathogens of<br />

Eucalyptus spp. commonly used in exotic plantations is known;<br />

however the susceptibility of many Eucalyptus spp. found only in<br />

natural ecosystems in Australia is unknown. There are two main<br />

uses of sentinel plantations. Firstly, tree species known to be<br />

susceptible to different pathogens can be planted within the<br />

natural environment to try and trap pathogens from their<br />

surroundings. In Australia, taxa trials planted in different<br />

environments act as sentinel plantings. By surveying these taxa<br />

trials we have collected and described a number of new eucalypt<br />

pathogens and reported the presence in Australia of Kirramyces<br />

destructans. The second use for sentinel planting is where many<br />

tree species are planted in a region known to harbour certain<br />

pathogens. In this manner the susceptibility of the different tree<br />

species can be determined.<br />

MATERIALS AND METHODS<br />

Taxa trials and adjoining natural vegetations have been surveyed<br />

in tropical and sub‐tropical Australia. This involves collection of<br />

diseased leaf and canker material, isolating the fungus using<br />

standard techniques and identification of the fungi using classic<br />

taxonomy and molecular phylogeny<br />

We have established sentinel trials of 25 eucalypt species in<br />

Vietnam, China and Thailand in regions known to harbour<br />

Kirramyces destructans and Colletogloeopsis zuluensis. To date<br />

only the trial in Vietnam has been surveyed for impact of leaf<br />

pathogens and insects. In addition, trees have been inoculated<br />

with Colletogloeposis zuluensis and lesion formation and lesion<br />

length measured. A matching trial has also been planted in<br />

northern Australia.<br />

not previously known to be affected. The trials in China and<br />

Thailand will be assessed in July 2009 prior and results available<br />

prior to the APPS conference in October<br />

Figure 1. (A) Leptocybe damage on young petioles, (B) Quambalaria spp.<br />

forming white spots on leaves<br />

These sentinel trials, established in Asia, will provide valuable<br />

information on the susceptibility to some of the keystone<br />

tropical Eucalyptus spp. to various exotic pathogens.<br />

ACKNOWLEDGEMENTS<br />

The work was supported by an ARC Discovery project<br />

DP0664334. We thank Great Southern Limited for provided land<br />

and maintaining trial on the Tiwi Islands.<br />

REFERENCES<br />

1. Burgess TI, Wingfield MJ, 2002. Impact of fungi in natural forest<br />

ecosystems; A focus on Eucalyptus. In: K Sivasithamparam, KW<br />

Dixon, RL Barrett (eds), Microorganisms in <strong>Plant</strong> Conservation and<br />

Biodiversity, Kluwer Academic Publishers, Dordrecht. pp. 285–306.<br />

2. Cortinas M‐N, Burgess TI, Dell B, Xu D, Wingfield MJ, Wingfield BD,<br />

2006. First record of Colletogloeopsis zuluense comb. nov., causing<br />

a stem canker of Eucalyptus spp. in China. Mycological Research.<br />

110: 229–236.<br />

3. Glen M, Alfenas AC, Zauza EAV, Wingfield MJ, Mohammed C, 2007.<br />

Puccinia psidii: a threat to the Australian environment and<br />

economy‐a review. <strong>Australasian</strong> <strong>Plant</strong> <strong>Pathology</strong>. 36: 1–16.<br />

4. Wingfield MJ, Slippers B, Roux J, Wingfield BD, 2001. Worldwide<br />

movement of exotic forest fungi especially in the tropics and<br />

Southern Hemisphere. Bioscience. 51: 134–140.<br />

Session 7B—Quarantine and exotic pathogens<br />

RESULTS AND DISCUSSION<br />

We have focused our sampling in northern Australia on fungi<br />

causing disease; on leaves the Mycosphaerellaceae<br />

predominated, especially Kirramyces species, the dominant<br />

pathogens in cankers belong to the Botryosphaeriaceae. Many of<br />

the species found have been described on eucalypts either in<br />

Australia, but often elsewhere where eucalypts are grown.<br />

Several new fungal species have been described. In the trial in<br />

Northern Australia only eucalypt species already known to be<br />

susceptible to K. destructans developed Kirramyces leaf Blight.<br />

The other 20 eucalypt species did not develop symptoms<br />

The trials in Vietnam have been monitored with the main finding<br />

being an expansion of the host range of the gall wasp Leptocybe<br />

invasa and the leaf pathogen Quambalaria eucalypti and the<br />

discovery of a new Quambalaria sp (Figure 1). The greatest<br />

lesion length after inoculation with Colletogloeopsis zuluensis<br />

was observed for Eucalyptus saligna and E. pellita, two species<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 111


Session 7B—Quarantine and exotic pathogens<br />

Methyl bromide alternatives for quarantine and pre‐shipment and other purposes—<br />

future perspectives<br />

J.E. Oliver{ XE "Oliver, J.E." }<br />

A Office of the Chief <strong>Plant</strong> Protection Officer, <strong>Plant</strong> Division, Biosecurity Services Group, GPO Box 858, Canberra, 2601, ACT<br />

INTRODUCTION<br />

The switch to methyl bromide alternatives for fumigation<br />

purposes has largely hinged on the ozone depleting status of this<br />

chemical.<br />

This paper explores the major plant and plant product<br />

commodities treated with methyl bromide for quarantine and<br />

pre‐shipment use on an Australian and international basis.<br />

The underlying domestic and international policy context,<br />

technical, regulatory, legislative, market access and trade and<br />

environmental factors, are all key drivers or impediments to<br />

adoption of alternatives.<br />

Yet there is no ‘silver bullet’ to replace methyl bromide<br />

treatment, or is there? Or is a paradigm shift required with how<br />

quarantine risk is assessed or when, how or if treatments are<br />

applied?<br />

Some of the key alternatives for plant and plant product<br />

commodities are identified, as are the underlying issues that<br />

determine the adoption or otherwise of alternatives.<br />

MATERIALS AND METHODS<br />

The <strong>Plant</strong> Health Committee consisting of plant protection<br />

officers from the Australian government and state biosecurity<br />

agencies commissioned an investigation of what methyl bromide<br />

alternative treatments were available for quarantine and preshipment<br />

and other plant uses. Specifically, what alternatives<br />

were under research, under registration and or in use in<br />

Australia and/or in use and/or registered internationally.<br />

To ensure data accessibility and an ongoing commitment to<br />

evaluating alternatives, an information system, the Methyl<br />

Bromide Alternatives Information System was created.<br />

International <strong>Plant</strong> Protection Convention standards for<br />

commodity classification and acceptance of a new phytosanitary<br />

treatment were adopted to ensure that the dataset would meet<br />

national and international requirements.<br />

of alternatives. Adoption by a country of alternatives is also<br />

dependent on quarantine services approving a treatment for<br />

use, and for trade using the alternatives being acceptable to<br />

both importing and exporting countries.<br />

The Methyl Bromide Alternatives Information System was<br />

designed to capture these factors and to provide a transparent<br />

and internationally acceptable repository for data on<br />

alternatives.<br />

The Methyl Bromide Alternatives Information System has over<br />

150 registered users and contains over 480 records. Initially<br />

registration was limited to within Australia. Since December<br />

2008 membership has been expanded to international<br />

subscribers through the IPPC network and by inviting further<br />

data input from the Quads countries—United States, Canada and<br />

New Zealand.<br />

Alternatives will be discussed using a case study of Australia’s<br />

highest volume uses for QPS and the pros and cons of<br />

alternatives that have been identified.<br />

Pests of quarantine concern where no efficacious alternatives<br />

have been identified, or for which there are no approved<br />

Australian quarantine treatments other than methyl bromide<br />

will also be discussed.<br />

The presentation will include a live demonstration of MBAIS and<br />

conference delegates will be encouraged to become registered<br />

users and data contributors to the system.<br />

REFERENCES<br />

1. Methyl Bromide Alternatives Information System<br />

www.daff.gov.au/mbais<br />

Whilst collating data on alternatives, a series of themes<br />

emerged. Through this process the major drivers and<br />

impediments to adopting alternatives were identified.<br />

RESULTS AND DISCUSSION<br />

The efficacy of alternative chemical or non‐chemical methods of<br />

fumigation was found not to be the primary driver for the switch<br />

to a methyl bromide alternative.<br />

Instead the underlying context of consumer demand for<br />

chemical free alternative treatments, occupational and<br />

environmental health concerns, more intense scrutiny for<br />

chemical registration and the European Union’s decision to<br />

deregister methyl bromide from March 2010 were the most<br />

significant contributing factors for adopting methyl bromide<br />

alternatives.<br />

The time between research and commercialisation of<br />

alternatives, the requirements for registration and variance in<br />

data protection laws contributed to different countries adoption<br />

112 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Fruit extracts of Azadirachta indica induces systemic acquired resistance in tomato<br />

against Pseudomonas syringae pv tomato<br />

INTRODUCTION<br />

V. Bhuvaneswari, Parul Tyagi, Sandeep Soni, Satyakam Pugla and P.K. Paul{ XE "Paul, P.K." }<br />

Amity Institute of Biotechnology, Amity University, Express Highway, Sector ‐125, Noida, Uttar Pradesh ‐201303, India<br />

Last couple of decades has witnessed a spurt in use of plant<br />

extracts as biocontrol agents for controlling a wide range of crop<br />

pathogens. However mechanism of action of such extracts has<br />

not been well understood.<br />

MATERIALS AND METHODS<br />

In the present study 6 week old plants of two cultivars of tomato<br />

i.e. wild and a F1 hybrid were treated with aqueous fruit extracts<br />

of A.indica (neem). The treated plants were inoculated with<br />

spores of Ps Syringae either 24 hrs prior to or after neem<br />

treatment. Leaf samples were collected from inoculated noninoculated<br />

treated and control plants at intervals of 24 hours for<br />

6 days, for estimating the activity of phenylalanine ammonia<br />

lyase, tyrosine ammonia lyase, polyphenol oxidase and contents<br />

of total phenol, mRNA and proteins. Disease intensity was<br />

recorded after 3rd and 10th week of inoculation. Neem treated<br />

and control plants were also treated with inhibitors of<br />

transcription and translation.<br />

Session 7C—Alternatives to chemical control<br />

RESULTS AND DISCUSSION<br />

Wild cultivar after treatment has substantially lower infection<br />

after neem treatment as compared to F 1 hybrid and controls. In<br />

both cultivars there was a sharp increase in activity of PAL, TAL,<br />

PPO and concentration of total phenols. Treated plants had new<br />

mRNA and high concentration of some proteins. Treatment with<br />

inhibitors of transcription resulted in absence of new mRNA and<br />

reduced concentration of some proteins. Translational inhibitors<br />

inhibited the increase in concentration of proteins. SDS‐PAGE<br />

analysis had revealed increased concentration of PAL, TAL, PPO.<br />

Disease incidence of neem treated plants after 10 weeks of<br />

inoculation was substantially less than control. Results reveal<br />

that neem extracts control pathogen on tomato through<br />

induction of SAR. SAR induction is a function of plant genome as<br />

evidenced from difference in results of wild and F1 hybrids.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 113


Session 7C—Alternatives to chemical control<br />

Fungal foliar endophytes induce systemic protection in cacao seedlings against<br />

Phytophthora palmivora<br />

INTRODUCTION<br />

C. Blomley{ XE "Blomley, C." } A , E.C.Y. Liew B and D.I. Guest A<br />

A Faculty of Agriculture Food and Natural Resources, The University of Sydney, 2006, NSW<br />

B Royal Botanic Gardens Trust, The Royal Botanic Gardens, Sydney, NSW, 2000<br />

The fungal endophytes of cacao (Theobroma cacao) comprise a<br />

diverse assemblage including pathogens and saprophytes. The<br />

interaction between these endophytes and their host is not yet<br />

understood, although some fungal endophytes may protect their<br />

host against pathogens 1 .<br />

Black pod, caused Phytophthora spp. is responsible for losses to<br />

global cocoa production of around 20‐30% annually 2 . In Papua<br />

New Guinea, the main methods of controlling black pod caused<br />

by P. palmivora are to use resistant cultivars and cultural<br />

practices, as fungicides have limited efficacy and are generally<br />

not cost‐effective for small‐holder farmers. Biological control of<br />

cocoa diseases using endophytic fungi may provide another tool<br />

in the Integrated Pest and Disease Management toolbox. The<br />

purpose of this study was to identify common fungal endophytes<br />

present in Australia and Papua New Guinea capable of reducing<br />

disease severity of disease caused by P. palmivora. Such<br />

knowledge is an essential first step in the development of a<br />

biological control agent and will help to solve unanswered<br />

questions regarding the ecological role of fungal foliar<br />

endophytes.<br />

MATERIALS AND METHODS<br />

Endophytes were sampled from leaves and pods of cacao<br />

growing in five locations in Australia and Papua New Guinea.<br />

Common endophyte taxa were screened for the ability to reduce<br />

the growth of P. palmivora in vitro via dual culture and precolonised<br />

plate assays. The same endophyte taxa were tested for<br />

the ability to reduce severity of foliar disease caused by<br />

P. palmivora. Cacao seeds were germinated and either exposed<br />

to ambient endophyte spores in the glasshouse or were kept<br />

endophyte free by raising seedlings in sterile soil in a clean room.<br />

Two leaves on each seedling were infected with a 5.6 x 10 6<br />

propagules/ml suspension of a single endophyte taxon or of all<br />

taxa combined. Endophyte infection was tested 17 days later by<br />

harvesting one leaf from each seedling, surface sterilising leaf<br />

fragments and then plating them onto growth media. 18 days<br />

after endophyte inoculation, one endophyte infected and one<br />

endophyte free leaf on each seedling was inoculated with<br />

P. palmivora by applying a 30µl drop of zoospore suspension<br />

(400,000 zoospores /ml) to the leaf midvein which had been<br />

pierced with a sterile hypodermic needles. Control seedlings<br />

were mock inoculated with sterile water. Ten seedling replicates<br />

were prepared for each treatment and control. Disease severity<br />

was measured five days later as the length of the necrotic lesion<br />

along the mid vein.<br />

inoculation of one cacao leaf on a seedling with either<br />

Phomopsis or Diplodia resulted in uninoculated leaves on the<br />

same plant being less susceptible to disease compared to<br />

completely endophyte free seedlings. In contrast, addition of<br />

endophyte inoculum to seedlings that had already been infected<br />

with ambient endophytes in the glasshouse did not result in a<br />

reduction in disease severity caused by P. palmivora.<br />

DISCUSSION<br />

Results from these experiments demonstrate that common, noncoevolved<br />

endophyte taxa can increase the resistance of their<br />

host plants to pathogens. In some cases, endophyte mediated<br />

protection is systemic and therefore likely to be due to induction<br />

of a non‐specific defence response in the host. Harnessing this<br />

apparently beneficial interaction between endophytes and their<br />

host to manage plant disease may not be easily achievable as<br />

our results show that addition of more endophytes to already<br />

infected seedlings did not decrease disease severity. However,<br />

these results suggest that fungal foliar endophytes have a<br />

protective interaction with their host in a broader ecological<br />

sense.<br />

REFERENCES<br />

1. Arnold AE, Mejia LC, Kyllo D, Rojas WI, Maynard Z, Robbins N, Herre<br />

EA (2003) Fungal endophytes limit pathogen damage in a tropical<br />

tree. Proc. Nat. Acad. Sci. 100, 15649‐15654<br />

2. Guest DI (2007) Black Pod: Diverse pathogens with a global impact<br />

on cocoa yield. Phytopathology 97, 1650‐1653<br />

RESULTS<br />

Five of the common endophyte taxa tested inhibited the growth<br />

of P. palmivora when co‐cultured on corn meal agar compared<br />

to P. palmivora grown alone. Subsequent experiments showed<br />

that one Xylaria morphotype inhibited the growth of<br />

P. palmivora via antibiosis whilst the other endophytes were<br />

inhibitory via competition for resources. Two endophyte taxa<br />

that were inhibitory in vitro, as well as three other common<br />

endophyte taxa reduced disease severity in seedlings that had<br />

been grown in an endophyte free environment. Additionally,<br />

114 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Effectiveness of the rust Puccinia myrsiphylli in reducing populations of the invasive<br />

plant bridal creeper in Australia<br />

INTRODUCTION<br />

Bridal creeper (Asparagus asparagoides) is a dense scrambling<br />

vine that smothers large areas of vegetation and threatens<br />

native biodiversity in Australia (1). In winter rainfall areas, it<br />

begins to grow in late summer – early autumn, produces fruit in<br />

mid to late spring and above‐ground foliage naturally senesces<br />

at the beginning of summer. Bridal creeper has been the target<br />

of a biological control program since the 1990s. Three agents of<br />

South African origin have since been released: the leafhopper<br />

Zygina sp. in 1999, the rust fungus Puccinia myrsiphylli in 2000<br />

and the leaf beetle Crioceris sp. in 2002 (2). Both the leafhopper<br />

and rust fungus have established widely on bridal creeper<br />

populations across temperate Australia. The rust fungus<br />

however, is the most effective agent, probably due to its major<br />

indirect impact on the plant’s below‐ground biomass (3).<br />

We report on results from two different approaches used to<br />

measure the effectiveness of P. myrsiphylli in reducing<br />

populations of bridal creeper across Australia.<br />

MATERIALS AND METHODS<br />

Before and after release comparisons. One to three years<br />

before releasing the rust, support structures (or trellises—2 m in<br />

height and 90 cm wide) were set up at the edge of 3 m 2 plots<br />

(three to four plots per site) in bridal creeper infestations at 15<br />

sites across southern Australia. Growth and reproductive<br />

parameters of bridal creeper in a 1 m 2 quadrat within each plot<br />

and climbing on the trellis were recorded in mid‐spring each year<br />

for up to 8 years after the release. Incidence of the rust was also<br />

measured annually. This long‐term experiment was performed in<br />

partnership with a range of collaborators across Australia.<br />

Fungicide exclusion experiments. Permanent 1 m 2 quadrats (10<br />

per site) with a central 100 cm long stake were set up in bridal<br />

creeper infestations at three sites in NSW and three sites in WA,<br />

where the rust fungus was the dominant agent. Sites in WA were<br />

managed by CSIRO Entomology staff based in Perth. Half of the<br />

quadrats at each site were maintained rust‐free using monthly<br />

fungicide applications during the growing season for 4 years. The<br />

remaining quadrats were sprayed with water only. Percentage<br />

cover of bridal creeper and other plant species within quadrats,<br />

as well as other bridal creeper growth and reproductive<br />

parameters, and disease incidence and severity were measured<br />

each year.<br />

L. Morin{ XE "Morin, L." } and A.M. Reid<br />

CSIRO Entomology, GPO Box 1700, Canberra, ACT 2601, Australia<br />

rust‐infected quadrats, bridal creeper was replaced by both<br />

native and exotic species, although other weeds were more<br />

prevalent at disturbed sites.<br />

Foliage dry weight (g) or<br />

seedling or shoot number / m 2<br />

[ln(y+1)]<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

Figure 1. Changes in bridal creeper growth parameters in permanent 1<br />

m 2<br />

quadrat established at 15 sites across southern Australia and<br />

monitored 1–2 year before and up to 8 years after the release of P.<br />

myrsiphylli.<br />

DISCUSSION<br />

Before<br />

release<br />

These studies demonstrate that P. myrsiphylli has major negative<br />

impacts on bridal creeper populations, although some sites may<br />

need to be carefully managed to facilitate native species<br />

recovery. For example, only a few plant species were present in<br />

rust‐infected quadrats by the end of the 4‐year fungicide<br />

exclusion experiment. More than four consecutive seasons of<br />

severe rust epidemics on bridal creeper may thus be required to<br />

sufficiently reduce populations and trigger major recruitment by<br />

other plants.<br />

ACKNOWLEDGEMENTS<br />

3-year period immediately<br />

after release<br />

Monitoring period<br />

Late after<br />

release<br />

Foliage<br />

Seedlings<br />

Shoots<br />

We gratefully acknowledge the contribution to these<br />

experiments of our collaborators from The Departments of<br />

Primary Industries in NSW and Victoria, The Department of<br />

Water, Land and Biodiversity Conservation in South Australia<br />

and CSIRO Entomology based in Canberra and Perth. We also<br />

wish to thank Bob Forrester of CSIRO Entomology for his<br />

assistance with statistical analyses. This work was supported by<br />

CSIRO, Weeds CRC, the Australian Government (NHT &<br />

Defeating the Weed Menace initiative) and the respective<br />

organisation of our collaborators.<br />

147.4<br />

53.6<br />

19.1<br />

6.4<br />

1.7<br />

0<br />

Back-transformed scale<br />

Session 7C—Alternatives to chemical control<br />

RESULTS<br />

Before and after release comparisons. After the release of the<br />

rust, bridal creeper seedling and shoot numbers and aboveground<br />

biomass in the permanent 1 m 2 quadrats steadily<br />

declined at all sites (Fig. 1). In contrast, the impact of the rust on<br />

bridal creeper climbing onto trellises varied considerably<br />

between sites.<br />

Fungicide exclusion experiments. Bridal creeper cover, aboveground<br />

biomass, and shoot, fruit and seedling numbers were<br />

substantially lower in rust‐infected compared to rust‐free<br />

quadrats across all sites over the years. The cover of bare ground<br />

and leaf litter was consistently higher in rust‐infected quadrats<br />

than in rust‐free quadrats. When other plants did establish in<br />

REFERENCES<br />

1. Morin L, Batchelor KL, Scott JK (2006) The biology of Australian<br />

weeds. Asparagus asparagoides (L.) Druce. <strong>Plant</strong> Protection<br />

Quarterly 21, 46–62.<br />

2. Morin L, Neave M, Batchelor KL, Reid A (2006) Biological control: a<br />

promising tool for managing bridal creeper in Australia. <strong>Plant</strong><br />

Protection Quarterly 21, 69–77.<br />

3. Morin L, Reid AM, Willis AJ (2006) Impact of the rust fungus<br />

Puccinia myrsiphylli on the below‐ground biomass of bridal<br />

creeper. In ‘Proceeding of 15th Australian Weeds Conference’ (Eds<br />

C Preston, JH Watts, ND Crossman) p 608. (Weed Management<br />

<strong>Society</strong> of South Australia).<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 115


Session 7C—Alternatives to chemical control<br />

Evaluation of essential oils and other plant extracts for control of soilborne pathogens<br />

of vegetable crops<br />

INTRODUCTION<br />

C.A. Scoble{ XE "Scoble, C.A." } A,B , E.C. Donald A , K.M. Plummer B and I.J. Porter A<br />

A Department of Primary Industries, 621 Burwood Hwy, Knoxfield, Victoria, 3180<br />

B Botany Department, La Trobe University, Bundoora, Victoria, 3086<br />

Soilborne plant pathogens including Pythium spp, Fusarium spp,<br />

and Rhizoctonia spp, can cause important diseases such as root<br />

rot and damping off, resulting in heavy crop losses in vegetables<br />

farms. Control of these diseases is problematic because these<br />

pathogens have a wide host range and survive in soil as<br />

oospores, chlamydospores and melanised hyphae, respectively,<br />

for long periods. Compounds derived from plant extracts have<br />

been proposed as potential control treatments for soilborne<br />

pathogens due to their antimicrobial activity in laboratory<br />

studies (1). For instance, essential oils can contain phenolic and<br />

terpenoid compounds which have antimicrobial properties (2).<br />

The antimicrobial activity of thyme oil has been shown to cause<br />

hyphal collapse by membrane disruption (1). In Australia, very<br />

few studies have investigated the effects of antimicrobial volatile<br />

compounds in essential oils on survival of soilborne pathogens.<br />

Our work is therefore investigating the antimicrobial activity of<br />

compounds derived from a range of essential oils and other<br />

plant extracts against key soilborne pathogens isolated from<br />

vegetable crops. Preliminary results from in vitro experiments<br />

are reported here.<br />

MATERIALS AND METHODS<br />

A series of in vitro experiments were conducted to investigate<br />

the effects of plant extracts on mycelial growth of soilborne<br />

pathogen isolates. Treatments tested included the active<br />

constituents (eugenol, thymol, carvacrol and geraniol) of some<br />

essential oils as well as 14 essential oils (thyme, clove bud,<br />

peppermint, geranium, eucalyptus, tea tree, origanum,<br />

rosemary, orange sweet, cardamon, sweet fennel, pine, black<br />

pepper and basil). The pathogenic isolates tested included<br />

Pythium sulcatum, P. aphanidermatum, P. irregulare, Fusarium<br />

oxysporum and a Rhizoctonia sp. Solutions of the actives and the<br />

oils were added to sterile suitable selective media at<br />

concentrations of 500, 1000, and 2500 ppm. A 5 mm mycelial<br />

plug was then plated onto the amended media. Plates with<br />

Pythium were incubated at 20ºC and those with Fusarium and<br />

Rhizoctonia at room temperature. Mycelial growth, expressed as<br />

colony diameter, was measured until mycelium in unamended<br />

plates reached the edge of the plate. After this, plugs that did<br />

not grow were transferred to fresh unamended media to<br />

determine whether the treatments were fungistatic or<br />

fungicidal. Examples of results are given for some of the isolates<br />

tested to illustrate the suppressive and biocidal effects of some<br />

of the oil treatments tested.<br />

RESULTS<br />

All concentrations of the four plant actives significantly (p


Use of grid weather forecast data to predict rice blast development in Korea<br />

Wee Soo Kang 1 , Kyu Rang Kim 2 , and Eun Woo Park{ XE "Park, E.W." } 1<br />

1 Department of Agricultural Biotechnology, Seoul National University, Seoul 151‐921, Korea<br />

2 Applied Meteorological Research Lab., National Institute of Meteorological Research, Korea Meteorological Administration, Seoul<br />

156‐720, Korea<br />

Keynote address<br />

Timely warnings on plant disease development are useful<br />

information for farmers to determine when to spray fungicides<br />

to control plant diseases. Real‐time weather data monitored by<br />

automated weather stations are often used to generate disease<br />

forecast information. However, when observed weather data are<br />

used, the time window for effective fungicide sprays after the<br />

disease warnings could be too short and farmers may fail to<br />

control the disease in time even though accurate disease<br />

forecast is available. In order to minimise the time limitations<br />

associated with real‐time disease forecasting, the weather<br />

forecast data need to be used for disease forecasting. At every<br />

3 hours, the Korea Meteorological Administration (KMA)<br />

releases 48‐hour weather forecasts at 3‐hour intervals on air<br />

temperature, relative humidity, and probability of precipitation<br />

and at 12‐hour intervals on precipitation based on the outputs<br />

from numerical weather prediction models. The spatial<br />

resolution of weather forecasts is 5 km. Using the grid weather<br />

forecasts, rice blast disease forecasting was conducted. The grid<br />

forecast data at 3 hour intervals are interpolated to produce<br />

hourly data. Hourly wetness period was estimated from a simple<br />

relative humidity model and a CART model using temperature,<br />

relative humidity, precipitation, and wind speed. Based on the<br />

hourly weather data, daily risk levels of rice blast infection were<br />

determined at the spatial resolution of 5 km in the map image.<br />

The results suggested that estimation of hourly leaf wetness<br />

needs to be improved to enhance the accuracy in forecasting<br />

infection periods.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 117


Session 8A—Future directions<br />

Investigating the impact of climate change on plant diseases<br />

J.E. Luck{ XE "Luck, J.E." } A , J.P. Aurambout B , S. Chakraborty C , K.J. Finlay A , W. Griffiths D , G.J. Hollaway D G. O’Leary D , A. Freeman D , K.<br />

Powell E , R. Norton F , D. Kriticos G and P. Trebicki D<br />

A<br />

Biosciences Research Division, Department of Primary Industries, Knoxfield, Private Bag 15, Ferntree Gully DC, 3156, Victoria<br />

B Future Farming Systems Research, Department of Primary Industries, Parkville, PO Box 4166, Parkville, 3030, Victoria<br />

C CSIRO <strong>Plant</strong> Industries, 306 Carmody Rd, St Lucia, 4067, Queensland<br />

D Biosciences Research Division, Department of Primary Industries, Horsham, Private Bag 260 Horsham, 3401, Victoria<br />

E Biosciences Research Division, Department of Primary Industries, Rutherglen, RMB 1145 Chiltern Valley Road, Rutherglen, 3685, Victoria<br />

F The University of Melbourne, Private Bag 260, Horsham, 3401, Victoria<br />

G CSIRO Entomology, Black Mountain, 2601 ACT<br />

INTRODUCTION<br />

Climate change is recognised as a major threat to agricultural<br />

systems and will likely alter the risks associated with the<br />

biosecurity and market access of its agricultural products (1). The<br />

potential effects on pest and disease threats to these changing<br />

systems are not well understood. Specifically, there is a critical<br />

lack of empirical data on how increasing atmospheric carbon<br />

dioxide (CO 2 ) will impact on pest and pathogen populations and<br />

crop production. Our approach to investigate the impact of<br />

climate change on disease threats is two‐fold. We developed<br />

models to analyse vector‐borne diseases of a number of<br />

endemic and exotic diseases and their hosts. Further, we have<br />

embarked upon a real time field study of the effects of elevated<br />

(e)CO 2 on pathogens of wheat, using the Free‐Air CO 2<br />

Enrichment (FACE).<br />

METHODS<br />

Modelling. We used a bioclimatic model (CLIMEX) to investigate<br />

the potential distribution of citrus canker (Xanthomonas citri pv<br />

citri). A second modelling approach (using STELLA) combined<br />

dynamic sub‐models of host‐plant physiology, vector population<br />

growth and climatic data to investigate the effect of climate<br />

change on the exotic Asiatic citrus psyllid (Diaphorina citri) which<br />

vectors the citrus disease huanglongbing (citrus greening). This<br />

approach was also used to develop an integrative model for the<br />

bird cherry‐oat aphid (Rhopalosiphum padi) under increasing<br />

temperatures which vectors Barley yellow dwarf virus in wheat.<br />

FACE. Three wheat pathogens were targeted to examine the<br />

influence of CO 2 on their biology and interaction with their host,<br />

Puccinia striiformis (wheat stripe rust), Fusarium<br />

pseudograminearum (crown rot) and Barley yellow dwarf virus<br />

(BYDV). Wheat stripe rust severity, latent period, fecundity and<br />

host resistance was assessed under ambient and 550ppm CO 2 .<br />

Crown rot severity on Tamaroi (very susceptible variety) and 2–<br />

49 (a partially resistant breeding line variety) were also assessed.<br />

The RPV strain of BYDV was isolated and maintained in oat and<br />

perennial ryegrass by serial aphid transfer using R.padi and<br />

transmitted to wheat growing under eCO 2 . DNA analysis of soil<br />

from ambient (aCO 2 ) and eCO 2 plots for soil‐borne pathogen<br />

detection was also completed.<br />

RESULTS<br />

Modelling. CLIMEX results revealed that the predicted citrus<br />

canker distribution would shift to southern coastal and inland<br />

regions under increasing temperatures, based on its current<br />

climatic range overseas. The STELLA model indicated earlier,<br />

shorter development times and a population shift southwards<br />

for the potentially invasive Asiatic citrus psyllid (Diaphorina citri),<br />

under increasing temperatures. However, an overall decrease in<br />

psyllid numbers is predicted due to reduce availability of new<br />

growth flushes on which the psyllid feeds and reproduces (2).<br />

FACE. There was no significant effect of eCO 2 on stripe rust<br />

progress, latent period, fecundity and disease resistance but a<br />

significant increase in the severity of crown rot symptoms and<br />

fungal biomass under eCO 2 . The methodology for analysing<br />

BYDV movement in wheat under eCO 2 has been established. In<br />

2007, visual scoring of BYDV symptom expression between eCO 2<br />

and aCO 2 treatments showed no significant difference. The 2008<br />

BYDV inoculation results are being analysed.<br />

DISCUSSION<br />

Modelling tools are available to predict geographic ranges of<br />

plants and their pests and pathogens but often do not take into<br />

consideration the specific biology of the organisms and hostpathogen/pest<br />

interactions. The value of this approach is<br />

confirmed by the Asiatic citrus psyllid model which would have<br />

only predicted shorter development times and an increase risk<br />

of potential distribution based on climatic data alone. Such<br />

improved models could be incorporated into plant disease<br />

management plans and contingency planning. The predictive<br />

power of the models will be increased with real‐time data<br />

generated by the FACE experiments. Limited chamber and field<br />

studies have shown an increase in aggressiveness of some fungal<br />

pathogens and enhanced host susceptibility (3,4) which was<br />

supported an increase severity of crown rot symptoms and<br />

fungal biomass under eCO 2 . For wheat stripe rust, cultivar<br />

resistance and seasonal rainfall have more significant effects on<br />

disease progress than increased atmospheric (550ppm) CO 2 .<br />

Further analysis is required to make an assessment of eCO 2<br />

effects on BYDV titre and movement in wheat and aphid<br />

vectoring capacity.<br />

REFERENCES<br />

1. Aurambout J‐P, Finlay KJ, Luck JE, Beattie, GAC (2009) A concept<br />

model to estimate the potential distribution of the Asiatic citrus<br />

psyllid (Diaphorina citri Kuwayama) in Australia under climate<br />

change—a means for assessing biosecurity risk. Ecological<br />

Modelling. In press.<br />

2. Aurambout J‐P, Constable F, Finlay KJ, Luck JE Sposito V (2006) The<br />

impacts of climate change on plant biosecurity. Primary Industries<br />

Research Victoria, Melbourne, 42pp.<br />

3. Chakraborty S, Datta S (2003) How will plant pathogens adapt to<br />

host plant resistance at elevated CO 2 under a changing climate?<br />

New Phytologist 159, 733–742.<br />

4. Karnosky DF, Percy KE, Xiang B, Callan B, Noormets A, Mankovska<br />

B, Hopkin A, Sober J, Jones W, Dickson RE, Isebrands JG (2002)<br />

Interacting elevated CO 2 and tropospheric O 3 predisposes aspen<br />

(Populus tremuloides Michx.) to infection by rust (Melampsora<br />

medusae f. sp. tremuloidae). Global Change Biology 8, 329–338<br />

118 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Impact of climate change in relation to blackleg on oilseed rape and blackspot on field<br />

pea in Western Australia<br />

M.U. Salam{ XE "Salam, M.U." } A , W.J. MacLeod A , M.J. Barbetti A,B , R. Khangura A , J. Galloway A , M. Amjad A , M. Seymour A , K.P. Salam A , T.<br />

Maling A , I. Foster A , D. Bowran A<br />

A Department of Agriculture and Food, Western Australia, Locked bag 4, Bentley Delivery Centre, WA 6983, Australia<br />

B School of <strong>Plant</strong> Biology, The University of Western Australia, 35 Stirling Highway, Crawley, WA 6009, Australia<br />

INTRODUCTION<br />

Worldwide, including Western Australia (WA), blackleg and<br />

blackspot are the most important diseases affecting production<br />

of oilseed rape (or canola) and field pea, respectively.<br />

Synchronisation of ascospore release with the seedling stage<br />

often results in particularly severe canker formation and<br />

correspondingly major yield losses for oilseed rape. Therefore,<br />

delay in onset of release of ascospores until the susceptible<br />

seedling stage has passed could be beneficial to oilseed rape. In<br />

contrast, field pea crops are susceptible to blackspot throughout<br />

all growth stages, therefore, it is recommended that growers<br />

delay crop establishment to avoid the peak of release of<br />

ascospores. In order to estimate the impacts of climate change<br />

on the severity of these diseases, it is important to identify the<br />

potential shifts in the pattern(s) of ascospore release for these<br />

diseases under the anticipated future climate conditions.<br />

slightly favour canker formation in oilseed rape (data not<br />

shown), with the potential for ideal conditions for canker<br />

formation becoming more frequent.<br />

On the contrary, the peak of the release of blackspot ascospores<br />

is predicted to be about four weeks earlier than under current<br />

climatic conditions; relative to break of the season, the release<br />

of blackspot ascospores will take place in about five weeks<br />

earlier (Fig. 1). That scenario would be highly beneficial to field<br />

pea growers, allowing earlier sowing dates to reap the greater<br />

agronomic yield potentials while being exposed to relatively low<br />

blackspot disease pressure.<br />

Session 8A—Future directions<br />

METHODS<br />

The ‘Blackleg Sporacle’ model was used to determine the timing<br />

of onset of release of blackleg ascospores and the ‘Blackspot<br />

Manager’ model was used to estimate the peak of release of<br />

blackspot ascospores in eight locations across the grain‐belt of<br />

WA, viz., Badgingarra, Corrigin, Dalwallinu, Esperance, Lake<br />

Grace, Merredin, Wagin, and Wandering. The models were run<br />

with statistically downscaled CSIRO Mk3 GCM simulation<br />

weather data based on the IPCC SRES A2 emission scenario<br />

(2036–2060) (1). This future scenario was compared with<br />

recorded climate data for the period of 1976–2004 inclusive.<br />

Table 1. Predicted timing of onset of blackleg ascospore release under<br />

current and future climates. DOY denotes day of the year (as of Julian<br />

day, 1 being 1 January and 366 being 31 December).<br />

Location<br />

Onset of ascospore release<br />

Current<br />

(DOY)<br />

Future<br />

(DOY)<br />

Badgingarra 170 181 11<br />

Corrigin 168 179 11<br />

Dalwallinu 183 196 13<br />

Esperance 104 116 12<br />

Lake Grace 163 177 14<br />

Merredin 177 191 14<br />

Wagin 150 168 18<br />

Wandering 149 164 15<br />

Average 158 172 13<br />

RESULTS AND CONCLUSIONS<br />

Difference<br />

(day)<br />

Model runs with these climate scenarios show that, on average,<br />

the onset of release of blackleg ascospores is likely to be delayed<br />

by about two weeks (Table 1). Whereas, the opening seasonal<br />

rains (also known as ‘break of the season’) were delayed by<br />

about one week (data not shown). Hence, the onset of blackleg<br />

ascospores, relative to the break of the season, will occur only<br />

one week later than currently. Therefore, the risk of<br />

synchronisation of major blackleg ascospore showers with<br />

seedling establishment appears to be similar under future<br />

climates. The predicted future increase in temperature may<br />

Figure 1. Predicted temporal pattern of blackspot ascospore release<br />

under current and future climate Esperance and Dalwallinu regions of<br />

Western Australia.<br />

ACKNOWLEDGEMENTS<br />

We thank the Australian Grains Research and Development<br />

Corporation (GRDC) and the Department of Agriculture and Food<br />

Western Australia for supporting this research.<br />

REFERENCES<br />

1. Gordon, H.B., Rotstayn, L.D., McGregor, J.L., Dix, M.R., Kowalczyk,<br />

E.A., O’Farrell, S.P., Waterman, L.J., Hirst, A.C., Wilson, S.G., Collier,<br />

M.A., Watterson, I.G. & Elliott, T.I., (2002) The CSIRO Mk3 Climate<br />

System Model. Technical Paper No. 60. CSIRO Atmospheric<br />

Research: Aspendale, Victoria, Australia 130 pp.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 119


Session 8A—Future directions<br />

Approaches to training in plant pathology capacity building projects in developing<br />

countries<br />

L.W. Burgess{ XE "Burgess, L.W." } A , J.L. Walsh B , H.T. Phan C and E.C.Y. Liew D<br />

A University of Sydney, A05 McMillan Building, Sydney, 2006, NSW<br />

B <strong>Plant</strong> Protection Centre, PO Box 811, Vientiane, Lao PDR<br />

C National Institute of Medicinal Materials, 3B Quang Trung, Hanoi, Vietnam<br />

D Botanic Gardens Trust, Mrs Macquaries Road, Sydney, 2000, NSW<br />

<strong>Plant</strong> diseases continue to be an impediment to the socioeconomic<br />

progress of developing countries. The ability of plant<br />

pathologists in these countries to accurately diagnose and<br />

provide advice on integrated disease management is often<br />

limited by an overall low level of training and lack of diagnostic<br />

specialists in country. Laboratory facilities are frequently<br />

inadequate and ongoing access to consumable laboratory<br />

materials and maintenance of equipment is difficult due to<br />

limited financial resources.<br />

Australia is an active donor to agricultural capacity building<br />

projects in many developing countries through government<br />

agencies and other funding bodies. Because capacity building<br />

projects focus on sustainable development through skills<br />

transfer, training of counterparts in developing countries is an<br />

inherent component of any capacity building project. Here we<br />

outline what we consider to be the three most important<br />

elements of training where many capacity building projects can<br />

fail: ‘English in Context’ training, suitability of Australian<br />

postgraduate courses and extended periods of in‐country<br />

training. Our recommendations are based on our experience<br />

working on plant pathology capacity building projects in<br />

Vietnam, Indonesia, Laos, China and Tunisia over the past 15<br />

years.<br />

‘English in Context’ training. Language can be a barrier to<br />

effective training for counterparts from non‐English speaking<br />

countries. A level of sufficiency in technical English used in plant<br />

pathology is required in order for counterparts to communicate<br />

with Australian mentors and interpret for international visitors,<br />

access internet resources and seek information from literature<br />

and disease compendia. ‘English in Context’ refers to the use of<br />

English as a second language in the workplace. We recommend<br />

that all capacity building projects in non‐English speaking<br />

countries include an intensive ‘English in Context’ training<br />

component at the outset with ongoing training provided through<br />

the life of the project. This approach has been used to great<br />

effect in Vietnam (1).<br />

In‐country training. For a more complete understanding of plant<br />

diseases, it is important for counterparts to follow the full<br />

diagnostic process from field surveys, through isolation and<br />

identification of a pathogen, followed by pathogenicity testing to<br />

complete Koch’s postulates. This requires the availability of the<br />

facilitator or mentor to be in‐country for extended periods of<br />

time. Although this is often difficult, it can be more cost effective<br />

than holding short workshops in Australia and can potentially<br />

reach more counterparts. The presence of a mentor or facilitator<br />

in country allows real‐life examples currently affecting farmers<br />

to be used in training programs. It also forces an<br />

acknowledgement of local diseases, agronomic practices and<br />

laboratory conditions by the facilitator. We also recommend a<br />

focus on interactive teaching, encouraging student participation.<br />

This approach facilitates ‘learning by doing’.<br />

The Australian Youth Ambassador for Development (AYAD)<br />

program is an ideal program for placing skilled young Australians<br />

in developing countries for up to 12 months. We encourage<br />

project leaders to consider including an AYAD in their project<br />

plan and for plant pathology graduates to seek such<br />

assignments.<br />

ACKNOWLEDGEMENTS<br />

We gratefully acknowledge funding from the Australian Centre<br />

for International Agricultural Research and the ATSE Crawford<br />

Fund as well as the contribution of our many colleagues with<br />

whom we have collaborated over the last 15 years, both in<br />

Australia and in developing countries.<br />

REFERENCES<br />

1. Burgess, LW, Burgess, JS (2009) Capacity building in plant<br />

pathology: Soil‐borne diseases in Vietnam, 1993–2009. <strong>Australasian</strong><br />

<strong>Plant</strong> <strong>Pathology</strong>: In press.<br />

Suitability of Australian postgraduate courses. Many<br />

counterparts aim to obtain scholarships to undertake<br />

postgraduate training in Australia in association with capacity<br />

building projects. Modern Australian plant pathology courses are<br />

becoming increasingly focused on molecular biology, with a<br />

decline in classical diagnostic training. However, in developing<br />

countries the greatest need is for field and laboratory training in<br />

basic plant pathology skills. Providing students from developing<br />

countries with advanced training in molecular techniques is<br />

illogical and ill conceived. Identification of a pathogen to species<br />

level or below is often not necessary for effective integrated<br />

disease management practices to be implemented. Supervisors<br />

need to consider the suitability of coursework undertaken by<br />

students more carefully as well as their ability to continue<br />

mentoring the student on their return.<br />

120 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Increasing global regulations on fumigants stimulates new era for plant protection<br />

and biosecurity<br />

INTRODUCTION<br />

I. Porter{ XE "Porter, I." } A , S. Mattner A , J. Edwards A and P. Fraser B<br />

A Department of Primary Industries, Knoxfield Centre, PMB 15, Ferntree Gully Delivery Centre, 3156, Vic<br />

B CSIRO Marine and Atmospheric Research, Aspendale, 3195,Vic<br />

Restrictions on fumigant chemicals due to environmental<br />

concerns (eg. Montreal Protocol) and tropospheric pollution (eg.<br />

new volatile organic compound regulations) has caused a surge<br />

in new strategies to control soilborne pathogens, but are<br />

sustainable practices being adopted? Since the early 1990’s<br />

when methyl bromide (MB) was shown to be responsible for<br />

ozone layer degradation, a massive global research effort has<br />

been undertaken to find alternative disinfestation strategies. It<br />

was anticipated that growers would readily adopt practices which<br />

conserved biodiversity and the principles of biological equilibrium. In<br />

reality this has not happened because pathologists have not yet<br />

developed a sufficient understanding of the relationships between<br />

soil biology and plant yield or the mechanisms of disease<br />

suppression in the absence of synthetic chemicals.<br />

After MB phase out, many sectors still use other fumigant<br />

chemicals (Fig 1), as other technologies have not always<br />

provided the same advantages afforded by soil fumigation, (i.e. a<br />

high level of pest and disease control with low risk, and good<br />

quality and high yielding crops). However, a bigger set of factors<br />

is now influencing crop production and crop protection; climate<br />

change, user and bystander safety, increasing prices of water, oil<br />

and inorganic fertilisers and concern over soil health is finally<br />

being recognised. This is causing a shift in grower approaches to<br />

soilborne disease control in crop protection.<br />

% Response relative to MBr-Pic (67:33)<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

92.3893<br />

95.78335 0<br />

TC35MNa<br />

PicFosDev<br />

MB67<br />

MI60<br />

TC35EC<br />

Strawberry Estimates with LSIs by Chemical Applied<br />

(Treatments with greater than five observations)<br />

MB50<br />

TC35<br />

PicEC<br />

TC35ECMNa<br />

PicECMNaFos<br />

Pic<br />

PicMNa<br />

8 6 123 18 51 24 63 28 21 9 62 13 13 13 10 14 30 2 3 17 77 87<br />

MB30<br />

Figure 1. The efficacy relative to MB/Pic (67:33) of a large number of<br />

crop protection practices on yield of strawberry plants in a metaanalysis<br />

of over 100 international studies (1). Elipse ‐ MB/Pic standard<br />

Also, whilst new strategies are being developed for sustainable<br />

control of soilborne pathogens very little attention is being given<br />

to products or strategies which eradicate soilborne pathogens<br />

and protect industries from incursions from exotic pathogens. In<br />

fact, there is a poor knowledge on the ability of any strategies to<br />

effectively eradicate soilborne microorganisms. Techniques, such<br />

as solarisation and steaming whilst effective on a limited scale<br />

are not totally effective in the field, biofumigants decrease risk<br />

but do not eradicate propagules and chemical fumigants do not<br />

guarantee total eradication. Development of soilless systems<br />

which exclude soilborne pathogens are a key future practice to<br />

eliminate disease, and use of grafting and resistant varieties can<br />

reduce effects of disease (Table 1).<br />

PicECMNa<br />

SolFert<br />

Treatment Group and Number of Trials<br />

MNaSol<br />

Daz<br />

BioFum<br />

Compost<br />

Sol<br />

MNa<br />

NoTr<br />

Table 1. Some of the future technologies that will be relied upon to<br />

replace control of soilborne plant pathogens by chemical fumigants C: ‐<br />

commodity treatments<br />

Present Disinfestants<br />

Telone C35 EC<br />

Chloropicrin EC<br />

Methyl iodide<br />

Dimethyl disulphide<br />

Solarisation<br />

Streaming<br />

C:Sulphuryl fluoride<br />

IPM Strategies to replace disinfestants<br />

Grafting and plant resistance<br />

Biorationals (AG3, Voom)<br />

Biofumigants<br />

Endophytes?<br />

Biocontrols?<br />

Strategic pesticides and herbicides<br />

Soilless systems: Substrates and hydroponics<br />

C: phosphine/CO 2 MB Recapture and recycling<br />

IMPACT OF FUMIGANT USE FOR BIOSECURITY<br />

Fumigation of imported and exported commodities is a key<br />

activity for quarantine and preshipment especially to satisfy<br />

phytosanitary requirements of the importing country. MB is the<br />

key fumigant used for QPS and is presently exempt from phase<br />

out under the rules of the Montreal Protocol, although the<br />

European Community has decided to phase out all uses by<br />

March 2010. Although MB is predominantly used to eradicate<br />

insect pests from commodities, it may not have any significant<br />

control of fungal pathogens, however this is not clearly<br />

understood. A key alternative for QPS commodity treatments,<br />

sulphuryl fluoride, also has concerns for use as it has a very high<br />

global warming potential (GWP ~ 4000). Therefore the key to<br />

successful QPS treatments for Australia would be to maintain a<br />

systems approach which only allows imports from regions where<br />

the diseases (eg. Phytophthora ramorum, Guava rust, etc.) are<br />

not known to occur.<br />

The Australian and international scientific community require<br />

answers to some key questions to minimise the impact of<br />

changes to availability of fumigants. For instance, in the future,<br />

will we have strategies to eradicate soilborne pathogens in the<br />

event of an incursion? Is it worth the investment to try to<br />

eradicate a soilborne organism? Why is MB being retained for<br />

use in nursery industries worldwide where high health is<br />

paramount, when its fungicidal properties may be insufficient?<br />

Has Australia identified the major exotic soilborne (and airborne)<br />

pathogen risks? Is Australia really prepared to cope with an<br />

outbreak of a serious exotic soilborne pathogen?<br />

This paper will further discuss some of the future challenges<br />

facing industries when considering adoption of new<br />

bioprotection, biosecurity or crop management practices.<br />

REFERENCES<br />

1. Porter IJ, Trinder L, Partington D (2006) Validating the performance<br />

of alternatives to methyl bromide for preplant fumigation. Report<br />

of the TEAP, May 2006, United Nations Environment Program,<br />

Nairobi, 91 pp.<br />

(http://www.unep.org/ozone/teap/Reports/index.asp)<br />

Session 8A—Future directions<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 121


Keynote address<br />

A world of possibilities: the importance of international linkages<br />

The issues facing our societies and professions are global in<br />

nature. <strong>Plant</strong> pathology as a discipline is no different since plant<br />

pathogens know no geographic or political boundaries. Through<br />

the efforts of individual scientists, global networks were<br />

established for scientific disciplines and this has been occurring<br />

through the ages. Many of our professional societies even when<br />

national based have international members indicating the<br />

importance of international networks. The American<br />

Phytopathological <strong>Society</strong> began their efforts in international<br />

programs in the 1940s and refined their goals in 1983 and again<br />

in 2005. However these efforts did not result in international<br />

linkages that bring together scientific societies. For there to be<br />

linkages across societies there must be assurance of each society<br />

being successful and that ventures undertaken across societies<br />

must complement each other’s strengths and fill in potential<br />

weaknesses. Linkages must seek common goals and interests.<br />

Possible items that could generate linkages across our societies<br />

will be explored. In order to move forward in the 21st century, it<br />

is time to embrace networks of alliances to advance the<br />

knowledgebase and this requires international linkages.<br />

Barbara J. Christ{ XE "Christ, B." }<br />

The Pennsylvania State University, University Park, PA USA<br />

122 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Poster abstracts<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 123


Posters<br />

List of posters<br />

Poster<br />

no. Theme Title Page<br />

1 Disease and epidemiology Identification and characterisation of phytoplasma pathogen associated with alfalfa diseases in Al<br />

Hasa, Saudi Arabia<br />

Khalid Alhudaib and Y. Arocha<br />

2 Disease and epidemiology A Phytophthora sp. is the cause of jackfruit decline in the philippines<br />

L.M. Borines, R. Danieland D. Guest<br />

3 Disease and epidemiology The effects of calcium chloride and calcium carbonate on germination and growth of Colletotrichum<br />

acutatum and Penicillium expansum<br />

K.S.H. Boyd‐Wilson and M. Walter<br />

4 Disease and epidemiology A sensitive PCR test for detecting the potato cyst nematode (Globodera rostochiensis) in large volume<br />

soil samples<br />

S.J. Collins, X.H. Zhang, G.I. Dwyer, J.M. Marshall and V.A.Vanstone<br />

5 Disease and epidemiology Management strategies to economically control blackspot and maximise yield in new improved field<br />

pea cultivars<br />

J.A. Davidson, L. McMurray and M. Lines<br />

6 Disease and epidemiology Bacterial canker of tomato: Australian diversity of Clavibacter michiganensis subsp michiganensis<br />

L.M. Forsyth, T. Crowe, A. Deutscher and L. Tesoriero<br />

7 Disease and epidemiology A new report on Pseudomonas syringae pv. mori causal agent of bacterial blight of mulberries in<br />

Australia<br />

H. Golzar and P. Mather<br />

8 Disease and epidemiology Efficacy of pre‐seeding fungicides for control of barley loose smut<br />

K.W. Jayasena, G. Thomas, W. J. MacLeod, K. Tanaka and R. Loughman<br />

9 Disease and epidemiology Specific genetic fingerprinting of Pseudomonas syringae pv. syringae strains from stone fruits in Iran<br />

with REP sequence and PCR<br />

S. Ketabchi<br />

10 Disease and epidemiology Can additional isolates of the Noogoora burr rust fungus be sourced to enhance biocontrol in<br />

northern Australia?<br />

L. Morin, M. Piper, R. Segura and D. Gomez<br />

11 Disease and epidemiology Boneseed rust: a highly promising candidate for biological control<br />

L. Morin and A.R. Wood<br />

12 Disease and epidemiology Helicotylenchus nematode contributing to turf decline in Australia<br />

L. Nambiar, J M. Nobbsand M. Quader<br />

13 Disease and epidemiology Biological control of Uncinula necator by mycophagous mites<br />

N. Panjehkeh, S. Ramroodi andA.R. Arjmandinezhad<br />

14 Disease and epidemiology In vitro screening of potential antagonists of Xanthomonas translucens infecting pistachio<br />

A. Salowi, D. Giblot Ducray and E.S. Scott<br />

15 Disease and epidemiology Characterisation of the causal agent of pistachio dieback as a new pathovar of Xanthomonas<br />

translucens, x. Translucens pv. pistaciae pv. nov.<br />

D. Giblot Ducray, A. Marefat, N.M. Parkinson, J.P. Bowman, K. Ophel‐Keller, E.S. Scott<br />

16 Disease and epidemiology Investigation of the effect of three essential oils, alone and in combination, on the in vitro growth of<br />

Botrytis cinerea<br />

S.M. Stewart‐Wade<br />

17 Disease and epidemiology Survival of the pistachio dieback bacterium in buried wood<br />

T.A. Vu Thanh, D. Giblot‐Ducray, M.R. Sosnowski and E.S. Scott<br />

18 Disease and epidemiology Discovery of a Ceratocystis sp. associated with wilt disease of two native leguminous tree hosts in<br />

Oman and Pakistan<br />

A.O. Al Adawi, I. Barnes, A.A. Al Jahwari, M.L. Deadman, B.D. Wingfield and M.J. Wingfield<br />

19 Disease and epidemiology In vitro study on the effect of NanoSilver (Nanosid) on Sclerotinia sclerotiorum fungi the causal agent<br />

of rapeseed white stem rot<br />

A. Zaman Mirabadi, K. Rahnama, R. Mehdi Alamdarlou and A. Esmaailifar<br />

20 Disease and epidemiology Study on the effect of number of spraying with fungicides on rapeseed sclerotinia stem rot control<br />

R. Mehdi Alamdarlou, A. Zaman Mirabadi, A. Esmaailifar and K. Foroozan<br />

21 Disease management First report of Leveillula taurica on Ficus carica (matrix nova)<br />

Javad Abkhoo and Alireza Arjmandi Nezhad<br />

22 Disease management Pulse virus surveys from Victoria and South Australa in 2007<br />

M. Aftab, A. Freeman and J. Davidson<br />

23 Disease management Mango sudden death syndrome assessment in various mango growing districts of Punjab, Pakistan<br />

F.S. Fateh, M.R. Kazmi, C.N. Akem, A. Iqbal and G. Bhar<br />

134<br />

137<br />

138<br />

141<br />

143<br />

149<br />

156<br />

165<br />

168<br />

183<br />

182<br />

184<br />

193<br />

199<br />

202<br />

210<br />

218<br />

220<br />

224<br />

225<br />

128<br />

131<br />

133<br />

124 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Poster<br />

no. Theme Title Page<br />

24 Disease management Integrated management of mango diseases using inoculum reduction strategies with fungicide spray<br />

treatments<br />

C.N. Akem, G. MacManus, P. Boccalatte, K. Stockdale, D. Lakhesar and R. Holmes<br />

25 Disease management Effect of avocado crop load on postharvest anthracnose and stem end rot, and cations and phenolic<br />

acid levels in peel<br />

E.K. Dann,L.M. Coates, J.R. Dean, L.A. Smith, A.W. Cooke and K.G. Pegg<br />

26 Disease management In vitro fungicide sensitivity of Botryosphaeriaceae species associated with ‘bot canker’ of grapevine<br />

R. Huang, W.M. Pitt, C.C. Steel and S. Savocchia<br />

27 Disease management The value of combined use of genetic resistance and fungicide application for management of stripe<br />

rust<br />

K.W. Jayasena, G. Thomas, R. Loughman, K. Tanaka and W. J. MacLeod<br />

28 Disease management Molecular identification of Pythium isolates of ginger from Fiji and Australia<br />

M.F. Lomavatu, J. Conroy and E. Aitken<br />

29 Disease management Development of techniques to measure SAR induction in broccoli for clubroot disease resistance<br />

D. Lovelock, A. Agarwal, E.C. Donald, I.J. Porter, R. Faggian and D.M. Cahill<br />

30 Disease management Incorporating host‐plant resistance to Fusarium crown rot into bread wheat<br />

D.J. Herde and C.D. Malligan<br />

31 Disease management Management and distribution of huanglongbing in Pakistan<br />

Shahid Nadeem Chohan, Obaid Aftab, Raheel Qamar, Shazia Mannan, Muhammad Ibrahim, Iftikhar<br />

Ahmed, M. Kausar Nawaz Shah, Paul Holford, G. Andrew C. Beattie<br />

32 Disease management Investigating the potential of in‐field starch accumulation tests for targeted citrus pathogen<br />

surveillance in Australia<br />

A.K. Miles, N. Donovan, P. Holford, R. Davis, K. Grice, M. Smith and A. Drenth<br />

33 Disease management Aerial photography—a tool to monitor Mallee onion stunt<br />

S.J. Pederick, J.W. Heap, T.J. Wicks, S. Anstis and G.E. Walker<br />

34 Disease management Diatrypaceae species associated with grapevines and other hosts in New South Wales<br />

W.M. Pitt, R. Huang, C.C. Steel and S. Savocchia<br />

35 Disease management First report of a eucalypt yellowing disease in Syria and its similarity to Mundulla yellows<br />

D. Hanold, B. Kawas and J.W. Randles<br />

36 Disease management New host records for ‘Candidatus Phytoplasma aurantifolia’ in Australia<br />

J.D. Ray<br />

37 Disease management A comparative study of methods for screening chickpea and wheat for resistance to root‐lesion<br />

nematode Pratylenchus thornei<br />

R.A. Reen, J.P. Thompson<br />

38 Disease management Interactions between Leptosphaeria maculans and fungi associated with canola stubble<br />

B. Naseri, J.A. Davidson and E.S. Scott<br />

39 Disease management Seed‐borne concerns with wheat streak mosaic virus in 2008<br />

S. Simpfendorfer and D. Nehl<br />

40 Disease management A single plant test for resistance to two species of root‐lesion nematodes and yellow spot in wheat<br />

J.P. Thompson, T.G. Clewett, J.G. Sheedy, S.H. Jones and P.M. Williamson<br />

41 Disease management Sources of resistance to root‐lesion nematode (Pratylenchus thornei) in wheat from West Asia and<br />

North Africa<br />

J.P. Thompson, T.G. Clewett and M.M. O’Reilly<br />

42 Disease management A single plant test for resistance in wheat to crown rot and root‐lesion nematode (Pratylenchus<br />

thornei)<br />

J.P. Thompson and R.B. McNamara<br />

43 Disease management Effect of irrigation method on disease development in a carrot seed crop<br />

R.S. Trivedi, J.M. Townshend, M.V. Jaspers, H.J. Ridgway, and J.G. Hampton<br />

44 Disease management First report of rapeseed blackleg caused by pathogenicity group T (PGT) of Leptosphaeria maculans in<br />

Mazandaran province of Iran<br />

A. Zaman Mirabadi, K. Rahnama, R.M. Alamdalou and A. Esmaailifar<br />

45 Host‐parasite interactions Fungal endophytes of the Boab species Adansonia gregorri and other native tree species<br />

M.L. Sakalidis, G.E.StJ. Hardy and T.I. Burgess<br />

46 Host‐parasite interactions Two new books: Diseases of Fruit Crops in Australia and Diseases of Vegetable Crops in Australia<br />

Tony Cooke, Denis Persley and Susan House, Cherie Gambley<br />

47 Host‐parasite interactions Through chain assessment and integrated management of brown rot risks in stonefruit<br />

R. Holmes, O. Villalta, S. Kreidl, M. Hossain and C. Gouk<br />

48 Host‐parasite interactions Effects of temperature on mixed bunch rot infections of grapes<br />

L.A. Greer, S. Savocchia, S. Samuelian and C.C. Steel<br />

132<br />

145<br />

163<br />

167<br />

172<br />

173<br />

175<br />

176<br />

180<br />

189<br />

192<br />

194<br />

195<br />

196<br />

203<br />

205<br />

212<br />

214<br />

211<br />

216<br />

223<br />

139<br />

151<br />

157<br />

158<br />

Posters<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 125


Posters<br />

Poster<br />

no. Theme Title Page<br />

49 Host‐parasite interactions Development of nationally endorsed diagnostic protocols for plant pests<br />

B.H. Hall, J. Moran, J. Plazinski, M. Whattam, S. Perry, V. Herrera, P. Gray, D. Hailstones, M. Glen, J. La<br />

Salle<br />

50 Host‐parasite interactions The impact and diversity of Mycosphaerella leaf disease isolated from Eucalyptus globulus in western<br />

australia<br />

S.L. Jackson, A. Maxwell, S.L. Collins, M.C. Calver, G.E.StJ. Hardy and B. Dell<br />

51 Host‐parasite interactions Impact of Phytophthora cinnamomi on native vegetation in South Australia<br />

S.F. McKay, K.H. Kueh, A.J. Able, R.M.A. Velzeboer, J.M. Facelli and E.S. Scott<br />

52 Host‐parasite interactions Reducing the carbon footprint in Riverland vineyards: assessing the efficacy and efficiency of control<br />

for powdery mildew by evaluating growers’ spray diaries<br />

P.A. Magarey, R.W. Emmett, T.S. Smythe, M.M. Moyer, J.R. Dixon, A.J. Pietsch and P.M. Burne<br />

53 Host‐parasite interactions The inhibitory effect of sumac stem extract on some fungal plant pathogens<br />

N. Panjehkeh, M. Abdolmaleki, M. Salari, S. Bahraminejad<br />

54 Host‐parasite interactions Evaluation of commercial cultivars for control of white blister rust in Brassica rapa and Brassica<br />

oleracea vegetables<br />

J.E. Petkowski, F. Thomson, E.J.Minchinton and C. Akem<br />

55 Host‐parasite interactions Fertilisation with N, P and K above critical values required for adequate plant growth influences plant<br />

establishment of cotton varieties in fusarium infested soil<br />

L.J. Smith and J.K. Lehane<br />

56 Host‐parasite interactions Eradication of Elsinoe ampelina by burning infected grapevine material<br />

M.R. Sosnowski, R.W. Emmett, T.A. Vu Thanh, T.J. Wicks and E.S. Scott<br />

57 Pathogens and diagnostics New records of Erysiphaceae (Ascomycota: Erysiphales) for Iran mycoflora<br />

Javad Abkhoo and Alireza Arjmandi Nezhad<br />

58 Pathogens and diagnostics Survey of propinquity among Erysiphe, Leveillula, Phyllactinia, Podosphaera, Sphaerothca, Uncinula<br />

and Uncinuliella based on analysis of morphological characters<br />

Javad Abkhoo and Alireza Arjmandi Nezhad<br />

59 Pathogens and diagnostics Efficient transformation of Colletotrichum capsici, the causal agent of chilli pepper anthracnose by<br />

Agrobacterium<br />

A.S.M. Auyong, R. Ford and P.W.J. Taylor<br />

60 Pathogens and diagnostics Infection process of endophytic Colletotrichum gloeosporioides on cacao leaves<br />

C. Blomley, E.C.Y Liew and D.I. Guest<br />

61 Pathogens and diagnostics An emerging nematode pest on bananas?<br />

J.A. Cobon, and T. Pattison<br />

62 Pathogens and diagnostics Phellinus noxius: brown root rot is increasing in importance in the Australian avocado industry<br />

E.K. Dann, L.A. Smith, M.L. Grose, G.S. Pegg and K.G. Pegg<br />

63 Pathogens and diagnostics Dispersal potential of Gibberella zeae ascospores<br />

P.A.B. Davies, L.W. Burgess, R. Trethowan, R. Tokachichu, D. Guest<br />

64 Pathogens and diagnostics Hosts of citrus scab, brown spot and black spot in coastal NSW<br />

N.J. Donovan, P. Barkley and S. Hardy<br />

65 Pathogens and diagnostics Nitrogen form affects Spongospora subterranea infection of potato roots<br />

Richard E. Falloon, Denis Curtin, Ros A. Lister, Ruth C. Butler, Catherine L. Scott and Nigel S. Crump<br />

66 Pathogens and diagnostics Relationships between Spongospora subterranea DNA in field soil and powdery scab in harvested<br />

potatoes<br />

Farhat A. Shah, Richard E. Falloon, Ros A. Lister, Ruth C. Butler, Alan McKay, Kathy Ophel‐Keller and<br />

Ikram Khan<br />

67 Pathogens and diagnostics Detection of Mycosphaerella fijiensis in the skin of ‘Cavendish’ banana<br />

R.A. Fullerton and S.G. Casonato<br />

68 Pathogens and diagnostics Rapid and robust identification of fungi associated with Acacia mangium root disease using DNA<br />

analyses<br />

M. Glen, V. Yuskianti, A. Francis, L. Agustini, A. Widyatmoko, A. Rimbawanto and C.L. Mohammed<br />

69 Pathogens and diagnostics Survey of the needle fungi associated with Spring Needle Cast in Pinus radiata<br />

I. Prihatini, M. Glen, A.H. Smith, T.J. Wardlaw and C.L. Mohammed<br />

70 Pathogens and diagnostics Disease‐management strategies for the rural sector that help deliver sustainable wood production<br />

from exotic plantations<br />

C. Beadle, A. Rimbawanto, A. Francis, M. Glen, D. Page, C.L. Mohammed<br />

71 Pathogens and diagnostics Infection and host responses in interactions between melon and Colletotrichum lagenarium<br />

Yonghong Ge and David Guest<br />

72 Pathogens and diagnostics Genetic diversity of Iranian Fusarium oxysporum f. sp. ciceris by RAPD molecular markers<br />

Sara Haghighi, Saeed Rezaee, Bahar Morid, Shahab Hajmansoor<br />

73 Pathogens and diagnostics The cause of the barley leaf rust in Western Australia is a typical Puccinia hordei<br />

Y. Anikster, K.W. Jayasena, T. Eilamand J. Manisterski<br />

160<br />

164<br />

170<br />

174<br />

188<br />

191<br />

207<br />

208<br />

129<br />

130<br />

135<br />

136<br />

140<br />

142<br />

144<br />

146<br />

147<br />

148<br />

150<br />

154<br />

155<br />

153<br />

152<br />

159<br />

166<br />

126 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


Poster<br />

no. Theme Title Page<br />

74 Pathogens and diagnostics Genetic diversity and population structure of Australian and South African Pyrenophora teres isolates<br />

A. Lehmensiek, R. Prins, G. Platz, W. Kriel, G.F. Potgieter and M.W. Sutherland<br />

75 Pathogens and diagnostics The Fusarium oxysporum f. sp cubense tropical race 4 vectoring ability of the banana weevil borer<br />

(Cosmopolites sordidus)<br />

R.A. Meldrum, A.M. Daly and L.T.T. Tran‐Nguyen<br />

76 Pathogens and diagnostics Genetic diversity of Pseudocercospora macadamiae populations by PCR‐RFLP<br />

A.K. Miles, O.A. Akinsanmi, E.A. Aitken and A. Drenth<br />

77 Pathogens and diagnostics Priming for resistance against pathogens: cellular responses of Arabidopsis to UV‐C radiation<br />

S.J.L. Mintoff, P.T. Kay and D.M. Cahill<br />

78 Pathogens and diagnostics The effect of phosphonate on the accumulation of camalexin following challenge of Arabidopsis by<br />

Phytophthora palmivora<br />

Zoe‐Joy Newby, Rosalie Daniel and David Guest<br />

79 Pathogens and diagnostics Characterisation of Phytophthora capsici isolates from black pepper in Vietnam<br />

N.V. Truong, L.W. Burgess and E.C.Y. Liew<br />

80 Pathogens and diagnostics Characterisation of Phytophthora capsici Isolates from chilli in Vietnam<br />

N.V. Truong, L.W. Burgess and E.C.Y. Liew<br />

81 Pathogens and diagnostics Survey of viruses infecting sweet potato crops in New Zealand<br />

Z.C. Perez‐Egusquiza, L.I. Ward, J.D. Fletcher and G.R.G. Clover<br />

82 Pathogens and diagnostics Multi‐locus sequence typing of isolates of Pseudomonas syringae pv. actinidiae, a biosecurity risk<br />

pathogen<br />

J. Rees‐George, I.P.S. Pushparajah and K.R. Everett<br />

83 Pathogens and diagnostics Uniform distribution of powdery mildew conidia using an improved spore settling tower<br />

Z. Sapak, V. Galea, D. Joyce and E. Minchinton<br />

84 Pathogens and diagnostics The effect of high nutrient loads on disease severity due to Phytophthora cinnamomi in urban<br />

bushland<br />

Kelly Scarlett, Zoe‐Joy Newby, David Guest and Rosalie Daniel<br />

85 Pathogens and diagnostics Non‐host resistance and pathogen virulence: an important role of toxic and infection‐inducing<br />

compound(s) from spore germination fluid of Botrytis cinerea<br />

N.N. Khanam, K. Toyoda, H. Yoshioka, Y. Narusaka and T. Shiraishi<br />

86 Pathogens and diagnostics First report of tomato yellow leaf curl virus in pepper (Capsicum annum) fields in Iran<br />

M. Shirazi, J. Mozafari, F. Rakhshandehrooand M. Shams‐Bakhsh<br />

87 Pathogens and diagnostics Effect of white rust infection, bion and phosphonate on glucosinolates in brassica crops<br />

Astha Singh, Les Copeland and David Guest<br />

88 Pathogens and diagnostics Recent plant virus incursions into Australia<br />

J.E. Thomas, V. Steele, A.D.W. Geering, D.M. Persley, C.F. Gambley, and B.H. Hall<br />

89 Pathogens and diagnostics Sugarcane downy mildew: development of molecular diagnostics<br />

N. Thompson and B.J. Croft<br />

90 Pathogens and diagnostics Role of nematodes and zoosporic fungi in poor growth of winter cereals in the northern grain region<br />

J.P. Thompson, T.G. Clewett J.G. Sheedyand K.J. Owen<br />

91 Pathogens and diagnostics Pathogenicity of Radopholus similis on ginger in Fiji<br />

U. Turaganivalu, G.R. Stirling, S. Reddy and M. K. Smith<br />

92 Pathogens and diagnostics The effect of dryland salinity on the diversity of arbuscular mycorrhizal fungi<br />

B.A. Wilson, G.J. Ash and J.D.I. Harper<br />

93 Pathogens and diagnostics Evaluation of plant extracts for control of sclerotinia pathogens of vegetable crops<br />

D. Wite, O. Villaltaand I.J. Porter<br />

94 Pathogens and diagnostics First report of Macrophominia phaseolina on rapeseed stem in some provinces of Iran<br />

A. Zaman Mirabadi, A. Esmaailifar, A. Alian and R. M. Alamdalou<br />

95 Pathogens and diagnostics DNA barcoding to support biosecurity decisions<br />

K. Pan, G.F. Bills, M.K. Romberg, W.H. Ho, and B.J.R. Alexander<br />

96 Pathogens and diagnostics Subcommittee on <strong>Plant</strong> Health Diagnostic Standards<br />

M.A. Williams, B.H. Hall, J. Plazinski, P. Gray, J. Moran, P. Stephens, S. Perry<br />

97 Pathogens and diagnostics What is laboratory accreditation and what will it mean for me and my laboratory?<br />

M.A. Williams, P. Gray, N. Kelly, J. Cunnington, P. Stephens, R. Makin and S. Peterson, B.H. Hall<br />

98 The biology and management of chestnut rot in south‐eastern Australia<br />

L. Shuttleworth, D. Guest, E. Liew<br />

171<br />

178<br />

179<br />

181<br />

185<br />

186<br />

187<br />

190<br />

197<br />

200<br />

201<br />

169<br />

204<br />

206<br />

209<br />

215<br />

213<br />

217<br />

219<br />

221<br />

222<br />

198<br />

161<br />

162<br />

227<br />

Posters<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 127


Posters<br />

21 First report of Leveillula taurica on Ficus carica (matrix nova)<br />

Javad Abkhoo{ XE "Abkhoo, J." } A and Alireza Arjmandi Nezhad B<br />

A,B Department of <strong>Plant</strong> Protection Research, Agriculture and Natural Resoureces Research Center of Sistan, Iran‐Zabol<br />

INTRODUCTION<br />

Leveillula taurica Lev. Belong to Erysiphaceae (Erysiphales,<br />

Ascomycota) that causing powdery mildew on over from 50<br />

plant families(1).<br />

MATERIALS AND METHODS<br />

During the summer 2007, typical symptoms of powdery mildew<br />

were observed in several fig fields assessed in Kereman Province,<br />

Iran. Samples were stained with Lactofushin(3) and<br />

morphological characteristics of fungus investigated by Olympus<br />

microscope (Modle: BH2) and drawn by drawing tube connected<br />

on microscope.<br />

RESULTS AND DISCUSSION<br />

Morphological characters of this fungus on ficus carica is as<br />

follow:<br />

Diseased leaves displayed typical powdery mildew signs<br />

consisting of whitish masses of conidia and conidiophores.<br />

Mycelial growth was thick, forming irregulare white patches,<br />

sometimes effused to cover the whole leave surface and usually<br />

not present appressoria. Conidiofores erect, foot cells cylindric,<br />

40–126 (‐148) × 4/5–7/8 µm usually followed by (1‐) 2–3 (‐4)<br />

shorter and different length cells. Conidia formed singly, primery<br />

conidia lancaeolate, 31–67 (80) × 12–20 µm and secondary<br />

conidia ellipsoid to cylandric, 33–76 × 13–22 µm (fig 1).<br />

Ascomata found on leaves as embedded in the mycelial felt,<br />

were gregarious to scattered and measured 145–250 µm in<br />

diameter. Appendages were myceliod, arising from the lower<br />

half of ascomata, brown, paler upward. Asci 20–30 (‐45) in each<br />

cleistothecia, clvate, stalked, 77–120 (‐135) × 25–42 µm.<br />

Ascospores (1‐) 2 (‐4) in each ascus, ellipsoid‐ovoid shaped, (20‐)<br />

25–40 (‐45) × 15–22 µm (fig 2).<br />

Figure 2. Leveillula taurica (telemorph). (A) asci, (B) ascospores.<br />

The fungus caused significant losses, estimated at more than<br />

35% of production, at the location studied by making infected<br />

plants unsuitable for use in propagation.<br />

REFERENCES<br />

1. Braun U (1987) A monograph of the Erysiphales (powdery<br />

mildews). Beih. Nova Hedwigia 89:1–700.<br />

2. Braun U (1995) The Powdery Mildews (Erysiphales) of Europe. Jena,<br />

Germany: VEB G. Fischer Verlag.<br />

3. Carmichael, J W (1955) Lacto‐fuchshin: A new medium for<br />

mounting fungi. Mycologia 47: 611–619.<br />

On the basis of morphological characters of the anamorph and<br />

telemorph, this fungus was identified as Leveillula taurica (1).<br />

This is also the first report of genus Leveillula on Morceae in<br />

world and Morceae is a new host family for Leveillula<br />

taurica(1,2).<br />

Figure 1. Leveillula taurica (anamorph). (A) conidiophores, (B) primery<br />

conidia, (C) secondary conidia, (D) germinated conidium.<br />

128 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


57 New records of Erysiphaceae (Ascomycota: Erysiphales) for Iran mycoflora<br />

Javad Abkhoo{ XE "Abkhoo, J." } A and Alireza Arjmandi Nezhad B<br />

A,B Department of <strong>Plant</strong> Protection Research, Agriculture and Natural Resoureces Research Center of Sistan, Iran‐Zabol<br />

Posters<br />

INTRODUCTION<br />

Powdery mildew fungi belong to the family Erysiphaceae<br />

(Ascomycta: Erysiphales) and infect a wide range of<br />

angiosperms. This family consists of 18 genera and about 435<br />

species(1). Ershad (4) has provided the best list of powdery<br />

mildews and their hosts in Iran.<br />

MATERIALS AND METHODS<br />

During the years 2007 and 2008 Surveys were carried out to<br />

determine species composition and host range of powdery<br />

mildews(Erysiphaceae) in Sistan region, Iran. Samples were<br />

stained with Lactofushin(3) and morphological characteristics of<br />

fungus investigated by Olympus microscope(Modle: BH2) and<br />

drawn by drawing tube connected on microscope. Observation<br />

of conidial germ tubes was carried out using the method of<br />

Hirata (5). Identification of species were carried out by reliable<br />

references(1,2).<br />

RESULTS AND DISCUSSION<br />

In this study fourteen taxa were identified which according<br />

Ershad(4) Among this taxa, Erysiphe australiana is new to Iran<br />

mycoflora and other thirteen taxa viz. Erysiphe convolvuli, E.<br />

cruciferarum, E. lycopsidis, E. necator, E. polygoni,<br />

Golovinomyces cichoraceaerum, G. orontii, Leveillula saxaouli, L.<br />

taurica, Phyllactinia moricola, Podosphaera leucotricha, P.<br />

pannosa and Blumeria graminis have previously been recorded<br />

from Iran.<br />

Morphology of appressoria, conidiophores, conidia as well as the<br />

host species, fit the description for Erysipphe(Uncinuliella)<br />

australiana provided by Braun (1). conidia with multilobed<br />

appressoria and the host species determined that Oidium yeneii<br />

donot occurs on Lagerstromia indica.<br />

So, for your more information also six plant species viz. Alhagi<br />

persarum(a host for L. taurica), Althaea officinalis(a host for L.<br />

taurica), Gaillardia sp.(a host for G. cichoraceaerum), Haloxylon<br />

persicum(a host for L. saxaouli), Malva microcarpa(a host for L.<br />

taurica) and Phalaris paradoxa(a host for B. graminis) are<br />

recorded as a new hosts for Iranian powdery mildew fungi.<br />

REFERENCES<br />

1. Braun U (1987) A monograph of the Erysiphales (powdery<br />

mildews). Beih. Nova Hedwigia 89:1–700.<br />

2. Braun U and Takamatsu S (2000) Phylogeny of Erysiphe,<br />

Microspaera, Uncinula (Erysipheae) and Cystotheca, Podosphaera,<br />

Sphaeroteca (Cystotheceae) inferred from RDNA ITS sequences‐<br />

Some taxonomic consequences. Schlechtendalia 4: 1–33.<br />

3. Carmichael J.W. (1955) Lacto‐fuchshin: A new medium for<br />

mounting fungi. Mycologia 47: 611–619.<br />

4. Ershad D (1995) Fungi of Iran. Pest and diseases research institute,<br />

department of botany, publication no. 10, 874 pp.<br />

5. Hirata K (1942) On the shape of the germ tubes of Erysipheae Bull<br />

Chiba Coll. Hortic. 5: 34–49.<br />

Morphological characters of Erysipphe australiana on<br />

Lagerstromia indica is as follow:<br />

Mycelium on leaves, effused, cover the whole leaves surface,<br />

nonpersistance, appressoria are multilobed. Conidiofores erect,<br />

foot cells cylindric occasionally ruffle, (23‐) 25–36 (‐50) (5‐) 6×‐<br />

9/5 µm followed by (1‐) 2 (‐3) different length cells. Conidia<br />

formed singly, ellipsoid to cylandric, (25‐) 28–43 (‐49) × (11/5‐)<br />

12/48–18 (19/5) µm. This funus didn’t form telemorph(fig 1).<br />

Figure 1. Erysipphe australiana on Lagerstromia indica: A—<br />

conidiophores, B—conidia, C—appressoria<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 129


Posters<br />

58 Survey of propinquity among Erysiphe, Leveillula, Phyllactinia, Podosphaera,<br />

Sphaerothca, Uncinula and Uncinuliella based on analysis of morphological characters<br />

Javad Abkhoo{ XE "Abkhoo, J." } A and Alireza Arjmandi Nezhad B<br />

A,B Department of <strong>Plant</strong> Protection Research, Agriculture and Natural Resoureces Research Center of Sistan, Iran‐Zabol<br />

INTRODUCTION<br />

Powdery mildew fungi belong to the family<br />

Erysiphaceae(Ascomycta: Erysiphales) which cause serious<br />

diseases in a variety of cultivated plants such as cereals,<br />

vegetables and ornamental plants. This family consists of 18<br />

genera and about 435 species(1).<br />

Phylogenetic relationships among the genera of powdery<br />

mildews have been proposed by some authors(1, 2, 3).<br />

This research is carried out in order to investigate propinquity<br />

and affinity among seven genera of powdery mildews based on<br />

morphological data.<br />

MATERIALS AND METHODS<br />

We used 18 morphological characters. Each character contains 2<br />

to 6 status. Varies status of characters take numbers 1 to 6. The<br />

data were analyzed using the Distance method by PAUP v.4.0b4a<br />

(4). Neighbour‐Joining(NJ) tree was obtained. The strength of the<br />

internal branches from the resulting trees were tested by<br />

bootstrap analysis (5).<br />

RESULTS AND DISCUSSION<br />

The results showed that all taxa are divided into five groups,<br />

which corresponded well to new mitosporic taxa. Clade 1<br />

consisted of Erysiphe section Erysiphe, Uncinula and Uncinuliella,<br />

all of which have single conidia, lobad apprasorium an Oidium<br />

subgenus Pseudoidium mitosporic state. Clade 2 consisted of E.<br />

galeopsidis, E. cichoracearum and E. orontii, which have without<br />

fibrosin badies catanate cinidia and Oidium subgenera<br />

Striatoidium and Reticuloidium mitosporic states. Clade 3<br />

consisted of of Leveillula and Phyllactinia, which have<br />

endophytic mycelia and same suface patterns of conidia and<br />

Oidiopsis and Ovulariopsis mitosporic states, respectively. Clade<br />

4 consisted of Podosphaera and Sphaerotheca, which have<br />

fibrosin badies catanate cinidia and singl ascual ascocarpes an<br />

Oidium subgenus Fibroidium mitosporic state. Clade 5 consisted<br />

of Blumeria graminis, which has digitat haustoria, bollbous<br />

sewellig of foot cell, unique suface patterns of conidia an Oidium<br />

subgenus Oidium mitosporic state. This resuts coincides with<br />

molocular analyses(1,3)<br />

Figure 1. Neighbour‐Joining tree inferred from morphological data.<br />

Branch support was determined by 1000 bootstrap replication, shown<br />

above the branches. Bootstrap values below 50% are not shown.<br />

REFERENCES<br />

1. Blumer S (1933) Die Erysiphaceen Mitteleuropas mit Besonderer<br />

Berucksichtigung der Schweiz. Beitr. Kryptogameflora Schweiz, 7:<br />

1–483.<br />

2. Braun U (1987) A monograph of the Erysiphales (powdery<br />

mildews). Beih. Nova Hedwigia 89:1‐ 700.<br />

3. Braun U and Takamatsu S (2000) Phylogeny of Erysiphe,<br />

Microspaera, Uncinula (Erysipheae) and Cystotheca, Podosphaera,<br />

Sphaeroteca (Cystotheceae) inferred from RDNA ITS sequences‐<br />

Some taxonomic consequences. Schlechtendalia 4: 1–33.<br />

4. Felsenstein J (1985) Confidence limits on hylogenies; an approach<br />

using the bootstrap. Evolution 39: 783–791.<br />

5. Swofford DL (2000) PAUP: Phylogenetic analysis using parsimony<br />

and other methods (PAUP Version4). Illinois Natural History Survey,<br />

Champain, Illinois, U. S. A.<br />

130 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


22 Pulse virus surveys from Victoria and South Australa in 2007<br />

M. Aftab{ XE "Aftab, M." } A , A. Freeman A and J. Davidson B<br />

A Department of Primary Industries Victoria Private Bag 260, Horsham VIC 3400<br />

B South Australian Research and Development Institute, GPO Box 397, Adelaide SA 5001<br />

Posters<br />

INTRODUCTION<br />

Surveys of pulse crops in 2007 (field pea, faba bean, lentil, lupin<br />

and chickpea) were undertaken in Victoria and South Australia to<br />

determine the occurrence and incidence of eight pulse viruses:<br />

Alfalfa mosaic virus (AMV), Bean yellow mosaic virus (BYMV),<br />

Cucumber mosaic virus (CMV), Pea seedborne mosaic virus<br />

(PSbMV), Bean leafroll virus (BLRV), Beet western yellows virus<br />

(BWYV), Tomato spotted wilt virus (TSWV) and Subterranean<br />

clover stunt virus (SCSV).<br />

MATERIALS AND METHODS<br />

In Victoria in October 2007, random samples were taken from 45<br />

crops, comprising 13 field pea, eight lentil, 15 faba bean, five<br />

chickpea and four lupin from 31 locations (Ararat, Berriwillock,<br />

Boyeo, Clear lake, Culgoa, Dimboola, Dooen, Douglas,<br />

Gymbowen, Inverleigh, Jung, Kaniva, Kerang, Lake Boga, Lalbert,<br />

Linton, Marnoo, Minimay, Mininera, Minyip, Murtoa, Natimuk,<br />

Nhill, Noradjuha, Rockwood, Rupanyup, Skipton, Treso,<br />

Warracknabeal, Woorinen and Ultima). In South Australia in<br />

October 2007, random samples were taken from 49 crops,<br />

comprising 20 faba bean, 11 field pea, nine lentil, two chickpea,<br />

five lupin and two vetch from 27 locations (Agery, Arthurton,<br />

Blyth, Bordertown, Brecon, Cockle Beach, Coonalpyn, Cummins,<br />

Freeling, Giles Corner, Glen Park, Hart, Karkoo, Keith, Kybunga,<br />

Minlaton, Monta, Mundulla, Owen, Rhynie, Riverton, Rogers<br />

Corner, Tarlee, Saddleworth, Warooka, Willalooka, Yeelanna).<br />

One hundred, randomly selected petioles, tendrils or shoots<br />

were collected from each crop and bundled into groups of ten.<br />

The bundles were blotted onto nitrocellulose membranes and<br />

processed using tissue blot immunoassay. Within‐crop virus<br />

incidence for each crop was estimated from the number of<br />

positive samples.<br />

RESULTS<br />

In Victoria, the most serious virus problems were BWYV in pea,<br />

lentil and chickpea, CMV in lentil and PSbMV in pea (Table 1).<br />

BWYV occurred in 100% of pea and chickpea and 88% of lentil<br />

and 67% of bean crops and the within crop virus incidence was<br />

highest in lentils (up to 61%) and lowest in beans (up to 8%).<br />

CMV occurred in all crop types with low within crop incidence<br />

(


Posters<br />

24 Integrated management of mango diseases using inoculum reduction strategies<br />

with fungicide spray treatments<br />

C.N. Akem{ XE "Akem, C.N." }, G. MacManus, P. Boccalatte, K. Stockdale, D. Lakhesar and R. Holmes<br />

Horticulture and Forestry Science; Queensland Primary Industries and Fisheries, DEEDI, PO Box 15, Ayr, Qld 4807<br />

INTRODUCTION<br />

Anthracnose caused by Colletotrichum gloeosporoides (Penz.)<br />

and stem‐end‐rots caused by Neofusicoccum parvum<br />

(Botryospheria spp and Lasiodiplodia theobromae) are the main<br />

postharvest diseases of mango in all tropical and sub‐tropical<br />

environments where mangoes are grown.<br />

Managing these diseases effectively is the key for producing<br />

quality mango fruits with a long shelf life. The use of pre‐ and<br />

post‐harvest fungicide treatments has been the main mechanism<br />

of trying to achieve this objective (1). There are environmental<br />

and residue concerns on the overuse of these fungicides and<br />

therefore a push to limit their use. To do this, other strategies<br />

need to be integrated with minimal fungicide use to overcome<br />

such concerns and still achieve effective disease control. The<br />

main objective of this study was to determine the effectiveness<br />

of integrating field inoculum reduction strategies with minimal<br />

fungicide sprays in managing mango postharvest rots, especially<br />

anthracnose and stem end.<br />

MATERIALS AND METHODS<br />

The trials were conducted on a uniform block of 18‐year old<br />

Kensington Pride mango trees at Ayr Research Station of the<br />

Queensland Primary Industries and Fisheries, during 2007 and<br />

2008 seasons. The block was divided into two sub‐blocks to<br />

represent partial and optimal inoculum reduction levels. On the<br />

partial reduction sub‐block a one‐time removal of dead twigs,<br />

branches and leaves from the tree canopy was undertaken soon<br />

after mechanical pruning, while on the optimal reduction subblock<br />

dead twigs, branches and leaves were removed from<br />

within and underneath the trees soon after pruning, and were<br />

followed up with monthly repeats of the same exercise.<br />

The following fungicides were applied to the treatment trees in<br />

different combinations at strategic times in each season:<br />

Mankocide (Mc), Mancozeb (Mz), Octave (Oc), Amistar (Am),<br />

Bravo (Br), Aero (Ae) and Tilt (Tt). This was to determine their<br />

integrated effects with the inoculum reduction strategies on<br />

mango postharvest diseases. The treatment trees in each subblock<br />

were arranged in a 6 x 4 RCBD and standard mango<br />

industry tree husbandry practices for irrigation, fertilisation and<br />

insect pest control were implemented.<br />

At harvest, 35 fruits were randomly picked from each treatment<br />

tree from which 25 more uniform ones were selected, desapped,<br />

washed and then placed in boxes and stored in a cool<br />

room at ~20–22 ° C. Fruits were assessed for postharvest rots<br />

disease incidence 14 days after incubation.<br />

RESULTS AND DISCUSSION<br />

All fungicide spray combinations in 2007 and 2008 were<br />

significantly (P=0.05) better than the control in suppressing<br />

postharvest rots incidence on the fruits (Figs 1 and 2). In 2008<br />

there were additional treatment differences within the fungicide<br />

treatment combinations (Fig. 2). Significant differences (P=0.05)<br />

between partial and optimal inoculum reductions on fruit rots<br />

were observed on most treatments in 2008 but not 2007. The<br />

repeat of the inoculum reduction exercise on the same sub‐block<br />

for two seasons significantly reduced the level of inoculumcarrying<br />

dead materials within and underneath the treatment<br />

trees resulting in this accumulated significant effect in 2008.<br />

Fungicide treatment combinations used in both seasons ranged<br />

from a minimum of 3 to a maximum of 7 sprays. This was<br />

significantly less than the current industry practice of up to 12 or<br />

more sprays per season, to achieve the same level of disease<br />

control as compared to the low levels from the optimal inoculum<br />

reduction sub‐block.<br />

These trial results demonstrate the role that basic orchard<br />

hygiene can play in field management of mango postharvest<br />

diseases, especially when integrated with minimal fungicide<br />

spray treatments.<br />

Percentage<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

c c<br />

ab<br />

ab<br />

Ctrl<br />

Mc/Oc/Mz/Am<br />

Mc/Mz/Oc/Mz/Am/Mz/Am<br />

a ab<br />

Mc/Br/Oc/Br/Am/Br/Am<br />

a<br />

a<br />

Treatments<br />

b<br />

b<br />

McTt/Mz/Oc/Mz/Am<br />

b<br />

Mc/Tt/Mz/Tt/Oc/Mz/Am<br />

abc<br />

Partial IR<br />

Optimal IR<br />

{Ctrl = Control, Mc = Mankocide, Oc = Octave, Mz = Mancozeb, Am = Amistar, Br =<br />

Bravo, Tt = Tilt }<br />

Figure 1. Effect of inoculum reduction (IR) and limited fungicide sprays<br />

on the incidence of postharvest rots—2007<br />

Percentage<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

d<br />

Ctrl<br />

e<br />

c<br />

Oc/Mz/Am<br />

d<br />

ab<br />

Oc/Mz/Am/Mz/Am<br />

ab<br />

Oc/Br/Am/Br/Am<br />

ab<br />

bc<br />

Treatments<br />

Oc/Mz/Ae<br />

b<br />

cd<br />

a<br />

Oc/Mz/Ae/Ae<br />

a<br />

Partial IR<br />

Optimal IR<br />

Figure 2. Effect of inoculum reduction (IR) and limited fungicide sprays<br />

on incidence of postharvest rots—2008.<br />

ACKNOWLEDGEMENTS<br />

We gratefully acknowledge HAL, AMIA and DPI&F for funding<br />

this research.<br />

REFERENCES<br />

1. Akem, N. Chrys (2006). Mango Anthracnose Disease: Present Status<br />

and Future Research Priorities. <strong>Plant</strong> Path J. 5(3):266–273.<br />

132 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


23 Mango sudden death syndrome assessment in various mango growing districts of<br />

Punjab, Pakistan<br />

Posters<br />

F.S. Fateh 1 , M.R. Kazmi 1 , C.N. Akem{ XE "Akem, C.N." } 2 , A. Iqbal 1 and G. Bhar 1<br />

1 Nat‐IPM Programme, IPEP, National Agricultural Research Centre, Park Road, Islamabad, Pakistan<br />

2 Queensland Department of Primary Industries and Fisheries, Ayr, Australia<br />

INTRODUCTION<br />

Mango is an important fruit in Pakistan enjoyed by all age groups<br />

in the country. The continuous production of this fruit is being<br />

threatened by a number of diseases. Among these is the mango<br />

sudden death syndrome (MSDS), a complex caused by a number<br />

of biotic and abiotic factors. Botryodiplodia theobromae and<br />

Ceratocystis fimbriata have been identified as the key fungal<br />

pathogens in Pakistan that lead to an increase in the severity of<br />

the disease. Typical symptoms of this disease include drooping<br />

and browning of leaves, bark splitting, gum oozing, and in most<br />

cases these symptoms are followed by the sudden wilting and<br />

death of the tree.<br />

The disease was first detected in Pakistan in 1995 from the<br />

Muzaffar Garh district of Punjab and its effects only became<br />

popularised in 2005 when it was reported to be spreading at<br />

epidemic proportions with serious effects on productivity (1).<br />

There has been a dire need to monitor the prevalence of the<br />

disease and identify factors associated with its rapid spread<br />

through out the mango production districts of Pakistan. The<br />

main objective of this study was to monitor the spread and<br />

distribution of the disease and collect data associated with its<br />

epidemiology as a first step towards the development of<br />

management strategies to stop or slow down its spread.<br />

MATERIALS AND METHODS<br />

A survey was conducted in the following mango growing districts<br />

of Punjab: Faisalabad, Jhang, Khanewal, Multan and Muzaffar<br />

Garh. The number of orchards visited in these districts was 3, 6,<br />

3, 8 and 4 respectively. Sampling was done from the twigs,<br />

branches, bark, stem at the collar regions and roots of the<br />

affected trees. The assessment was made on the basis of a mean<br />

disease severity rating of 0–5 reflecting the percentage of<br />

symptoms such as gummosis, bark splitting and bark beetle<br />

holes observed on the parts sampled, where 0=healthy trees;<br />

1=1–10%; 2=11–20%; 3=21–30%; 4=31–50% and 5= more than<br />

50% diseased area.<br />

The infected samples were cut into small pieces with a sterilised<br />

scissors and disinfected with 10% commercial bleach for one<br />

minute followed by three rinses in distilled water. After drying,<br />

the pieces were aseptically plated on potato dextrose agar<br />

medium and incubated at 25°C. After 7 days of incubation<br />

resulting fungal colonies were microscopically identified based<br />

on spore morphological characteristics. The colonisation<br />

frequency of each sample was also determined using the<br />

following formula:<br />

Colonisation (%) = No. of pieces colonised by a pathogen x 100<br />

Total No. of pieces<br />

RESULTS AND DISCUSSION<br />

The maximum of 27% mean disease incidence was found in<br />

Multan samples followed by 22% in Muzaffar Garh, 18% in Jhang<br />

and 15% in Khanewal. The minimum of 12% disease incidence<br />

was found in Faisalabad. A maximum severity rating of 4 was<br />

also observed on Multan samples followed by a 3 rating for both<br />

Khanewal and Muzaffar Garh and a 2 rating for Jhang samples. A<br />

minimum rating of 2 was recorded on Faisalabad samples (Fig 1).<br />

These results clearly demonstrate that mango sudden death<br />

disease prevails in all the mango growing districts of Punjab<br />

surveyed. The common prevalence of the disease may be<br />

associated with the large number of abandoned orchards that<br />

are receiving little or no management attention. It was also<br />

common to observe adjacent orchards with high levels of<br />

infection, especially in the Multan district, suggesting the ease of<br />

pathogen movement between such orchards. The colonisation<br />

percentage of individual fungi from the different orchards<br />

sampled shows that B. theobromae was the most common<br />

pathogen while C. fimbriata was the least (Fig 2) This survey<br />

results suggest the need for more emphasis on the importance<br />

of orchard sanitation and improved cultural practices to reduce<br />

the prevalence of different fungal pathogens causing sudden<br />

death of mango in Punjab<br />

Mean Disease Incidence (%) and Mean Disease<br />

Severity Rating (0-5)<br />

30<br />

25<br />

4.5<br />

4<br />

3.5<br />

20<br />

3<br />

15<br />

2.5<br />

2<br />

10<br />

1.5<br />

5<br />

1<br />

0.5<br />

0<br />

0<br />

Faisalabad Jhang Khanewal M ultan M uzaffar<br />

Locations<br />

Garh<br />

Figure 1. Mean incidence and severity of sudden death in different<br />

districts of Punjab.<br />

45<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

1 2 3 4 3<br />

Disease severity rating<br />

N<br />

Ceratocystis fimbriata<br />

Botryodilodia<br />

theobromae<br />

Fusarium sp.<br />

Nattrassia mangiferae<br />

Figure 2. Mean disease incidence of different fungi colonising mango<br />

trees with sudden death symptoms in Punjab<br />

ACKNOWLEGEMENTS<br />

Funding for this work was provided by Etiology and Management<br />

of Mango Project and ASLP Mango Production Project.<br />

REFERENCE<br />

Munawar, R.K., F.S. Fateh, K. Majeed, A.M. Kashkhely, I. Hussain, I.<br />

Ahmad and A. Jabeen. 2005. Incidence and etiology of mango<br />

sudden death phenomenon. Pak. J. Phytopathology. 17(2): 154–<br />

158.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 133


Posters<br />

1 Identification and characterisation of phytoplasma pathogen associated with alfalfa<br />

diseases in Al Hasa, Saudi Arabia<br />

Khalid Alhudaib{ XE "Alhudaib, K." } and Y. Arocha<br />

<strong>Plant</strong> Protection Department, King Faisal University PO Box 55009 Alhasa 31982<br />

National Center for Animal and <strong>Plant</strong> Health (CENSA), Havana, Cuba<br />

INTRODUCTION<br />

Alfalfa (Medicago sativa L.) is cultivated as a forage crop in many<br />

countries and is distinguished from other agricultural crops in<br />

having a perennial habit. However, phytoplasmas diseases have<br />

been reported to cause very significant economic losses in<br />

several countries worldwide. In Eastern province Saudi Arabia,<br />

alfalfa is the most important forage crop.<br />

A number of phytoplasma diseases have been reported to be<br />

associated with alfalfa plants that result from drastic reduction<br />

in forage yield. Average annual loss of alfalfa in the Sultanate of<br />

Oman due to witches’‐broom disease is approximately 25% of<br />

green hay, an estimated loss of US$30 million (1).<br />

Recently a phytoplasma of 16SrI, Ca. Phytoplasma asteris, has<br />

been associated with the Al‐Wijam disease of date palm disease,<br />

in Al Hasa, Saudi Arabia, and leafhoppers belonging to<br />

Cicadellidae family were identified as potential vectors (2). This<br />

paper reports results of a PCR‐based phytoplasma survey on<br />

diseased alfalfa grown in Eastern province of Saudi Arabia.<br />

MATERIALS AND METHODS<br />

Field‐collection of plants, leafhoppers Alfalfa (M. sativa L.)<br />

surveys were conducted from April to September 2008 in<br />

Eastern province of Saudi Arabia (Fig. 1). A total of 76 samples<br />

showing typical symptoms of witches’ broom disease. 254<br />

leafhoppers were collected from alfalfa field. Type specimens<br />

were identified at the National Museum of Wakes, Cardiff as<br />

follows: Empoasca decipiens (Paoli) and Cicadulina bipunctata<br />

(Melichar).<br />

16SrII. The 16SrII group was found in alfalfa and all insect species<br />

tested. In the Gulf region, the 16SrII group has been identified in<br />

lime alfalfa (1).<br />

Amplified DNA was sequenced to determine relationships<br />

between phytoplasma isolates. The work confirmed that<br />

phytoplasmas are infecting alfalfa crops in Saudi Arabia, and<br />

progress was made in identifying the phytoplasma groups<br />

present and information gained on their potential spread by<br />

vectors.<br />

E. decipiens and C. bipunctata, the most abundant insects<br />

collected from fields, were carrying phytoplasmas from 16SrI<br />

and 16SrII in proportions of 10:33, and 6:33, respectively. Results<br />

suggest that E. decipiens and C. bipunctata, are the major insect<br />

candidates to vector phytoplasmas from 16SrI and 16SrII groups.<br />

It is known that many vectors can transmit more than one type<br />

of phytoplasma and that many plants can harbour two or more<br />

distinct phytoplasmas. Vector‐host‐plant interactions play an<br />

important role in determining the spread of phytoplasmas. It is<br />

very likely that due to their abundance and capability to carry<br />

phytoplasmas, E. decipiens and C. bipunctata mainly contributed<br />

to the spread phytoplasma diseases, in alfalfa, date palm,<br />

decline in lime and the disease in papaya, by cycling from the<br />

alternative reservoirs to the crops, so that, the spread of<br />

diseases is a consequence of the vector‐phytoplasma‐plant three<br />

way interaction.<br />

PCR and RFLP analysis DNA was extracted from leaf tissue and<br />

insects. Aliquots of final DNA preparations were used as<br />

template for a nested PCR (nPCR) assay with phytoplasma 16S<br />

rDNA primers R16mF2/R16mR1 for the first round, and either<br />

R16F2n/R16R2 and fU5/rU3 for the nested reaction. Nested PCR<br />

products (10 ml) were digested with restriction endonucleases<br />

AluI, HpaII, Hea III and Sau3A I.<br />

16S rDNA sequencing and phylogenetic analysis Phytoplasma<br />

rDNA amplified by PCR using the primer pair P1/P7 was purified.<br />

The PCR products were sequenced in both directions using<br />

primer pair P1/P7 and the 16S rDNA sequences of phytoplasmas<br />

identified in our study were compared with others in Genbank<br />

by BLAST.<br />

RESULTS AND DISCUSSION<br />

Crop samples showing typical symptoms of phytoplasma<br />

infection were collected from different areas of Al Hasa (Fig 1).<br />

Leafhoppers were also trapped by netting for examination as<br />

potential vectors of the disease. <strong>Plant</strong> samples and leaf hoppers<br />

were analysed by DNA extraction and amplification with<br />

phytoplasma‐specific primers.<br />

Phytoplasmas were detected in 43/76 alfalfa samples, and from<br />

16/33 batches of all leafhopper species tested. No PCR products<br />

were obtained for asymptomatic plant samples. RFLPs were used<br />

to partially characterise isolates from plants and insects. Based<br />

on RFLP and sequencing analysis, phytoplasmas from group<br />

Figure 1. Alfalfa witches’‐broom symptoms<br />

ACKNOWLEDGEMENTS<br />

We thank BAE Systems and the British Council for funding the<br />

Post doctoral Summer Research Program for Dr Khalid Alhudaib.<br />

We thank Dr Mike Wilson at the national Museum of Wales,<br />

Cardiff, UK, for the finer identification of the leafhopper species.<br />

REFERENCES<br />

1. Khan A, Botti S, Al‐Subhi A, Gundersen‐Rindal D, Bertaccini A. 2002.<br />

Molecular identification of a new phytoplasma associated with<br />

alfalfa witches’ broom in Oman. Phytopathology 92: 1038–47.<br />

2. Alhudaib K, Arocha Y, Wilson M, Jones P. 2007. First report of a<br />

16SrI, Candidatus Phytoplasma asteris group phytoplasma<br />

associated with a date palm disease in Saudi Arabia. <strong>Plant</strong><br />

<strong>Pathology</strong> NDR 15: Feb. 2007.<br />

134 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


59 Efficient transformation of Colletotrichum capsici, the causal agent of chilli pepper<br />

anthracnose by Agrobacterium<br />

Posters<br />

A.S.M. Auyong{ XE "Auyong, A.S.M." }, R. Ford and P.W.J. Taylor<br />

BioMarka/Centre for <strong>Plant</strong> Health, Melbourne School of Land and Environment, The University of Melbourne, 3010, Victoria<br />

INTRODUCTION<br />

Anthracnose disease of chilli pepper is caused by a complex of<br />

Colletotrichum spp. with C. capsici being the most severe in<br />

South East Asia (1). Knowledge of the mechanisms for host<br />

resistance and pathogenicity is crucial for developing effective<br />

and durable disease control. Several putative pathogenicity<br />

genes involved in C. capsici infection of chilli pepper have been<br />

identified and partially cloned (Auyong, unpublished). A fungal<br />

transformation system is required to prove the function of these<br />

putative genes in the infection process. In this study an efficient<br />

transformation system was successfully developed to serve as a<br />

platform towards understanding chilli pepper‐C. capsici<br />

interactions. Agrobacterium tumefaciens carrying a hygromycin<br />

phosphotransferase gene (hph) and a green fluorescent protein<br />

(GFP) gene was used to transform the conidiospores of C.<br />

capsici. Transformation efficiency was correlated with<br />

conidiospores density, ratio of conidiospores to bacterial cells,<br />

type of Agrobacterium strains and plasmid, presence or absence<br />

of acetosyringone, co‐cultivation time and co‐cultivation<br />

temperature.<br />

MATERIALS AND METHODS<br />

Fungal culture. Colletotrichum capsici, BRIP 26974 isolated from<br />

Capsicum annuum was supplied by the Department of Primary<br />

Industry (DPI), Queensland, Australia, and maintained on potato<br />

dextrose agar (PDA).<br />

Fungal transformation. Conidial suspension was prepared and<br />

adjusted to 10 2 , 10 4 , 10 6 and 10 8 conidiospores per ml and mixed<br />

at different ratio (1:3, 1:5, 1:1, 3:1 and 5:1) with Agrobacterium<br />

(AGL1 or LBA4404) containing either pJF1, pPK2 or pKHt plasmid.<br />

The mixture was plated onto filter paper on solid induction<br />

medium, either amended or non‐amended with 200 μM<br />

acetosyringone. Following co‐cultivation for 1, 2, 3, 4, and 5 days<br />

at co‐cultivation temperature of 24°C, 28°C, 32°C or 36°C, the<br />

fungal‐bacterial cells on the filter paper were transferred to PDA<br />

amended with hygromycin B and cefotaxime to eliminate the A.<br />

tumefaciens cells. Individual transformants were transferred<br />

after 4 to 6 days to PDA amended with hygromycin B. In all<br />

experiments, C. capsici conidiospores co‐cultivated with<br />

uninoculated induction medium were included as a negative<br />

control. All experiments were replicated and results were<br />

analysed using ANOVA.<br />

Analysis of transformation events. The frequency and<br />

randomness of T‐DNA integration in the fungal genome was<br />

determined by PCR and Southern blot.<br />

RESULTS AND DISCUSSION<br />

Agrobacterium was successfully used to transform C. capsici<br />

conidiospores and mycelium (Figure 1). The biological<br />

differences among fungal species can influence transformation<br />

efficiencies in different filamentous fungi (2). Following<br />

optimisation, high transformation efficiencies were routinely<br />

obtained for C. capsici.<br />

Conidiospore density. Transformation efficiency was consistently<br />

found to be optimum at the conidiospores density of 10 6 and 10 8<br />

conidiospores per ml. Hence, subsequent transformations were<br />

carried out using 10 6 conidiospores per ml.<br />

Ratio of fungal spores to bacterial cells. The highest<br />

transformation efficiency was obtained with equal volume of the<br />

mixture.<br />

Agrobacterium strains. A. tumefaciens AGL1 strain produced<br />

more transformants (16.2% more) than LBA4404 regardless of<br />

the binary vector used.<br />

Plasmid type. pJF1 and pPK2 plasmids provided similar<br />

transformation efficiencies. In contrast, pKHt plasmid produced<br />

significantly less transformants (p


Posters<br />

60 Infection process of endophytic Colletotrichum gloeosporioides on cacao leaves<br />

C. Blomley{ XE "Blomley, C." } A , E.C.Y Liew B and D.I. Guest A<br />

A Faculty of Agriculture Food and Natural Resources, The University of Sydney, 2006, NSW<br />

B Royal Botanic Gardens Trust, The Royal Botanic Gardens, Sydney, NSW, 2000<br />

INTRODUCTION<br />

Colletotrichum species are commonly isolated as endophytes<br />

from leaves and fruits of tropical plants. In preliminary surveys<br />

Colletotrichum spp. accounted for 29–48% of isolates sampled as<br />

endophytes from leaves of the cacao tree (Theobroma cacao) in<br />

four sites in Australia and Papua New Guinea. By definition<br />

endophytes do not cause disease symptoms at the time they are<br />

isolated from plant tissue. The infection process and subsequent<br />

tissue colonisation has been elucidated for only a few endophyte<br />

host interactions, none of which include tropical plants.<br />

Colletotrichum species can penetrate plant tissue through<br />

wounds, natural openings such as stomata or by penetration of<br />

the plant cuticle 1 . They are often categorised into three groups:<br />

intracellular hemibiotrophs, subcuticular intramural colonisers<br />

and those that display a combination of the two infection<br />

strategies 1 . Intracellular hemibiotrophs first grow biotrophically<br />

in host tissue before switching to a necrotrophic stage which<br />

results in symptom development. Subcuticular intramural<br />

pathogens grow beneath the cuticle and cause dissolution of the<br />

epidermal cell walls. The aim of this research was to investigate<br />

the infection process of an endophytic isolate of C.<br />

gloeosporioides on T. cacao leaves.<br />

MATERIALS AND METHODS<br />

An isolate of C. gloeosporioides was isolated from apparently<br />

healthy leaf tissue of T. cacao in Far North Queensland,<br />

Australia. Young and mature leaves of cacao were sprayed with a<br />

1x10 5 conidia/mL suspension to runoff. Leaf tissue was sampled<br />

for observations every 2h for 16h, at 24h and then every 24h for<br />

6 days. Tissue was cleared for 4h at 60˚C followed by 20h at<br />

room temperature in a solution of 0.15% trichloroacetic acid in<br />

3:1 ethanol:chloroform. Tissue was immersed in 0.025% aniline<br />

blue in lactoglycerol for 1h at 60˚C followed by 23h at room<br />

temperature in order to stain fungal hyphae. Experiments were<br />

repeated at least three times.<br />

RESULTS<br />

Conidia began germinating within 6 hours post inoculation (hpi),<br />

usually giving rise to one and rarely two germ tubes. Appressoria<br />

were produced at 8–10 hpi, either directly or at the end of a<br />

short germ tube and became melanised by 12 hpi. Infection pegs<br />

were produced predominantly over cell walls at 12–16 hpi in<br />

both young and mature leaves. Stomatal penetration was never<br />

observed. Infection vesicles were visible at 3 days post<br />

inoculation (dpi) in young leaves and appeared as thick, highly<br />

lobed hyphae which filled the epidermal cell directly beneath the<br />

infection peg. At 4–5 dpi infection vesicles had branched into<br />

narrow secondary hyphae which penetrated cell walls and grew<br />

inter‐ and intra‐cellularly in young leaves (Fig. 1). Infection<br />

vesicles formed in mature leaves 4–5 dpi and had a similar<br />

appearance to those in young leaves (Fig 2). In mature leaves,<br />

infection was restricted to the initial cell in which the infection<br />

vesicle formed over the 6 days of observation.<br />

Figure 1. Primary hyphae of C. gloeosporioides colonising epidermal cells<br />

of T. cacao leaves 4 dpi. Bar = 10µm<br />

Figure 2. Primary hyphae of C. gloeosporioides restricted to the<br />

epidermal cell directly below appressoria (5 dpi). Bar = 20um).<br />

DISCUSSION<br />

Endophytic C. gloeosporioides on T. cacao leaves can be<br />

categorised as an intercellular hemibiotroph. Infection was<br />

observed in epidermal cells directly beneath the appressorium<br />

and no subcuticular intramural growth was observed. The<br />

infection process did not differ on young and mature T. cacao<br />

leaves in the first 3 dpi. Following this, colonisation was more<br />

rapid in young leaves and led to the production of disease<br />

symptoms. Infection in mature leaves remained biotrophic and<br />

fungal growth appeared to cease after infection and colonisation<br />

of one epidermal cell. The length of the biotrophic,<br />

asymptomatic phase has been correlated the redox state 2 and<br />

pH of the host tissue 3 in other Colletotrichum‐host interactions.<br />

Factors affecting the infection process in T. cacao are currently<br />

being investigated.<br />

REFERENCES<br />

1. Bailey JA, O'Connell RJ, Pring RJ & Nasby C (1992). Infection<br />

strategies of Colletotrichum species. In ‘Colletotrichum: Biology,<br />

<strong>Pathology</strong> and Control’ (Eds JA Barley & MJ Jeger) pp. 88–120.<br />

(C.A.B. International: Wallingford)<br />

2. Wei YD, Byer KN, Goodwin, PH (1997) Hemibiotrophic infection of<br />

round‐leaves mallow by Colletotrichum gloeosporioides f.sp.<br />

malvae in relation to leaf senescence and reducing agents.<br />

Mycological Research 101, 357–364<br />

3. Kramer‐Haimovitch H, Servi E, Katan T, Rollins J, OkonY, & Prusky,<br />

D. (2006) Effect of Ammonia production by Colletotrichum<br />

gloeosporioides on pelB activation, pectate lyase secretion and fruit<br />

pathogenicity. Applied and Environmental Microbiology 72,1034–<br />

1039.<br />

136 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


2 A Phytophthora sp. is the cause of jackfruit decline in the philippines<br />

L.M. Borines{ XE "Borines, L.M." } A , R. Daniel B and D. Guest B<br />

A Department of Pest Management, Visayas State University, Visca, Baybay, 6521 Leyte, Philippines<br />

B University of Sydney, Sydney, NSW 2006 Australia<br />

Posters<br />

INTRODUCTION<br />

Jackfruit is a very popular domestic fruit in the Philippines. It has<br />

a wide distribution and is cultivated throughout the country. A<br />

survey of jackfruit growers in the Eastern Visayas indicated that<br />

wilt disease was the main constraint to improved productivity. In<br />

some areas up to 90% of jackfruit trees are affected by wilt<br />

disease, manifested by leaf yellowing, defoliation, girdling stem<br />

lesions and rot. Previous attempts to identify the pathogen<br />

yielded a range of fungal, nematode or bacterial isolates, none<br />

of which proved pathogenic. Accurate identification of the cause<br />

of the decline syndrome is imperative for the control of the<br />

disease. This study seeks to isolate and identify the pathogen<br />

causing jackfruit wilt and to evaluate a range of disease<br />

management strategies through participatory action research.<br />

MATERIALS AND METHODS<br />

Affected roots, stem canker lesions and soil from near infected<br />

trees was suspended in water and baited with flower petals 1 .<br />

Lesions that developed within 2 days were surface sterilised and<br />

plated on Potato Dextrose Agar, Carrot Agar and Onion Agar,<br />

supplemented with benomyl, nystatin and streptomycin. Pure<br />

cultures were re‐introduced to flower baits to induce sporangia,<br />

zoospore and chlamydospore formation for inoculation of<br />

detached jackfruit leaves and seedlings.<br />

RESULTS<br />

Wilt disease was recorded in all the fields examined within Leyte<br />

and Samar islands, at an incidence of 5–90% of trees. Areas with<br />

very high incidence were typically subject to periodic heavy<br />

flooding, particularly during the rainy season. Yield losses were<br />

estimated to be range from 5–80%. Field visits and farmer<br />

interviews showed that almost all of the farmers were unaware<br />

of the cause of the disease or appropriate management<br />

strategies.<br />

zoospores through a vesicle before they separate and swim<br />

away (Figs 2a and 2b). The isolated Phytophthora species<br />

produces abundant intercalary and terminal chlamydospores<br />

when re‐introduced to flower baits.<br />

a<br />

Figure 2. a) Phytophthora sporangia, b) zoospores exiting via spherical<br />

vesicles.<br />

Participatory action research (PAR) disease management trials<br />

are being established by researchers, extension officers and<br />

jackfruit farmers in Leyte and Samar Islands. Nine PAR trials have<br />

been established in Leyte and Samar to test a range of<br />

management options including field sanitation, organic<br />

amendments, improved drainage, good nursery practices and<br />

chemical control in managing wilt disease.<br />

The identification of the pathogen associated with the symptoms<br />

of decline and wilt will enable the development of more<br />

effective, targeted management strategies.<br />

ACKNOWLEDGMENTS<br />

This research is funded by ACIAR HORT/2006/067/2<br />

REFERENCES<br />

Drenth A. & Guest DI. 2006. Biology and Management of Phytophthora<br />

diseases in the tropics. ACIAR Monograph 114.<br />

b<br />

A Phytophthora species was consistently isolated from affected<br />

jackfruit roots and canker lesions, and from soil collected near<br />

infected plants. Pathogenicity was confirmed when the isolates<br />

produced typical wilting symptoms on inoculated plants (Fig. 1a)<br />

and leaf lesions (Fig.1b).<br />

a<br />

b<br />

Figure 1. a. Un‐inoculated (leftmost) and inoculated jackfruit seedlings<br />

showing different degrees of wilting. b. The isolated pathogen causes<br />

leaf lesions.<br />

In pure culture the mycelium of the pathogen is white with a<br />

stellate growth pattern. Sporangia are seldom produced in PDA,<br />

Carrot or Onion Agar, but readily produced when pure cultures<br />

were re‐introduced to flower petal baits. The pathogen<br />

produced spherical to ovoid sporangia with an average length of<br />

41.5 µm and breadth of 26.5 µm. Sporangia have a relatively<br />

long pedicels, are semi‐papillate to papillate and release<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 137


Posters<br />

3 The effects of calcium chloride and calcium carbonate on germination and growth<br />

of Colletotrichum acutatum and Penicillium expansum<br />

K.S.H. Boyd‐Wilson{ XE "Boyd‐Wilson, K.S.H." } A and M. Walter A<br />

A The New Zealand Institute for <strong>Plant</strong> and Food Research Limited, PO Box 51, Lincoln 7640, New Zealand<br />

INTRODUCTION<br />

Calcium chloride (CaCl 2 ) and calcium carbonate (CaCO 3 ) have<br />

been found to enhance the biocontrol activity of yeasts against a<br />

range of diseases. A number of factors including inhibition of the<br />

pathogen by the compounds may account for this (1). The aim of<br />

this research was to investigate whether these compounds<br />

inhibited germination and growth of Penicillium expansum, the<br />

causal agent of blue mould of apples, and of Colletotrichum<br />

acutatum, which causes bitter rot of pome fruit (2,3).<br />

MATERIALS AND METHODS<br />

Germination (%) and germ‐tube length (µm) were assessed after<br />

20–24 h at 20ºC for three C. acutatum isolates made up in 0, 10,<br />

or 20 mg/ml CaCl 2 . P. expansum requires exogenous nutrients to<br />

germinate well, thus germination assays were conducted in 0,<br />

12.5 and 25% apple broth with 20 mg/ml CaCl 2 for three P.<br />

expansum isolates. For each combination, the per cent<br />

germination of 150 conidia and germ‐tube length of 30 conidia<br />

was recorded.<br />

Because suspended CaCO 3 made it difficult to observe conidia in<br />

germination tests, the effects of CaCO 3 and CaCl 2 (each 20<br />

mg/ml) on the three isolates of P. expansum and two of C.<br />

acutatum were also investigated by dilution plating on 0, 12.5<br />

and 25% apple broth agar with and without CaCl 2 and counting<br />

colony forming units (cfu) after 4–7 days.<br />

All experiments were conducted twice and data were analysed<br />

using analysis of variance. Germ‐tube lengths were log 10<br />

transformed and colony counts were square‐root transformed<br />

before analysis. P=0.05 was used to assess significance.<br />

On apple broth agar, no differences between treatments were<br />

observed and therefore results are given only for agar with no<br />

apple broth.<br />

For all P. expansum isolates, the addition of CaCO 3 to the agar<br />

resulted in significantly fewer cfu than in the CaCl 2 and water<br />

only treatments. The addition of calcium carbonate did not<br />

significantly affect cfu counts of C. acutatum isolates.<br />

In conclusion, neither CaCl 2 nor CaCO 3 reduced germination,<br />

germ‐tube growth or cfu counts for C. acutatum. This suggests<br />

that some other mode of action contributes to enhancement of<br />

bitter rot control when these compounds are combined with<br />

yeasts. In contrast, the germination, germ‐tube growth and cfu<br />

counts for P. expansum were all reduced by the addition of CaCl 2<br />

and CaCO 3 to the growth medium suggesting that this direct<br />

inhibition could contribute to the improved blue mould control<br />

in apples when yeasts are combined with these compounds.<br />

ACKNOWLEDGEMENTS<br />

Thanks to the New Zealand Foundation for Research, Science<br />

and Technology for funding this project (C06X0302).<br />

REFERENCES<br />

1. Everett KR, Vanneste JL, Hallett IC, Walter M (2005) Ecological<br />

alternatives for disease management of fruit rot pathogens. New<br />

Zealand <strong>Plant</strong> Protection 58, 55–61.<br />

2. Rosenberger DA (1997) Blue mold. In ‘Compendium of apple and<br />

pear diseases’. (Eds A L Jones, HS Aldwinckle) pp 54–55. (APS Press:<br />

USA)<br />

3. Sutton TB (1997) Bitter rot. In ‘Compendium of apple and pear<br />

diseases’. (Eds A L Jones, HS Aldwinckle) pp 15–16. (APS Press: USA)<br />

RESULTS AND DISCUSSION<br />

Increasing concentrations of CaCl 2 increased the germination and<br />

germ‐tube length of C. acutatum, although there was a<br />

significant interaction between the factors studied (Table 1).<br />

Table 1. Mean per cent germination and germ‐tube length (log 10<br />

transformed) for each Colletotrichum acutatum isolate and CaCl 2<br />

concentration.<br />

Germination (%) Germ‐tube length (µm log 10 )<br />

CaCl 2<br />

(mg/ml)<br />

CaCl 2<br />

(mg/ml)<br />

Isolates 0 10 20 0 10 20<br />

C4 57 57 85 0.56 0.45 0.34<br />

C7 41 69 100 0.39 0.67 0.70<br />

C8 71 51 85 0.45 0.27 0.70<br />

s.e.d. interaction 17.2 s.e.d concentration 0.097<br />

P. expansum conidia did not germinate in water alone. In apple<br />

broth, the addition of CaCl 2 significantly reduced mean<br />

germination of P. expansum from 37 to 16% and germ‐tube<br />

length from 0.382 to 0.160 compared with the control. There<br />

was no difference in germination and germ‐tube length between<br />

the two concentrations of apple broth.<br />

138 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


45 Fungal endophytes of the Boab species Adansonia gregorri and other native tree<br />

species<br />

Posters<br />

M.L. Sakalidis AB , G.E.StJ. Hardy B and T.I. Burgess{ XE "Burgess, T.I." } B<br />

A monique.sakalidis@hotmail.com<br />

B Faculty of Sustainability, Environmental and Life Sciences, Murdoch University, Murdoch 6150, WA, Australia<br />

INTRODUCTION<br />

In southern Africa dying Boabs have been reported and<br />

subsequent surveying of these trees has indicated the presence<br />

of the fungal pathogen Lasiodiplodia which may be the cause of<br />

the decline. Due to the close genetic relationship of the African<br />

and Australian Boabs and the fact that these two continents<br />

share are large amount of floral families, they may subsequently<br />

also share the pathogens of many of these plants. The surveying<br />

of the otherwise healthy Australian Boab and surrounding tree<br />

species deemed a prudent course of action.<br />

In this study Boabs were surveyed in 25 sites in the Kimberley<br />

region and material was also taken from surrounding tree<br />

species at 3 sites. Endophytic fungi that were isolated from these<br />

samples were identified using both molecular and morphological<br />

data and seven new species were described (2). The<br />

pathogenicity of identified species to Boabs was determined.<br />

This is the first study to identify endophytes of the Adansonia<br />

and to conduct pathogenicity trials on these trees.<br />

At the extreme margin of the lesions the wood was cut away<br />

using a knife in order to establish the extent of interior lesion<br />

development.<br />

RESULTS AND DISCUSSION<br />

433 fungal isolations were made, 282 of these consisted of<br />

isolates belonging to Botryosphaeriaceae including species of<br />

Neofussicoccum, Pseudofusicoccum, Lasiodiploidia Dothiorella<br />

and Neoscytalidium.<br />

For the trial with tap roots, isolates from the Lasiodiplodia<br />

theobromae complex produced the largest lesions.<br />

Neofusiccocum ribis and Neoscyltalidium novaehollandia caused<br />

moderate lesions and isolates of Lasiodiploida crassispora,<br />

Dothiorella longicollis, Pseudofusicoccum adansoniae and<br />

Fusicoccum ramosum all caused minor lesions indicating low<br />

virulence.<br />

MATERIALS AND METHODS<br />

Stem and leaf material was collected from a range of sites across<br />

the Kimberlys, Western Australia. Material was taken from<br />

Adansonia gregorrii and a range of native flora in the same area.<br />

Endophytes were isolated using standard protocols (1).<br />

Lesion development in seedling tap roots. 24 isolates that<br />

represented the genetic diversity of samples collected were used<br />

to inoculate the tap root of four‐month‐old Boab seedlings. They<br />

were inoculated by using a sterile scalpel blade to make a small<br />

lateral incision along the middle of the carrot. Into which a 1 cm 2<br />

agar plug colonised with mycelium was inserted face up. This<br />

was then lightly wrapped with parafilm. There were 24 isolates<br />

plus controls (10 replicates of each). Tap roots from each<br />

replicate were placed in random order onto wooden racks inside<br />

plastic. The containers were then sealed with aluminium foil and<br />

tape and placed into a 25̊C room and left for 4–5 days. After four<br />

days lesion development in the tap roots were measured. The<br />

lesions presented as a rotted mass that could easily be scraped<br />

out of the tap root. The inoculated tap root was weighed, the<br />

lesion was scraped out and the carrot was re‐weighed<br />

immediately. The lesion length and width was also measured.<br />

Lesion development in young trees. 2–3 year old Boab trees<br />

were harvested in Kununurra from commercial Boab growers<br />

“Boabs in the Kimberlies.”. They were planted within 2 weeks of<br />

initial removal into one meter long PVC pipes in a potting<br />

medium of 1/3 coarse river sand and 2/3 potting mix and were<br />

watered twice a day for ten minutes by an automatic dripping<br />

system. Nine isolates were selected from the tap root trial. Boab<br />

stems were inoculated in the same manner as the roots. There<br />

were 5 replicates for each of the 9 isolates and also noninoculated<br />

controls. After 6 months trees were assessed for leaf<br />

cover and stems with lesions were harvested. The width, length<br />

and depth of lesions were measured using callipers and a ruler.<br />

The stems were cut in half at the centre of the initial mycelium<br />

plug insertion in order to determine the depth of lesion<br />

development.<br />

Figure 1. Mean length(cm) of lesions in 2–3 year old boab trees.<br />

Standard deviations are represented by the error bars.<br />

The results of the tree trial (Figure 1) confirm the results of the<br />

preliminary investigation, L. theobromae was found to be<br />

significantly more pathogenic then other species considered in<br />

the study. The lesions produced from inoculation of boab stems<br />

by L. theobromae resulted in the lesions lengths ranging from 3<br />

cm to 25cm (mean= 10.68cm). N. ribis and Neoscytalidium<br />

novaehollandia both exhibited similar lesion severity (means=<br />

3.46 cm and 3.54 cm respectively).<br />

This trial indicates the potential threat that L. theobromae<br />

presents to the iconic Boab trees. Recently a dying Boab in<br />

Broome was reported with similar disease symptoms those of<br />

dying Boabs in South Africa and similarly, L. theobromae was the<br />

only pathogen isolated from the cankers. As shown in this trial,<br />

endophytes such as L. theobromae are capable of causing<br />

disease and killing Boabs in Australia.<br />

REFERENCES<br />

1. Taylor, K., Barber, PA, Giles E. St J. Hardy and Burgess. TI (2009)<br />

Botryosphaeriaceae from tuart (Eucalyptus gomphocephala)<br />

woodland, including descriptions of four new species. Mycological<br />

Research 113: 337–353<br />

2. Pavlic D, Barber PA, Hardy GESJ, Slippers B, Wingfield MJ, Burgess<br />

TI, 2008. Seven new species of the Botryosphaeriaceae discovered<br />

on baobabs and other native trees in Western Australia. Mycologia.<br />

100: 851–866.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 139


Posters<br />

61 An emerging nematode pest on bananas?<br />

J.A. Cobon{ XE "Cobon, J.A." } A , and T. Pattison B<br />

A Queensland Primary Industries and Fisheries, 80 Meiers Rd, Indooroopilly, 4068, Queensland<br />

B Queensland Primary Industries and Fisheries, PO Box 20, South Johnstone, 4859, Queensland<br />

INTRODUCTION<br />

The focus of nematode control on bananas in Australia and<br />

worldwide has been on Radopholus similis (burrowing<br />

nematode) as the key nematode pest. International research and<br />

breeding programs have been active in managing and selecting<br />

cultivars for resistance to this nematode. However, recent<br />

surveys of African bananas have found Pratylenchus goodeyi P.<br />

coffee, Meloidogyne spp. and Helicotylenchus multicinctus as key<br />

nematode species.<br />

P. goodeyi (lesion nematodes) is considered to be indigenous to<br />

Africa where its distribution has been limited to the cooler<br />

highland growing areas including Central, Eastern and West<br />

Africa. It is considered that P. goodeyi has the potential to<br />

become an important pest of bananas where they are grown in<br />

cooler climatic zones of the Mediterranean and Middle Eastern<br />

countries. Bananas, as well as a number of other tropical crops,<br />

have been planted in the subtropical regions of South America,<br />

southern Africa, the Mediterranean basin, Australia and<br />

southern China. P. goodeyi has been recorded in the Canary<br />

Island, Crete and Egypt and Australia (1). However, it has<br />

recently also been found in warmer banana production areas of<br />

Africa (2).<br />

P. goodeyi can invade and feed in the root and corm tissue of<br />

banana plants. It causes similar symptoms and destruction as<br />

caused by R. similis including root lesions, stunted growth,<br />

reduced bunch weight and toppling of the bunching<br />

pseudostem.<br />

There is concern that the distribution of P. goodeyi may spread<br />

throughout subtropical banana production areas and it may also<br />

move into warmer, tropical production areas in Australia,<br />

following the recent experience in Africa.<br />

MATERIALS AND METHODS<br />

Root samples were submitted from several banana growing<br />

properties from northern NSW for the extraction of R. similis for<br />

molecular analysis. Roots were sliced open lengthwise and<br />

placed in a misting chamber for 5 days for the extraction of<br />

nematodes (3). Nematodes were caught by passing the washing<br />

from the mister over a 38 µm sieve. The washings were then<br />

examined under a compound microscope for the presence of<br />

nematodes and positive identifications made. For this accurate<br />

identification, nematodes were picked from solution using an<br />

eyelash and placed under higher magnification to further<br />

distinguish between R. similis and P. goodeyi.<br />

As P. goodyei was thought to have originated from the African<br />

highlands it is still unclear how this nematode arrived in<br />

Australia. However, with the number of sites with P. goodeyi<br />

increasing in NSW and the experience in Africa of P. goodeyi<br />

moving into the hotter production areas, the Australian banana<br />

industry needs to be aware of the distribution of these<br />

nematodes to avoid a repeat of the African experience.<br />

The movement of infested planting material and soil could<br />

increase the spread of P. goodeyi to additional banana growing<br />

areas within Australia, therefore, growers need to adhere to the<br />

clean planting material policy and be aware of the possibility of<br />

spread with infested soil and machinery.<br />

Accurate identification of the nematode species is essential to<br />

develop management options such as crop rotation and use of<br />

resistant cultivars. Surveys need to be undertaken to establish<br />

the spread of this nematode, as well as resistance screening of<br />

rotation crops and of new banana cultivars.<br />

ACKNOWLEDGEMENTS<br />

Funding for this work was provided by Queensland Primary<br />

Industries and Fisheries.<br />

REFERENCES<br />

1. Bridge, J., Fogain, R., Speijer, (1997) (1981) The root Lesion<br />

Nematodes of Banana. Musa pest fact sheet No 2.<br />

2. Coyne, D. & L. Waeyenberge (2008) <strong>Plant</strong>‐parasitic Nematodes<br />

Affecting Banana and <strong>Plant</strong>ain in Africa: A Shifting Focus?<br />

Proceedings of the 5th International Congress of Nematology,<br />

Brisbane, Australia. pp 64<br />

3. Hooper, D. J. (1986) Extraction of nematodes from plant material.<br />

IN SOUTHEY, J. F. (Ed.) Laboratory methods for work with <strong>Plant</strong> and<br />

Soil Nematodes. London, Her Majesty's Stationary Office.51–58<br />

4. Moody, E. H., Lownsbery, B. F. & Ahmed, J. M. (1973) Culture of the<br />

Root‐Lesion nematode Pratylenchus vulnus on Carrot Disks. Journal<br />

of Nematology. 5, 225–226.<br />

Single mature female nematodes of both R. similis and P.<br />

goodeyi were placed on a sterile carrot (Daucus carota) disc in<br />

order to initiate a single genotype isolate (4) for further<br />

experimental work.<br />

RESULTS AND DISCUSSION<br />

The root samples from these farms were found with a high<br />

incidence of P. goodeyi. Furthermore, some farms had<br />

exclusively P. goodeyi, and not R. similis as was believed. This<br />

suggested that P. goodeyi may be increasing in numbers and<br />

importance within banana plantations of NSW and south‐east<br />

Queensland.<br />

140 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


4 A sensitive PCR test for detecting the potato cyst nematode (Globodera<br />

rostochiensis) in large volume soil samples<br />

Posters<br />

S.J. Collins{ XE "Collins, S.J." } A , X.H. Zhang A , G.I. Dwyer A , J.M. Marshall B and V.A.Vanstone A<br />

A<br />

Department of Agriculture and Food Western Australia, South Perth, 6151, Western Australia<br />

B JM Advisory NZ Ltd, Christchurch, 8052, New Zealand<br />

INTRODUCTION<br />

Potato cyst nematode (PCN) Globodera rostochiensis, a<br />

devastating plant pathogen worldwide, impacts potato<br />

production and affects market access. PCN was detected<br />

between 1986 and 1989 on six properties (a total of 15ha) in the<br />

metropolitan area of Perth, Western Australia (WA). A strict<br />

quarantine and eradication program was immediately<br />

implemented, and no PCN has been detected anywhere in the<br />

state since. With almost 20 years since the last detection, WA is<br />

now in an excellent position to re‐claim Area Freedom from PCN.<br />

We are developing a sensitive PCR test to enable<br />

presence/absence of PCN to be determined directly from large<br />

soil samples for confirmation of Area Freedom for the state. PCR<br />

offers an alternative to traditional microscopic detection of PCN,<br />

which is time‐consuming and prone to operator error,<br />

particularly if cysts are present in low numbers.<br />

MATERIALS AND METHODS<br />

Field survey. Proving a ‘negative’ is always a challenge. With this<br />

in mind, survey methods were tailored to generate data to show<br />

with the highest possible confidence that PCN no longer occurs<br />

in WA. At all survey sites, 50g soil samples to a depth of 15 cm<br />

were collected on a 5 x 5 m grid pattern across entire fields. This<br />

resulted in collection of approx. 20 kg/ha, all of which was<br />

processed (without sub‐sampling) by the Fenwick method for<br />

total organic matter extraction. This sampling regime is far more<br />

intensive than any standard worldwide. All organic matter<br />

samples will be assessed using the PCR test under development.<br />

PCR test. There are numerous technical challenges when<br />

amplifying DNA extracted directly from soil (e.g. incomplete<br />

cyst/egg lysis, DNA adsorption to soil, co‐purification of PCR<br />

inhibitors, and degradation of target DNA). To reduce inputs, it is<br />

necessary to develop methods that maximise sample area per<br />

test without compromising assay integrity. PCR analysis of soil<br />

has usually been done with samples of only 1 to 15g. In contrast,<br />

we are developing a novel strategy to test 20kg pooled soil<br />

samples (each representing assessment of 1ha sampled on a 5 x<br />

5m grid) for presence/absence of PCN.<br />

Due to quarantine against the use of PCN, we are developing<br />

methodologies using Cereal Cyst Nematode, Heterodera avenae<br />

(CCN). The goal is to develop the technology for routine<br />

detection of 10 cysts in a 20kg soil sample. Once optimised,<br />

detection methodologies will be validated in blind studies using<br />

PCN‐infested soil in New Zealand.<br />

RESULTS AND DISCUSSION<br />

Results are encouraging, with as little as 1 CCN cyst detected in<br />

20kg of soil (Fig. 1). Since the average PCN cyst contains approx.<br />

400 eggs, this is equivalent to detection of approx. 0.02 eggs/g<br />

soil which is a detection level that could identify extremely low<br />

levels of infestation.<br />

—Lanes: —1 — —2 — — 3 — — 4 — — 5 — — 6 — –— 7 —– — 8—–—9<br />

Figure 1. PCR results from samples with 1‐200 CCN cysts. Lane 1: 1 cyst;<br />

Lane 2: 5 cysts; Lane 3: 10 cysts; Lane 4: 20 cysts; Lane 5, 50 cysts; Lane<br />

6: 100 cysts; Lane 7: 200 cysts; Lane 8: blank; Lane 9: H 2 O.<br />

Although we have been able to detect only 1 CCN cyst in 20kg of<br />

soil, reliability of the test is more consistent for 10 cysts/20kg of<br />

soil (Fig. 2). This represents detection of approx. 0.2 eggs/g soil,<br />

which is far below national and international standards.<br />

——Lanes— —1—— 2— —3— — 4 — — 5— — 6— — 7 — 8 —— 9—— 10<br />

Figure 2. PCR results using target DNA extracted from samples with 10<br />

CCN cysts. Lanes 1–4: 5uL target DNA; Lanes 5–8: 10uL target DNA; Lane<br />

9: +ve control; Lane 10: H 2 O.<br />

Currently, the test is being refined to eliminate the effects of<br />

contaminants and inhibitors in the soil. Ways to increase<br />

reliability of the PCR test in different soil types are also being<br />

assessed. For example (Fig. 3), results from a pilot trial have<br />

shown that detection of CCN DNA from Albany soil (Humic<br />

Podzols) was more reliable than from Busselton soil (Jindong<br />

Sandy Loam).<br />

— Lanes: — 1— 2— 3— 4— 5— 6— 7— 8— 9— 10 —11 —12 —13 —14<br />

Figure 3. PCR results using 5ul of target DNA extracted from 10 CCN cysts<br />

in two different soil types (Albany and Busselton). Lanes 1 and 2: ‐ve<br />

control for Albany soil; Lanes 3–6: Albany soil; Lanes 7 and 8: ‐ve control<br />

for Busselton soil; Lanes 9–12: Busselton soil; Lane 13: +ve control; Lane<br />

14: H 2 O.<br />

Once optimised, this test could have potential application for<br />

detection of other pests and pathogens that can be found in soil<br />

organic matter.<br />

ACKNOWLEDGEMENTS<br />

Horticulture Australia Ltd and the Potato Growers’ Association<br />

WA provide funding. WA growers allowed us to sample fields.<br />

Larry Hegarty (Potato Marketing Corporation) supplied planting,<br />

harvest data and maps. Dyane Jardine and Ali Bhatti provided<br />

field and lab technical support. Peter Philippe (WA Quarantine<br />

Inspection Service) allowed access to historical PCN sampling<br />

data.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 141


Posters<br />

62 Phellinus noxius: brown root rot is increasing in importance in the Australian<br />

avocado industry<br />

E.K. Dann{ XE "Dann, E.K." }, L.A. Smith, M.L. Grose, G.S. Pegg and K.G. Pegg<br />

Queensland Primary Industries and Fisheries, Department of Employment, Economic Development and Innovation, 80 Meiers Rd,<br />

Indooroopilly, QLD 4068<br />

INTRODUCTION<br />

Phellinus noxius is an indigenous wood decay basidiomycete,<br />

common in tropical and subtropical rainforests of eastern<br />

Australia, Asia, the Pacific, central America and Africa. Phellinus<br />

noxius causes root and lower stem rot (“brown root rot”) disease<br />

of native and introduced trees planted on former rainforest<br />

sites. Overseas, it has caused significant losses through tree<br />

deaths in hosts such as rubber, mahogany, teak, cocoa, longan,<br />

litchi, pear, persimmon and Acacia mangium (widespread<br />

plantings in south east Asia for pulpwood).<br />

Infection takes place when roots contact infested woody matter<br />

present in the soil, and thus spread is most likely via root‐root<br />

contact. Trees can suffer a rapid decline, and foliage may<br />

transform from green and healthy to wilted and dead within a<br />

few weeks (Plate 1). Decline in older trees can be more gradual,<br />

with some mature infected trees surviving for many years. One<br />

key diagnostic feature is the presence of a thick brown mycelial<br />

matt with a white actively growing margin that melanises with<br />

age, which can be found growing on the root and stem surfaces<br />

(Plate 2). Fruiting bodies, although uncommon, occur in two<br />

forms. The resupinate form is seen on the underside of fallen<br />

logs, and between buttress roots of Ficus spp., and the bracket<br />

form is more often seen on dead trees in higher rainfall areas<br />

such as northern Queensland.<br />

MATERIALS AND METHODS<br />

Several orchards across the Atherton Tablelands and<br />

Childers/Bundaberg production regions, representing areas of<br />

approximately 650 and 300 km 2 , respectively, were visited.<br />

Selected additional properties in other areas of SE QLD or<br />

northern NSW were also visited. Samples from actively growing<br />

infection stockings were collected and plated onto malt extract<br />

agar containing 1%w/v streptomycin and 1ppm benomyl.<br />

Cultures were identified by morphological features of the<br />

hyphae.<br />

RESULTS<br />

P. noxius was confirmed on 17 out of 18 properties visited on the<br />

Atherton Tablelands, including in mango at 2 sites. It was also<br />

confirmed from 3 (and suspected on a further 2) orchards in the<br />

Childers/Bundaberg area, where 2 properties visited were<br />

apparently free of the disease. It has been confirmed on one<br />

orchard at Maleny and 2 orchards in northern NSW. Losses were<br />

particularly severe (approx. 10% tree death in affected blocks) in<br />

at least 4 orchards visited, and attempts to replant in infested<br />

soil failed. To date, no fruiting bodies have been found on<br />

avocado.<br />

The disease leads to significant losses in hoop pine plantations in<br />

Queensland, and in broadleaf hosts (eg. Ficus spp.) in urban<br />

parks and gardens (2). Death of avocado trees successively down<br />

rows was first noted on the Atherton Tablelands in QLD in 2001.<br />

The first positive identification of P. noxius causing tree death in<br />

avocado occurred in 2002 from the Maleny district of the<br />

Sunshine Coast hinterland in Queensland.<br />

This paper reports on a scoping study undertaken to assess the<br />

spread and severity of brown root rot in major avocado growing<br />

areas of north eastern Australia.<br />

Plate 2. The characteristic infection “stocking” at the base of avocado<br />

trunk<br />

DISCUSSION<br />

Effective management relies on complete removal of dead and<br />

dying trees and their roots, and one apparently healthy tree on<br />

either side. Root barriers can then be installed to prevent roots<br />

from uninfected trees coming into contact with infected debris<br />

in soil. There is currently no effective chemical control.<br />

ACKNOWLEDGEMENTS<br />

We acknowledge M. Weinert (QPIF, Mareeba) and E. Dunn (Crop<br />

Tech, Bundaberg), for assistance in organising the orchard visits<br />

and numerous avocado growers for their cooperation and<br />

access. The project is funded by Avocados Australia Ltd. and HAL.<br />

Plate 1. Quick decline of an avocado tree, laden with fruit.<br />

REFERENCES<br />

1. Bolland L (1984) Phellinus noxius: cause of a significant root‐rot in<br />

Queensland hoop pine plantations. Australian Forestry 47:2–10.<br />

2. Hood I (2003) “An introduction to fungi on wood in Queensland”<br />

pp. 312–315. (University of New England: Armidale)<br />

142 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


5 Management strategies to economically control blackspot and maximise yield in<br />

new improved field pea cultivars<br />

Posters<br />

J.A. Davidson{ XE "Davidson, J.A." } A , L. McMurray B and M. Lines B<br />

A South Australian Research and Development Institute (SARDI), GPO Box 397, Adelaide, 5001, SA<br />

B South Australian Research and Development Institute (SARDI), GPO Box 822, Clare, 5453, SA<br />

INTRODUCTION<br />

Field pea production in South Australia has remained constant at<br />

approximately 120,000 ha since the mid 1990s, although<br />

plantings have increased in medium to low rainfall areas.<br />

Blackspot, caused by a complex of fungi i.e. Mycosphaerella<br />

pinodes, Phoma medicaginis var. pinodella, Ascochyta pisi and<br />

Phoma koolunga (1), is the most common disease in field peas.<br />

Research on blackspot in the 1990s, based in traditional areas<br />

and on traditional late‐maturing trailing type peas i.e. cv. Alma,<br />

found that foliar fungicides were uneconomic but delaying<br />

sowing minimised blackspot infection from airborne spores.<br />

Delayed sowing is still a major recommendation for the pea<br />

industry across South Australia (2). Given the expansion into low<br />

rainfall areas and increasing frequency of low rainfall seasons,<br />

the potential yield loss through delayed sowing is often now<br />

greater than the loss from blackspot. Furthermore, fungicide<br />

costs have reduced and this practice may now be economic in<br />

some environments. The pea industry has also adopted higher<br />

yielding cultivars including early maturing erect semi‐leafless<br />

types i.e. cv. Kaspa. Agronomic trials were conducted in 2007<br />

and 2008 to identify economic strategies to control blackspot in<br />

new improved pea cultivars, and to identify optimum sowing<br />

dates in low to medium rainfall areas for these cultivars.<br />

MATERIALS AND METHODS<br />

Trials were sown at three sites each season, viz. high rainfall<br />

(450mm per annum) at Kingsford in 2007 and Turretfield in<br />

2008; medium rainfall (400mm per annum) at Hart in both<br />

seasons; low rainfall (325 mm per annum) at Minnipa in both<br />

seasons. The high and medium rainfall sites had three sowing<br />

times and the low rainfall site had two sowing times. First sowing<br />

occurred within a week of the break of the season (first week of<br />

May) at each site and subsequent sowing times were at intervals<br />

of three weeks. Trials were split plot design, with time of sowing<br />

as the main block, with three replicates. Cultivars included the<br />

conventional trailing types Alma and cv. Parafield (the latter at<br />

Minnipa only), and the new erect semi‐leafless types including<br />

the current commercial cultivar Kaspa and advanced breeding<br />

lines WAPEA2211 and OZP0602 (the latter in 2008 only).<br />

Fungicide treatments were the seed treatment P‐Pickel T ®<br />

(thiram plus thiabendazole, 200 ml/100kg seed), a foliar<br />

application of mancozeb (2 kg/ha) at 9 node growth stage, foliar<br />

applications of mancozeb at 9 nodes plus early flowering growth<br />

stage, P‐Pickel T ® seed dressing plus a foliar application of<br />

mancozeb at 9 nodes, foliar applications of chlorothalonil (2<br />

L/ha) every fortnight (i.e. disease control) and an untreated<br />

control. Disease was assessed regularly, 2 or 3 weeks apart,<br />

throughout the growing season, and recorded as % leaf area<br />

diseased (%LAD) in the early stages of the epidemic, or as % of<br />

nodes infected (%ND) in the later stages of the epidemic. Plot<br />

yields were recorded as tonnes per hectare. Significant<br />

differences identified by analyses of variance were separated on<br />

P


Posters<br />

63 Dispersal potential of Gibberella zeae ascospores<br />

P.A.B. Davies{ XE "Davies, P.A.B." } A , L.W. Burgess B , R. Trethowan A , R. Tokachichu A , D. Guest B<br />

A<br />

<strong>Plant</strong> Breeding Institute, University of Sydney, PMB 11, Camden, NSW, 2750<br />

B<br />

Faculty of Agriculture, Food and Natural Resources, University of Sydney, NSW, 2006<br />

INTRODUCTION<br />

Fusarium head blight (FHB) of wheat, caused by the fungus<br />

Gibberella zeae (anamorph Fusarium graminearum) is a disease<br />

that occurs sporadically in the Liverpool Plains region of<br />

Northern NSW. The fungus is also a pathogen of maize, causing<br />

Gibberella stalk and ear rots, and an asymptomatic endophyte of<br />

sorghum. The pathogen survives in the residues of these hosts,<br />

and in spring and autumn, perithecia form on these residues and<br />

forcibly discharge ascospores into the air (1).<br />

spores away from the source of inoculum closely fitted an<br />

exponential curve (R 2 =0.97) (Figure 1).<br />

The potential for long distance dispersal of these ascospores has<br />

been examined in North America (1, 2, 3), where spores have<br />

been recovered at least 3km from the nearest inoculum source<br />

and at 60m above the earth’s surface (2). This suggests that<br />

where there is a significant regional source of inoculum,<br />

localised control of infested residue through rotation or tillage<br />

practices may not effectively reduce the risk of FHB in individual<br />

fields (1).<br />

While inoculum levels and potential for dispersal are<br />

traditionally greater in North America compared to Australia,<br />

due to more favourable climatic conditions and the greater<br />

presence of maize within the farming system, evidence to<br />

support longer distance dispersal has been observed in the<br />

Liverpool Plains during 2005, when wheat crops free of inoculum<br />

had moderate levels of FHB.<br />

To determine the potential for long distance dispersal of<br />

ascospores under Australian conditions, a spore trapping<br />

experiment was established during October, 2008.<br />

MATERIALS AND METHODS<br />

A centre pivot irrigation field (80ha) in Spring Ridge (latitude 31°<br />

31'2.1''S longitude 150°14'6.3''E) was identified as a source of<br />

inoculum due to significant amounts of G. zeae perithecia on 6<br />

month old maize residue and high levels of FHB and perithecia<br />

on a cv. Beaufort wheat crop.<br />

Spore traps, standing 1m in height were placed at 50m intervals<br />

in a north easterly direction into a field 12 months fallow from<br />

Chickpeas, to a distance of 250m from the inoculum. Traps were<br />

also placed at 50m intervals into the wheat crop to a distance of<br />

250m into the crop. Traps consisted of four 90mm petri dishes<br />

containing Fusarium‐selective medium with increased rates of<br />

antibiotics, exposed to the atmosphere from sunset to sunrise<br />

the following morning. Exposure of the plates was timed to<br />

follow an irrigation event to the wheat crop of equivalent to<br />

15mm of rainfall 24 hours prior.<br />

Plates were recovered and incubated for 3 days under<br />

alternating light and dark conditions with temperatures at 24°C<br />

and 22°C respectively. A random subset of the colonises were<br />

subcultured from each plate and identified morphologically.<br />

Spore counts were taken from each plate and used to determine<br />

the number of G. zeae ascospores intercepted.<br />

RESULTS<br />

Ascospores of G. zeae were recovered at all locations and ranged<br />

from 90 cfu per plate at 250m from the inoculum source to 750<br />

cfu per plate within the wheat crop. The pattern of dispersal of<br />

Figure 1. G. zeae ascospore deposition away from inoculum source. The<br />

pattern of deposition closely follows the exponential curve y = 1.52 +<br />

56.99 x 0.99 x R 2 = 0.97<br />

DISCUSSION<br />

The pattern of ascospore dispersal agrees with previous reports<br />

of the incidence of disease away from an inoculum source being<br />

described by an exponential model (3). The results also suggest<br />

that recovery of spores at distances greater than 250m is likely.<br />

Extrapolation of the model to 500m suggests that 4000<br />

spores/m 2 would be deposited nightly. Whether this level of<br />

deposition is sufficient to initiate disease however is yet to be<br />

established.<br />

This experiment demonstrated that spore release events can be<br />

triggered by overhead irrigation events. The timing of irrigation<br />

events on wheat crops following maize should attempt to avoid<br />

irrigating during anthesis, at which wheat is susceptible to<br />

infection. Residue management may also be necessary to reduce<br />

the risk of FHB in such situations.<br />

ACKNOWLEDGEMENTS<br />

The research was completed with the assistance of a GRDC<br />

Grains Research Scholarship.<br />

REFERENCES<br />

1. Schmale, D.G., III, E.J. Shields, and G.C. Bergstrom, Nighttime<br />

spore deposition of the fusarium head blight pathogen,<br />

Gibberella zeae, in rotational wheat fields. Canadian Journal<br />

of <strong>Plant</strong> <strong>Pathology</strong>, 2006. 28(1).<br />

2. Maldonado‐Ramirez, S.L., et al., The relative abundance of<br />

viable spores of Gibberella zeae in the planetary boundary<br />

layer suggests the role of long‐distance transport in regional<br />

epidemics of Fusarium head blight. Agricultural and Forest<br />

Meteorology, 2005. 132(1/2): p. 20–27.<br />

3. Paulitz, T.C., et al., A generalized two‐dimensional Gaussian<br />

model of disease foci of head blight of wheat caused by<br />

Gibberella zeae. Phytopathology, 1999. 89(1): p. 74–83.<br />

144 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


25 Effect of avocado crop load on postharvest anthracnose and stem end rot, and<br />

cations and phenolic acid levels in peel<br />

Posters<br />

E.K. Dann, L.M. Coates, J.R. Dean{ XE "Dean, J.R." }, L.A. Smith, A.W. Cooke and K.G. Pegg<br />

Queensland Primary Industries and Fisheries, Department of Employment, Economic Development and Innovation, 80 Meiers Rd,<br />

Indooroopilly, QLD 4068<br />

INTRODUCTION<br />

Crop load (or tree yield), rootstock and mineral nutrition in the<br />

flesh can influence quality of ‘Hass’ avocado fruit in terms of<br />

anthracnose and internal flesh disorders (1,2). While fruit from<br />

higher yielding trees are often smaller, they have been reported<br />

to have less anthracnose (Colletotrichum gloeosporioides) and<br />

higher Ca (2). Ca is thought to strengthen cell walls making them<br />

more resistant to fungal pectolytic enzymes. The balance<br />

between N and Ca is also critical as excessive nitrogenous<br />

fertiliser can result in increased photosynthetic activity in leaves,<br />

outcompeting the developing fruit for water and Ca.<br />

Phenolic compounds are abundant in plants and have important<br />

roles as/in cellular support materials, eg. lignins, detoxification,<br />

components of flower and fruit colour (eg. anthocyanins),<br />

protection against herbivore predators, signal molecules, and as<br />

phytoalexins. Thus, they contribute to disease resistance<br />

mechanisms of plants.<br />

We investigated the effect of crop load on anthracnose and stem<br />

end rot (caused primarily by Botryosphaeria spp. but also C.<br />

gloeosporioides) postharvest diseases in ‘Hass’ avocado in two<br />

field seasons, and measured cation concentrations (particularly<br />

Ca and N) and total soluble phenolic acid levels in peel to<br />

determine associations with disease levels.<br />

MATERIALS AND METHODS<br />

“Hass” avocado fruit was harvested from trees in commercial<br />

orchards in northern NSW determined to have ‘high’ or ‘low’<br />

crop loads, in 2007 and 2008. Peel samples were collected from<br />

sub‐samples for cation analyses, and at harvest, ‘sprung’ (when<br />

fruit first start to soften) and ‘eating ripe’ for total soluble<br />

phenolic acid contents. Other fruit samples were maintained in a<br />

controlled environment room (22–23°C, 65% RH), and assessed<br />

at eating ripe stage for anthracnose and stem end rot diseases.<br />

Dried peel samples were finely ground and analysed for major<br />

cations by SGS Agritech. Phenolic acid contents were determined<br />

by using the Folin‐Ciocalteau reagent on samples extracted with<br />

50% v/v methanol, and compared against a gallic acid standard<br />

curve.<br />

RESULTS AND DISCUSSION<br />

In both years, the incidence of fruit with anthracnose disease<br />

was significantly less when harvested from trees with high crop<br />

loads (Table 1). Severity of disease was also less, but not<br />

significantly. Conversely, stem end rot, caused primarily by<br />

Botryosphaeria spp., was more severe in fruit from high crop<br />

bearing trees (significant in 2008, Table 2). The trees were<br />

drought stressed, which is thought to exacerbate stem end rot<br />

diseases in mango and avocado, and the greater crop load most<br />

likely added to this stress. There was a higher percentage of<br />

marketable fruit from high crop load trees (data not shown).<br />

Table 1. Severity and incidence of anthracnose in “Hass” fruit from high<br />

and low crop bearing trees in 2007 and 2008<br />

Crop load<br />

% severity<br />

anthracnose<br />

% incidence anthracnose<br />

2007 2008 2007 2008<br />

High 11.1 34.2 27.1 b 78.8 b<br />

Low 19.5 53.6 50.3 a 90.0 a<br />

Table 2. Severity and incidence of stem end rot in “Hass” fruit from high<br />

and low crop bearing trees in 2007 and 2008<br />

% severity<br />

stem end rot<br />

% incidence<br />

stem end rot<br />

Crop load 2007 2008 2007 2008<br />

High 10.5 3.83 a 33.5 16.5<br />

Low 4.8 1.42 b 27.7 8.8<br />

Cation analyses show that peel from fruit harvested in 2008 from<br />

high crop load trees had significantly higher calcium, lower N:Ca<br />

ratio, and higher Ca+Mg:K ratio than from fruit from low crop<br />

load trees. This is consistent with what was previously known, ie.<br />

that high Ca and low N is associated with better quality fruit (2).<br />

Table 3. Effect of crop load on major cations and their ratios in ‘Hass’<br />

avocado peel at harvest, 2008<br />

Crop Content (% dry wt of fruit peel) N:Ca Ca+Mg:K<br />

load N Ca Mg K ratio ratio<br />

High 0.792 0.036 a 0.082 1.078 22.3 b 0.110 a<br />

Low 0.868 0.028 b 0.078 1.232 31.8 a 0.087 b<br />

The results for soluble phenolics in these trials did not indicate<br />

that they were influenced by crop load, and no clear associations<br />

can be made between severity and incidence of postharvest<br />

disease and total soluble phenolic acid content. There was,<br />

however, a clear association between phenolics and disease<br />

reaction and rootstock type in another study (unpublished).<br />

Fruit quality can thus be improved by optimising tree yield and<br />

nutrient concentrations, and reducing drought stress.<br />

REFERENCES<br />

1. Willingham SL, Pegg KG, Cooke AW, Coates LM, Langdon PWB,<br />

Dean JR (2001) Rootstock influences postharvest anthracnose<br />

development in ‘Hass’ avocado. Australian Journal of Agricultural<br />

Research 52: 1017–1022.<br />

2. Hofman PJ, Vuthapanich S, Whiley AW, Klieber A, Simons D (2002)<br />

Tree yield and fruit minerals concentrations influence ‘Hass’<br />

avocado fruit quality Scientia Horticulturae 92:113–123.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 145


Posters<br />

64 Hosts of citrus scab, brown spot and black spot in coastal NSW<br />

N.J. Donovan{ XE "Donovan, N.J." } A , P. Barkley B and S. Hardy C<br />

A NSW Department of Primary Industries, PMB 8, Camden, 2570, NSW<br />

B Citrus Australia Ltd, PO Box 46, Mulgoa, 2745, NSW<br />

c NSW DPI, Locked Bag 26, Gosford, 2250, NSW<br />

INTRODUCTION<br />

Characterisation of endemic pathogens, including knowledge of<br />

host range and pathotyping, is important for disease control<br />

programs and trade negotiations.<br />

Citrus scab is a serious leaf and fruit disease of lemons in coastal<br />

areas of NSW and Qld. Previous studies have reported six Elsinoë<br />

fawcettii pathotypes on citrus worldwide (1, 2). In Australia, the<br />

Tryon’s and “Lemon” pathotypes have been described (1) but<br />

the pathotype range may not have been well‐represented as all<br />

of the isolates studied were from one lemon producing area of<br />

NSW.<br />

Citrus brown spot (Alternaria alternata) affects fruit and foliage<br />

of mandarins, tangelos and tangors in the humid coastal regions<br />

of Australia and is rarely found on grapefruit. In Florida, isolates<br />

sampled from grapefruit and the hybrid cv. Nova were<br />

genetically distinct from isolates sampled from other hybrid<br />

cultivars including Minneola tangelo and Murcott tangor (3).<br />

Citrus black spot (Guignardia citricarpa) is a serious disease of<br />

Valencia and navel oranges in coastal areas of eastern Australia.<br />

A non‐pathogenic species G. mangiferae has a wider host range<br />

which includes citrus.<br />

The aim of this study was to observe the incidence of scab,<br />

brown and black spots in an old citrus germplasm collection<br />

containing some rare species. The trees had not been sprayed<br />

for several years. By expanding the host varieties observed<br />

additional pathotypes may be found.<br />

MATERIALS AND METHODS<br />

Surveys were conducted in 1999 and 2009 in a citrus arboretum<br />

at NSW DPI’s Gosford Horticultural Institute. Symptoms on fruit<br />

and foliage were recorded for scab, black spot and brown spot.<br />

Further work will be conducted to characterise and identify<br />

pathotypes.<br />

RESULTS AND DISCUSSION<br />

Survey findings are presented in Table 1. Citrus scab was not<br />

observed on Satsuma mandarin, in contrast to overseas studies<br />

(2). Wotton rough lemon was the only rough lemon clone not to<br />

show symptoms of scab. Leaves of this clone are typical of rough<br />

lemon, but the fruits are acharacteristic, suggesting that it is a<br />

hybrid.<br />

The surveys found no evidence of brown spot affecting<br />

grapefruit, even though symptoms have been seen on a red<br />

grapefruit tree in Qld adjacent to badly affected Minneola<br />

tangelo trees.<br />

Sour orange is reportedly not susceptible to black spot, but one<br />

clone of smooth seville (a sour orange hybrid) in the arboretum<br />

showed symptoms. Lemons are often a preferred host of G.<br />

citricarpa with infections in new regions often occurring first on<br />

lemons. Infection was severe on the foliage of a number of<br />

lemon varieties and hybrids and on the rootstocks Troyer<br />

citrange and Swingle citrumelo, but not on C. trifoliata. All<br />

species/clones except C. trifoliata showed melanose (Diaporthe<br />

citri) symptoms to varying degrees including the native finger<br />

lime C. australasica.<br />

REFERENCES<br />

1. Timmer LW, Priest M, Broadbent P, Tan MK (1996) Morphological<br />

and pathological characterisation of species of Elsinoë causing scab<br />

diseases of citrus. Phytopathology 86, 1032‐1038<br />

2. Hyun JW, Yi SH, MacKenzie SJ, Timmer LW, Kim KS, Kang SK, Kwon<br />

HM, Lim HC (2009) Pathotypes and genetic relationship of<br />

worldwide collections of Elsinoë spp. causing scab diseases of<br />

citrus. Phytopathology 99,721‐728<br />

3. Peever TL, Olsen L, Ibanez A, Timmer LW (2000) Genetic<br />

differentiation and host specificity among populations of Alternaria<br />

spp. causing brown spot of grapefruit and tangerine x grapefruit<br />

hybrids in Florida. Phytopathology 90, 407‐414<br />

Table 1. Survey findings for scab and other fungal pathogens in the citrus arboretum, Narara NSW<br />

Citrus species 1<br />

Common name<br />

Varieties with scab symptoms<br />

observed<br />

Varieties with brown spot<br />

symptoms<br />

Varieties with black spot<br />

symptoms<br />

C. reticulata Blanco mandarin tangerine ex China Shekwasha, Szinkom, Ladu, Cleopatra, Batanges<br />

Cleopatra, Batanges, Emperor,<br />

Sunki, Satsuma<br />

C. x microcarpa Bunge calamondin calamondin<br />

C. × insitorum Mabb citrange Troyer citrange, Swingle<br />

citrumelo<br />

C. × aurantium L.<br />

(= C. aurantium L. and<br />

C. sinensis (L.) Osbeck)<br />

sour, sweet, Valencia<br />

and navel oranges,<br />

grapefruit & King<br />

orange<br />

Taiwanica<br />

tangelo (Sampson, Minneola,<br />

Orlando, Yalaha, San Jacinto,<br />

Wekiwa, Seminole, Thornton,<br />

Sexton)<br />

smooth Seville (Waddell),<br />

San Jacinto tangelo, Shunkokan<br />

C. × limon (L.) Osbeck lemon lemon (Volkamer, Lisbon,<br />

Eureka, Lemonade, Yen Ben,<br />

Villafranca, Thornless, Meyer,<br />

Assam), Rangpur lime<br />

C. × taitensis Risso<br />

rough lemon<br />

rough lemon (Settree, Wilson,<br />

(= × jambhiri Lush.)<br />

Narara)<br />

C. × aurantiifolia (Christm.) lime<br />

lime (West Indian, Kusiae, acid,<br />

Swingle<br />

accession 3233)<br />

C. × indica Tanaka Indian wild orange Indian wild orange<br />

1. Classification is according to Mabberley DJ (1997) A classification for edible citrus. Telopea 7, 167‐72o<br />

rough lemon (Settree, Wilson,<br />

Narara, Wotton)<br />

Kusaie lime<br />

lemon (Volkamer, Lisbon,<br />

Eureka, Lemonade, Yen Ben,<br />

Villafranca, Thornless, Meyer),<br />

Rangpur lime<br />

rough lemon (Settree, Wilson,<br />

Wotton)<br />

146 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


65 Nitrogen form affects Spongospora subterranea infection of potato roots<br />

Richard E. Falloon{ XE "Falloon, R.E." } A,B , Denis Curtin A , Ros A. Lister A , Ruth C. Butler A , Catherine L. Scott A and Nigel S. Crump C<br />

A New Zealand Institute for <strong>Plant</strong> and Food Research Limited, PB 4704, Christchurch, New Zealand<br />

B Bio‐Protection Research Centre, PO Box 84, Lincoln University, Canterbury, New Zealand<br />

C DPI Victoria, Knoxfield Centre, Private Bag 15, Ferntree Gully Delivery Centre, Victoria 3156, Australia<br />

Posters<br />

INTRODUCTION<br />

Powdery scab of potato tubers (Solanum tuberosum) is caused<br />

by the plasmodiophorid pathogen Spongospora subterranea f.<br />

sp. subterranea. The disease is important where potatoes are<br />

grown under intensive management, as it causes severe<br />

reductions in quality of seed and ware potatoes from affected<br />

crops (1). The pathogen can also infect potato roots, causing<br />

root galls and reducing plant growth (1). Manipulation of soil<br />

nutrients could be part of integrated powdery scab<br />

management. Nitrogen (N)‐containing amendments have been<br />

shown to reduce (2) and increase (3) powdery scab in fieldgrown<br />

potatoes.<br />

We present results from an experiment that aimed to determine<br />

effects of different rates and types (nitrate or ammonium) of N<br />

compounds on infection of potato plant roots by S. subterranea.<br />

MATERIALS AND METHODS<br />

The experiment was carried out in a glasshouse compartment<br />

(17ºC ± 2ºC; 16 h light, 8 h dark). Tissue‐cultured potato<br />

plantlets (cv. Iwa; very susceptible to powdery scab) were<br />

planted into a 50:50 w:w mix of field soil and coarse sand (>1<br />

mm) in plastic pots (11 cm diam., 680 ml capacity). The soil in<br />

each pot was irrigated with deionised water (by weight) to 90%<br />

water holding capacity three times each week for 8 weeks.<br />

Two weeks after planting, treatments of three N compounds<br />

(ammonium nitrate (NH 4 NO 3 ), ammonium sulphate ((NH 4 ) 2 SO 4 ),<br />

calcium nitrate (CaNO 3 ) 2 .4H 2 O)) were each applied to separate<br />

pots as solutions at five different rates, calculated to apply<br />

equivalent amounts of N. The rates were 0.05, 0.10, 0.20, 0.40,<br />

or 0.60 g N pot ‐1 , equivalent to 62.5, 125, 250, 500 and 750 kg N<br />

ha ‐1 . At the same time the pots were each inoculated with<br />

suspensions of S. subterranea sporosori (30,000 pot ‐1 ). Two<br />

control treatments of no added N with or without inoculum<br />

were also applied. The experiment included 17 treatments<br />

(three N compounds, five rates of each, plus two controls), and<br />

was of randomised complete block design with seven replicates.<br />

The plants were harvested 8 weeks after planting. Each plant<br />

was carefully washed free of soil, the number of S. subterranea<br />

root galls was counted and root dry weight (10 h at 70ºC)<br />

determined. Data were transformed (square root) to stabilise<br />

variances and analysed with ANOVA.<br />

RESULTS AND DISCUSSION<br />

Fig. 1 summarises data of severity of S. subterranea galling on<br />

the roots of harvested plants. No galls were observed on<br />

uninoculated plants, while plants from the nil N inoculated<br />

treatment had a mean of 80.5 galls g ‐1 root. All of the N<br />

treatments reduced root galling. Increasing rates of N for both<br />

ammonium sulphate and ammonium nitrate gave decreasing<br />

numbers of root galls. For calcium nitrate, however, increasing<br />

rate had little effect on root galling. At the three lowest rates in<br />

N, of the three N‐containing compounds, ammonium sulphate<br />

gave the greatest reduction in root galling.<br />

These results indicate that ammonium‐N is more inhibitory to S.<br />

subterranea infection of potato roots than nitrate‐N. They also<br />

suggest that increased powdery scab in the field after addition of<br />

high rates of N fertiliser (3) were unlikely to be due to direct<br />

effects of N on the pathogen, but may have been caused by<br />

indirect host growth effects (e.g. increased root mass resulting in<br />

increased amounts of zoospore inoculum).<br />

Mean galls per g root dry weight<br />

80<br />

50<br />

20<br />

10<br />

0<br />

0.00 0.05 0.10 0.20 0.40 0.60<br />

N (g pot -1 )<br />

Inoculated Control<br />

Uninoculated Control<br />

NH 4 NO3<br />

(NH 4 )2SO4<br />

Ca(NO 3 )2<br />

Figure 1. Mean numbers of Spongospora subterranea root galls on<br />

potato plants grown in pots treated with different amounts of N‐<br />

containing compounds. Bar = LSD (P=0.05) for visual comparison of the<br />

means (square root transformed scale).<br />

Ammonium‐N is usually converted to nitrate by soil microorganisms<br />

soon after application (4). It is likely, therefore, that<br />

the inhibitory effect of ammonium on S. subterranea occurred<br />

during early host infection stages, possibly affecting zoospore<br />

release from sporosori and/or infection of host roots.<br />

These results suggest that ammonium‐N may usefully reduce S.<br />

subterranea infection of potato. This should be confirmed in<br />

field evaluations in crops of potatoes grown in soil naturally<br />

infested with the pathogen.<br />

ACKNOWLEDGEMENTS<br />

The NZ Foundation for Research, Science and Technology and<br />

HAL (through the Australian Processing Potato Research<br />

Programme) funded this research.<br />

REFERENCES<br />

1. Merz U, Falloon RE (2009) Review: powdery scab of potato—<br />

increased knowledge of pathogen biology and disease<br />

epidemiology for effective disease management. Potato Research<br />

52: 17–37.<br />

2. Falloon RE (2008) Control of powdery scab of potato; towards<br />

integrated disease management. American Journal of Potato<br />

Research 85: 253–260.<br />

3. Falloon RE et al. (2007) Nitrogen fertiliser increases powdery scab<br />

incidence and severity; work in progress. 2nd European Powdery<br />

Scab Workshop, Langnau, Switzerland, 29–31 Aug, 2007.<br />

www.spongospora.ethz.ch/EUworkshop07/index.htm<br />

5. Haynes RJ, Williams PH (1992) Changes in soil solution composition<br />

and pH in urine‐affected areas of pasture. Journal of Soil Science<br />

43: 323–334.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 147


Posters<br />

66 Relationships between Spongospora subterranea DNA in field soil and powdery<br />

scab in harvested potatoes<br />

Farhat A. Shah A , Richard E. Falloon{ XE "Falloon, R.E." } A,B , Ros A. Lister A , Ruth C. Butler A , Alan McKay C , Kathy Ophel‐Keller C and Ikram<br />

Khan A<br />

A New Zealand Institute for <strong>Plant</strong> and Food Research, PB 4704, Christchurch, New Zealand<br />

B Bio‐Protection Research Centre, PO Box 84, Lincoln University, Canterbury, New Zealand<br />

C South Australian Research and Development Institute, GPO Box 397, Adelaide 5001, South Australia<br />

INTRODUCTION<br />

Powdery scab (caused by Spongospora subterranea f. sp.<br />

subterranea) is an important disease of potato (Solanum<br />

tuberosum). This disease is difficult to control, partly because S.<br />

subterranea can survive in soil for many years (1). Molecular<br />

detection and quantification of the pathogen in soil are possible<br />

components of disease management, to indicate pre‐planting S.<br />

subterranea inoculum levels and powdery scab risk (1).<br />

per ha.). In this study, where large pre‐planting quantities of S.<br />

subterranea DNA occurred in soil, DNA quantification did not<br />

accurately predict incidence or severity of powdery scab in<br />

harvested tubers.<br />

100<br />

90<br />

A<br />

We measured S. subterranea DNA levels in soil from a naturally<br />

infested field at planting and powdery scab in subsequently<br />

harvested tubers, over two growing seasons. Relationships<br />

between pre‐planting soil DNA and disease on harvested tubers<br />

were examined.<br />

MATERIALS AND METHODS<br />

The field area (0.36 ha) for this study had been previously used<br />

as a trial during the 2006/07 growing season. Twelve treatments<br />

(two cropping histories, three nitrogen fertiliser application<br />

rates, two irrigation regimes) were applied to potatoes grown in<br />

96 plots, each 5 × 5 m. The trial was of split split plot design with<br />

eight replicates (2). The treatments resulted in different levels of<br />

powdery scab in each plot (April 2007). During two subsequent<br />

growing seasons (2007/08 and 2008/09), the same plots were<br />

planted with cv. Agria (very susceptible to powdery scab) in<br />

October, in rows (2 m long; eight tubers/row) centrally in the 5 ×<br />

5 m plots. Soil samples were taken from each row at planting<br />

and analysed for S. subterranea DNA, using quantitative PCR<br />

techniques (3). Resulting tubers were harvested in April, washed<br />

free of soil and individually assessed for powdery scab severity (0<br />

= no disease, 1 = 5% tuber surface affected, 2 = 20%, 3 = 46%, 4 =<br />

60%). Relationships between the S. subterranea DNA in soil and<br />

powdery scab incidence and severity were explored graphically<br />

and with linear correlations (Pearson’s r).<br />

Mean score<br />

% Tubers infected<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

B<br />

RESULTS AND DISCUSSION<br />

2007/08 growing season. Powdery scab incidence in the plots<br />

varied from 30 to 100%, and mean severity score varied from 0.3<br />

to 2.7 (equivalent to 2 to 40% of tuber surface area affected).<br />

The relationships between S. subterranea DNA quantities in soil<br />

and powdery scab incidence (r = 0.53) and severity (r = 0.63) are<br />

illustrated in Figure 1.<br />

2008/09 growing season. Powdery scab incidence in the plots<br />

varied from 4 to 83%, mean severity score varied from 0.04 to1.6<br />

(equivalent to 0.2 to 16% of tuber surface area affected) and<br />

amount of S. subterranea DNA varied from 58 to 1997 pg g ‐1 soil.<br />

The relationships between amount of DNA in soil and powdery<br />

scab incidence and severity were very poor (r = 0.02 and 0.07<br />

respectively).<br />

This study has shown moderate to poor correlations between S.<br />

subterranea DNA in soil sampled at the time of sowing and<br />

powdery scab in harvested tubers. These results were from a<br />

field where soil DNA quantities and powdery scab were assessed<br />

from 96 evenly spaced positions in a 0.36 ha area (≈ 270 samples<br />

1.8 2.0 2.2 2.4 2.6 2.8 3.0 3.2 3.4 3.6<br />

log 10<br />

pg DNA g -1 soil<br />

Figure 1. Relationships between S. subterranea DNA in soil at planting<br />

and powdery scab incidence (A) and mean severity score (B) for potatoes<br />

grown in field plots.<br />

ACKNOWLEDGEMENTS<br />

The NZ Foundation for Research Science and Technology and<br />

HAL (through the Australian Potato Research Program) funded<br />

this research.<br />

REFERENCES<br />

1. Merz U, Falloon RE (2009) Review: powdery scab of potato—<br />

increased knowledge of pathogen biology and disease<br />

epidemiology for effective disease management. Potato Research<br />

52: 17–37.<br />

2. Falloon RE, FA Shah, et al. (2007) Nitrogen fertiliser increases<br />

powdery scab incidence and severity; work in progress.<br />

Proceedings of the 2nd European Powdery Scab Workshop.<br />

www.spongospora.ethz.ch/EUworkshop07/index.htm<br />

148 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


3. Ophel‐Keller K, McKay A, Hartley D, Herdina, Curran J (2008)<br />

Development of a routine DNA‐based testing service for soilborne<br />

diseases in Australia. <strong>Australasian</strong> <strong>Plant</strong> <strong>Pathology</strong> 37: 243–253.<br />

Posters<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 149


Posters<br />

6 Bacterial canker of tomato: Australian diversity of Clavibacter michiganensis subsp<br />

michiganensis<br />

L.M. Forsyth{ XE "Forsyth, L.M." }, T. Crowe, A. Deutscher and L. Tesoriero<br />

NSW Department of Primary Industries, Elizabeth Macarthur Agricultural Institute, Woodbridge Rd, Menangle 2568<br />

INTRODUCTION<br />

Bacterial canker of tomato is an important disease in Australian<br />

tomato production, especially amongst the greenhouse industry.<br />

The disease is caused by systemic vascular infection of the<br />

bacteria Clavibacter michiganensis subsp. michiganensis (Cmm).<br />

Canker infection can result in yield reductions of up to 100%.<br />

Currently there are no effective chemical or biological control for<br />

canker, the only effective methods are the quarantine and<br />

eradication of infected material. The external symptoms of<br />

bacterial canker have altered over the past decades. Whether<br />

this is due to the use of newer tomato cultivars which react<br />

differently to infection or due to the introduction of new genetic<br />

strains of the bacteria is not known.<br />

It is possible that there are new strains present in Australia since<br />

Cmm can be seed borne and imported seed is generally<br />

untreated since the relaxation of Australian quarantine<br />

requirements in the early 1990s.<br />

MATERIALS AND METHODS<br />

Isolate collection. Isolates were collected from tomato growing<br />

areas around Australia with a particular focus on greenhouse<br />

tomatoes. International isolates have been sourced from the<br />

Belgium culture collection (BCCM) and from America.<br />

Genetic diversity. DNA was extracted from the isolates using the<br />

Qiagen DNeasy kit, before quantification and dilution. DNA<br />

fingerprinting was undertaken using the ERIC, BOX and REP PCR<br />

(1). Further analysis of the genomic internal transcribed region<br />

(ITS) was undertaken on all isolates using a combination of<br />

sequencing and PCR‐RFLPs (2).<br />

Pathogenic diversity. A range of tomato cultivars were used to<br />

compare pathogenicity of Cmm isolates selected based upon the<br />

genetic diversity results. Isolates were also screened against<br />

other solanaceous crop plants commonly grown including<br />

eggplant and capsicum.<br />

RESULTS<br />

Genetic diversity. Preliminary screening results using BOX, ERIC<br />

and REP primers have revealed differences amongst Australian<br />

Cmm isolates. Sequencing analyses of the ITS region of four<br />

selected isolates has shown some base changes allowing the<br />

development of PCR‐RFLP.<br />

Pathogenic diversity. All tomato cultivars examined showed high<br />

levels of susceptibility to Cmm, though symptom expression<br />

appears to be cultivar dependent. Of the isolates examined the<br />

majority were pathogenic, though there was at least one isolate<br />

which appears to be avirulent.<br />

DISCUSSION<br />

Cmm diversity. Early results from the DNA fingerprinting of the<br />

Australian isolates of Cmm has revealed genetic diversity.<br />

Further comparisons with international isolates and isolates<br />

from field‐grown tomatoes will help understanding of whether<br />

this diversity is reflected in the international diversity or<br />

localised population drift potentially due to the high selection<br />

pressure within the greenhouse environments.<br />

Preliminary analyses of the genetic diversity of the avirulent<br />

isolate has not revealed any distinct differences. The avirulent<br />

isolate is being further tested using pulse field gel<br />

electrophoresis to examine the presence of pathogenicity<br />

conferring plasmids and virulence genes previously described<br />

(3).<br />

Only limited pathogenic diversity has been observed amongst<br />

isolates of Cmm. Although there was some variation between<br />

tomato cultivars in symptom expression, all tomato cultivars<br />

assessed showed high levels of susceptibility to the majority of<br />

isolates and eventually died due to application of Cmm. The<br />

localised lesions which developed on the capsicum leaves did<br />

not spread systemically in these experiments, implying that in a<br />

controlled environment only direct contact with Cmm on the<br />

leaves will lead to disease. Further testing using Cmm isolates<br />

isolated from capsicum plants will be undertaken to determine<br />

whether more severe disease could develop.<br />

ACKNOWLEDGEMENTS<br />

This work was funded by ACIAR, NSW DPI, Horticulture Australia<br />

Ltd, AusVeg and the tomato consortium. Acknowledgement for<br />

technical assistance from J. Collins, K. Turton, and S. Austin.<br />

REFERENCES<br />

1. Versalovich J, Schneider M, De Bruijn FJ, Lupski JR (1994) Genomic<br />

fingerprinting of bacteria using repetitive sequence based<br />

polymerase chain reaction. Methods in Molecular and cellular<br />

biology 5, 25–40.<br />

2. Fegan M, Croft BJ, Teakle DS, Hayward AC, Smith GR (1998)<br />

Sensitive and specific detection of Clavibacter xyli subsp. xyli,<br />

causal agent of ratoon stunting disease of sugarcane, with a<br />

polymerase chain reaction‐based assay. <strong>Plant</strong> pathology, 47, 495–<br />

504.<br />

3. Meletzus D, Bermpohl A, Crier J, Eichenlaub R (1993) Evidence for<br />

plasmid encoded virulence factors in the phytopathogenic<br />

bacterium Clavibacter michiganensis subsp. michiganensis<br />

NCPPB382. Journal of Bacteriology, 175, 2131–2136.<br />

Experiments examining a wider host range of Cmm revealed that<br />

pathogenic isolates were able to infect the two capsicum<br />

cultivars examined resulting in small localised lesions on the leaf<br />

lamina where the inoculum was initially applied. No systemic<br />

infection was observed within the capsicum plants. No<br />

symptoms were observed on the eggplant cultivar used.<br />

150 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


67 Detection of Mycosphaerella fijiensis in the skin of ‘Cavendish’ banana<br />

R.A. Fullerton{ XE "Fullerton, R.A." } A and S.G. Casonato A<br />

A The New Zealand Institute for <strong>Plant</strong> and Food Research Limited, Private Mail Bag 92169, Auckland Mail Centre 1142, New Zealand<br />

Posters<br />

INTRODUCTION<br />

Mycosphaerella fijiensis, cause of black leaf streak, has a very<br />

slow incubation period compared with many other pathogens.<br />

The fungus infects leaves as they emerge from the plant. The<br />

first symptoms (rusty streaks) appear in the highly susceptible<br />

‘Cavendish’ types in 2–3 weeks, extending to over 35 days for<br />

the more resistant ‘Ducasse’, ‘Pahang and ‘Pisang Mas’. In some<br />

genotypes symptoms may not be expressed until the leaf begins<br />

to senesce (1). These patterns of host response show that the<br />

pathogen has the ability to survive without symptoms in the leaf<br />

for extended periods. There are no records of infection of the<br />

fruit of dessert bananas. A record of infection in plantain (2)<br />

shows that the fungus has the capacity to invade the skin of<br />

fruit. This study aimed to investigate whether the fungus can be<br />

present without symptoms in the skin of ‘Cavendish’ bananas.<br />

MATERIALS AND METHODS<br />

Source of fruit. Green, fully developed fruit were obtained from<br />

diseased ‘Cavendish’ plants in the field in Samoa. Additionally,<br />

‘green‐mature’ and ripe fruit were obtained from the local<br />

market.<br />

is a constant association between the red fleck symptom and<br />

M. fijiensis.<br />

In many cases the presence of M. fijiensis may have remained<br />

undetected because of overgrowth of the skin pieces by other<br />

faster growing fungi such as Colletotrichum spp., Cordana<br />

musae, Penicillium sp., Cladosporium sp., Curvularia sp.,<br />

Fusarium sp., Rhizopus sp., Trichoderma sp., and Pestalotiopsis<br />

sp.<br />

This study has shown that M. fijiensis can infect and survive<br />

without symptoms in the skin of ‘Cavendish’ banana. While only<br />

a very low recovery rate was achieved in this study, the<br />

incidence in nature may be much higher.<br />

–—– A ––– B – – C – – D –– E –– F – – G –– H – – I – – J<br />

Isolation protocol. Pieces of skin tissue approximately 5 mm<br />

square and 0.5–1.0 mm thick were excised aseptically and plated<br />

onto Potato Dextrose Agar (PDA) or V8 agar, both modified with<br />

streptomycin and penicillin (100 µg/ml of each). Plates were<br />

incubated under continuous white/near‐UV light and were<br />

examined microscopically after five days<br />

Overall, a total of 1040 skin pieces were taken from 60 fruit from<br />

13 different sources. Five pure cultures of putative M. fijiensis<br />

were returned to New Zealand under a Biosecurity New Zealand<br />

permit to confirm their identities.<br />

Identification of isolates. The identities of the cultures returned<br />

to New Zealand were confirmed by spore morphology,<br />

polymerase chain reaction (PCR) using species‐specific primers<br />

provided on a confidential basis by the Cooperative Research<br />

Centre (CRC) for Tropical <strong>Plant</strong> Protection, and by the<br />

sequencing of the internal transcribe spacer region of rDNA.<br />

RESULTS AND DISCUSSION<br />

Of the five fruit isolates returned to New Zealand, two were<br />

positively identified as M. fijiensis, with concurring results from<br />

spore morphology, PCR, and sequencing (100% homology to<br />

sequences of M. fijiensis in GenBank). PCR results are shown in<br />

Figure 1. Both isolates were obtained from different skin<br />

samples from the same fruit. Microscopical examination (in<br />

Samoa) of a cluster of hyphae on skin of a different fruit revealed<br />

dark hyphae strongly resembling the early stages of a stroma of<br />

M. fijiensis. The structure was insufficiently developed to obtain<br />

a positive identification based on morphology, and overgrowth<br />

by other fungi prevented its isolation into pure culture. In all<br />

three cases, (two confirmed M. fijiensis, one suspected), the<br />

fungus was associated with minute (~1 mm diameter) red,<br />

necrotic flecks on the surface of the skin. Red flecks were<br />

relatively common on many of the test fruit. Most did not yield<br />

any fungi and others were overgrown by various fungal species<br />

before the slow growing M. fijiensis would have had time to<br />

emerge. It cannot be determined from this study whether there<br />

Figure 1. Amplified Mycosphaerella fijiensis products of approximately<br />

1050 bp. Lane A: 100 bp ladder B: 88a; C: 88b; D: Mfb; E: Mfc; F: Myc#1;<br />

G: 589 yellow Sigatoka; H: 748 M. fijiensis (positive control); I: blank<br />

(negative control); J: 100 bp ladder. Arrow indicates 600 bp on 100 bp<br />

ladder.<br />

ACKNOWLEDGEMENTS<br />

We gratefully acknowledge the Samoa Ministry of Agriculture,<br />

Fisheries, Forests and Meteorology for access to the laboratories<br />

of the Nuu Crop Research Centre, and assistance in this study.<br />

We thank Dr Juliane Henderson of CRC for Tropical <strong>Plant</strong><br />

Protection for supplying primer sequences.<br />

REFERENCES<br />

1. Cedeño L, Carrero C, Quintero K. 2000. Identification of<br />

Mycosphaerella fijiensis as cause of specks on fruits of plantain cv<br />

Harton in Venezuela. Fitopatologia‐Venezolana. 1:6–10.<br />

2. Fullerton RA, Olsen TL. 1995. Pathogenic variability in<br />

Mycosphaerella fijiensis Morelet, cause of black Sigatoka disease in<br />

banana and plantain. New Zealand Journal of Crop and<br />

Horticultural Science. 23:39–48.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 151


Posters<br />

46 Two new books: Diseases of Fruit Crops in Australia and Diseases of Vegetable<br />

Crops in Australia<br />

Tony Cooke, Denis Persley and Susan House, Cherie Gambley{ XE "Gambley, C." }<br />

Queensland Primary Industries and Fisheries, Department of Employment, Economic Development and Innovation, 80, Meiers Road,<br />

Indooroopilly Qld 4068, Australia<br />

Accurate information on disease diagnosis and management is<br />

essential for sustainable crop production. Two new books,<br />

Diseases of Fruit Crops in Australia and Diseases of Vegetable<br />

Crops in Australia, provide comprehensive coverage of<br />

important diseases affecting the broad range of fruit and<br />

vegetables grown throughout Australia. Written in a practical,<br />

straight forward style, the text explains how to identify and<br />

manage each disease, describing the symptoms of the disease,<br />

its importance, the means of infection and spread, and disease<br />

management.<br />

Based on the highly regarded early 1990 editions of Diseases of<br />

Fruit Crops and Diseases of Vegetable Crops published by the<br />

Department of Primary Industries and Fisheries, Queensland,<br />

these new books have been extensively revised and expanded.<br />

Emphasis is placed on integrated disease management and<br />

diseases that are biosecurity threats to Australian fruit and<br />

vegetable production.<br />

The text is supported by quality colour images. The books will<br />

become new standard references in applied plant pathology in<br />

Australia for fruit and vegetable crops.<br />

FEATURES<br />

• Chapters are authored by experienced plant pathologists<br />

from throughout Australia.<br />

• Written in a straightforward style with a minimum of<br />

scientific terms.<br />

• Provides accurate information about significant diseases<br />

affecting major and specialty fruit crops and vegetable<br />

crops in Australian tropical and temperate regions.<br />

• Each disease is extensively illustrated with high quality<br />

colour photographs.<br />

• Contains a comprehensive glossary and provides up‐to‐date<br />

sources of further information.<br />

• Describes key exotic diseases that are biosecurity risks to<br />

Australian fruit and vegetable growers.<br />

Both books are being published by CSIRO Publishing (Landlinks<br />

Press) and are due for release in October 2009<br />

REFERENCES<br />

1. Persley Denis (1993) Diseases of Fruit Crops. Department of<br />

Primary Industries, Queensland.<br />

2. Persley Denis (1994) Diseases of Vegetable Crops. Department of<br />

Primary Industries, Queensland.<br />

152 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


71 Infection and host responses in interactions between melon and Colletotrichum<br />

lagenarium<br />

Posters<br />

Yonghong Ge{ XE "Ge, Y." } and David Guest<br />

Faculty of Agriculture, Food and Natural Resources, The University of Sydney, Sydney, 2006, NSW AUSTRALIA<br />

INTRODUCTION<br />

Melon is an economically important horticultural crop that is<br />

susceptible to anthracnose caused by Colletotrichum<br />

lagenarium. Two types of infections are caused by<br />

Colletotrichum species: intracellular hemibiotrophic invasion or<br />

subcuticular intramural necrotrophic invasion (1). However, the<br />

infection process of melon anthracnose caused by C. lagenarium<br />

remains unknown. This study of the compatible interaction<br />

between C. lagenarium and melon leaves investigated the<br />

infection process and monitored defence responses of the<br />

melon plant.<br />

MATERIALS AND METHODS<br />

Preparation of infected tissues. Seeds of rockmelon cv. Galaxy<br />

and Ultra sweet Miami (Terranova seeds Pty Limited, NSW,<br />

Australia) were grown in 10 cm plastic pots filled with UC potting<br />

mix in the glasshouse at 24oC, with illumination for 16 h. <strong>Plant</strong>s<br />

were watered daily and fertilised with Aquasol® weekly. Three<br />

week old seedlings were inoculated on the abaxial surface of the<br />

first leaf with a suspension of 106 conidia mL‐1 of C. lagenarium.<br />

Inoculated plants were maintained at 25oC, 100% relative<br />

humidity for 24 h, then returned to the glasshouse.<br />

Conidia attached and germinated on the leaf surface 6 hai (Fig.<br />

1A), and differentiated a germ tube at one tip 12 hai (Fig. 1B).<br />

Melanised appressoria were first observed 24 hai, sometimes<br />

formed directly from one tip of the conidium (Fig. 1C), or from<br />

the tip of the germ tube. Penetration pegswere observed 48 hai<br />

(Fig. 1D). By 72hai, epidermal cells of melon leaves had been<br />

penetrated and contained intracellular fungal structures<br />

comprising swollen, saccate infection vesicles with elongated<br />

neck regions (Fig. 1E). Infection vesicles enlarged and formed<br />

primary hyphae (Fig. 1E), and at this stage of host‐pathogen<br />

interaction, infected melon leaves were symptomless. Beyond<br />

72 hai, secondary hyphae developed from the primary hyphae<br />

and invaded surrounding tissues (Fig. 2F), and the infected<br />

melon leaves developed visible anthracnose symptoms. The<br />

results also indicated that the resistant and susceptible cultivars<br />

use the same infection process.<br />

Callose deposition around the infection sites was noted 48 hai in<br />

susceptible and resistant cultivars (Fig. 2, Fig. 3). Callose<br />

deposition was brighter and more intense in the resistant<br />

cultivar.<br />

Light microscopy. Leaf samples were collected at 6, 12, 24, 48,<br />

72, 96 hrs after inoculation. Decolourised sections were<br />

immersed in lactophenol for 1min and then stained with 0.025%<br />

aniline blue for 30 min. After staining, the tissues were rinsed<br />

(2!2min) in lactophenol and mounted in fresh lactophenol on<br />

glass slides for microscopy(2). Callose was visualized under UV<br />

after staining with aniline blue (3).<br />

RESULTS AND DISCUSSION<br />

Figure 2. (A) Light and (B) UV micrographs showing the accumulation of<br />

callose 48 hai of the first leaves of susceptible melon with C. lagenarium.<br />

(C) Light and (D) UV micrographs showing the accumulation of callose 48<br />

hai of resistant melon.<br />

a=appressorium; Ca=callose<br />

These results indicate that the infection process of C. lagenarium<br />

was intracellular hemibiotrophic invasion.<br />

ACKNOWLEDGEMENTS<br />

This research was supported by the Australian Centre for<br />

International Agricultural Research (ACIAR). We thank Suneetha<br />

Medis for technical assistance.<br />

Figure 1. Infection structures of C. lagenarium in the first leaves of<br />

rockmelon. (A) Ungerminated conidia on the leaf surface with stomata 6<br />

hai. (B) Conidia with germ tubes 12 hai. (C) Melanised appressoria 24 hai.<br />

(D) Appressorium with penetration peg 48 hai. (E) Formation of infection<br />

vesicle and primary hyphae 72 hai. (F) Formation of secondary hyphae 96<br />

hai.<br />

c=conidia;s=stomata; gt=germ tube; a=appressorium; pp=penetration<br />

peg; ph=primary hyphae;sh=secondary hyphae; iv=infection<br />

vesicle;hai=hour after inoculation.Bars=20μm.<br />

REFERENCES<br />

1. Bailey JA, O’Connell RJet al. (1992) Infection strategies of<br />

Colletotrichum species. In Colletotrichum: Biology, <strong>Pathology</strong> and<br />

Control (JA Bailey, and MJ Jeger, Eds), CAB International<br />

2. Latunde‐Dada AO, Bailey JAet al.(1997) Infection process of<br />

Colletotrichumdestructivum O’Gara from lucerne (Medicagosativa<br />

L.). European Journal of <strong>Plant</strong> <strong>Pathology</strong> 103, 35‐41.<br />

3. Borden S, Higgins VJ (2002) Hydrogen peroxide plays a critical role<br />

in the defence response of tomato to Cladosporium fulvum.<br />

Physiological and Molecular <strong>Plant</strong> <strong>Pathology</strong> 61, 227‐236.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 153


Posters<br />

70 Disease‐management strategies for the rural sector that help deliver sustainable<br />

wood production from exotic plantations<br />

C. Beadle A , A. Rimbawanto B , A. Francis C , M. Glen{ XE "Glen, M." } A , D. Page C , C.L. Mohammed A,C<br />

A CSIRO Sustainable Ecosystems, Private Bag 12, Hobart, Tasmania 7001, Australia<br />

B Centre for Biotechnology and Tree Improvement, Yogyakarta, Indonesia<br />

C Tasmanian Institute of Agricultural Science and School of Agricultural Science, University of Tasmania, PB 54, Hobart, TAS 7001,<br />

Australia<br />

INTRODUCTION<br />

A training workshop and post‐workshop field trip was held in<br />

May 2009 in Indonesia to advance knowledge and understanding<br />

in disease management. It was opened by the Minister for<br />

Forestry and attended by a wide range of participants from the<br />

forest, oil palm and rubber industries, universities, and research<br />

and government agencies. The workshop was supported by<br />

international experts from South Africa, the United Kingdom and<br />

Australia.<br />

Case studies were used at the workshop to provide training and<br />

exposure for participants in concepts of forest pathology,<br />

biosecurity and forest health surveillance and their application<br />

towards developing strategies for disease management. These<br />

case studies focused on disease issues and threats of immediate<br />

relevance to tree crops in Indonesia, for example, fungal rot in<br />

hardwood plantations, rubber and oil palm, rust galling in<br />

Paraserianthes falcata, and the significance of a guava rust<br />

incursion. An exercise was carried out in the field to train in the<br />

basic concepts of ground based forest health surveillance. The<br />

field trip in Sumatra examined demonstration sites and<br />

experiments in acacia and eucalypt areas most severely affected<br />

by root rot, and included hands‐on experience in disease<br />

assessment and novel ways to examine disease risk.<br />

OUTPUTS OF TRAINING EXERCISE<br />

Position paper. The position paper focuses on using all available<br />

information about root and basal stem rot in plantation‐based<br />

industries to assist in capturing both the current status of forest<br />

disease management capacity in Indonesia and what type of<br />

capacity will be required to combat some very serious diseases<br />

of plantation crops. It first considers the background that has led<br />

to root rot and basal stem rot becoming diseases that have a<br />

significant economic effect on plantation‐based industries. A<br />

summary of the current size of the plantation estates in the oil<br />

palm, pulpwood and rubber industries is then provided. These<br />

sections give us an idea of the possible returns from investing in<br />

building capacity.<br />

The paper then collates information that has been collected<br />

from professional staff working in both the private and public<br />

sector in roles that are connected to disease management for oil<br />

palm, forestry (primarily pulpwood) species and rubber, and<br />

supports this information with that from published literature. A<br />

separate section considers the concept of ecosystem<br />

management—this research focus lies primarily in the public<br />

sector. Next there is a dissertation on biological control that<br />

examines the potential characteristics and development of<br />

control agents and the challenges that must be overcome to<br />

make them work. These three sections assist in highlighting the<br />

types of research and operational disease management capacity<br />

required.<br />

policy directions that are relevant to both this education and the<br />

application of disease management.<br />

Proceedings and DVD Disease Management Strategies in<br />

<strong>Plant</strong>ations. The Proceedings will summarise the information<br />

from the workshop from the various sessions (Introduction to<br />

Disease Management; Morphological and Molecular<br />

Identification Tools; Forest Health Surveillance; Biosecurity;<br />

Chemical, Genetic and Biological Control; Silvicultural and Risk<br />

Management; Integrated and Ecosystem Management; Policy<br />

Development).<br />

The DVD which contains all the talks from the Workshop is<br />

available on request and the Proceedings will be available in<br />

September.<br />

Field guide. A field guide to the identification of basidiomycete<br />

root rot diseases in tree crops will be published at the end of<br />

2009. This will include crown and root symptoms associated with<br />

the various stages of root rot disease and a description of the<br />

sporocarps associated with the various fungal pathogens capable<br />

of causing root rot disease.<br />

SUMMARY<br />

As in Australia there is little specific University training in forest<br />

pathology or disease management. <strong>Plant</strong> pathology education is<br />

comparatively well resourced in Indonesia, especially in Bogor,<br />

and the industries can draw from this pool of graduates. Barriers<br />

to building expertise in forest disease management lie in the fact<br />

that young people do not wish to live in remoter regions and<br />

there are often organisational barriers to the sharing of<br />

expertise and a collaborative approach to solving problems, even<br />

within the same industry.<br />

The Government of Indonesia has no formal approach to disease<br />

management in its forest policy. However the Minister has<br />

acknowledged the problem of disease, especially root‐rot<br />

disease (which is probably the most serious pest problem faced<br />

by the hardwood plantation industry) and is actively encouraging<br />

a cooperative approach to disease management. Substantial<br />

funding is potentially available to promote this collaboration and<br />

this supports the case for more open communication as was<br />

achieved by the workshop and field trip. Root rot disease has<br />

been a surprising catalyst for opening pathways of<br />

communication.<br />

ACKNOWLEDGEMENTS<br />

We thank the AUSAID Public Sector Linkage Programme for<br />

funding this activity. We also thank the many numerous<br />

Indonesian colleagues who participated in this activity.<br />

The current capacity to deliver professional services in disease<br />

management is then examined. The paper concludes with a<br />

consideration of the part of Indonesia’s higher education system<br />

that delivers training in <strong>Plant</strong> Protection and <strong>Plant</strong> <strong>Pathology</strong> and<br />

154 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


68 Rapid and robust identification of fungi associated with Acacia mangium root<br />

disease using DNA analyses<br />

Posters<br />

M. Glen{ XE "Glen, M." } A , V. Yuskianti B , A. Francis C , L. Agustini C,D , A. Widyatmoko B , A. Rimbawanto B and C.L. Mohammed A,C<br />

A CSIRO Sustainable Ecosystems, Private Bag 12, Hobart, Tasmania, Australia, 7001<br />

B Centre for Biotechnology and Tree Improvement, Yogyakarta, Indonesia<br />

C IAR, University of Tasmania, Private Bag 12, Hobart, Tasmania, Australia, 7001<br />

D Forest and Nature Conservation Research and Development Centre, FORDA, Bogor, Indonesia<br />

INTRODUCTION<br />

Indonesia, like many developing and developed countries, lacks<br />

people with experience in identifying root rot pathogens, both as<br />

sporocarps but more particularly, in culture. This increases the<br />

difficulty of managing Acacia mangium plantations which suffer<br />

severe economic losses due to root rot. This type of disease is<br />

caused by a variety of primary pathogens in Indonesian<br />

plantations. Existing forestry research laboratory facilities<br />

already use DNA techniques in plant genetics research and this<br />

capability was adapted to identification of fungi isolated from<br />

diseased roots of Acacia mangium.<br />

Several species of Ganoderma are associated with root rot in<br />

Acacia mangium (1). As part of ACIAR project FST/2003/048 a<br />

large number of fungal cultures were isolated from the roots of<br />

A. mangium and sporocarps collected from A. mangium<br />

plantations, with a view to determining the most prevalent and<br />

damaging root pathogens, elucidating their mode of dispersal<br />

and developing strategies for their management. Accurate<br />

isolate identification is a prerequisite for success in these aims.<br />

MATERIALS AND METHODS<br />

DNA was extracted and the rDNA ITS was amplified and<br />

sequenced (1). Isolates were grouped into Operational<br />

Taxonomic Units (OTUs) based on DNA sequence similarity. The<br />

OTU was identified by DNA sequence identity with herbarium<br />

collections where possible.<br />

Development of species‐specific primers targeting the rDNA ITS<br />

allowed faster and cheaper identification of two of the most<br />

prevalent species associated with root rot in Acacia mangium,<br />

Ganoderma philippii and G. mastoporum. Subsequent to the<br />

development of these specific primers all isolates isolated from<br />

roots or sporocarps were screened with these primers. This<br />

allowed faster identification and reduced the number of isolates<br />

sent for sequencing.<br />

RESULTS AND DISCUSSION<br />

Over 200 root rot isolates were confirmed as G. philippii or G.<br />

mastoporum either by DNA sequencing or species‐specific PCR<br />

(Figure 1). Another 120 cultures were grouped into 43<br />

operational taxonomic units (OTUs) by DNA sequence similarity.<br />

Eighteen OTUs were linked by DNA sequences to sporocarp<br />

collections, facilitating the morphological verification of culture<br />

identification.<br />

Figure 1. Species‐specific PCR for identification of Ganoderma philippii.<br />

Upper panel, PCR with primers Gphil2f/Gphil6r; lower panel, PCR with<br />

primers Gphil3f/Gphil4r. Lanes contain: 1, DNA size marker; 2–10, G.<br />

philippii isolates; 11–22, other Ganoderma spp.; 23–24, positive controls<br />

(G. philippii); 25, negative control (no DNA).<br />

G. philippii and G. mastoporum, G. aff. australe, G. aff.<br />

steyaertanum, G. subresinosum, G. aff. subresinosum, G.<br />

colossum, G. weberianum, Amauroderma rugosum and Phellinus<br />

noxius were isolated from A. mangium plantations. Isolates from<br />

diseased roots are predominantly G. philippii, with a low<br />

incidence of G. mastoporum and Phellinus noxius.<br />

Sporocarps of Fomes, Irpex, Phlebia, Trametes spp. have been<br />

collected and formally identified by DNA analysis. This study also<br />

discovered another fungal species that warrants investigation as<br />

a potential root‐rot biocontrol. Some cultures from roots were<br />

identified as belonging to the genus Phlebiopsis. Phlebiopsis<br />

gigantea has been demonstrated to be an effective prophylactic<br />

biocontrol for root rot caused by Heterobasidion annosum.<br />

Maintaining a large number of fungal isolates in tropical regions<br />

poses a challenge. Confident and rapid identification of fungal<br />

isolates has reduced the work required to maintain fungal<br />

isolates by allowing non‐target fungi to be discarded. This also<br />

reduces the risk of culture contamination by ‘weedy’ species.<br />

ACKNOWLEDGEMENTS<br />

This work was funded by ACIAR project FST/2003/048.<br />

REFERENCES<br />

1. Glen M, Bougher NL, Francis A, Nigg SQ, Lee SS, Irianto R, Barry KM,<br />

Beadle CL & Mohammed CL. (2009) Ganoderma and Amauroderma<br />

species associated with root‐rot disease of Acacia mangium<br />

plantation trees in Indonesia and Malaysia. <strong>Australasian</strong> <strong>Plant</strong><br />

<strong>Pathology</strong> 38, 345–356.<br />

Identification of the remaining OTUs is based on DNA sequence<br />

similarity to sequences from public DNA databases. Only one<br />

Ganoderma isolate has not been linked to a sporocarp collection.<br />

33 OTUs are linked to species/genus information in public DNA<br />

databases, providing an indication, at various taxonomic levels of<br />

species affinities.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 155


Posters<br />

69 Survey of the needle fungi associated with Spring Needle Cast in Pinus radiata<br />

I. Prihatini A , M. Glen{ XE "Glen, M." } B , A.H. Smith A , T.J. Wardlaw C and C.L. Mohammed A,B<br />

A Tasmanian Institute of Agricultural Science and the School of Agricultural Science, University of Tasmania, Private Bag 54, Hobart,<br />

Tasmania 7001, Australia<br />

B CSIRO Sustainable Ecosystems, Private Bag 12, Hobart, Tasmania 7001, Australia<br />

C Forestry Tasmania, 79 Melville St, Hobart, Tasmania 7000, Australia<br />

INTRODUCTION<br />

Spring needle cast (SNC) is currently classified as a serious<br />

disease of Pinus radiata growing in closed‐canopy stands on high<br />

altitude, wet sites in Tasmania (1). SNC affects about 30% of the<br />

Pinus radiata estate in Tasmania and causes the premature<br />

casting of needles at the end of their first year. This leads to<br />

growth reductions in direct proportion to the amount of<br />

defoliation. Stands with moderate or severe SNC can be<br />

expected to suffer potential losses in clearfall volume of 30–50%<br />

(1). Unlike other serious needle cast diseases elsewhere in<br />

Australia and New Zealand such as Dothistroma septospora or<br />

Cyclaneusma, SNC in Tasmania is not considered to be a classical<br />

needle blight disease caused by a primary fungal pathogen. It is<br />

thought to be caused by a suite of endophytic fungi that are<br />

triggered into secondary pathogenic activity by an<br />

environmental stress. Three fungal species are considered to<br />

play a role in Spring Needle Cast in Tasmania: Cyclaneusma<br />

minus, Lophodermium pinastri and Strasseria geniculata (2).<br />

The Pinus radiata Spring Needle Cast Marker Aided Selection<br />

(MAS) trial was planted in June 1999 by Forestry Tasmania in<br />

Oonah, North West Tasmania (annual rainfall: 1655mm/yr; mean<br />

daily temperature: 9.9 °C; altitude 450 metres). There are three<br />

full sib families with known breeding values for SNC (3).<br />

The objective of this study was to characterise the fungal<br />

communities associated with needles on trees scored for SNC<br />

damage in the MAS trial.<br />

MATERIAL AND METHODS<br />

The needle samples were collected in spring 2007 from the<br />

Oonah SNC MAS trial. The trees in this trial were scored for SNC<br />

severity immediately before sample collection. Trees were given<br />

a score ranging from 1 (no disease) to 4 (severe disease). For<br />

each disease score within each family, 3 trees were sampled for<br />

needles. Three different ages of needle were collected from a<br />

tree.<br />

DNA was extracted from needles and fungal DNA was amplified<br />

by PCR (4). PCR products from the same aged needles of trees in<br />

the same disease class and family (i.e. 3 trees) were pooled then<br />

cloned using a commercial kit. Thirty‐two colonies from each<br />

cloning reaction were randomly selected and screened using<br />

PCR‐RFLP to reduce the number of samples for sequencing.<br />

Approximately 12–16 clones from each set were sequenced.<br />

Sequences of high similarity were retrieved from public<br />

databases.<br />

RESULTS AND DISCUSSION<br />

PCR, cloning and DNA sequencing has been completed for<br />

samples in disease categories 1 to 3 (Table 1).<br />

Table1. Prevalence of fungal species detected in disease categories 1 (no<br />

disease) to 3 (moderate disease) for all families combined.<br />

Disease Category<br />

Species 1 2 3 4<br />

Allantophomopsis sp 2 1 2 0 na 2<br />

Catenulostroma sp 3 3 3 na<br />

Cyclaneusma sp 4 5 2 na<br />

Mycosphaerella sp 1 8 6 4 na<br />

Mycosphaerella sp 2 8 6 5 na<br />

Mycosphaerella pini 5 6 2 na<br />

Lophodermium pinastri 0 1 4 na<br />

Phoma sp 1 2 1 na<br />

Tumularia sp 5 3 0 na<br />

1 the number of samples out of 9 in which a species was detected<br />

2 Data to be presented on poster.<br />

In this study Mycosphaerella species 1 and and 2 were common<br />

to all three disease categories. Mycosphaerella pini was common<br />

to classes 1 and 2 but less frequently detected in class 3.<br />

Cyclaneusma sp. was not clearly correlated with disease<br />

incidence in the three disease categories analysed.<br />

Lophodermium pinastri was found more frequently in needles<br />

from trees with a higher disease severity.<br />

From the data so far collated, no clear association of any fungal<br />

species with disease incidence is evident. Further data analysis<br />

will be needed to study the correlation of fungal species with the<br />

host genetics.<br />

ACKNOWLEDGEMENTS<br />

Australian Research Council, Forestry Tasmania, Rayonier,<br />

Taswood Growers, Norske‐Skog, Forests NSW, Hoskings Ltd.<br />

New Zealand. Istiana Prihatini is the recipient of a John Allwright<br />

Fellowship, ACIAR.<br />

REFERENCE<br />

1. Dick MA, Somerville JG, Gadgil, PD (2001). Variation in the fungal<br />

population in Cyclaneusma needle‐cast in New Zealand. In 'Forest<br />

Reseach Bulletin'. (Ed. LS Bulman) pp. 12–20)<br />

2. Wardlaw T (2008) A review of the outcomes of a decade of forest<br />

health surveillance of state forests in Tasmania. Australian Forestry<br />

71, 254–260.<br />

3. Kube P, Piesse M (1999) 'Pinus radiata Spring Needle Cast Marker<br />

Aided Selection Trial.' Forestry Tasmania, RP251/9.<br />

4. Glen M, Tommerup IC, Bougher NL, and O' Brien PA (2002) Are<br />

Sebacinaceae common and widespread ectomycorrhizal associates<br />

of Eucalyptus species in Australian forests? Mycorrhiza 12, 243–<br />

247.<br />

156 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


7 A new report on Pseudomonas syringae pv. mori causal agent of bacterial blight of<br />

mulberries in Australia<br />

Posters<br />

H. Golzar{ XE "Golzar, H." } and P. Mather<br />

Department of Agriculture and Food Western Australia, Bentley Delivery Centre, WA 6983, Australia<br />

INTRODUCTION<br />

Pseudomonas syringae pv. mori (Boyer & Lambert) Young et al.<br />

causes a leaf spot and blight of young shoots on mulberry. It has<br />

been a common disease of mulberry worldwide. Symptoms<br />

appear as small water‐soaked leaf spots, turning brown or black,<br />

sometimes with a yellow halo. Spots on the midribs and vines<br />

are sunken. Infected leaves are often become distorted and<br />

bacterial ooze may be extruding from lenticels. Young shoots<br />

may show rapid necrosis and cankers (Fig. 1). The disease has<br />

been reported on mulberry (1) however, there is no official<br />

record of P. syringae pv. mori in Australia.<br />

MATERIALS AND METHODS<br />

Leaf spots and blight of young shoots were observed on<br />

mulberry trees in east of Perth in the summer of 2008 (Fig.1).<br />

Samples of leaf and infected branches were collected. Isolations<br />

were made from lesions on the leaf and stem tissues. Isolates<br />

with positive hypersensitivity on tobacco (Nicotiana glutinosa)<br />

were used for further tests. The bacterial isolates were identified<br />

based on biochemical tests (2) and using the Biolog identification<br />

system based on the carbon utilisation microplate assay (Biolog<br />

MicroLog 4.0 System, Biolog Inc., Hayward, CA).<br />

Pathogenicity tests. To confirm identification of the bacterial<br />

strains, pathogenicity tests with two isolates were performed on<br />

immature lemon, pear fruits, young bean pods and tomato<br />

seedlings (3). Isolates were grown on sucrose peptone agar (SPA)<br />

for 24 h and then suspended in sterile water and diluted to a<br />

concentration of 10 6 CFU/ml. Fruits and bean pods were surface<br />

sterilised with alcohol, washed with sterile water and inoculated<br />

by placing drops of an aqueous bacterial suspension on the<br />

surface and pricked through the drops using sterile needles.<br />

Controls were inoculated with sterile water. After inoculation,<br />

fruits and bean pods were incubated in the moist trays at 25°C<br />

for 7 days. Four‐week‐old healthy tomato plants were inoculated<br />

using the same bacterial inocula, controls and techniques. <strong>Plant</strong>s<br />

were placed under mist for 48 h and then moved to the<br />

growthroom chamber at 22 ± 1°C. Disease symptoms were<br />

checked 7 days post‐inoculation.<br />

To test pathogenicity of the same P. syringae pv. mori isolates on<br />

Malus alba, young shoots and detached leaves were inoculated<br />

by placing drops of bacterial suspension (10 6 cfu/ml) on freshly<br />

wounded shoot and midrib tissues. Controls were inoculated in<br />

the same way using sterile water and then incubated in moist<br />

trays at 25°C.<br />

RESULTS<br />

A bacterial blight was found on mulberries in east of Perth in the<br />

summer of 2008. White‐coloured and fluorescent bacterial<br />

colonies were consistently isolated from the leaf and stem<br />

tissues. Isolates were gram negative, fluorescent on King's<br />

medium B, oxidase negative, catalase positive, potato soft rot<br />

negative, arginine dihydrolase negative and tobacco HR positive.<br />

The representative isolates were tested using the biolog system<br />

and were identified as P. syringae pv. mori with a probability<br />

range of 96 to 100%.<br />

The isolates caused necrosis of the shoots and tissue along the<br />

midribs of the mulberry leaves 7 days after inoculation (Fig.2).<br />

The isolates also caused water soaked lesions on the bean pods,<br />

pear and lemon fruits, although they did not produce disease<br />

symptoms. In leaves of tomato inoculated by pricking,<br />

chlorosis areas were seen 7 days after inoculation. Koch's<br />

postulates were fulfilled and reisolated bacterial colonies were<br />

identified as P. syringae pv. mori. Culture of P. syringae pv. mori<br />

has been deposited in the WA culture collection as WAC 13254.<br />

To our knowledge, this is the first official report of P. syringae pv.<br />

mori on mulberry in Australia.<br />

Figure 1. Disease symptoms; leaf spots (a) and blight of a young shoot<br />

(b). Bars = 1cm.<br />

Figure 2. Pathogenicity test; Mulberry leaves showing necrosis<br />

symptoms (a), in comparison with a healthy leaf (b)<br />

REFERENCES<br />

a<br />

a<br />

1. Janse JD (2006) Pseudomonas syringae pv. mori. In<br />

‘Phytobacteriology Principles and Practice’. pp. 212–213. (CAB<br />

Publishing. UK)<br />

2. Lelliott RA, Stead DE (1987) Methods in <strong>Plant</strong> <strong>Pathology</strong> Vol. 2:<br />

Methods for the Diagnosis of Bacterial Diseases of <strong>Plant</strong>s. Blackwell<br />

Scientific Publications. Oxford, UK.<br />

3. Scheck HJ, Canfield ML, Pscheidt JW, Moore LW (1997) Rapid<br />

evaluation of pathogenicity in Pseudomonas syringae pv. syringae<br />

with a lilac tissue culture bioassay and syringomycin DNA probes.<br />

<strong>Plant</strong> Dis 81, 905–910.<br />

b<br />

b<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 157


Posters<br />

47 Through chain assessment and integrated management of brown rot risks in<br />

stonefruit<br />

R. Holmes, O. Villalta, S. Kreidl, M. Hossain and C. Gouk{ XE "Gouk, C." }<br />

Department of Primary Industries, Knoxfield, Private Bag 15 Ferntree Gully DC, Vic, 3156 Australia<br />

INTRODUCTION<br />

Brown rot caused by Monilinia fructicola (G. Wint.) Honey is the<br />

major disease challenge for stonefruit growers and supply chain<br />

businesses in Australia. Crop losses are attributed to blossom<br />

blight and fruit rots, and occur despite fungicidal sprays applied<br />

by growers. Infection can be controlled with well‐timed,<br />

effective fungicides (1); however, growers lack access to sitespecific<br />

weather data and disease models to support control.<br />

M. fructicola may infect during flowering, fruit development and<br />

after harvest, thus a through chain approach to disease<br />

management is required. There are many key management<br />

strategies including reducing inoculum potential, predicting<br />

infections, optimal timing of protectant and curative fungicides,<br />

understanding changes in host tissue susceptibility, controlling<br />

pests which vector the disease or assist infection and<br />

understanding the potential for postharvest disease. In this<br />

paper, we summarise our research in these areas and discuss the<br />

potential for their integration into a brown rot management<br />

strategy.<br />

MATERIALS AND METHODS<br />

Field sites. Trials were established in commercial orchards in the<br />

main Victorian stonefruit districts: 4 sites in 06/7, 7 in 07/8 and<br />

10 in 08/9. Each site had 6 plots of 10 trees. A weather station<br />

was placed in the centre of the 4th plot at each site. These<br />

recorded day length, rainfall, leaf wetness inside and outside the<br />

canopy, RH and air temperature.<br />

Inoculum for primary infection. In the first and second seasons,<br />

sites were surveyed during bloom for the presence of dried or<br />

mummified fruit in the trees and on the ground. Samples of<br />

these were collected (up to 20/plot) and moist incubated to<br />

detect M. fructicola.<br />

Weather based prediction of infection risk. A weather‐driven<br />

infection risk model (G Tate pers. comm.) was evaluated using<br />

data collected by the weather stations over 3 seasons. In the first<br />

season, surface wetness duration, a critical factor for infection<br />

risk was compared inside and outside tree canopies.<br />

Effectiveness of fungicide programs. The predicted occurrence<br />

and severity of infection periods were examined in relation to<br />

fungicides applied by growers and brown rot incidence after<br />

harvest. Over the three seasons, growers made incremental<br />

changes towards spraying in response to predicted infection<br />

periods. The success of this was evaluated.<br />

Influence of Carpophilus beetle populations on infection risk.<br />

At two sites, canning peach (var T204) blocks were treated with<br />

the carpophilus attract and kill system. Beetle populations and<br />

postharvest brown rot incidences were compared against<br />

untreated blocks.<br />

Phenological influence on infection risk. In the 08/09 season,<br />

peach fruit (vars Golden Queen and T204) were spray inoculated<br />

at early shuck fall, post pit hardening and 1–2 weeks prior to<br />

harvest. Brown rot development was monitored during the<br />

growing season and postharvest.<br />

A postharvest predictor of rot risk. At each site 20 fruit per plot<br />

were harvested at commercial maturity and moist incubated at<br />

21°C for 7 and 12 days to establish the level of latent infections<br />

leading to rots.<br />

RESULTS AND DISCUSSION<br />

The abundance of mummified fruit infected with M. fructicola<br />

did not exclusively explain the incidence of postharvest rot (2).<br />

Tate’s infection model was convenient for identifying periods of<br />

weather conducive to infection. However to make best use of<br />

the model it is necessary to understand a) the susceptibility of<br />

the crop at different phenological stages and b) the inoculum<br />

potential. In the first season of trials, inoculation of developing<br />

flowers and fruit at different growth stages did not reveal<br />

differences in tissue susceptibility and therefore, more<br />

comprehensive trials are planned.<br />

There was a strong positive relationship between the number of<br />

moderate and severe infection risk events in the two weeks<br />

before harvest and the postharvest rot incidence. Well timed<br />

fungicides during this period appeared to have suppressed<br />

infections.<br />

Flat plate sensors outside tree canopies generally recorded<br />

longer wetness intervals than sensors inside canopies, for both<br />

rain and dew events. This agrees with Henshall et al. (3) who<br />

showed this was the case in vineyards. Therefore wetness<br />

duration measured outside tree canopies will estimate greater<br />

disease risks (2).<br />

Controlling carpophilus significantly reduced postharvest rot<br />

incidence and more work is required to determine how this fits<br />

into a disease control strategy.<br />

Moist incubating samples of fruit collected a few days before<br />

commercial harvest can be used to estimate the risk of rots<br />

developing during storage, transport and marketing. Thus<br />

packers and distributors can appropriately treat and market high<br />

risk fruit into short supply chains, minimising wastage.<br />

ACKNOWLEDGEMENTS<br />

This project is supported by the Victorian Government,<br />

Summerfruit Australia, the Canned Fruit Industry Council and<br />

Horticulture Australia.<br />

REFERENCES<br />

1. Tate, K.G., Wood, P.N. and Manktelow, D.W. (1995). Development<br />

of an improved spray timing system for process peaches in Hawke’s<br />

Bay. Proceedings 48th N.Z. <strong>Plant</strong> Protection Conference, 101–106.<br />

2. Holmes, R., Villalta, O., Kreidl, S., Partington, D., Hodson, A. and.<br />

Atkins T A (2007) Weather‐based Model Implemented in HortPlus<br />

MetWatch with Potential to Forecast Brown Rot Infection Risk in<br />

Stone Fruit. Acta Hort. 803:19–27<br />

3. Henshall, W.R., Beresford, R.M., Chynoweth, R.W. and Ramankutty,<br />

P. 2005. Comparing surface wetness inside and outside grape<br />

canopies for region‐wide assessment of plant disease risk. New<br />

Zealand <strong>Plant</strong> Protection 58:80–83.<br />

158 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


48 Effects of temperature on mixed bunch rot infections of grapes<br />

L.A. Greer{ XE "Greer, L.A." }, S. Savocchia, S. Samuelian and C.C. Steel<br />

National Wine and Grape Industry Centre, Charles Sturt University, Locked Bag 588, Wagga Wagga, NSW 2678<br />

Posters<br />

INTRODUCTION<br />

Although bunch rot of grapes is frequently associated with<br />

Botrytis cinerea or grey mould, this pathogen can be absent from<br />

bunch‐rot affected vineyards under some climatic conditions. Of<br />

note is the occurrence of Ripe Rot and Bitter Rot caused by<br />

Colletotrichum acutatum and Greeneria uvicola respectively, in<br />

sub‐tropical vineyards that experience warm and wet conditions<br />

close to harvest. In Australia, Ripe Rot and Bitter Rot have been<br />

recorded in coastal regions such as the Hunter Valley (NSW),<br />

Kingaroy (QLD) and Carnarvon (WA).<br />

Previous studies have revealed that C. acutatum and G. uvicola<br />

were the predominant bunch rot pathogens isolated from<br />

berries collected at different phenological stages in the Hastings<br />

Valley (mid north coast NSW) (1), whereas isolation of B. cineria<br />

was infrequent. Both C. acutatum and G. uvicola can occur<br />

concurrently in the one vineyard and even on the one berry. In<br />

an attempt to explore factors leading to the absence of B.<br />

cinerea from some vineyards, we investigated the ability of C.<br />

acutatum, G. uvicola and B. cinerea to co‐infect berries at either<br />

20°C or 27°C.<br />

MATERIALS AND METHODS<br />

Detached, disease‐free Cabernet Sauvignon berries (22.4° Brix)<br />

were surface sterilised, rinsed in sterile water and placed into 24<br />

well microtitre plates. Berries were inoculated either singularly<br />

or with combinations of B. cinerea, C. acutatum and G. uvicola<br />

(10 μL droplet on the distal apex of the berry, 10 6 spores/mL)<br />

using three replicates per isolate with 24 berries per replicate.<br />

Berries were incubated for five days at either 20°C or 27°C in the<br />

dark at 100% RH. Berry colonisation was assessed by plating<br />

grape berries onto potato dextrose agar. Results were expressed<br />

as the mean percentage of berries infected.<br />

RESULTS AND DISCUSSION<br />

Grape berries were susceptible to infection by all three of the<br />

bunch rot pathogens examined. A higher percentage of berries<br />

were infected by B. cinerea at 20°C than at 27 ºC, while G.<br />

uvicola infection was favoured at 27ºC. There was little<br />

difference in the infection of grape berries by C. acutatum at<br />

either temperature (Table 1).<br />

The colonisation of grape berries by B. cinerea was not affected<br />

by co‐inoculation with either C. acutatum or G. uvicola at 20ºC<br />

but was reduced at 27ºC. Conversely the growth of C. acutatum<br />

and G. uvicola was reduced by co‐inoculation with B. cinerea at<br />

20ºC and not at 27ºC. G. uvicola failed to colonise any berries at<br />

20ºC when co‐inoculated with B. cinerea. C. acutatum also<br />

reduced berry infection by G. uvicola when co‐inoculated at<br />

either temperature. G. uvicola had no effect on berry<br />

colonisation by C. acutatum at either of the temperatures<br />

examined.<br />

vinifera cv. Cabernet Sauvignon berries (22.4º Brix). Control berries were<br />

water inoculated.<br />

Treatment<br />

% berries infected with<br />

Inoculum Temp ºC B. cinerea C. acutatum G. uvicola<br />

Control 20 0 0 0<br />

27 0 0 0<br />

Bc 20 92 ‐ ‐<br />

27 65 ‐ ‐<br />

Ca 20 ‐ 96 ‐<br />

27 ‐ 100 ‐<br />

Gu 20 ‐ ‐ 57<br />

27 ‐ ‐ 99<br />

Bc + Ca 20 82 65 ‐<br />

27 29 97 ‐<br />

Bc + Gu 20 95 ‐ 0<br />

27 46 ‐ 99<br />

Ca + Gu 20 ‐ 95 12<br />

27 ‐ 96 55<br />

Bc + Ca + + 20 88 28 0<br />

Gu 27 45 96 45<br />

B. cinerea is frequently associated with bunch rot of grapes in<br />

cool climates. Our results support earlier observations on the<br />

optimum climatic conditions for grey mould development (2).<br />

The sub‐tropical climatic conditions of regions experiencing Ripe<br />

Rot and Bitter Rot are likely to pre‐dispose berries to these<br />

diseases. Our additional observations (unpublished data) on the<br />

relative growth rates of the three pathogens on PDA, at a range<br />

of temperatures, support this hypothesis and may partially<br />

explain the absence of grey mould in sub‐tropical vineyards.<br />

ACKNOWLEDGEMENTS<br />

This work was supported by the Winegrowing Futures Program,<br />

a joint initiative of the Grape and Wine Research and<br />

Development Corporation and the National Wine and Grape<br />

Industry Centre.<br />

REFERENCES<br />

1. Steel CC, Greer LA, Savocchia S (2007) Studies on Colletotrichum<br />

acutatum and Greeneria uvicola: Two fungi associated with bunch<br />

rot of grapes in sub‐tropical Australia. Australian Journal of Grape<br />

and Wine Research 13, 23–29.<br />

2. Broome JC, English JJ, Marois BA, Latorre BA & Aviles JC (1995)<br />

Development of an infection model for Botrytis bunch rot of grapes<br />

based on wetness duration and temperature. Phytopathology 85,<br />

97–102<br />

These observations were further confirmed by inoculating grape<br />

berries with all three bunch rot pathogens at the same time. At<br />

20ºC B. cinerea was the pre‐dominant pathogen while at 27ºC C.<br />

acutatum predominated.<br />

Table 1. Effect of co‐inoculating Colletotrichum acutatum (Ca), Greeneria<br />

uvicola (Gu) and Botrytis cinerea (Bc) on disease expression in Vitis<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 159


Posters<br />

72 Genetic diversity of Iranian Fusarium oxysporum f. sp. ciceris by RAPD molecular<br />

markers<br />

Sara Haghighi{ XE "Haghighi, S." } 1 , Saeed Rezaee 2 , Bahar Morid 2 , Shahab Hajmansoor 2<br />

1 Islamic Azad University, Damghan Branch, Iran, sara_haghighi1026@yahoo.com<br />

2 Islamic Azad University, Science and Research Branch Tehran, Iran<br />

INTRODUCTION<br />

Fusarium oxysporum Schlecht, Emend (Snyder & Hansen) is one<br />

of the most important soilborn plant pathogens with a<br />

worldwide distribution. One of the most important crops in Iran<br />

is the Iranian chickpea (Cicer arietinum L.) with the annual<br />

production of 260,000 tons from 755,000 hectares. Fusarium<br />

wilt of chickpea is a devastating disease in chickpeas growing in<br />

different regions of Iran. This fungal disease is caused by<br />

Fusarium oxysporum f.sp. ciceri. Characteristic symptoms of<br />

disease are leaves necrosis, yellowing, vascular wilt and<br />

damping‐off (2).<br />

Iran is the world’s fourth important chickpea producing<br />

countries and, this pathogen can reduce yield about 15%, so an<br />

investigation of genetic diversity of this pathogen in the regions<br />

seems to be of great significance.<br />

MATERIALS AND METHODS<br />

Thirty isolates of Fusarium oxysporum f.sp. ciceri with different<br />

geographical origins were chosen for genetic diversity studies. In<br />

vitro pathogenicity tests were performed using a root‐dip assay,<br />

cluster analysis of the isolates classified into three categories of<br />

highly, moderate and weakly virulent groups. DNA extraction<br />

was performed using Readers & Borda method with few<br />

modifications (3). For RAPD analysis thirty random primers were<br />

screened and ten primers producing the highest number of<br />

scorable bands were selected for the final analysis (Table 1). 20<br />

ng of genomic DNA from each isolate was amplified with the<br />

selected primers. Amplified DNA was cluster analysed using<br />

MVSP software and UPGMA method with jaccard coefficient.<br />

Table 1. Sequence of primers used in this study.<br />

primers Sequences 5 –3<br />

1 CCG GCC TTA G<br />

2 ACC GGG TTT C<br />

3 GGG GGG ATC A<br />

4 CCT GGC GGT A<br />

5 CCT GTG CTT A<br />

6 CCT GGG CTT G<br />

7 CCT GGG GGT T<br />

8 CCT GGG CTT C<br />

9 CCT GGG CCT A<br />

10 CCT GGG TTC C<br />

RESULTS AND DISCUSSION<br />

Results showed that there is a high genetic diversity among F.<br />

oxysporum isolates. Honnareddy & Dubey (2005) probed the<br />

genetic diversity of the aforementioned fungus utilising RAPD<br />

technique and isolates were classified into seven categories (1).<br />

Singh (2006) through the investigations performed on 30 isolates<br />

of F. oxysporum f.sp. ciceri collected from North India, observed<br />

little genetic variability and classified the isolates into three<br />

clusters (4).<br />

Through our investigation of polymorphic bands, 15 bands were<br />

observed. Considering a 70% similarity on dendrogram diagram<br />

genotypes were classified into 8 clusters (figure 1). Our results of<br />

RAPD‐PCR demonstrated the existence of polymorphism in the<br />

fungi populations, and a high genetic diversity was also observed<br />

among isolates under investigation. According to existence or<br />

non‐existence of bands, the genotypes classification has not<br />

matched geographical localisation. With respect to the fact that<br />

there is no significant correlation between the geographical<br />

origin of isolates and polymorphic bands, the occurrence of such<br />

a condition could be the result of seed exchange between the<br />

farmers.<br />

It seems that the more the polymorphic bands are the more is<br />

the possibility of recombination and genetic diversity in<br />

pathogens which is in turn due to their ability to mutate and<br />

anastomosis with other isolates. This will result in resistance<br />

break down against the pathogen in resistant cultivars. Due to<br />

the fact that resistant cultivars are used to control this disease,<br />

when genetic characteristic of the pathogen population changes<br />

continuously, we should prevent resistance break down by<br />

relentless reviewing of the genetic diversity on the one hand and<br />

searching for new resistant cultivars on the other hand.<br />

UPGMA<br />

F 25<br />

F 28<br />

F 22<br />

F 6<br />

F 5<br />

F 29<br />

F 23<br />

F 26<br />

F 21<br />

F 8<br />

F 30<br />

F 24<br />

F 20<br />

F 3<br />

F 14<br />

F 16<br />

F 27<br />

F 7<br />

F 4<br />

F 12<br />

F 18<br />

F 11<br />

F 19<br />

F 9<br />

F 15<br />

F 10<br />

F 2<br />

F 17<br />

F 13<br />

F 1<br />

0.04 0.2 0.36 0.52 0.68 0.84 1<br />

Jaccard's Coefficient<br />

Figure 1. Dendogram derived from RAPD analysis of Fusarium oxysporum<br />

f.sp. ciceri by UPGMA.<br />

REFERENCES<br />

1. Honnareddy N and Dubey SC (2005) Pathogenic and molecular<br />

characterization of Indian isolates of Fusarium oxysporum<br />

F.sp.ciceris causing chickpea wilt. Current science 5, 662–666.<br />

2. Nelson P (1981) Life cycle and epidemiology of Fusarium<br />

oxysporum. In ‘Fungal wilt diseases of plants’. (Eds ME Mace, AA<br />

Bell, CH Beckman) pp. 51–80. (Academic Press: New York).<br />

3. Reader V and Borda P (1985) Rapd preparation of DNA from<br />

filamentous fungi. Lett. Appl. Microbiology 1, 17–20.<br />

4. Singh BP, Saikia R, Yadav M, Singh R, Chauhan VS, and Arora DK<br />

(2006) Molecular characterization of Fusarium oxysporum<br />

F.sp.ciceris causing wilt of chickpea. African Journal of<br />

Biotechnology 5, 497–502.<br />

160 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


49 Development of nationally endorsed diagnostic protocols for plant pests<br />

B.H. Hall{ XE "Hall, B.H." } A , J. Moran B , J. Plazinski C , M. Whattam D , S. Perry E , V. Herrera F , P. Gray C , D. Hailstones G , M. Glen H , J. La Salle I<br />

A South Australian Research and Development Institute, GPO Box 397, Adelaide 5001, South Australia<br />

B Department of Primary Industries, Victoria Private Bag 15, Ferntree Gully DC 3156, Victoria<br />

C Office of the Chief <strong>Plant</strong> Protection Officer, GPO Box 858, Canberra 2601, ACT<br />

D AQIS, Private Bag 19, Ferntree Gully 3156, Victoria<br />

E Biosecurity Queensland, GPO Box 46, Brisbane 4001, Qld<br />

F MAF Biosecurity New Zealand, PO Box 2049, Auckland 1140, New Zealand<br />

G CRC Diagnostics Research Program, NSW DPI, PMB 8 Camden 2570, NSW<br />

H CSIRO Sustainable Ecosystems, Private Bag 12, Hobart 7001, Tasmania<br />

I CSIRO Entomology, GPO Box 1700, Canberra 2601, ACT<br />

Posters<br />

INTRODUCTION<br />

In 2005, a new “committee on plant diagnostics and laboratory<br />

accreditation” was formed as a subcommittee of <strong>Plant</strong> Health<br />

Committee (PHC). Called the Subcommittee for <strong>Plant</strong> Health<br />

Diagnostic Standards (SPHDS), the primary goal was to<br />

“establish, implement and monitor professional and technical<br />

standards within plant health diagnostic laboratories through<br />

the development and maintenance of an accreditation system<br />

and national diagnostic standards”.<br />

The Diagnostic Standards Working Group (DSWG) of SPHDS has<br />

developed a set of reference standards (1) to assist potential<br />

authors in developing diagnostic protocols for the detection and<br />

identification of plant pests, particularly the 253 organisms<br />

categorised as being of high importance under the Emergency<br />

<strong>Plant</strong> Pest Response Deed (2) and Industry Biosecurity plans. The<br />

standards are consistent with the international standard for<br />

diagnostic protocols for regulated pests. (3)<br />

PROTOCOL DEVELOPMENT<br />

National Diagnostic Protocols are defined as “A PHC endorsed<br />

Australian document containing detailed information about a<br />

specific plant pest or group of plant pests relevant to its<br />

diagnosis”. They are designed to assist diagnosticians in the<br />

identification of a specific pest and include data on the pest, its<br />

hosts and taxonomy, methods for detection and identification,<br />

acknowledgements, references and contacts for further<br />

information.<br />

New protocols are often developed with the assistance of a<br />

scholarship to work in an overseas laboratory, as it is not always<br />

possible to bring positive controls into Australia. Information for<br />

many of the pests of concern is available on the <strong>Plant</strong> Biosecurity<br />

Toolbox (4), part of the Pest and Disease Image Library (PaDIL).<br />

Draft protocols are developed by authors utilising information<br />

from the <strong>Plant</strong> Biosecurity Toolbox and then submitted to SPHDS<br />

for assessment.<br />

ASSESSMENT PROCESS<br />

When protocols are submitted to SPHDS, the DSWG form an<br />

Assessment Panel, comprising members of DSWG and other<br />

“experts” as deemed necessary by SPHDS. The protocols are<br />

assessed using the criteria outlined in Reference Standard (RS)<br />

No. 3 (1). In collaboration with the author, the Assessment Panel<br />

facilitates verification and peer review of the protocol according<br />

to RS No. 4 (1). Verification is undertaken by an independent<br />

laboratory with the aim of demonstrating whether the<br />

diagnostic procedures can be followed. Peer review is where an<br />

expert of the pest area reviews the accuracy and currency of the<br />

scientific information provided in the submitted diagnostic<br />

protocol, similar to a journal review. Once both Verification and<br />

Peer Review reports are received by SPHDS, the Assessment<br />

Panel reconvenes and determines whether the protocol has<br />

been deemed acceptable, or more revision is required.<br />

Once accepted, the Assessment Panel recommends to SPHDS<br />

that the completed protocol be submitted to PHC with a<br />

recommendation for endorsement as a National Diagnostic<br />

Standard.<br />

ENDORSED NATIONAL DIAGNOSTIC PROTOCOLS<br />

Protocols endorsed by PHC are placed as a version controlled<br />

document on the SPHDS website (5) to be used as part of a<br />

national response to emergency plant pest incidents for specific<br />

pest species. In some instances they may also be suitable for use<br />

in surveys to demonstrate evidence of absence to enable market<br />

access of Australian produce.<br />

Currently there are two endorsed protocols:<br />

NP1—Apple Brown Rot (Monilinia fructigena)<br />

NP2—Plum Pox Virus<br />

The protocols are reviewed every three years and if necessary<br />

are subject to rewriting and resubmission to SPHDS.<br />

FUTURE WORK<br />

DSWG is in the process of facilitating the verification and peer<br />

review of another 15 protocols. It is anticipated that at least 10<br />

of these will be completed and endorsed by the end of 2009.<br />

With 253 important pests on the list to do, not counting other<br />

high risk regulated pests and others that may appear<br />

unexpectedly, there is still a lot of work ahead.<br />

REFERENCES<br />

1. SPHDS Reference Standards. http://www.daffa.gov.au/animalplanthealth/plant/committees/sphds/national_diagnostic_protocol_gui<br />

delines_and_reference_standards (viewed 30 April 2009).<br />

2. Emergency <strong>Plant</strong> Pest Response Deed (EPPRD).<br />

http://www.planthealthaustralia.com.au/project_documents/displ<br />

ay_documents.asp?category=15 (viewed 30 April 2009).<br />

3. International Standards for Phytosanitary Measures No. 27<br />

Diagnostic Protocols for Regulated Pests (2006) www.ippc.int<br />

4. PaDIL Toolbox. http://www.padil.gov.au/pbt/ (viewed 30 April<br />

2009)<br />

5. SPHDS National Diagnostic Protocols.<br />

http://www.daffa.gov.au/animal‐planthealth/plant/committees/sphds/national_diagnostic_protocol_gui<br />

delines_and_reference_standards/national_diagnostic_protocols_f<br />

or_emergency_plant_pests (viewed 30 April 2009).<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 161


Posters<br />

96 Subcommittee on <strong>Plant</strong> Health Diagnostic Standards<br />

M.A. Williams A , B.H. Hall{ XE "Hall, B.H." } B , J. Plazinski C , P. Gray C , J. Moran D , P. Stephens E , S. Perry F<br />

A Department of Primary Industries and Water 13 St John’s Ave, New Town, 7008, Tas<br />

B South Australian Research and Development Institute, GPO Box 397, Adelaide, 5001, SA<br />

C ffice of the Chief <strong>Plant</strong> protection Officer, Department of Agriculture, Fisheries and Forestry, GPO Box 858, Canberra, 2601, ACT<br />

D Department of Primary Industries, Victoria, Private Bag 15, Ferntree Gully, 3156, Vic<br />

E Department of Regional Development, Primary Industry, Fisheries and Resources, PO Box 3000, Darwin, 0801, NT<br />

F Department of Employment, Economic Development and Industry, GPO Box 46, Brisbane, 4001, QLD<br />

INTRODUCTION<br />

In 2005, a new “committee on plant diagnostics and laboratory<br />

accreditation” was formed as a subcommittee of <strong>Plant</strong> Health<br />

Committee (PHC). Called the Subcommittee for <strong>Plant</strong> Health<br />

Diagnostic Standards (SPHDS), the primary goal was to<br />

“establish, implement and monitor professional and technical<br />

standards within plant health diagnostic laboratories through<br />

the development and maintenance of an accreditation system<br />

and national diagnostic protocols”. This is all part of a push to<br />

facilitate activities that will enhance Australia’s plant biosecurity.<br />

WORK IN PROGRESS<br />

Laboratory accreditation: Several different models were<br />

explored with accreditation to the international standard AS<br />

ISO/IEC 17025 (1) adopted as the way forward. Steps in<br />

development of a Field Application Document (FAD) included:<br />

draft FAD incorporating plant health diagnostic testing with the<br />

requirements for Veterinary Testing, independent FAD for the<br />

field of <strong>Plant</strong> Health Diagnostic Testing and, most recently,<br />

incorporation of plant health testing requirements into the<br />

Biological Testing FAD.<br />

A revised edition of the Biological Testing FAD incorporating a<br />

<strong>Plant</strong> Health Diagnostic Testing Annex should appear on the<br />

NATA website shortly (5).<br />

Diagnostic Protocols. The Diagnostic Standards Working Group<br />

(DSWG) of SPHDS has developed a set of reference standards (2)<br />

to assist potential authors in developing diagnostic protocols for<br />

the detection and identification of plant pests, particularly those<br />

categorised as being of high importance. These reference<br />

standards provide a standardised format for protocols and<br />

describe a process for assessment involving verification and peer<br />

review (2). Most, but not all, of the organisms on the list for<br />

protocol development come from the Emergency <strong>Plant</strong> Pest<br />

Response Deed (4) and from industry biosecurity planning<br />

processes.<br />

Currently two National Diagnostic Protocols have been endorsed<br />

by <strong>Plant</strong> Health Committee and can be found on the SPHDS<br />

website (3).<br />

• National Diagnostic Strategy. PHC has charged SPHDS with<br />

developing a National Diagnostic Strategy for plant health.<br />

This is additional to the National <strong>Plant</strong> Health Strategy<br />

developed by <strong>Plant</strong> Health Australia.<br />

• Training Workshops. One of SPHDS tasks is to prioritise and<br />

coordinate training in the diagnostic community.<br />

FUTURE ACTIVITIES<br />

Enhancing plant biosecurity for Australia is integral to successful<br />

plant health management. Having readily available, soundly<br />

based, diagnostic protocols for plant pests plays an important<br />

role. The magnitude of the task facing DSWG is illustrated by the<br />

over 230 plant pests listed in Table 4 of the National <strong>Plant</strong> Health<br />

Status Report(6) that require protocol development. Laboratory<br />

accreditation assures quality and integrity of the results<br />

produced and provides confidence to biosecurity administrators<br />

in managing biosecurity emergencies. Information on plant<br />

health laboratory accreditation and what it means for plant<br />

health diagnosticians will be prepared.<br />

ACKNOWLEDGEMENTS<br />

The guidance of NATA staff in preparation of the FAD documents<br />

is gratefully acknowledged.<br />

REFERENCES<br />

1. AS ISO/IEC 17025: 2007 General requirements for the competence<br />

of testing and calibration laboratories. referred to generally as ‘ISO<br />

17025’<br />

2. SPHDS Reference Standards. http://www.daffa.gov.au/animalplanthealth/plant/committees/sphds/national_diagnostic_protocol_gui<br />

delines_and_reference_standards (viewed 30 April 2009).<br />

3. SPHDS www.daff.gov.au/sphds (viewed on 18 May 2009)<br />

4. Emergency <strong>Plant</strong> Pest Response Deed (EPPRD).<br />

http://www.planthealthaustralia.com.au/project_documents/displ<br />

ay_documents.asp?category=15 (viewed 30 April 2009).<br />

5. NATA http://www.nata.asn.au/ (viewed 18 May 2009)<br />

6. <strong>Plant</strong> Health Australia, 2009 The National <strong>Plant</strong> Health Status<br />

Report (07/08), Canberra, ACT<br />

Diagnostic Services. Diagnostic service capability and capacity in<br />

Australia is a critical issue. SPHDS is involved in multiple ways of<br />

highlighting the issues and developing strategies. These include:<br />

• DAFF Training scholarships, administered by SPHDS, help to<br />

increase diagnostic capacity by sending diagnosticians<br />

overseas for training. There is an expectation that a draft<br />

diagnostic protocol would be an outcome of a scholarship.<br />

Scholarships have been awarded each year since 2004.<br />

Thirty‐eight scholarships had been awarded by the end of<br />

2008.<br />

162 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


97 What is laboratory accreditation and what will it mean for me and my laboratory?<br />

M.A. Williams A , P. Gray B , N. Kelly C , J. Cunnington D , P. Stephens E , R. Makin F and S. Peterson G , B.H. Hall{ XE "Hall, B.H." } H<br />

A Department of Primary Industries and Water, 13 St John’s Ave, New Town, 7008, Tasmania<br />

B Office of the Chief <strong>Plant</strong> protection Officer, Department of Agriculture, Fisheries and Forestry, GPO Box 858, Canberra, 2601, ACT<br />

C NSW Department of Primary Industries, PMG 8, Camden, 2570, NSW<br />

D Department of Primary Industries, Victoria, Private Bag 15, Ferntree Gully, 3156, Vic<br />

E Department of Regional Development, Primary Industry, Fisheries and Resources, PO Box 3000, Darwin, 0801, NT<br />

F National Association of Testing Authorities, PO Box 7507, Silverwater, 2128, NSW<br />

G <strong>Plant</strong> Health Australia, 5/4 Phipps Close, Deakin, 2600, ACT<br />

H South Australian Research and Development Institute, GPO Box 397, Adelaide, 5001, SA<br />

Posters<br />

INTRODUCTION<br />

Accreditation is seen as a key plank in supporting the<br />

improvement of capability and capacity in Australian plant<br />

health diagnostic laboratories to respond to biosecurity<br />

emergencies and is a key part of the business of Subcommittee<br />

on <strong>Plant</strong> Health Diagnostic Standards (1).<br />

There are two types of accreditation that could apply to plant<br />

health laboratories. The first of these examines the ability of a<br />

laboratory to secure and contain plant pests while undertaking<br />

diagnostic procedures and is administered by Australian<br />

Quarantine Inspection Service (AQIS). To this end a number of<br />

Quarantine Containment or QC Levels are recognised. This<br />

scheme is not discussed here.<br />

The second type of accreditation is an internationally recognised<br />

system for Quality Assurance based on the international<br />

standard AS ISO/IEC 17025 (2) administered in Australia by<br />

National Association of Testing Authorities (NATA) (3). This paper<br />

details the development of this scheme and how it will affect<br />

you.<br />

ACKNOWLEDGEMENTS<br />

The support of NATA, <strong>Plant</strong> Health Australia and <strong>Plant</strong> Health<br />

Committee, the parent body of SPHDS, is gratefully<br />

acknowledged; together with the input from staff of already<br />

accredited laboratories who have given examples of what it<br />

means to be accredited from their perspective.<br />

REFERENCES<br />

1. Williams MA, Hall BH, Plazinski J, Gray P, Moran J, Stephens P 2009<br />

Subcommittee on <strong>Plant</strong> Health Diagnostic Standards. Submitted for<br />

‘<strong>Plant</strong> Health Management: An Integrated Approach’ APPS 2009.<br />

2. AS ISO/IEC 17025: 2007 General requirements for the competence<br />

of testing and calibration laboratories. Referred to generally as ‘ISO<br />

17025’.<br />

3. NATA www.nata.asn.au<br />

MATERIALS AND METHODS<br />

Several plant health diagnostic testing laboratories are currently<br />

accredited to AS ISO/IEC 17025 under the Biological Testing Field<br />

Application Document (FAD). An additional Annex to this FAD<br />

has been developed to specifically cover plant health diagnostic<br />

laboratories.<br />

To get to this stage, SPHDS explored several different models,<br />

with accreditation to the international standard AS ISO/IEC<br />

17025 (2) adopted as the way forward. Steps in development of<br />

a FAD included: drafting plant health diagnostic testing<br />

requirements for incorporation with the Veterinary Testing FAD,<br />

drafting an independent FAD for the field of <strong>Plant</strong> Health<br />

Diagnostic Testing and, most recently, incorporation of plant<br />

health testing requirements into the Biological Testing FAD.<br />

RESULTS<br />

A revised edition of the Biological Testing FAD incorporating a<br />

<strong>Plant</strong> Health Diagnostic Testing Annex should appear on the<br />

NATA website shortly (3).<br />

DISCUSSION<br />

There are significant advantages in developing and operating<br />

according to an internationally recognised quality assurance<br />

system. It is recognised that there are also some disadvantages<br />

including the cost associated with developing and maintaining<br />

the system.<br />

What will it mean to laboratories and individuals working in<br />

them? This will be discussed in the context of examples from<br />

laboratories already accredited.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 163


Posters<br />

26 In vitro fungicide sensitivity of Botryosphaeriaceae species associated with ‘bot<br />

canker’ of grapevine<br />

R. Huang{ XE "Huang, R." }, W.M. Pitt, C.C. Steel and S. Savocchia<br />

National Wine and Grape Industry Centre, Charles Sturt University, Locked Bag 588, Wagga Wagga, NSW, 2678<br />

INTRODUCTION<br />

Botryosphaeriaceae species are commonly associated with the<br />

grapevine trunk disease, ‘Bot canker’. This disease is a serious<br />

threat to the productivity and longevity of vineyards in Australia<br />

and begins when pruning wounds are infected by these fungi.<br />

Consequently, damage to the vascular system of the vine limits<br />

vegetative growth and reduces yield. Control strategies<br />

emphasise the protection of wounds against infection, and a<br />

number of chemicals have been screened in vitro for inhibition<br />

of the Botryosphaeriaceae (1). To date, there are still no<br />

fungicides registered for the management of ‘Bot canker’ in<br />

Australia. Based on a previous study (1), 10 fungicides were<br />

selected and further evaluated for their activity on four<br />

additional Botryosphaeriaceae species recently isolated from<br />

diseased grapevines in New South Wales and South Australia.<br />

The aims of this research were to determine the sensitivity of<br />

these species to chemical fungicides and to identify potential<br />

agents for management of the Botryosphaeriaceae.<br />

MATERIALS AND METHODS<br />

The active ingredients of 10 fungicides were evaluated in‐vitro<br />

for mycelial inhibition of five isolates each of Diplodia mutila,<br />

Neofusicoccum australe, Dothiorella viticola, and Dothiorella<br />

iberica (Table 1). Agar plugs (5 mm diameter) from the margins<br />

of actively growing four‐day‐old fungal cultures were transferred<br />

to fungicide‐amended‐agar plates. Fungicides were dissolved in<br />

acetone (


50 The impact and diversity of Mycosphaerella leaf disease isolated from Eucalyptus<br />

globulus in western australia<br />

Posters<br />

S.L. Jackson{ XE "Jackson, S.L." } A , A. Maxwell B , S.L. Collins C , M.C. Calver A , G.E.StJ. Hardy A and B. Dell A<br />

A School of Biological Sciences and Biotechnology, Murdoch University, Perth, Western Australia 6150<br />

B Australian Quarantine Service PO Box 606 Welshpool Western Australia 6986<br />

C ITC Ltd, PO Box 1421 Albany Western Australia 6331<br />

INTRODUCTION<br />

The most serious foliar disease of eucalypt plantations in WA is<br />

Mycosphaerella leaf disease (MLD) (1). Since the<br />

commencement of the plantation industry, several fungal<br />

species contributing to MLD, previously known only in eastern<br />

Australia or overseas, have been reported on E. globulus in WA.<br />

Initially only three species were identified (2). More recently,<br />

five new records from WA (M. aurantia, M. ellipsoidea, M.<br />

mexicana and M. fori) have been identified that have not been<br />

recorded elsewhere in Australia (1, 3). Currently, 13 species of<br />

Mycosphaerella have been recorded in WA from Eucalyptus (3).<br />

Re‐examination of cultures adds six new species that have yet to<br />

be described from E. globulus in WA. The impact of MLD on<br />

growth of E. globulus plantations in WA was examined in a<br />

chemical exclusion trial at two plantations in the Albany region.<br />

MATERIALS AND METHODS<br />

<strong>Plant</strong>ations were surveyed over the period 2001–2006 for MLD<br />

and associated pathogens. Lesions from infected leaves were<br />

soaked 20–60 mins before being blotted dry and place on the lid<br />

of a Petri dish. Single spore isolations were made according to<br />

(1). Thirty spore measurements of each species were made at<br />

1000x magnification. The internal transcribed spacer region (ITS)<br />

was sequenced to confirm the identity of each species.<br />

To examine the effect of controlling MLD, an experiment was<br />

conducted on two (A and B) one‐year‐old commercial E. globulus<br />

plantations and consisted of four spray treatments (fungicide,<br />

insecticide, fungicide plus insecticide and non‐treated controls),<br />

replicated 5 times with 50 trees per replicate. This regime was<br />

designed to determine whether controlling pest and fungal<br />

diseases for 2–3 yrs increases above‐ground biomass at 2 and 5<br />

yrs. The systemic fungicide benomyl (Benlate ® , DU PONT<br />

Australia Ltd), and chlorothalonil (Bravo ® 500 DU PONT Australia<br />

Ltd) or chlorothalonil/ethylene glycol (Rover ® 500 Flowable,<br />

NUFARM Australia Ltd), were used alternately to ensure<br />

fungicide resistance would not occur. Alphacypermethrin<br />

(Dominex ® 100, Crop Care Australasia Pty Ltd) was applied<br />

regularly to control insects. Tree height and stem diameter were<br />

measured prior to the experiment (1 yr) and twice thereafter (3<br />

and 5 yrs). Volume was calculated using ITC’s standard equation<br />

and standardised for statistical analysis.<br />

RESULTS AND DISCUSSION<br />

The current study documents an increase in the number of<br />

Mycosphaerella species associated with E. globulus plantations<br />

in WA from 13 to 19. There are a number of important<br />

implications that arise from these detections including the<br />

potential impact on plantations in WA; biosecurity implications<br />

of the origin and spread of eucalypt diseases; and the ecological<br />

function of the diverse Mycosphaerella assemblage that is<br />

associated with Eucalyptus forests and plantations.<br />

volumes by 2.9%–13.5%. The critical question from a<br />

management viewpoint is whether the demonstrated increases<br />

in standardised tree volume were sufficient to warrant the cost<br />

of fungicide and insecticide treatments of the trees. Significantly,<br />

the plantations experienced a very low incidence of disease and<br />

pest attack during the trial period. Even so, the results clearly<br />

showed a significant difference between treatment types and<br />

disease outcome. This suggests that the use of chemical<br />

treatments may be useful in controlling disease outbreaks.<br />

However, the treatments most likely would have to be ongoing.<br />

Although MLD in WA occurs at relatively low levels compared to<br />

other states in Australia, the fact that the diversity of species has<br />

not yet stabilised is a concern. The number of new species<br />

isolated is steadily increasing, however, our knowledge of the<br />

biology and epidemiology of these organisms remains largely<br />

unchanged. M. cryptica and M. nubilosa are the two most<br />

important species found in WA. However, with the increasing<br />

number of species being recorded in WA, the chance of finding<br />

other significant pathogenic species is high. The industry should<br />

not remain complacent, and a concerted effort should be made<br />

to remain vigilant. Although field diagnosis remains problematic,<br />

monitoring plots across the state of varying ages should be<br />

established and outbreaks investigated in detail. The efficacy of<br />

Forest Stewardship Council accredited fungicides on MLD should<br />

be investigated in case of severe outbreaks in the future.<br />

ACKNOWLEDGEMENTS<br />

This work was part of an ARC Linkage (LP0219585). ITC Ltd was<br />

the industry partner and their financial and in kind support is<br />

gratefully acknowledged.<br />

REFERENCES<br />

1. Maxwell A, Dell B, Neumeister‐Kemp HG and Hardy GEStJ (2003)<br />

Mycosphaerella species associated with Eucalyptus in southwestern<br />

Australia: new species, new records and a key. Mycological<br />

Research 107 pp. 351–359.<br />

2. Carnegie AJ, Keane PJ and Podger FD (1997) The impact of three<br />

species of Mycosphaerella newly recorded on Eucalyptus in<br />

Western Australia. <strong>Australasian</strong> <strong>Plant</strong> <strong>Pathology</strong> 26 pp. 71–77.<br />

3. Jackson SL, Maxwell A, Burgess TI, Dell B and Hardy GEStJ (2008)<br />

Incidence and new records of Mycosphaerella species within a<br />

Eucalyptus globulus plantation in Western Australia. Forest Ecology<br />

and Management 255 pp. 3931–3937.<br />

While site differences had the greatest effect on standardised<br />

tree volumes of blue gums between 2002 and 2004 in the<br />

chemical trials, there were also significant treatment effects. The<br />

application of fungicides and insecticide increased wood<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 165


Posters<br />

8 Efficacy of pre‐seeding fungicides for control of barley loose smut<br />

K.W. Jayasena{ XE "Jayasena, K.W." } A , G. Thomas B , W. J. MacLeod B , K. Tanaka A and R. Loughman B<br />

A Department of Agriculture and Food, 444 Albany Hwy, Albany, 6330, Western Australia<br />

B Department of Agriculture and Food, 3 Baron‐Hay Court, South Perth, 6151, Western Australia<br />

INTRODUCTION<br />

In recent years, loose smut caused by Ustilago nuda (Jensen)<br />

Kellerman & Swingle has been seen widely and caused yield<br />

losses in barley (Hordeum vulgare L.) crops along the south coast<br />

of Western Australia (WA). Barley is an increasingly important<br />

crop in this region, which has an environment suited to spread<br />

and development of this disease. All the major malting barley<br />

varieties grown in WA, including Baudin which is widely adopted<br />

in this region, are susceptible to loose smut. Increased barley<br />

cropping area in disease favourable environments, widespread<br />

utilisation of susceptible varieties and changes in seeding<br />

fungicide usage towards fertiliser applied fungicide to manage<br />

diseases such as powdery mildew has raised concern amongst<br />

WA south coast barley producers over the re‐emergence of<br />

loose smut. The aim of the current study was to evaluate the<br />

efficacy of some pre‐seeding fungicides available in the<br />

Australian market for loose smut control.<br />

MATERIALS AND METHODS<br />

At three geographically separate locations, field trials were<br />

carried out using naturally infected seed of Baudin barley. The<br />

fungicide rates used are shown in Table 1. All trials were<br />

randomised block designs with four replicates. Assessments<br />

were made at each site of tiller counts, loose smut incidence and<br />

grain yield.<br />

RESULTS AND DISCUSSION<br />

Significant reductions in disease transmission were evident from<br />

most of the treatments, however none of the tested products<br />

completely eradicated transmission of loose smut (Table 1). The<br />

level of disease transmission and the relative efficacy of some<br />

fungicide products varied between experimental sites, as<br />

previously reported by Loughman et al. (1). Triadimenol<br />

(Baytan), triticonazole (Real) and tebuconazole (Raxil) based<br />

products significantly reduced loose smut at all sites. At the rates<br />

used in these experiments, fluquinconazole (Jockey) and<br />

difenoconazole (Dividend) gave variable responses, being<br />

effective at only some of the experimental sites. Triadimefon<br />

(Triad IF) applied to fertiliser and banded with seed was<br />

ineffective at all sites. Yield responses to fungicide applications<br />

were noted at Gibson, ranging from 4.6 to 5.1 t/ha and Mt<br />

Barker 3.9 to 4.4 t/ha respectively. Increased yield at these two<br />

sites does not appear related to loose smut control but was<br />

possibly due to reductions in foliar diseases such as powder<br />

mildew. The yield at Avondale ranging from 2.6 to 2.8 t/ha and<br />

treatment difference were not significant (data not shown).<br />

management strategies to combat loose smut and other<br />

diseases. Increased availability of varietal resistance to control<br />

loose smut would assist in the management of the disease and<br />

could simplify control and management options where loose<br />

smut has traditionally occurred with other diseases of barley.<br />

Table 1. Effect of seed dressing and in‐furrow fungicides on loose smut<br />

incidence in Baudin barley at Avondale, Mt Barker and Gibson, 2005.<br />

Incidence infected heads<br />

(% heads infected) **<br />

Treatments * Avondale Mt Barker Gibson<br />

Untreated 0.39 (3.5) a† 0.15 (2.2) a 0.07 (1.5) a<br />

Dividend LO 0.30 (3.0) ab 0.08 (1.5) a 0.10 (1.8) a<br />

Jockey LO 0.13 (1.9) c 0.08 (1.5) a 0.06 (1.4) a<br />

Baytan LO 0.15 (2.2) bc 0.00 (0.0) b 0.01 (0.4) b<br />

Real LO 0.10 (1.7) c 0.01 (0.2) b 0.00 (0.2) b<br />

Raxil L 0.07 (1.4) c 0.01 (0.4) b 0.02 (0.6) b<br />

Triad IF O 0.42 (3.7) a 0.17 (2.2) a 0.07 (1.3) a<br />

p


73 The cause of the barley leaf rust in Western Australia is a typical Puccinia hordei<br />

Y. Anikster A , K.W. Jayasena{ XE "Jayasena, K.W." } B , T. Eilam A and J. Manisterski A<br />

A Institute for Cereal Crops Improvement, Tel Aviv University, Ramat Aviv, 69978, Israel<br />

B Department of Agriculture and Food, 444 Albany Hwy, Albany, 6330, Western Australia<br />

Posters<br />

INTRODUCTION<br />

Leaf rust of barley caused by Puccinia hordei exists in most areas<br />

in which barley is grown. The Australian populations of this rust<br />

fungus, especially those of Western Australia (WA), are isolated<br />

from mainland Asian and South African populations by<br />

thousands of kilometers.<br />

It is of great interest to compare characters (other than<br />

pathotypes) of the WA rust population to the Israeli, because<br />

Israel is located in the centre of the cultivated barley origin. The<br />

wild ancestor—Hordeum spontaneum and the alternate host—<br />

Ornithogalum spp. still exist in the area. The sexual stage of<br />

barley leaf rust is found annually all over the northern part of<br />

Israel.<br />

MATERIALS AND METHODS<br />

Teliospore germination and inoculation of alternate aecial host.<br />

Telia originated on cultivated barley fields from the southern<br />

region of WA and from wild barley (H. spontaneum) in central<br />

Israel, were used for inducing teliospore germination and<br />

inoculate O. eigii plants in the greenhouse (1).<br />

DNA content of pycniospore nuclei. Pycniospores were<br />

harvested from pycnial clusters on O. eigii. Stained for 2h with<br />

propidium iodide in TRIS‐HC1 buffer containing RNAse and<br />

TritonX‐100. Relative DNA content was determined by flow<br />

cytometer (FACS). Fluorescence intensity was measured (1).<br />

Teliospores morphology. Teliospores were mounted in 50%<br />

glycerol on glass slides. Images were obtained with a digital<br />

camera. Spore dimensions were analysed using image analysis<br />

software.<br />

Staining of Substomatal Vesicles (SSV). Segments taken from<br />

inoculated barley leaves were microwaved in 0.03% trypan‐blue<br />

in lactophenol‐ethanol for 60 s, cleared in chloral hydrate, and<br />

mounted in lactophenol for microscope examination. Images<br />

were taken with digital camera (1).<br />

RESULTS AND DISCUSSION<br />

Teliospore germination and inoculation of the alternate host.<br />

Figure 1 shows the pycnial and aecial clusters of P. hordei on O.<br />

eigii. The WA isolates proved to be infectable (after inducing<br />

teliospore germination) to the alternate host—O. eigii, pycnial<br />

and aecial clusters were found (Fig. 1B). The aeciospores were<br />

infectable on barley seedlings, giving rise to uredinial sori<br />

(greenhouse experiments).<br />

Crosses between WA and Israeli isolates. Crosses of WA isolates<br />

and Israeli isolates in both directions of nectar transfer could be<br />

achieved, and gave rise to aecial clusters (Fig. 1A). Analysis of<br />

differential host range of the hybrids in comparison to their<br />

parents is not of the scope of this abstract (2).<br />

Figure 1. Pycnial and aecial clusters<br />

of P. hordei on O. eigii, after<br />

artificial inoculation in the<br />

greenhouse (21±2 ºC). A). Single<br />

aecial cluster (at the center of it a<br />

few pycnia may be seen).<br />

Inoculation with Israeli isolate<br />

#1930, fertilisation of the pycnial<br />

cluster with pycniospores of WA<br />

isolate #22507. B). Pycnia and aecia can be seen along the rachis and fruit after<br />

inoculation with WA isolate #22507.<br />

Figure 2. Histograms showing<br />

number of nuclei of given<br />

fluorescence intensities obtained<br />

by flow cytometry for propidium<br />

iodide—stained pycniospores of<br />

the leaf rust isolates of A). Israeli<br />

#1946 B). WA #22507. About 10000<br />

pycniospores were measured in<br />

each FACS run. The similar position<br />

of the two histograms on the X axis<br />

points out a similar content of<br />

nuclear DNA in the pycniospores of both isolates.<br />

Figure 3. Teliospores of P.<br />

hordei isolates from A). Israeli<br />

#1963 and B). WA #22507 and<br />

dimensions. Bar = 20 µm for<br />

both figures. Dimensions: area<br />

(µm 2 ), length (µm) and width (µm); A): 774±101, 47±5 and 21±2; B): 776±96, 45±5<br />

and 22±2.<br />

Figure 4. Uredinial<br />

substomatal vesicles (SSV)<br />

formed in barley cultivar L‐94<br />

by: A). Israeli isolate #1930,<br />

and B). WA isolate #22507.<br />

SSV—substomatal vesicle; ms<br />

– median septum; h‐ hyphae; st—invaded stoma. The SSV of the two isolates are<br />

very similar in shape and in dimension. Bar = 20µm.<br />

Our results give rise to the opinion that despite geographic<br />

isolation, the WA population of P. hordei is taxonomically similar<br />

to isolates from outside Australia. It may have either reached<br />

Australia quite recently (in an evolutionary sense a few hundred<br />

years ago) with a very low rate of changes or there is some way<br />

of connection (winds, or another way) of the WA population and<br />

international populations.<br />

ACKNOWLEDGEMENTS<br />

Authors wished to thank Drs. Robert Park, University of Sydney<br />

and Robert Loughman, Department of Agriculture and Food<br />

Western Australia for initial reviewing of the abstract.<br />

REFERENCES<br />

1. Anikster Y, Eilam T, Manisterski J, Leonard KJ (2003) Self‐fertility<br />

and other distinguishing characteristics of a new morphotype of<br />

Puccinia coronata pathogenic on smooth brome grass. Mycologia<br />

95, 87–97.<br />

2. Anikster Y (1984) Parasitic specialization of Puccinia hordei in Israel.<br />

Phytopathology 74, 1061–1064.<br />

Nuclear DNA content spore morphology and SSV. Comparison<br />

of the isolates from WA and Israeli by DNA content, spore<br />

morphology and SSV show very close similarities (Figs. 2, 3, 4).<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 167


Posters<br />

27 The value of combined use of genetic resistance and fungicide application for<br />

management of stripe rust<br />

K.W. Jayasena{ XE "Jayasena, K.W." } A , G. Thomas B , R. Loughman B , K. Tanaka A and W. J. MacLeod B<br />

A Department of Agriculture and Food, 444 Albany Hwy, Albany, 6330, Western Australia<br />

B Department of Agriculture and Food, 3 Baron‐Hay Court, South Perth, 6151, Western Australia<br />

INTRODUCTION<br />

Stripe rust (yellow rust) of wheat (Triticum aestivum L.) caused<br />

by Puccinia striiformis f. sp. tritici was first detected in Western<br />

Australia (WA) in 2002. Regional outbreaks have caused<br />

considerable yield losses, particularly in susceptible wheat<br />

varieties. Wheat varieties grown in WA range from susceptible<br />

to resistant, with many exhibiting varying degrees of partial<br />

resistance. The most effective use of fungicides in combination<br />

with these varying levels of resistance is poorly understood. The<br />

aim of the present study was to determine how varieties with<br />

different levels of rust resistance respond to fungicide for the<br />

control of stripe rust.<br />

MATERIALS AND METHODS<br />

Varieties with a range of stripe rust resistance, EGA Bonnie Rock<br />

(S‐VS), Carnamah (MS‐S), Wyalkatchem (MS), Janz (MR‐MS) and<br />

GBA Ruby (R) were tested in combination with 3 fungicide<br />

treatments, being either nil, partial or full fungicide control,<br />

during 2007 and 2008. Trial design was split plot with four<br />

replications. Full control consisted of tebuconazole (Folicur<br />

430SC) @ 290 mL/ha applied at early stem elongation (Z31), flag<br />

leaf emergence (Z39/40), ear emergence (Z55) and late<br />

flowering (Z68) to provide maximum disease protection and<br />

yield potential. Partial fungicide control consisted of a single<br />

application at ear emergence (Z55) in 2007 or two applications<br />

commencing with the first sign of the stripe rust (Z32) and again<br />

at ear emergence (Z55) in 2008. In 2007, the trial was sown on 5<br />

July, adjacent to susceptible wheat (cv. Harrismith) inoculated<br />

twice on 30 July and 23 August to generate inoculum. In 2008<br />

the trial was sown 20 June adjacent to susceptible wheat (cv.<br />

Westonia) that was inoculated with stripe rust on 24 July.<br />

RESULTS AND DISCUSSION<br />

In both years, the stripe rust was evident between stem<br />

extension and flag leaf emergence. In 2007, the stripe rust<br />

severity ranged from 4 to 96% in untreated control plots of the<br />

five varieties whereas in 2008, it varied from 6 to 93% (Figure<br />

1a). GBA Ruby had no response to fungicide for stripe rust<br />

control. Application of fungicides either as single, double or<br />

multiple sprays reduced the stripe rust levels in all other wheat<br />

varieties, in both years. Under these experimental<br />

circumstances, where the varieties were subject to continuous<br />

high disease pressure from nearby infected susceptible wheat,<br />

partial fungicide control was less effective than full control with<br />

multiple fungicide sprays in Carnamah, EGA Bonnie Rock, Janz,<br />

and Wyalkatchem.<br />

Over two years, extreme yield loss (87–94%) was observed in<br />

EGA Bonnie Rock (Figure 1b). In Janz and Wyalkatchem, partial<br />

resistance reduced the impact of stripe rust however yield losses<br />

of 27–54% were still observed. Partial control combined with<br />

partial resistance reduced yield losses to 17–30%, depending on<br />

variety. Application of fungicides significantly increased the yield<br />

and hectolitre weight in all the varieties tested except for GBA<br />

Ruby.<br />

In 2007, screenings varied from 3.3 to 18.4% among the<br />

untreated varieties whereas in 2008, it was 0.8 to 7.9% (data not<br />

shown). EGA Bonnie Rock had higher screenings in both years.<br />

These experiments demonstrate the effect of single major gene<br />

resistance. Under high disease pressure, stripe rust infection in<br />

GBA Ruby was very low and maximum yield was achieved<br />

without fungicide protection. However, Australian and<br />

international experience is that single major gene resistance to<br />

stripe rust, though highly effective, is not durable while multiple<br />

gene resistance is more robust. GBA Ruby carries Yr27, which is<br />

currently fully effective in WA but recent reports indicate<br />

development of Yr27 virulence in the stripe rust population in<br />

eastern Australia.<br />

Under very high disease pressure, the varieties with partial<br />

resistance genes such as Janz and Wyalkatchem showed high<br />

levels of infection; however yield was significantly greater than<br />

in susceptible types. With the partial fungicide protection,<br />

significant yield benefits were obtained. In environments which<br />

are less conducive to stripe rust, the value of partial resistance<br />

would be expected to be greater.<br />

Major gene resistance provides maximum protection from stripe<br />

rust, however many of the varieties preferred by Western<br />

Australian grain producers utilise some level of partial resistance<br />

rather than single gene resistance. In general, partial resistance<br />

to stripe rust combined with strategic fungicide application can<br />

be used to minimise yield losses and restrict epidemic<br />

development.<br />

a<br />

% stripe rust severity<br />

b<br />

Yield (t/ha)<br />

(av. Fnec to av. F-1ne<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

Untre ate d P ar tial C o n tr o l Fu ll C o n tro l<br />

V1 V2 V3 V4 V5 V1 V2 V3 V4 V5<br />

lsd 5% = 14 w he at varie tie s<br />

lsd 5% = 10<br />

lsd 5% = 0 .4<br />

2007<br />

2007 2008<br />

Untre ate d Partial Control Full Control<br />

2008<br />

V1 V2 V3 V4 V5 V1 V2 V3 V4 V5<br />

lsd lsd 5% 5% = 0.4 w he at varie tie s<br />

lsd 5% = 0.6<br />

lsd 5% = 0.6<br />

Figure 1. Response of five wheat varieties (V1—EGA Bonnie Rock; V2—<br />

Carnamah; V3—Wyalkatchem; V4—Janz; V5—GBA Ruby) with different<br />

resistance levels to fungicide application for control of stripe rust. a)<br />

average stripe rust severity assessed on two top leaves which showing<br />

necrosis due stripe rust infection at milk development stage and, b)<br />

yield, in 2007 and 2008 at Manjimup, WA.<br />

ACKNOWLEDGEMENTS<br />

Many thanks to Mr Ian Guthridge, DAFWA, at Manjimup<br />

Horticulture Research Institute for help with field operations.<br />

This work was supported by Grains Research and Development<br />

Corporation (DAW00159).<br />

168 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


9 Specific genetic fingerprinting of Pseudomonas syringae pv. syringae strains from<br />

stone fruits in Iran with REP sequence and PCR<br />

Posters<br />

S. Ketabchi{ XE "Ketabchi, S." } A<br />

A Department of <strong>Plant</strong>pathology Shiraz Islamic Azad University, POX 71365‐364, Shiraz, Iran<br />

INTRODUCTION<br />

Bacterial canker and blast of stone fruit trees, caused by<br />

Pseudomonas syringae pv.syringae affects all commercially<br />

grown Prunus species in province of Shiraz in Iran. The<br />

relationship between P. syringae pv. syringae strains infecting<br />

Prunus species and strains that infect other crops such as,<br />

cereals, is unknown and needs to be elucidated Molecular<br />

analysis of genomic variability has been used to differentiate and<br />

classify bacterial strains below the level of species repetitive<br />

Extragenic palindromic (REP), which are short repetitive DNA<br />

sequences with highly conserved central inverted repeats that<br />

are dispersed throughout the genomes of diverse bacterial<br />

species (1), have been used to design universal PCR primers that<br />

generate highly reproducible, strain‐specific fingerprints that can<br />

differentiate bacterial strains below the level of species or<br />

subspecies. The objective of this study was to identify and<br />

characterise strains of P. syringae pv. syringae isolated from<br />

various Prunus species and other plant hosts by using Rep‐PCR<br />

analysis.<br />

MATERIALS AND METHODS<br />

Strain isolation. Samples of both healthy and diseased tissues<br />

from stone fruit trees were collected from different orchard of<br />

Iran<br />

Rep‐PCR. Rep primers The PCR conditions were as previously<br />

described (21, 32) DNA fragments in the gel were visualised by<br />

staining with ethidium bromide.<br />

RESULTS<br />

Twenty‐five strains of P. syringae pv. syringae collected<br />

from stone fruit orchard sites, wheat and sugar beet in the<br />

Shiraz, Tehran and another part of Iran were used in this study.<br />

The DNA fingerprints were determined by using PEP‐PCR. Most<br />

of P. syringae pv. syringae strains from stone fruits, shown<br />

similar pattern and are different white other hosts. Wheat and<br />

sugar beet strain have several common bands.<br />

DISCUSSION<br />

In this study, the P. syringae pv. syringae strains isolated from<br />

Prunus hosts in Iran generated similar genetic profiles in PEP‐PCR<br />

whereas most strains of P. syringae pv. syringae isolated from<br />

other hosts generated dissimilar patterns (3). This suggests a<br />

host specialisation of the stone fruit strains within the<br />

heterogeneous pathovar syringae. Specialisation of P. syringae<br />

pv. syringae strains toward a particular host has been observed<br />

in previous studies(4) REP PCR has been shown to be a rapid and<br />

reliable method to differentiate plant‐pathogenic bacteria at or<br />

below the pathovar level with highly reproducible results (5).<br />

Our results suggest that strains of P. syringae pv. syringae that<br />

are adapted to a specialised niche, such as Iran stone fruits, may<br />

be the result of a recent adaptation and/or genetic isolation,<br />

resulting in the genetically homogeneous population of<br />

P. syringae pv. syringae strains from stone fruits observed in this<br />

study, which formed a distinct group from strains isolated from<br />

other hosts.(3)<br />

REFERENCES<br />

1. De Bruijin, F.J. 1992. Use of repetitive (repetitive extragenic<br />

palindromic and enterobacteria repetitive intergenic consensus)<br />

sequences and the polimeras chain reaction to finger print the<br />

genoms of Rhizobium meliloti isolates and other soil bacteria. Appl.<br />

Environ. Microbiol. 8: 2180–2187<br />

2. Little, E.L., Bostock, R.M. and Kirkpatrick, B.C. 1998. Genetic<br />

characterization of Pseudomonas syringae pv syringae strain from<br />

stone fruits in California. Appl. Environ. Microbiol. 64: 3818–3823.<br />

3. Louws, F.J., Fulbright, D.W., Stephens, C.T. and De Bruiin, F.J. 1994.<br />

Specific genomic fingerprints of phytopathogenic Xanthomonas<br />

and Pseudomonas pathovars and strains generated with repetitive<br />

sequences and PCR. Apple. Environ. Microbiol. 80: 2286–2292.<br />

4. Versalovic, J., Koeuth, T., and Lupski, J.R. 1991. Distribution of<br />

repetitive DNA sequences in Eubacteria and application<br />

fingerprinting bacterial genomes. Nucleic. Acids. Res. 19: 6823–<br />

6831.<br />

Figure 1. REP fingerprints of several P. syringae pv. syringae strains<br />

isolated from various plant hosts, showing strain variability within the<br />

pathovar. Lanes: kb, the 1‐kb molecular marker; 1, Almond (pattern 1, 2);<br />

2, Walnut (pattern 3–5); 3, Apricot (pattern 6, 7); 4, Peach (pattern 8, 9);<br />

5, cherry (pattern 10_13); 7, wheat (pattern 14); 8, Sugar beet (pattern<br />

15); 9Negative control (pattern 16)<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 169


Posters<br />

85 Non‐host resistance and pathogen virulence: an important role of toxic and<br />

infection‐inducing compound(s) from spore germination fluid of Botrytis cinerea<br />

N.N. Khanam{ XE "Khanam, N.N." }, K. Toyoda, H. Yoshioka, Y. Narusaka and T. Shiraishi<br />

Okayama University, Tsushima Naka1‐1‐1, Okayama shi, 700‐8530, Japan<br />

INTRODUCTION<br />

<strong>Plant</strong>s are continually exposed to a vast number of potential<br />

pathogens and, as a result, they have evolved intricate defense<br />

mechanisms like hypersensitive response (HR), oxidative burst<br />

and increased expression of pathogenesis related protein. The<br />

HR appears to play a pivotal role in the success of B. cinerea. As a<br />

typical necotroph, it may produce multiple metabolites and<br />

proteins that determine its necrotrophic life style (van Kan,<br />

2006). One of the key mechanisms of Botrytis species is their<br />

ability to induce active cell death in their host plants in order to<br />

be pathogenic (van Baarlen et al. 2004). This study describes the<br />

action of spore germination fluid of B. cinerea (Bc) on<br />

pathogenecity and on host responses.<br />

MATERIALS AND METHODS<br />

Spore germination fluid (SGF) was obtained from watergerminated<br />

conidia from a highly virulent strain BC.02RO<br />

supplemented with 0.05 µg/ml glucose. An avirulent Alternaria<br />

alternata (15B) and a hypovirulent strain of B. cinerea<br />

(BC.236795) were evaluated on disease development and<br />

cellular response on Nicotiana benthamiana (Nb) with or<br />

without SGF. Disease development was monitored<br />

macroscopically by measuring the lesion area, and, cellular<br />

changes were monitored microscopically by histochemical<br />

staining. Freeze‐dried SGF solution was made as weight/volume.<br />

A 40 ug/ml was used as a standard concentration because of<br />

visible clear necrosis on Nb, kidney bean, barley, and so on.<br />

RESULTS<br />

A hypovirulent BC.236795 and an avirulent 15B did not infect<br />

and evoke visible lesion with sterilised water at 2 dpi on Nb.<br />

Both induced papilla formation and callose deposition. On the<br />

other hand, BC.236795 or 15B with SGF (40 µg/ml) from<br />

BC.02RO evoked lesion and cell death on Nb (Fig. 1) as BC.02RO<br />

did. The activities of infection‐induction and lesion formation are<br />

detected in a 10–30 kDa fraction of SGF. That is, BC.02RO‐SGF<br />

contained toxic and infection‐inducing factor(s). H 2 O 2 generation<br />

was observed 9 h after treatment with SGF (Fig. 1) while the<br />

generation was accelerated by inoculation with 15B or<br />

BC.236795. These effects were more significant with BC.02RO.<br />

The SGF alone induced cell death but not callose deposition.<br />

Though necrosis was induced at >20µg/ml of SGF, infection by<br />

15B or BC.236795 was established at >5µg/ml (Fig. 2). Proteinase<br />

K (PrK) negated apparently lesion formation and cell death<br />

induced by SGF with 15B. Prk also limited lesion formation by<br />

BC.02RO dose dependently.<br />

Figure 1. Effect of SGF (40 ug/ml) on lesion formation (5 dpi; left), cell<br />

death (4 dpi; middle), with or without 15B and H 2 O 2 generation (right; 9<br />

h upper, 12 h lower) on Nb.<br />

Figure 2. Effect of SGF on infection by 15B on N. benthamiana at 2 dpi.<br />

DISCUSSION<br />

As described above, we found that the SGF of a virulent strain of<br />

Bc induced necrosis and accessibility even to a hypovirulent Bc<br />

or to an avirulent 15B, and both pathogens showed similar<br />

activity with exogenous H 2 O 2 . The SGF also induced prominent<br />

accumulation of H 2 O 2 , which is required for cell death (van Kan<br />

2006). It was also reported that accumulated H 2 O 2 is necessary<br />

to achieve full virulence of Botrytis (van Baarlen et al. 2004).<br />

Taken together with these reports, we hypothesise that SGF<br />

plays a crucial role in establishment of infection or lesion<br />

formation by Bc or related necrotrophic fungi mediated by H 2 O 2<br />

generation. The preliminary experiment with PrK indicates that<br />

the principle for necrosis‐induction and accessibility‐induction in<br />

SGF seems to be a proteinous compound(s). In conclusion, SGF<br />

contains determinant(s) of pathogenicity of this necrptroph.<br />

ACKNOWLEDGEMENTS<br />

The research was funded by Japan <strong>Society</strong> for Promotion of<br />

Science.<br />

REFERENCES<br />

1. Van Kan, JAL (2006) Licensed to kill the lifestyle of a necrotroph<br />

plant pathogen. Trends <strong>Plant</strong> Sci 11, 247–253.<br />

2. Van Baarlen P, Staats M, van Kan JAL (2004) Induction of<br />

programmed cell death in lily by the fungal pathogen Botrytis<br />

elliptica. Molecular <strong>Plant</strong> <strong>Pathology</strong> 5(6), 559–574.<br />

170 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


51 Impact of Phytophthora cinnamomi on native vegetation in South Australia<br />

S.F. McKay A , K.H. Kueh{ XE "Kueh, K.H." } A , A.J. Able A , R.M.A. Velzeboer C , J.M. Facelli B and E.S. Scott A<br />

A School of Agriculture, Food and Wine, The University of Adelaide, Waite Campus, PMB 1, Glen Osmond, South Australia, 5064<br />

B School of Earth and Environmental Sciences, The University of Adelaide, North Terrace Campus, South Australia, 5005<br />

C Department for Environment and Heritage, 41 Victoria St, Victor Harbor, South Australia, 5211<br />

Posters<br />

INTRODUCTION<br />

Phytophthora dieback, caused by the soil‐borne Oomycete<br />

Phytophthora cinnamomi Rands (Pc), has been identified by the<br />

Australian Government as a key threat to native ecosystems. A<br />

National Threat Abatement Plan (NTAP) has been developed to<br />

limit damage to our native flora and fauna. In spite of the threat<br />

that Pc represents for South Australia (SA), basic information<br />

about the effect of the disease on native vegetation in SA is<br />

lacking. This project aims to increase understanding of the<br />

susceptibility of threatened and key plant species and ecological<br />

changes in plant communities due to Phytophthora dieback in<br />

SA.<br />

MATERIALS AND METHODS<br />

Susceptibility testing of threatened and key plant species.<br />

Testing of selected threatened species has commenced (Table<br />

1). These species were chosen on the basis of availability, ease of<br />

seed germination and handling in the greenhouse, and<br />

occurrence in moderate or high “risk of Phytophthora” area(s)<br />

(1). Other species, abundant at our field sites, will also be tested,<br />

e.g. Allocasuarina, Hakea and Hibbertia spp. Three‐month old<br />

plants will be inoculated with Pc using a method modified from<br />

Butcher et al (1984) and Shearer et al (2004) and monitored for<br />

disease symptoms and mortality.<br />

Table 1. Threatened plant species to be tested for susceptibility to Pc.<br />

Species<br />

Common name<br />

Allocasuarina robusta<br />

Mount Compass oak‐bush<br />

Brachyscome diversifolia<br />

Tall daisy<br />

Olearia pannosa<br />

Silver‐leaved daisy<br />

Austrodanthonia carphoides Short wallaby grass<br />

Acacia enterocarpa<br />

Jumping jack wattle<br />

Acacia pinguifolia<br />

Fat‐leaf wattle<br />

Glycine tabacina<br />

Variable glycine<br />

Correa calycina<br />

SA green correa<br />

Pomaderris halmaturina<br />

Kangaroo Is. pomaderris<br />

Prostanthera halmaturina<br />

Monarto mintbush<br />

Oreomyrrhis eripoda<br />

Australian carraway<br />

Spyridium parvifolium<br />

Dusty miller<br />

Spyridium spathulatum<br />

Spoon‐leaved spyridium<br />

Dynamics of Pc in the field. The rate and pattern of spread of Pc<br />

are being studied at two sites in the Mount Lofty Ranges (Mount<br />

Bold reservoir reserve and Scott Creek Conservation Park). The<br />

sites are open woodland, are floristically similar to one another<br />

and the presence of Pc has been confirmed. Permanent quadrats<br />

have been established and the following parameters measured<br />

and data collected in 2008:<br />

• soil and fine root samples; baited for Pc<br />

• numbers and health of key indicator species e.g.<br />

Xanthorrhoea semiplana and other vascular plants<br />

• percentage cover by vascular plants, leaf litter and bare<br />

ground<br />

• other data e.g. soil moisture, rainfall.<br />

These parameters will be measured again in autumn and spring<br />

of 2009 and 2010.<br />

Effect of companion plants on susceptibility. The hypothesis<br />

that the plant neighbourhood influences spread and expression<br />

of Phytophthora dieback is being examined in a series of pot<br />

experiments. In the first experiment, seeds of Acacia pycnatha<br />

and A. myrtifolia have been sown in pots containing 1‐year‐old<br />

plants of X. semiplana. Pots will be inoculated with Pc when<br />

acacias are 3 months old and symptoms assessed.<br />

Microbial antagonists. Rhizosphere soil from Pc tolerant plants,<br />

e.g. some Acacia spp., and from sites where Pc is present but not<br />

causing disease will be screened in vitro for antagonists of Pc, in<br />

particular streptomycetes. Preliminary work has yielded several<br />

species strongly antagonistic to Pc. Streptomycetes will be<br />

tested further for antagonism in planta. Results from these<br />

experiments may help to explain suppression of Pc root rot in<br />

some native ecosystems.<br />

RESULTS AND DISCUSSION<br />

Baseline data collected from the field sites in 2008 will be<br />

compared with data collected in 2009 and 2010 which will<br />

enable documentation of the rate and pattern of spread of the<br />

pathogen and disease over time. Information from field<br />

observations and glasshouse experiments about susceptibility of<br />

threatened and key species will enable improved management<br />

decisions regarding conservation of threatened plant species.<br />

Knowledge of companion plant interactions will increase<br />

understanding of the factors that affect the spread of the<br />

disease. Information from this project will facilitate the adoption<br />

of management strategies in line with NTAP objectives.<br />

ACKNOWLEDGEMENTS<br />

This research is funded by the ARC and has the following linkage<br />

partners: Adelaide‐Mt Lofty NRM Board, Adelaide Hills Council,<br />

City of Tea Tree Gully Council, Department for Environment and<br />

Heritage, Department of Transport, Energy and Infrastructure,<br />

Forestry SA, PIRSA Forestry, SA Murray Darling Basin NRM Board<br />

and SA Water. We thank the Sarawak State Government,<br />

Malaysia, for funding the PhD studies of Mr Kueh Kiong Hook.<br />

REFERENCES<br />

1. Velzeboer R, Stubbs W, West A, Bond A (2005) Threatened plant<br />

species at risk from Phytophthora in South Australia. (Department<br />

for Environment and Heritage, SA. Government of South Australia).<br />

2. Butcher TB, Stukely MJC, Chester GW (1984) Genetic variation in<br />

resistance of Pinus radiata to Phytophthora cinnamomi. Forest<br />

Ecology and Management 8, 197–220.<br />

3. Shearer BL, Crane CE, Cochrane A (2004) Quantification of the<br />

susceptibility of the native flora of the South‐West Botanical<br />

Province, Western Australia. Australian Journal of Botany 52, 435–<br />

443.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 171


Posters<br />

74 Genetic diversity and population structure of Australian and South African<br />

Pyrenophora teres isolates<br />

A. Lehmensiek{ XE "Lehmensiek, A." } A , R. Prins B , G. Platz C , W. Kriel D , G.F. Potgieter E and M.W. Sutherland A<br />

A Centre for Systems Biology, University of Southern Queensland, Toowoomba, 4350 QLD, Australia<br />

B CenGen (Pty) Ltd, 78 Fairbairn Street, Worcester, 6850, South Africa<br />

C DEEDI Primary Industries and Fisheries, Hermitage Research Station, Warwick, 4370 QLD, Australia<br />

D Department of <strong>Plant</strong> Sciences, University of the Free State, Bloemfontein, 9300, South Africa<br />

E South African Barley Breeding Institute, PO Box 27, Caledon 7230, South Africa<br />

INTRODUCTION<br />

Net blotch, caused by the fungus Pyrenophora teres, is a serious<br />

production problem for the barley (Hordeum vulgare L.) industry<br />

in Australia, South Africa and elsewhere (1, 2, 3, 4). Two forms of<br />

net blotch exist: one is the net form (NFNB) caused by P. teres f.<br />

teres (PTT) and the other is the spot form (SFNB) caused by P.<br />

teres f. maculata (PTM). Several Australian and international<br />

studies have used molecular markers, such as amplified<br />

fragment length polymorphisms (AFLP) to investigate the genetic<br />

structure of P. teres (3, 5, 6, 7). In contrast, while the incidence<br />

of net blotches on barley have increased recently in South Africa,<br />

local populations of the fungus have remained uncharacterised.<br />

To address this issue, PTT and PTM isolates were collected from<br />

the south‐western Cape region of South Africa. AFLP analysis<br />

was conducted on extracted DNA from these isolates and from a<br />

collection of Australian isolates to determine the genetic<br />

diversity and structure of South African populations and to<br />

determine their relatedness to Australian isolates.<br />

MATERIALS AND METHODS<br />

DNA extractions. Fungal mycelium were harvested from cultures<br />

grown on potato dextrose agarose plates at 25°C for one week.<br />

A CTAB DNA extraction method was used to extract the fungal<br />

DNA.<br />

AFLP analysis. The AFLP procedure was carried out using an<br />

Invitrogen AFLP Core Reagent kit. The EcoRI primers were hexlabelled.<br />

The samples were visualised using a Gel‐Scan 2000<br />

DNA fragment analyser (Corbett Life Sciences, Sydney,<br />

Australia).<br />

Scoring and data analysis. Both monomorphic and polymorphic<br />

bands were scored and used in the data analysis. Bands were<br />

scored independently by two people. The cluster analysis was<br />

performed using NTSYSpc V2.20f, whereas the program<br />

Structure V2.2 was used to determine the population structure.<br />

RESULTS<br />

AFLP analysis was conducted on DNA of 23 South African and 37<br />

Australian PTT isolates, 37 South African and 29 Australian PTM<br />

isolates, six Bipolaris sorokiniana isolates, two P. tritici‐repentis<br />

and two Drechslera rostrata isolates. Eight primer combinations<br />

were used to amplify AFLPs and on average 50 loci were<br />

produced with each primer combination. In total, 400 loci could<br />

be accurately scored across all samples and 168 of these loci<br />

were polymorphic in the P. teres samples.<br />

No genetic differentiation associated with locations within<br />

Australia or South Africa could be identified.<br />

The program Structure separated the PTT and PTM isolates into<br />

three and two groups, respectively.<br />

DISCUSSION<br />

Our study indicates that the genetic diversity among South<br />

African and Australian Pyrenophora isolates is low and that there<br />

is no clear geographical substructuring. These findings are similar<br />

to those of studies in other regions (3, 5, 6). Results produced by<br />

the two software packages NTSYS and Structure will be<br />

compared and discussed.<br />

ACKNOWLEDGEMENTS<br />

The authors would like to thank Dr Hugh Wallwork and Dr Sanjiv<br />

Gupta for the isolate samples provided by them. We also would<br />

like to thank Debbie Snyman, Denise Liebenberg and Lizaan<br />

Rademeyer for their technical help in the Cengen laboratory.<br />

This project was funded by the South African Winter Cereals<br />

Trust.<br />

REFERENCES<br />

1. Campbell GF, Crous PW (2003) Genetic stability of net x spot hybrid<br />

progeny of the barley pathogen Pyrenophora teres. <strong>Australasian</strong><br />

<strong>Plant</strong> <strong>Pathology</strong> 32, 283–287.<br />

2. Gupta S, Loughman R, Platz G, Lance RCM (2003) Resistance in<br />

cultivated barleys to Pyrenophora teres f. teres and prospects of its<br />

utilisation in marker identification and breeding. Australian Journal<br />

of Agricultural Research 54, 1379–1386.<br />

3. Leisova L, Minarikova V, Kucera L, Ovesna J (2005) Genetic diversity<br />

of Pyrenophora teres isolates as detected by AFLP analysis. Journal<br />

of Phytopathology 153, 569–578.<br />

4. Manninen O, Kalendar R, Robinson J, Schulman AH (2000)<br />

Application of BARE‐1 retrotransposon markers to the mapping of a<br />

major resistance gene for net blotch in barley. Molecular and<br />

General Genetics 264, 325–334.<br />

5. Bakonyi J, Justesen AF (2007) Genetic Relationship of Pyrenophora<br />

graminea, P‐teres f. maculata and P‐teres f. teres assessed by RAPD<br />

analysis. Journal of Phytopathology 155, 76–83.<br />

6. Jonsson R, Sall T, Bryngelsson T (2000) Genetic diversity for random<br />

amplified polymorphic DNA (RAPD) markers in two Swedish<br />

populations of Pyrenophora teres. Canadian Journal of <strong>Plant</strong><br />

<strong>Pathology</strong>‐Revue Canadienne De Phytopathologie 22, 258–264.<br />

7. Serenius M, Manninen O, Wallwork H, Williams K (2007) Genetic<br />

differentiation in Pyrenophora teres populations measured with<br />

AFLP markers. Mycological Research 111, 213–223.<br />

Cluster analysis separated the NFNB and SFNB isolates into two<br />

strongly divergent groups (similarity coefficient = 0.6). Low<br />

genetic differentiation was observed within the NFNB and SFNB<br />

groups (similarity coefficient = 0.9). Interestingly, the South‐<br />

African NFNB isolates clustered together with the Australia NFNB<br />

isolates whereas the South‐African SFNB isolates were grouped<br />

into a distinct cluster separate from the Australian SFNB isolates.<br />

172 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


28 Molecular identification of Pythium isolates of ginger from Fiji and Australia<br />

M.F. Lomavatu{ XE "Lomavatu, M.F." } A,B , J. Conroy B and E. Aitken B<br />

A<br />

Koronivia Research Station, PO Box 77, Nausori, Fiji<br />

B<br />

School of Biological Sciences, The University Of Queensland, St Lucia, Brisbane<br />

Posters<br />

INTRODUCTION<br />

Pythium myriotylum Drech is one of the main causal organisms<br />

for Pythium rhizome rot of ginger and is common worldwide. It<br />

was first recorded in 1907 by Butler. In a recent review (1)<br />

eleven species of Pythium were listed as causal agents of a<br />

rhizome rot of ginger. Pythium affects ginger throughout its<br />

growing stages and the main entry points of infection are the<br />

buds, roots, developing rhizomes and the collar region of the<br />

plant.<br />

In Fiji and Australia P. myriotylum is commonly associated with<br />

the disease and isolates of this species are capable of destroying<br />

ginger rhizomes in 1–2 weeks under appropriate environmental<br />

conditions (2). In Fiji, surveys during 2007 and 2008 found that<br />

Pythium rhizome rot was associated with hot, humid and high<br />

rainfall periods during the wet season. It was noted during the<br />

previous study (2) that additional species of Pythium may be<br />

associated with the disease in both countries. Some Fijian<br />

isolates grew at different rates on agar and showed variability in<br />

cultural characteristics, while preliminary work indicated<br />

variability in aggressiveness within isolates obtained from<br />

Australia. This abstract details further investigations into the<br />

molecular variability of Pythium isolates from ginger in Fiji and<br />

Australia.<br />

MATERIALS AND METHODS<br />

1. Culture isolation Diseased ginger rhizomes were collected<br />

from different ginger growing areas in Fiji (2) and in Queensland.<br />

The five Fijian isolates used in this study were imported on an<br />

AQIS permit and kept in an quarantine lab at the University of<br />

Queensland. Of the Australian Pythium isolates used, one was<br />

from capsicum and two isolated from diseased ginger rhizomes<br />

(Table 1).<br />

2. Molecular identification<br />

2.1 DNA extraction. After 5 days growth on potato dextrose<br />

broth, mycelia of each isolate were collected in Falcon tubes and<br />

lyophilised. Mycelia were ground in liquid nitrogen to fine<br />

powder and DNA extracted as per Lee & Taylor (3).<br />

2.2 Polymerase chain reaction and sequencing. The ITS region<br />

of the Pythium isolates were amplified using the universal<br />

primers ITS 1 (5’ – TCCGTAGGTGAACCTGCGG‐3’) and ITS 4 (5’‐<br />

TCCTCCGCTTATTGATATGC ‐3’) described by White et al. (4) and<br />

were sequenced at AGRF Brisbane.<br />

RESULTS<br />

Molecular identification BLAST analysis revealed that three<br />

Pythium species, P. myriotylum, P. vexans and P. graminocola,<br />

were identified each from three different ginger growing<br />

localities (Veikoba, Navua and upper Naitasiri, respectively). The<br />

two Australian ginger isolates revealed sequence alignment with<br />

P. myriotylum and P. zingiberis; the capsicum isolate was<br />

confirmed as P. myriotylum (Table 1).<br />

Table 1. Pythium isolates from Fiji (KRS suffix) and Queensland (BRIP<br />

suffix) showing location collected, species identification based on ITS<br />

sequence similarity and Genbank accession number where deposited. All<br />

isolates from ginger except* from capsicum.<br />

Sample location Pythium sp. GenBank<br />

KRS11 Navua P.vexans<br />

KRS13 Muainaweni P.graminicola<br />

KRS14 Veikoba P.myriotylum FJ797574<br />

KRS15 Waibau P.graminicola<br />

KRS17 Veikoba P.myriotylum FJ797575<br />

BRIP39907* Bundaberg P.myriotylum FJ797576<br />

BRIP52426 Templeton P.myriotylum FJ797577<br />

BRIP52427 Templeton P.zingiberis. FJ797578<br />

DISCUSSION<br />

This is the first record of P. vexans and P. graminocola from<br />

ginger in Fiji and the first putative record of P. zingiberis in<br />

Australia. According to Dohroo (2005) P.graminicolum is present<br />

in Sri Lanka while P.vexans in India. The two species have been<br />

found to cause problems during rainy weather. The surveys are<br />

still relatively limited and there may be more species present in<br />

the Fijian ginger growing areas, or indeed other species present<br />

in Fiji and Australia with the capacity to cause rhizome rot in<br />

ginger. Consequently, more survey work is warranted. P.<br />

zingiberis has been recorded in Japan and Korea (5). In Japan, it<br />

has been isolated from various parts of rotten ginger, especially<br />

from the basal part of terrestrial stem and rhizomes regardless<br />

of stages of disease development and locations. It has been<br />

isolated from soils where ginger is growing and from areas<br />

where ginger has been previously grown. Morphologically,<br />

P.zingerberis is very similar to P.myriotylum and at the molecular<br />

level only a single base pair difference in the ITS2 region of the<br />

rDNA separates the two species (6). P.zingiberis has never been<br />

recorded in Australia, and to validate this record further<br />

morphological analysis is required.<br />

ACKNOWLEDGEMENTS<br />

This work was supported in part by a John Allwright Fellowship<br />

provided to the presenting author. We thank G &AM Stirling and<br />

Mike Smith for provision of the Australian isolates and advice in<br />

this project.<br />

REFERENCES<br />

1. Dohroo NP (2005) Diseases of ginger. In ‘Ginger, the genus<br />

Zingiber’. (Eds PN Ravindran, K Nirmal Babu) pp. 305–340. (CRC<br />

Press: Boca Raton)<br />

2. Stirling GR, Turaganivalu U, Lomavatu, MF, Stirling AM, Smith MK<br />

(2009). <strong>Australasian</strong> <strong>Plant</strong> <strong>Pathology</strong> (in press).<br />

3. Lee & Taylor (1990), Isolation of DNA from fungal mycelia and<br />

single spores. PCR Protocols: A guide to methods and applications.<br />

Academic Press, Inc.<br />

4. White, T.J., T. Bruns, S. Lee and J. Taylor. 1990. Amplification and<br />

direct sequencing of fungal ribosomal RNA genes for phylogenetics.<br />

In: PCR protocols. Eds. M.A. Innis, D.H. Gelfand, J.J. Sninisky and T.J.<br />

White. Academic Press, San Diego. pp. 315–322.<br />

5 Ichitani T and Shinsu T (1980) Ann.Phytopath. Soc.Japan 46: 435–<br />

441<br />

6. Levesque, C.A. and de Cock A.W.A.M. (2004). Mycological Research<br />

108: 1363–1383<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 173


Posters<br />

29 Development of techniques to measure SAR induction in broccoli for clubroot<br />

disease resistance<br />

D. Lovelock{ XE "Lovelock, D." } A , A. Agarwal B , E.C. Donald B , I.J. Porter B , R. Faggian C and D.M. Cahill A<br />

A School of Life and Environmental Sciences, Deakin University, Geelong, 3217, Victoria<br />

B Department of Primary Industries, Private Bag 15, Ferntree Gully DC, 3156, Victoria<br />

C Department of Primary Industries, 32 Lincoln Square, North Carlton, 3052, Victoria<br />

INTRODUCTION<br />

The phytohormone, salicylic acid, (SA) is required for a number<br />

of physiological processes within plants but primarily it is an<br />

important signalling molecule in plant defence, at both cellular<br />

and tissue levels but also systemically (1). Salicylic acid is<br />

implicated as a signal in defence against pathogens via systemic<br />

acquired resistance (SAR), a mechanism of induced defence that<br />

confers long‐lasting protection against a broad spectrum of<br />

microorganisms (2). An increased concentration of endogenous<br />

SA prior to and during SAR has been correlated with an increase<br />

in the production of pathogenesis‐related proteins throughout<br />

the plant. We are investigating SA‐induced SAR in broccoli<br />

following inoculation with Plasmodiophora brassicae. Molecular<br />

and biochemical methods are being developed to measure SAR<br />

induction in broccoli for the first time. A real‐time reverse<br />

transcriptase quantitative PCR (RT‐qPCR) has been developed to<br />

measure chitinase gene expression in plant tissue. Extraction<br />

from broccoli tissue and High Performance Liquid<br />

Chromatography (HPLC) analysis are being optimised to quantify<br />

SA levels post induction.<br />

MATERIALS AND METHODS<br />

Roots of broccoli (cv. Marathon) seedlings (10 to 14 day old)<br />

grown in trays were dipped in SA solution (1–5 mM<br />

concentration) for 15 minutes. Root and leaf sample pairs<br />

collected from the same plant were harvested 24 hours post<br />

treatment. Total RNA was isolated from roots and leaves of each<br />

sample for 3 replicates and checked by spectrophotometry for<br />

its quality and quantity. RNA was treated with DNase 1 and then<br />

reverse transcribed into cDNA and quantified by<br />

spectrophotometry. RT‐qPCR was conducted in duplicate for<br />

each sample using primers for two different genes—Actin‐8<br />

(house‐keeping gene) and Chitinase (gene of interest). The Delta‐<br />

Delta C t method was used for calculating the relative fold change<br />

of chitinase gene expression in treated samples compared to the<br />

control. SA extraction from broccoli is currently being optimised<br />

based on published protocols (3) and quantified using HPLC.<br />

RESULTS AND DISCUSSION<br />

The chitinase gene was chosen for this study as it is a known<br />

marker gene for measuring SAR induction and secondly it was<br />

possible to design the primers for B. oleracea. Following<br />

treatment of roots with 1 mM SA chitinase gene expression was<br />

increased above controls and was uniform throughout the plant<br />

in each of the three replicates. At higher concentrations up to 5<br />

mM chitinase gene expression was less consistent. These results<br />

indicate that a low dose of SA might be enough to trigger a good<br />

SAR response in the whole plant. A higher dose of SA might not<br />

necessarily induce a higher SAR response rather it may alter the<br />

physiology of the plant.<br />

mAU<br />

Figure 1. A typical chromatogram of a broccoli root extract that<br />

illustrates the retention time of SA (arrow) based on a separation<br />

performed on a standard 250mm x 0.5µm C 18 HPLC column.<br />

These preliminary studies have confirmed that SA can be used as<br />

an SAR inducer in broccoli and that the molecular and<br />

biochemical techniques under development can be used to<br />

measure SAR induction. Future work will determine the link<br />

between SA and resistance to P. brassicae in broccoli and the<br />

optimum level of SA required for SAR induction and disease<br />

resistance.<br />

ACKNOWLEDGEMENTS<br />

This work has been funded by DPI Victoria and Horticulture<br />

Australia Limited (HAL) using the vegetable levy and matched<br />

funds from the Australian Government. We thank Dr Xavier<br />

Conlan Deakin University, Geelong, for providing assistance in<br />

the SA analysis. We also thank Prof. Jutta Ludwig‐Muller,<br />

Technical University, Dresden, for her valued expertise at the<br />

beginning of this project. D. Lovelock is funded by a DPI‐Deakin<br />

University Post‐Graduate Scholarship.<br />

REFERENCES<br />

Retention time (min)<br />

1. Ludwig‐Muller J, Schuller A (2008) What can we learn from<br />

clubroots: alterations in host roots and hormone homeostasis<br />

caused by Plasmodiophora brassicae. European Journal of <strong>Plant</strong><br />

<strong>Pathology</strong> 121, 291–302.<br />

2. Durrant WE, Dong X (2004) Systemic acquired resistance. Annual<br />

Review of Phytopathology 42, 185–209.<br />

3. Li X et al (1999) Identification and cloning of a negative regulator of<br />

systemic acquired resistance, SNI1, through a screen for<br />

suppressors of npr1‐1. Cell 98, 229–339.<br />

Initial HPLC analysis has revealed that SA can be detected at<br />

nanomolar levels in broccoli root and shoot extracts post<br />

treatment with 1 mM SA (Fig 1). We are now in a position to<br />

examine the levels of SA that are present post SAR induction and<br />

the response to P. brassicae.<br />

174 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


52 Reducing the carbon footprint in Riverland vineyards: assessing the efficacy and<br />

efficiency of control for powdery mildew by evaluating growers’ spray diaries<br />

Posters<br />

P.A. Magarey{ XE "Magarey, P.A." } A , R.W. Emmett B , T.S. Smythe C , M.M. Moyer D , J.R. Dixon E , A.J. Pietsch F and P.M. Burne F<br />

A South Australian R&D Institute, PO Box 411, Loxton, 5333, SA<br />

B Department of Primary Industries, PO Box 905, Mildura, Vic, 3502<br />

C Riverland Wine Industry Development Council, 6 Kay Avenue, Berri, SA, 5343<br />

D NY Ag Exp Station, Geneva, NY, USA 14456<br />

E SA Murray‐Darling Basin Research Information Centre, 6 Kay Avenue, Berri SA 5343<br />

F CCW Co‐op Ltd, PO Box 238, Berri, SA, 5343<br />

INTRODUCTION<br />

A single sulphur spray to control powdery mildew (Erysiphe<br />

necator) in the Riverland region, comprising ~21,000 ha of<br />

viticulture near Loxton, South Australia (SA), costs ~$1.2 million.<br />

Like many other industries, the Australian wine grape industry,<br />

competes in international markets. Cost efficiencies and a<br />

demonstrably clean and sustainable green image are required.<br />

Because recent SA legislation aims to ensure that carbon<br />

emission targets are met, the SA wine grape industry is taking<br />

the initiative by fostering adoption of the cheapest and<br />

environmentally best disease control strategies.<br />

In line with this, we analysed grape grower spray diaries to: 1)<br />

assess the efficiency of spraying practices used for powdery<br />

mildew control; 2) provide a benchmark for current practice; and<br />

3) identify the scope for introducing modified spraying practices<br />

for optimum control using advanced knowledge of disease<br />

epidemiology.<br />

MATERIALS AND METHODS<br />

Paper‐based spray diaries supplied by two local wineries from<br />

2006/07 were converted to electronic format for the rapid and<br />

uniform review of individual records. The records contained<br />

vineyard details, spray information (dates, products, application<br />

details etc), and winery pre‐harvest assessments of powdery<br />

mildew severity. A MS Access ® ‐based, vineyard spray program<br />

evaluator with a theoretical optimum spray strategy based on<br />

best practice spray timing, fungicide treatment, and application<br />

technique, was used to: 1) compare each spray record in relation<br />

to temporally adjacent sprays; and 2) calculate a disease<br />

management score for each record. The scores for each spray<br />

event in each diary were then combined for each patch to<br />

provide a simple but rapid means of evaluating spray programs<br />

used on more than 1,200 patches of Chardonnay, 1,450 of Shiraz<br />

and 13 of Verdelho. Of the former, 138 were examined in detail.<br />

RESULTS AND DISCUSSION<br />

Analyses of the spray diaries showed that some growers<br />

achieved excellent control with as few as 2–4 sprays while others<br />

sprayed ≥ 12 times and achieved poor control. Many applied too<br />

many sprays with little benefit. For instance, Chardonnay<br />

growers applied an average of 6–7 sprays (range 2–14)/season<br />

and 25% applied ≥ 8 sprays annually. Of the total number, 73%<br />

(81% of total area surveyed) had at pre‐harvest, insignificant<br />

amounts of mildew while only 6% (3% area) had levels of disease<br />

sufficient for the winery to reject the crop. Significantly, there<br />

was no correlation between the number of sprays and the<br />

winery’s pre‐harvest disease score (Figure 1). It was not the<br />

number of sprays but their timing that was critical in controlling<br />

disease.<br />

A comparison of the disease management scores with the<br />

theoretical optimum spray strategy indicated that about 45% of<br />

growers were applying well‐timed programs whereas 55% were<br />

performing poorly. Nearly all growers were applying the right<br />

fungicide treatments but a lack of spray diary data on sprayer<br />

calibration, and therefore the efficiency of spray coverage,<br />

prevented accurate assessment of spray technique as a factor in<br />

disease control. Similarly, the lack of records of inoculum<br />

reservoirs inhibited the study of the efficacy of the sprays<br />

applied.<br />

Winery’s Mildew Disease Rating<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

Figure 1. Plot of the number sprays applied for powdery<br />

mildew/patch/season for cv. Chardonnay, 2007/08.<br />

Spatial analyses of the spray diary records indicated clustering of<br />

vineyards with higher disease scores suggesting that disease<br />

control might be improved by addressing cultural, sociological<br />

and/or meso‐climatic factors. In addition, higher incidences of<br />

powdery mildew occurred in patches with under‐vine irrigation<br />

compared to those with drip irrigation.<br />

Our analyses showed that the vineyard spray program evaluator<br />

could compare disease management practices from any spray<br />

diary of the same format to: 1) rapidly review past records of<br />

sprays applied in other seasons or other regions so that industry<br />

bench‐marks for the use of fungicides to control powdery<br />

mildew can be devised; 2) evaluate planned spray schedules for<br />

effectiveness and weakness before sprays are applied in the<br />

vineyard; and 3) determine which spray schedules are effective<br />

and which are not. From this, the minimum number of effective<br />

sprays needed to improve control of powdery mildew can be<br />

determined, reducing fuel and chemical use, lowering costs and<br />

reducing carbon emissions in Australian vineyards.<br />

CONCLUSION<br />

There is scope to adopt improved spray programs for better<br />

control of powdery mildew. The vineyard spray program<br />

evaluator could be developed as an online module in<br />

CropWatchOnline.com for growers, vineyard managers,<br />

consultants and others to assess their own records.<br />

ACKNOWLEDGEMENTS<br />

More Sprays Don’t Give Better Control<br />

0<br />

0 2 4 6 8 10 12 14<br />

Total Number Sprays/Season<br />

This research was initiated by the Riverland Wine Industry<br />

Development Council and supported in part by a Grape and<br />

Wine Research and Development Corporation RITA grant.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 175


Posters<br />

30 Incorporating host‐plant resistance to Fusarium crown rot into bread wheat<br />

D.J. Herde and C.D. Malligan{ XE "Malligan, C.D." }<br />

Department of Employment, Economic Development and Innovation, Primary Industries and Fisheries, Leslie Research Centre,<br />

Toowoomba, 4350, QLD<br />

INTRODUCTION<br />

Crown rot, caused predominantly by Fusarium<br />

pseudograminearum (teleomorph Gibberella coronicola), is a<br />

major soilborne disease problem in the wheat and barley<br />

industries. The disease is widespread and causes losses in yield<br />

and quality in Queensland, New South Wales, Victoria, and South<br />

Australia. Losses are estimated to be up to $56M in bread wheat<br />

throughout Australia. In Queensland, losses have been<br />

estimated at up to 50% in some areas and losses of 20 to 30%<br />

occur regularly, while the disease can inflict yield loss of up to<br />

89% (1).<br />

Breeding for resistance to crown rot has been difficult, partly<br />

due to variability associated with disease measurement, but also<br />

due to an incomplete understanding of the nature of the<br />

genetics of resistance.<br />

Previous work (2) found complex models of inheritance<br />

controlling crown rot resistance. This knowledge is being used to<br />

direct a number of different approaches aimed at building<br />

disease resistance levels.<br />

MATERIALS AND METHODS<br />

Of the bread wheat genotypes studied, two (Puseas and<br />

Kennedy) are susceptible and seven (2–49, CPI133814, IRN497,<br />

Lang, QT10162, Sunco, and W21MMT70) have partial resistance.<br />

The parent 2–49 is considered one of the strongest sources of<br />

resistance to crown rot currently available (3).<br />

The seedlings were assessed for crown rot resistance in a<br />

glasshouse test, following a modification of the Wildermuth and<br />

McNamara method (4). This method is a three week duration<br />

experiment that closely mimics field infection, and is highly<br />

correlated with field results.<br />

The approaches we took were:<br />

• selection in targeted crosses with knowledge of the genetic<br />

model<br />

• selection without knowledge of the genetic model<br />

• gene pyramiding using half‐sib crosses<br />

• gene pyramiding using molecular tools (DArT).<br />

RESULTS AND DISCUSSION<br />

Having the genetic information available enables an informed<br />

decision to be made about the difficultly in working with<br />

particular crosses.<br />

found in a fixed line. Further work is under way to compare<br />

selection in crosses with different types of epistatic control.<br />

Selection without knowledge of the genetic control can still<br />

provide useful results, but until arriving at a fixed line there will<br />

be uncertainty about whether the disease resistance is real or<br />

the product of unfixable gene interactions.<br />

Half‐sib crosses (where the male and female have a parent in<br />

common) combining two different sources of resistance, in this<br />

case IRN497 and the synthetic wheat CPI133814, were crossed<br />

into a common agronomic background (Sunco). This can be a<br />

useful method of elevating resistance by combining diverse<br />

resistance genes. Resistance levels in the progeny are extremely<br />

high after three rounds of selection (F2 to F4), with the best<br />

showing ~30% less disease severity than 2–49.<br />

DArT markers have been used to direct intercrosses between<br />

resistant selections from a cross of CPI133814 and IRN497,<br />

which was identified as the optimal cross for strongest crown rot<br />

resistance (from the listed parent set). DArT genotyping can<br />

identify gene differences in individuals that show the same level<br />

of resistance, enabling crosses to be made to maximise the<br />

amount of resistance genes within an individual plant. This<br />

strategy is aiming to produce a parental line for further<br />

development with elevated resistance levels beyond those<br />

currently available, rather than a variety for release, as the<br />

parents lack adaptation characteristics.<br />

Pre‐breeding selection work has commenced with the better<br />

performing crosses that include an adapted parent in the cross.<br />

REFERENCES<br />

1. Klein TA, Burgess LW, Ellison FW (1991) The incidence and spatial<br />

patterns of wheat plants infected by Fusarium graminearum Group<br />

1 and the effect of crown rot on yield. Australian Journal of<br />

Agricultural Research 42, 399–407.<br />

2. Herde DJ, McNamara RB, Wildermuth GB (2008) Obtaining genetic<br />

resistance to Fusarium crown rot in bread wheat. 11th<br />

International Wheat Genetics Symposium, August 24–29, 2008,<br />

Brisbane.<br />

3. Wildermuth GB, McNamara RB, Quick JS (2001) Crown depth and<br />

susceptibility to crown rot in wheat. Euphytica 122, 397—405.<br />

4. Wildermuth GB, McNamara RB (1994) Testing wheat seedlings for<br />

resistance to crown rot caused by Fusarium graminearum Group 1.<br />

<strong>Plant</strong> Disease 78, 949–953.<br />

Many of the crosses in this study had complex epistatic models<br />

controlling crown rot resistance. A number were controlled<br />

through an additive gene model or additive x additive epistasis,<br />

which allows the resistance to be captured in a fixed line. A<br />

number of other crosses with strong resistance were controlled<br />

with dominance or dominance x dominance epistasis, meaning<br />

the resistance will not be able to be captured in a fixed line.<br />

This information is able to guide selection of populations to<br />

advance, and explains why resistance in parent lines or<br />

segregating material alone will not guarantee resistance will be<br />

176 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


31 Management and distribution of huanglongbing in Pakistan<br />

Shahid Nadeem Chohan 1,3 , Obaid Aftab 1 , Raheel Qamar 1 , Shazia Mannan{ XE "Mannan, S." } 2 , Muhammad Ibrahim 2 , Iftikhar Ahmed 4 , M.<br />

Kausar Nawaz Shah 1 , Paul Holford 3 , G. Andrew C. Beattie 3<br />

1 Department of Biosciences, COMSATS Institute of Information Technology, Chak Shahzad Campus, Islamabad, Pakistan<br />

2 Department of Biosciences, COMSATS Institute of Information Technology, Sahiwal Campus, 520‐B Civil Lines, Jail Road, Sahiwal,<br />

Pakistan<br />

3 Centre for <strong>Plant</strong> and Food Science, University of Western Sydney, Locked Bag 1797, Penrith South DC, NSW 1797, Australia<br />

4 National Agricultural Research Centre, Pakistan Agricultural Research Council, Park Road, Islamabad, Pakistan<br />

Posters<br />

INTRODUCTION<br />

Citrus as a crop provides living for individual farmers and foreign<br />

exchange for Pakistan as it is one the major agricultural export<br />

commodities. This fruit crop is susceptible to many biotic<br />

stresses including diseases of which huanglongbing is perhaps<br />

the most devastating of all. The presence of this disease has<br />

previously been reported in Pakistan, however, the molecular<br />

evidence that it is caused by ‘Candidatus Liberibacter asiaticus’<br />

(α‐Proteobacteria) was provided only recently (1). Both the<br />

prevalence and severity of this disease are, however, yet to be<br />

ascertained. This information is imperative to devise a<br />

management strategy at national level for containing the losses<br />

sustained by the farmers. Farmers, as well as extension workers,<br />

also need to be made aware of the presence of, and damage<br />

caused by, this disease and a concerted extension campaign is<br />

required for this purpose. Due to the technical difficulties in the<br />

diagnosis of the disease, a diagnostic service should be provided<br />

to the farmers at more than one place in the country. To enable<br />

the medium equipped laboratories a testing kit can be useful.<br />

These points have been addressed in the present study.<br />

METHODOLOGY<br />

The Survey and Sampling Strategy. To assess the prevalence and<br />

severity of the disease, a survey and sampling strategy has been<br />

devised which we term as “triple five”, to survey the citrus<br />

orchards country wide and collect the suspect samples based<br />

upon visual symptoms.<br />

Pakistan is administratively divided in to four provinces (Punjab,<br />

Sindh, Balochistan and Sarhad or North West Frontier Province)<br />

and four territories (Islamabad Capital Territory, Federally<br />

Administered Areas (FATA), the Northern Areas (FANA), and<br />

Azad Kashmir). The provinces of Sarhad and Balochistan each<br />

have Provincially Administered Areas (PATA) as well. Each<br />

province or territory is further subdivided in to districts and<br />

tehsils. The citrus orchards are spread around the country with a<br />

few areas with dense plantings and others with little production.<br />

Based upon existing statistics, districts of Pakistan containing<br />

citrus orchards covering at least 100 hectares were selected for<br />

the survey. At least five orchards in each tehsil were surveyed to<br />

select five trees on the bases of visual symptoms. Five<br />

symptomatic leaves were collected from each tree and tested<br />

for the presence of the disease. Projects are also under way in<br />

these areas to eventually uplift the farmers’ income from citrus.<br />

These samples have been tested by both PCR and the iodinestarch<br />

test (IS (2 and 3)).<br />

RESULTS AND DISCUSSION<br />

The Incidence of HLB. Survey of over 12 districts in Punjab has<br />

been completed and over 400 samples tested for the presence<br />

of HLB. Usually, the IS and the PCR test results were in<br />

agreement. However, some samples showing positive IS tests<br />

were found to be negative by PCR. The data will be provided<br />

later during the presentation.<br />

Multiplexing of primers for ‘Ca. Liberibacter asiaticus’ with those<br />

for ‘Ca. L. americanus’ has been successful; however, ‘Ca. L.<br />

americanus’ has not been detected.<br />

Huanglongbing Map of Pakistan. Currently, a citrus map of<br />

Pakistan is being prepared based upon the available statistics<br />

about the fruit crop. This citrus map of Pakistan will be upgraded<br />

to the huanglongbing map of Pakistan based upon the resultsof<br />

this study. Year‐round meteorological data will be incorporated<br />

showing highest temperature in the areas in Pakistan.<br />

Extension activity. Two brochures have been published for the<br />

awareness of the extension workers and the local farmers each<br />

in English and Urdu disseminated. The same brochure is now<br />

planned to be expanded to include all the prevalent diseases and<br />

insect pests of citrus as well as agronomic and postharvest<br />

recommendations which will eventually lift economic<br />

circumstances of the individual farmers.<br />

Building nation‐wide diagnostic capability and capacity. The<br />

presence of HLB is detected by the following methods: the<br />

resumptive IS test and by PCR. A kit has been prepared<br />

containing all the ingredients required to do these tests. The Part<br />

A of the kit has theingredients for the IS test. Part B of the kit has<br />

the ingredients used for DNA release and PCR analysis including<br />

a 1 tube PCR method. This kit is currently in the final stages of its<br />

testing and will be released to the interested parties soon. Any<br />

laboratory with a standard thermal cycler and gel<br />

electrophoresis facility would be able to test the presence of<br />

disease using this technique. Therefore, it will be a great boost<br />

for diagnostic capacity building in the country. Training<br />

workshops are also planned for this purpose.<br />

Huanglongbing Testing Services. One of the main sources of the<br />

disease is budwood from the infected mother blocks. Most of<br />

the budwood is provided from the government‐owned or<br />

administrated citrus orchards. Screening these blocks for the<br />

presence of this disease will greatly help in suppressing the<br />

spread of HLB through infected budwood material. Therefore,<br />

we have established a service for the testing facility of these<br />

mother blocks.<br />

The service is also offered to general farmers and<br />

documentation has been prepared to educate them with the<br />

procedures of taking and posting samples for testing.<br />

These services will remain free of charge as long as funds remain<br />

available for this purpose.<br />

Seasonal Variation. Extremely high temperatures may restrict<br />

proliferation of the pathogen in infected trees. Data is currently<br />

being obtained from the Meteorological Department of Pakistan<br />

to find out the highest temperature range in the country. This<br />

information will be incorporated in to the citrus map of Pakistan.<br />

This map will indicate relationship between the disease and high<br />

temperatures in the citrus growing areas of Pakistan.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 177


Posters<br />

Effect of Temperature on the Pathogen and its Remaining DNA<br />

in Infected Leaves. The leaves from the same infected twig are<br />

being subjected to heat treatments of different temperatures<br />

and times, then tested for the presences of pathogen DNA using<br />

standard and real‐time PCR protocols to ascertain the<br />

concentration of DNA within the leaves.<br />

REFERENCES<br />

1. Chohan SN, Qamar R, Sadiq I, Azam M, Holford P, Beattie GAC<br />

(2007) Molecular evidence for the presence of huanglongbing in<br />

Pakistan. <strong>Australasian</strong> <strong>Plant</strong> Disease Notes 2, 37–38.<br />

2. Takushi T, Toyozato T, Kawano S, Taba S, Taba K, Ooshiro A,<br />

Numazawa M, Tokeshi M, (2007) Scratch method for simple, rapid<br />

diagnosis of citrus huanglongbing using iodine to detect high<br />

accumulation of starch in the citrus leaves. Japanese Journal of<br />

Phytopathology 73, pp. 3–8.<br />

3. Eng L (2007) A presumptive field test for huanglongbing (citrus<br />

greening disease). Senior Officers’ Conference, Department of<br />

Agriculture Sarawak, 11–14 December 2007, Kuching, Sarawak.<br />

178 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


75 The Fusarium oxysporum f. sp cubense tropical race 4 vectoring ability of the<br />

banana weevil borer (Cosmopolites sordidus)<br />

Posters<br />

R.A. Meldrum{ XE "Meldrum, R.A." } A,B,C , A.M. Daly B and L.T.T. Tran‐Nguyen B<br />

A Cooperative Research Centre for National <strong>Plant</strong> Biosecurity<br />

B <strong>Plant</strong> Industries, NT Department of Regional Development, Primary Industry, Fisheries and Resources, GPO Box 3000, Darwin, 0801, NT<br />

C School of Biological Sciences, The University of Queensland, St Lucia, 4072, QLD<br />

INTRODUCTION<br />

Fusarium wilt of banana is regarded as one of the most<br />

widespread and destructive plant diseases in the recorded<br />

history of agriculture (1). Fusarium wilt is caused by the soilborne<br />

fungus Fusarium oxysporum f. sp. cubense, Foc. Foc is<br />

present throughout the world where bananas are grown. A<br />

particularly virulent strain capable of attacking Cavendish<br />

bananas in the tropics and referred to as tropical race 4 (Foc<br />

TR4), was discovered in 1997 in the Northern Territory. It has<br />

since led to the closure of several banana plantations (2). To<br />

date, Foc TR4 does not occur in any other state of Australia. Foc<br />

TR4 is capable of killing plants faster than any other strain and<br />

disease can build up rapidly without control measures (3). There<br />

is no method for eradicating the fungus. Current control<br />

measures involve limiting the spread of the pathogen within and<br />

between farms. However, these measures have often been<br />

ineffective and disease has continued to appear in new areas.<br />

The reasons behind this spread have not always been easy to<br />

explain.<br />

The banana weevil borer, Cosmopolites sordidus, is present in<br />

most banana production areas throughout the world. It causes<br />

damage to banana plants by boring into the rhizome and<br />

pseudostem to feed and lay eggs (4). Weevils are capable of<br />

crawling between banana plantations. Therefore, it is important<br />

to know if they are capable of spreading Foc TR4.<br />

This project aims to determine the presence of Foc TR4 on or in<br />

banana weevils. This is highly important, especially if this<br />

pathogen is detected in areas of Australia or other countries<br />

where it currently is not present. Revealing whether banana<br />

weevils are potential vectors of Foc TR4 will provide a greater<br />

understanding of disease epidemiology and could assist in<br />

limiting spread.<br />

MATERIALS AND METHODS<br />

Pseudostem traps were set at a Foc TR4 infested research site,<br />

the Coastal Plains Banana Quarantine Station (CPBQS), as well as<br />

a site free from Foc TR4. A total of 50 weevils were collected<br />

from the traps at CPBQS, as well as ten control weevils from the<br />

site not infested with Foc TR4. They were individually vortexed<br />

for in sterile distilled water (SDW) to loosen the fungal spores on<br />

the external section of the weevil. After vortexing the insects<br />

were surface sterilised before the internal sections were<br />

macerated in SDW. The SDW solutions from both the ‘external’<br />

and macerated ‘internal’ sections were spread onto plates of<br />

malachite green agar. Fungal growth was subcultured onto<br />

potato dextrose agar and single spore isolates were obtained.<br />

DNA was extracted from putative Foc TR4 fungal growth and<br />

analysed using Foc TR4 specific primers to confirm the fungus<br />

identity.<br />

of the ten control weevils contained Foc TR4 spores, either<br />

internally or externally (Table 1).<br />

Table 1. Presence of Foc TR4 from Cosmopolites sordidus<br />

Trapping site<br />

Foc TR4 infested 50<br />

Not infested<br />

(control)<br />

No. of<br />

weevils<br />

analysed<br />

10<br />

Body part(s)<br />

analysed<br />

Internal 0<br />

External 10<br />

Internal 0<br />

External 0<br />

Recovery of<br />

Foc TR4 (%)<br />

DISCUSSION<br />

While Foc TR4 was not successfully detected in the internal<br />

sections of the weevils, it was detected on the external sections.<br />

These preliminary results imply that C. sordidus can act as a<br />

carrier for Foc TR4 and possibly assist with its dispersal within<br />

and between plantations. C. sordidus may also vector other<br />

strains of Foc present in Australia and throughout the world.<br />

ACKNOWLEDGEMENTS<br />

The authors would like to acknowledge The Department of<br />

Agriculture, Fisheries and Forestry for providing the funds for<br />

this project.<br />

REFERENCES<br />

1. Stover, R.H., and Simmonds, N.W. (1987). ‘Bananas, third edition’,<br />

Longman Scientific and Technical, UK. pp 310–314.<br />

2. Condé, B.D., and Pitkethley, R.N. (2001). The discovery,<br />

identification and management of banana fusarium wilt outbreaks<br />

in the Northern Territory of Australia. In: ‘Banana Fusarium wilt<br />

management: Towards sustainable cultivation’ (Eds. A.B. Molina,<br />

N.H. Masdek, and K.W. Liew,). International Network for the<br />

Improvement of Banana and <strong>Plant</strong>ain‐Asia and the Pacific Network,<br />

Los Banos, Philippines. pp 260–265.<br />

3. Daly, A. and Walduck, G. (2006). Fusarium wilt of bananas (Panama<br />

disease), Department of Primary Industry, Fisheries and Mines, NT,<br />

Agnote number I51.<br />

4. Gold, C. S., Pena, J. E. and Karamura, E. B. (2001). Biology and<br />

integrated pest management for the banana weevil Cosmopolites<br />

sordidus (Germar) (Coleoptera: Curculionidae). Integrated Pest<br />

Management Reviews 6(2): 79–155.<br />

RESULTS<br />

All the macerated internal sections of the weevils collected from<br />

the CPBQS tested negative for Foc TR4 fungal growth. However,<br />

viable spores were detected from the suspension created by the<br />

external sections of five of the 50 weevils by PCR analysis. None<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 179


Posters<br />

76 Genetic diversity of Pseudocercospora macadamiae populations by PCR‐RFLP<br />

A.K. Miles{ XE "Miles, A.K." } A , O.A. Akinsanmi A , E.A. Aitken B and A. Drenth A<br />

A Tree <strong>Pathology</strong> Centre, The University of Queensland and Primary Industries and Fisheries, 80 Meiers Rd Indooroopilly, Brisbane, 4068,<br />

Queensland<br />

B School of Biological Sciences, The University of Queensland, St Lucia, 4072, Queensland<br />

INTRODUCTION<br />

Pseudocercospora macadamiae is known only to exist and cause<br />

husk spot of macadamia in commercial orchards in Australia (1,<br />

2). Field trials and observations suggest that the asexual conidia<br />

of the fungus are the most important infective propagule in the<br />

disease cycle (3). The teleomorphic state of P. macadamiae is yet<br />

to be observed in nature or culture (4). We hypothesise that the<br />

husk spot disease cycle lacks a teleomorphic state of P.<br />

macadamiae.<br />

Testing the hypothesis that a teleomorph is absent from the<br />

husk spot disease cycle by means of direct survey of orchards or<br />

native vegetation is challenging. However, if the teleomorph is<br />

absent from commercial orchards, it is expected that the lack of<br />

sexual reproduction would result in predominantly clonal<br />

populations of the fungus. Therefore, we investigated the<br />

genetic diversity of field populations of P. macadamiae to test<br />

our hypothesis.<br />

MATERIALS AND METHODS<br />

In order to determine the genetic diversity of P. macadamiae<br />

populations, 105 isolates were collected from diseased fruit from<br />

trees in three orchards located at Bundaberg (Qld), Glasshouse<br />

Mountains (Qld), and the Northern Rivers (NSW), respectively.<br />

DNA was extracted and PCR‐RFLP performed in duplicate for 6<br />

genes (actin, β‐tubulin, calmodulin, EFA, G3P and ITS), each<br />

digested with three restriction enzymes.<br />

RESULTS<br />

Results of the PCR‐RFLP study showed that more than 80% of the<br />

isolates were of the same genotype, with the predominant<br />

genotype occurring at frequencies of 64, 95, and 79% at<br />

Bundaberg, Glasshouse Mountains and Northern Rivers,<br />

respectively (Fig 1). The Northern Rivers population included the<br />

highest number of genotypes for a single location (5). The<br />

number of polymorphic alleles differentiating the identified<br />

genotypes was low. Of the six genes studied, actin, β‐tubulin,<br />

and EFA were the most polymorphic, whilst no polymorphisms<br />

were detected in the calmodulin, G3P or ITS genes.<br />

DISCUSSION<br />

We hypothesised that the husk spot disease cycle lacks a<br />

teleomorphic state of P. macadamiae. Our study shows the<br />

genetic diversity of P. macadamiae populations to be low, and<br />

that a single genotype predominates at all the collection sites.<br />

The lack of genetic diversity is supportive of our hypothesis, and<br />

evidence for any significant role of a teleomorph in the husk spot<br />

disease cycle remains elusive.<br />

Figure 1. Location of Pseudocercospora macadamiae collection sites and<br />

pie charts representing the genotype frequency within individual, and all,<br />

populations. Different segments of pies represent unique genotypes.<br />

ACKNOWLEDGEMENTS<br />

We gratefully acknowledge the financial support of Horticulture<br />

Australia Ltd., Australian Macadamia <strong>Society</strong> Ltd., and The<br />

University of Queensland.<br />

REFERENCES<br />

1. Beilharz V, Mayers PE, Pascoe IG, (2003) Pseudocercospora<br />

macadamiae sp. nov., the cause of husk spot of macadamia.<br />

<strong>Australasian</strong> <strong>Plant</strong> <strong>Pathology</strong> 32, pp. 279–282.<br />

2. Akinsanmi OA, Miles AK, Drenth A, (2007) Timing of fungicide<br />

application for control of husk spot caused by Pseudocercospora<br />

macadamiae in Macadamia. <strong>Plant</strong> Disease 91, pp. 1675–1681.<br />

3. Miles AK, Akinsanmi OA, Aitken EAB, Drenth A. The disease cycle of<br />

Pseudocercospora macadamiae on macadamia. in 16th Biennial<br />

<strong>Australasian</strong> <strong>Plant</strong> <strong>Pathology</strong> <strong>Society</strong> Conference 24–27 September.<br />

2007. Adelaide, South Australia.<br />

4. Akinsanmi OA, Miles AK, Drenth A, (2008) Alternative fungicides for<br />

controlling husk spot caused by Pseudocercospora macadamiae in<br />

macadamia. <strong>Australasian</strong> <strong>Plant</strong> <strong>Pathology</strong> 37, pp. 141–147.<br />

180 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


32 Investigating the potential of in‐field starch accumulation tests for targeted citrus<br />

pathogen surveillance in Australia<br />

Posters<br />

A.K. Miles{ XE "Miles, A.K." } A , N. Donovan B , P. Holford C , R. Davis D , K. Grice E , M. Smith F and A. Drenth A<br />

A Tree <strong>Pathology</strong> Centre, The University of Queensland and Primary Industries and Fisheries, 80 Meiers Rd Indooroopilly, Brisbane, 4068,<br />

Qld<br />

B NSW DPI Elizabeth Macarthur Agricultural Institute, PMB 8, Camden, 2567, NSW<br />

C Centre for <strong>Plant</strong> and Food Science, University of Western Sydney, Locked Bag 1797, Penrith South DC, 1797, NSW<br />

D Northern Australian Quarantine Strategy, Australian Quarantine and Inspection Service, PO Box 96, Cairns International Airport, 4870,<br />

Qld<br />

E Mareeba Centre for Tropical Agriculture, Primary Industries and Fisheries, 28 Peters St, Mareeba, 4880, Qld<br />

F Bundaberg Research Station, Primary Industries and Fisheries, 49 Ashfield Road, Bundaberg, 4670, Qld<br />

INTRODUCTION<br />

Australia needs to be prepared to undertake large‐scale<br />

surveillance for the devastating, exotic disease of citrus,<br />

huanglongbing (HLB) (‘Candidatus Liberibacter’ species). One<br />

immediate obstacle is the potential for endemic diseases to<br />

exhibit HLB‐like symptoms, such as Australian citrus dieback<br />

(ACD) (possibly phytoplasma), and tristeza (Citrus tristeza virus)<br />

(CTV). Similar leaf symptoms include asymmetric<br />

mottling/chlorosis, vein yellowing and corking. In addition,<br />

effects of nutrient deficiency, wounding and other maladies can<br />

also be similar, adding to potential confusion. Because of this,<br />

the number of specimens collected based on visual symptoms<br />

would likely overwhelm current diagnostic capacity, and a rapid,<br />

in‐field test to help select diagnostic samples is needed.<br />

Starch accumulation has been well correlated to HLB symptoms<br />

in some studies (1–3) and could be a useful tool to aid<br />

surveillance. However, starch can accumulate due to nutrient<br />

deficiencies, insect damage and other factors including diseases<br />

other than HLB. Therefore, a preliminary study was undertaken<br />

to: i) compare two established starch tests; and ii) postulate the<br />

specificity of the starch tests amongst endemic citrus diseases.<br />

MATERIALS AND METHODS<br />

Citrus trees (grapefruit, mandarin, lemon, pommelo) and one<br />

tree of Murraya sp. were surveyed in the Burnett Basin,<br />

Queensland. Leaves showing asymmetric mottling and/or<br />

chlorosis, vein yellowing and/or corking, and corresponding<br />

healthy leaves were sampled. Leaves were field tested for starch<br />

accumulation through its reaction with iodine. The iodine<br />

reactions were carried out using both the “scratch” (2) and “leaf<br />

cut” (3) methods. The leaf cut method immersed the freshly cut<br />

edge of a leaf into iodine solution, then required inspection of<br />

the cut edge with a hand lens to visualise any inky blue/black<br />

colour development. The scratch method involved abrading the<br />

leaf surface with sandpaper, and then transferred the sandpaper<br />

to a ziplock bag containing an iodine solution. Any colour change<br />

was observed in the bag.<br />

Leaf samples were tested in the laboratory for pathogens<br />

capable of causing the symptoms of interest. Direct tissue blot<br />

immunoassay was used to detect CTV and molecular methods<br />

were used to detect phytoplasmas (ACD) and ‘Ca. Liberibacter’<br />

species causing HLB.<br />

RESULTS<br />

In total 31 samples (20 symptomatic and 11 healthy) were<br />

collected and tested. No healthy samples were positive for<br />

starch accumulation by either method. Positive, partial, and<br />

negative results were found amongst the symptomatic samples<br />

(Table 1). Variation between methods was least amongst<br />

positive samples, according to either method, with 3 out of 5<br />

possible samples in agreement (60%). The greatest variation<br />

between the two methods was in the production of ambiguous<br />

results, with only 1 of 13 possible samples being in agreement<br />

(8%).<br />

Table 1. Starch accumulation Number of symptomatic samples positive,<br />

partial or negative for starch accumulation, determined by scratch or leaf<br />

cut methods, and the number of samples for which the methods agree.<br />

Starch reaction Scratch Leaf cut Agreement<br />

Positive 5 4 3<br />

Partial 2 13 1<br />

Negative 13 3 3<br />

The six positive starch reactions (by either method) could not be<br />

clearly attributed to any particular pathogen. However, one leaf<br />

sample was taken from a wounded branch, and another noted<br />

as possible early symptoms of Mycosphaerella citri. The causal<br />

agent of HLB, Ca. Liberibacter was not detected in any samples<br />

using the laboratory diagnostic assays. CTV was detected in 23<br />

leaf samples, 7 of which were asymptomatic, healthy leaves. An<br />

additional 3 leaves had symptoms typical of ACD, tested positive<br />

for phytoplasmas, but had negative/partial starch reactions.<br />

DISCUSSION<br />

Starch accumulation was detected in Australian citrus leaf<br />

samples using both methods and could not be attributed to any<br />

of the pathogens tested for. This suggests that the test may be<br />

of limited use for pre‐HLB incursion surveillance. If starch<br />

accumulation can be demonstrated to be useful post‐incursion<br />

(when HLB might be responsible for the majority of starch<br />

accumulation), the ‘scratch’ method is preferred due to its<br />

delivery of clear results and practicality for field use. Individual<br />

sampling kits for the scratch method can be easily pre‐prepared,<br />

the results photographed for record keeping, and the waste selfcontained<br />

for removal from the survey site and disposal.<br />

ACKNOWLEDGEMENTS<br />

Assistance of Cherie Gambley is gratefully acknowledged.<br />

REFERENCES<br />

1. Eng L. A presumptive field test for huanglongbing (citrus greening<br />

disease). in Senior Officers' Conference, Department of Agriculture<br />

Sarawak, 11–14 December. 2007. Kuching, Sarawak.<br />

2. Takushi T, Toyozato T, Kawano S, Taba S, Taba K, Ooshiro A,<br />

Numazawa M, Tokeshi M, (2007) Scratch method for simple, rapid<br />

diagnosis of citrus huanglongbing using iodine to detect high<br />

accumulation of starch in the citrus leaves. Japanese Journal of<br />

Phytopathology 73, pp. 3–8.<br />

3. Etxeberria E, Gonzalez P, Dawson W, Spann TM, (2007) An iodinebased<br />

starch test to assist in selecting leaves for HLB testing.<br />

UF/IFAS EDIS HS375.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 181


Posters<br />

77 Priming for resistance against pathogens: cellular responses of Arabidopsis to<br />

UV‐C radiation<br />

S.J.L. Mintoff{ XE "Mintoff, S.J.L." }, P.T. Kay and D.M. Cahill<br />

Deakin University, School of Life and Environmental Sciences, Geelong, Victoria 3217<br />

INTRODUCTION<br />

<strong>Plant</strong>s face many biotic and abiotic challenges in the<br />

environment including pathogen challenge and damage due to<br />

energetic wavelengths of light in the ultraviolet (UV) region,<br />

both of which have detrimental effects on plants.<br />

When Arabidopsis thaliana was irradiated with doses of UV‐C<br />

light between 0 and 500 J/m ‐2 , leaves showed resistance towards<br />

compatible isolates of the Oomycete Hyaloperonospora<br />

arabidopsis in a dose‐dependant manner (1). Previous research<br />

has strongly suggested that this priming response is linked to<br />

DNA damage and repair, both of which are invoked after UV‐C<br />

irradiation (2).<br />

0 J/m ‐2<br />

500 J/m ‐2<br />

0 h 24 h 48 h<br />

It is still unclear, however, which signals and biochemical events<br />

regulate this induced defence response and how they may relate<br />

to DNA damage/repair. Also, it is still not known if the same<br />

phenomenon can be used to induce resistance against other<br />

pathogens or whether UV radiation may be having more subtle<br />

non‐targeted effects on host cells. This study examines the<br />

cellular and tissue responses of Arabidopsis leaves following<br />

exposure to doses of ultraviolet radiation.<br />

MATERIALS AND METHODS<br />

UV‐C treatments. Arabidopsis Col‐0 plants were irradiated with<br />

different doses of UV‐C (500 and 1000 J/m ‐2 ). <strong>Plant</strong>s were then<br />

returned to normal growth conditions. Leaf tissue was harvested<br />

at 0 and 24 hours post irradiation.<br />

Callose and reactive oxygen species assays. Leaves were stained<br />

for callose deposition with aniline blue (0.5%) and visualised<br />

using fluorescence microscopy. The production of hydrogen<br />

peroxide was visualised using 3, 3 diaminobenzidine (DAB).<br />

0 J/m ‐2<br />

0 h 24 h<br />

1000 J/m ‐2<br />

Figure 2. Leaf tissue stained with aniline blue, no callose is present in<br />

controls but at 500 and 1000 J/m ‐2 callose was induced at 24 and 48h<br />

(bright areas in the figure).<br />

RESULTS AND DISCUSSION<br />

H 2 O 2 production increased following UV treatment (Figure 1).<br />

Lack of DAB staining in the control showed low levels of H 2 O 2 to<br />

be present. However, at 500 J/m ‐2 increased H 2 O 2 production at<br />

24 hours was observed. Increased H 2 O 2 accumulation could also<br />

be seen in leaves exposed to 1000 J/m ‐2 at 0 and 24 hours. H 2 O 2<br />

production was associated with collapsed cells that appear to<br />

have undergone cell death.<br />

UV‐C radiation stimulated the deposition of callose in cells and<br />

cell walls in both a time and dose dependent manner (Figure 2).<br />

Deposition of callose was most intense at 1000 J/m ‐2 but less so<br />

for treatment at 500 J/m ‐2 at both 24 and 48 hours. Future work<br />

aims to determine if the induced resistance seen in interactions<br />

of Arabidopsis with H. arabidopsis occurs in other interactions of<br />

pathogens with different lifestyles (eg. biotrophs, hemibiotrophs<br />

and necrotrophs).<br />

ACKNOWLEDGEMENTS<br />

We thank the Australian Research Council for funding.<br />

500J/m ‐2<br />

REFERENCES<br />

1. Kunz B.A, Cahill D.M, Mohr P.G, Osmond M.J, Vonrax E.J (2006)<br />

<strong>Plant</strong> Responses to UV Radiation and Links to Pathogen Resistance.<br />

International review of cytology 225: 1–40<br />

2. Kunz, B.A, Dando, P.K, Grice D.M, Mohr P.G, Schenk P.M, Cahill D.M<br />

(2008) UV‐induced DNA damage promotes resistance to the<br />

biotrophic pathogen Hyaloperonospora parasitica in Arabidopsis.<br />

<strong>Plant</strong> Physiology 148: 1021–1031<br />

1000J/m ‐2<br />

Figure 1. UV‐C treated leaf tissue stained with DAB, dark spots indicate<br />

the presence of H 2 O 2 within cells (solid arrows). Dead cells can be seen as<br />

stained collapsed cells (unfilled arrows).<br />

182 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


11 Boneseed rust: a highly promising candidate for biological control<br />

L. Morin{ XE "Morin, L." } A and A.R. Wood B<br />

A CSIRO Entomology, GPO Box 1700, Canberra, ACT 2601, Australia<br />

B Agricultural Research Council, <strong>Plant</strong> Protection Research Institute, P Bag X5017, Stellenbosch 7599, South Africa<br />

Posters<br />

INTRODUCTION<br />

The woody evergreen shrub boneseed (Chrysanthemoides<br />

monilifera ssp. monilifera), which originates from South Africa, is<br />

a major invasive plant of natural ecosystems in Victoria and<br />

South Australia. Small infestations also occur in Tasmania.<br />

Improving the effectiveness of the biological control program<br />

against boneseed has been identified as a high research priority<br />

in the National Strategy for this Weed of National Significance,<br />

as none of the six insect agents released so far have established<br />

in the field.<br />

The systemic South African rust fungus, Endophyllum<br />

osteospermi, is a highly promising biological control agent for<br />

boneseed because it reduces growth and reproduction of plants<br />

by causing extensive deformation of infected branches (witches’<br />

brooms). In South Africa the rust appears to be a primary cause<br />

for the decline and death observed in some local boneseed<br />

populations (1, 2). The systemic nature of the rust is a desirable<br />

characteristic for biological control purposes as once the fungus<br />

is established within the host the infection is retained until the<br />

death of the witches’ brooms. The boneseed rust is only<br />

recorded in South Africa on a small group of related plants of the<br />

genera Chrysanthemoides and Osteospermum (Calenduleae:<br />

Asteraceae). As there are no indigenous representatives of the<br />

Calenduleae in Australia, the non‐target plants most at risk from<br />

this rust fungus are the introduced, ornamental species<br />

belonging to this tribe.<br />

A novel approach had to be taken to test the host‐specificity of<br />

E. osteospermi, because it develops visible symptoms only 1–3<br />

years after infection of its host. We present in this paper results<br />

from host‐specificity tests so far.<br />

MATERIALS AND METHODS<br />

Host‐specificity tests. Tier 1 tests were performed on detached<br />

leaves of plant species of the approved test list to determine,<br />

using microscopy techniques, whether the rust was capable of<br />

penetrating epidermal cells of non‐target species. Additional Tier<br />

1 tests were also carried out on leaves still attached to plants of<br />

some of the species.<br />

Tier 2 tests on leaves still attached to plants of the species where<br />

penetration occurred in Tier 1 tests were performed to<br />

determine if the fungus was capable of colonising tissue of these<br />

species in the weeks following inoculation.<br />

Three series of Tier 2 tests were performed in South Africa and<br />

in the CSIRO quarantine facility in Canberra. No leaf colonisation<br />

was observed on boneseed and any of the other species tested.<br />

These results cast doubts on the belief that the fungus colonises<br />

plants via young leaves. It is possible that infection of axillary<br />

buds is essential for further colonisation. Alternatively,<br />

conditions during tests may have been slightly suboptimal and<br />

prevented infection to occur.<br />

DISCUSSION<br />

The difficulties encountered with Tier 2 tests prompted us to<br />

initiate a series of Tier 3 tests, whereby whole plants are<br />

repeatedly inoculated with the rust fungus over a few weeks and<br />

maintained for up to 2–3 years until witches’ broom symptoms<br />

developed. These tests were conducted in late winter 2008 and<br />

witches’ broom symptoms have not yet developed on boneseed<br />

or any other species inoculated.<br />

Results from host‐specificity tests will be used to fully assess the<br />

risk of significant impact on non‐target plant species, before<br />

deciding if an application for the release of this rust fungus in<br />

Australia should be submitted to the authorities.<br />

ACKNOWLEDGEMENTS<br />

We thank Ms Gwen Samuels (PPRI) and Melissa Piper (CSIRO) for<br />

technical assistance. Financial support from CSIRO, Australian<br />

and New Zealand Environment and Conservation Council<br />

(ANZECC) and Land and Water Australia (Defeating the Weed<br />

Menace R&D initiative) is gratefully acknowledged.<br />

REFERENCES<br />

1. Wood AR (2002) Infection of Chrysanthemoides monilifera by the<br />

rust fungus Endophyllum osteospermi is associated with a reduction<br />

in vegetative growth and reproduction. <strong>Australasian</strong> <strong>Plant</strong><br />

<strong>Pathology</strong> 31, 409–415.<br />

2. Wood AR, Crous PW (2005) Epidemic increase of Endophyllum<br />

osteospermi (Uredinales, Pucciniaceae) on Chrysanthemoides<br />

monilifera. Biocontrol Science and Technology 15, 117–125.<br />

3. Wood AR (2006) Preliminary host specificity testing of Endophyllum<br />

osteospermi (Uredinales, Pucciniaceae), a biological control agent<br />

against Chrysanthemoides monilifera ssp. monilifera. Biocontrol<br />

Science and Technology 16, 495–507.<br />

RESULTS<br />

In Tier 1 tests, successful penetration was observed on boneseed<br />

and its close relative species tested within the Calenduleae tribe<br />

(bitou bush [C. monilifera subsp. rotundata], Osteospermum<br />

spp., Dimorphotheca jucundum) (3). Penetration also occurred<br />

on four other species outside the Calenduleae (Gazania rigens,<br />

Gerbera jamesonii, Bedfordia arborescens, Eucalyptus<br />

cladocalyx). Additional Tier 1 tests carried out on leaves still<br />

attached to plants confirmed accuracy of results obtained with<br />

detached leaves. Penetration of epidermal cells however, does<br />

not necessarily imply that the infection process will continue and<br />

be successful.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 183


Posters<br />

10 Can additional isolates of the Noogoora burr rust fungus be sourced to enhance<br />

biocontrol in northern Australia?<br />

L. Morin{ XE "Morin, L." } A , M. Piper A , R. Segura B and D. Gomez A<br />

A CSIRO Entomology, GPO Box 1700, Canberra, ACT 2601, Australia<br />

B CSIRO Entomology, Mexican Field Station, Veracruz, Mexico<br />

INTRODUCTION<br />

Noogoora burr (Xanthium occidentale) is an invasive plant across<br />

northern Australia, mostly inhabiting riparian areas. The exotic<br />

rust fungus Puccinia xanthii, illegally or accidentally introduced<br />

to Australia in the mid‐1970s, has been highly effective in<br />

controlling Noogoora burr in south‐eastern Queensland, but has<br />

had limited impact in tropical northern Australia (1). The<br />

introduction of additional isolates of this rust fungus better<br />

adapted to tropical conditions has been suggested as an<br />

approach to improve impact of this biological control agent in far<br />

northern regions (2).<br />

We present results from 1) surveys carried out in tropical<br />

America, 2) pathogenicity tests with Australian and exotic<br />

isolates of P. xanthii performed on Australian accessions of<br />

Noogoora burr and other Xanthium spp. and 3) the identification<br />

of a diagnostic marker that differentiates between Australian<br />

and exotic rust isolates.<br />

MATERIALS AND METHODS<br />

Surveys. Surveys for exotic isolates of P. xanthii were<br />

undertaken in 2007 in Venezuela, Mexico and Dominican<br />

Republic, areas that climatically match those of northern<br />

Australia where the rust fungus has not been highly effective (2).<br />

Rust‐infected material collected was pressed and dried and then<br />

placed at 4°C until shipment to the CSIRO quarantine facility in<br />

Canberra.<br />

Pathogenicity tests. Noogoora burr plants (ex. Daly River, NT)<br />

were inoculated with telia (3) collected from each of the<br />

overseas sites and from two purified Australian rust isolates (ex.<br />

Daly River and Victoria River, NT). This trial was repeated once.<br />

Additional inoculations of other Australian accessions of<br />

Noogoora burr and other Xanthium spp. were also performed.<br />

Diagnostic marker. Primers were designed from characterised<br />

Simple Sequence Repeats (SSR) loci isolated from P. xanthii.<br />

Primers were screened on the DNA of 29 single‐telium isolates<br />

from Australia, Hungary, Brazil, Argentina, Mexico and<br />

Dominican Republic.<br />

RESULTS<br />

Surveys. P. xanthii was found at 9 of the 13 Xanthium‐infested<br />

sites surveyed in Dominican Republic, and at three of the four<br />

infested sites in Mexico. None of the plants of the two Xanthium<br />

populations found in Venezuela were infected by the fungus.<br />

Pathogenicity tests. None of the Australian accessions of<br />

Noogoora burr and other Xanthium spp. inoculated with the<br />

exotic rust isolates developed disease symptoms, even though<br />

germination tests indicated the inoculum was viable.<br />

Microscopic examination of cleared and stained leaf samples<br />

from these plants showed typical plant resistance responses<br />

following penetration by the rust. In all trials, plants inoculated<br />

with the two Australian rust isolates developed disease<br />

symptoms. Closer examination of Xanthium specimens collected<br />

in Dominican Republic and Mexico revealed that they were<br />

morphologically and genetically different to Australian<br />

accessions (data not shown).<br />

Diagnostic marker. Several of the eight SSR markers identified<br />

differentiated rust isolates from Australia, Mexico and<br />

Dominican Republic, with SSR px09 being the most reliable<br />

diagnostic marker.<br />

DISCUSSION<br />

This body of work indicated that Australian Noogoora burr<br />

plants, as well as other Xanthium spp., are resistant to P. xanthii<br />

isolates collected in Dominican Republic and Mexico. This came<br />

as a major surprise considering previous research showed that<br />

an Australian isolate of the rust was capable of infecting the four<br />

different Xanthium species found in the ‘Noogoora burr<br />

complex’ in Australia: X. occidentale, X. italicum, X. orientale and<br />

X. cavanillesii (3). As a result, it was not possible to establish<br />

cultures of the exotic rust isolates for subsequent host‐specificity<br />

tests.<br />

SSR marker px09, could be developed into a simple PCR‐based<br />

diagnostic tool to differentiate between exotic and Australian<br />

isolates of P. xanthii. Such a tool would be valuable for<br />

monitoring the establishment of additional isolates of P. xanthii<br />

from tropical America after their release in Australia, providing<br />

that isolates pathogenic to Australian Noogoora burr plants can<br />

be found in the future.<br />

A more extensive follow‐up project is required to deliver on the<br />

original goal of introducing additional isolates of P. xanthii better<br />

adapted to the climate of tropical northern Australia. The<br />

establishment of an outdoor experimental garden consisting of<br />

northern Australian accessions of Noogoora burr in tropical<br />

America would be required to source pathogenic rust isolates.<br />

ACKNOWLEDGEMENTS<br />

We are grateful to collaborators in Queensland, Northern<br />

Territory and NSW for collecting and sending seeds from various<br />

Noogoora burr infestations. Financial support from CSIRO,<br />

DAFWA and Land and Water Australia (Defeating the Weed<br />

Menace R&D initiative) is gratefully acknowledged.<br />

REFERENCES<br />

1. Morin L, Auld BA, Smith HE (1996) Rust epidemics, climate and<br />

control of Xanthium occidentale. In ‘Proceedings of the IX<br />

International Symposium on Biological Control of Weeds’ (Eds VC<br />

Moran, JH Hoffmann) pp. 385–391. (University of Cape Town).<br />

2. van Klinken RD, Julien MH (2003) Learning from past attempts:<br />

does classical biological control of Noogoora burr (Asteraceae:<br />

Xanthium occidentale) have a promising future? Biocontrol Science<br />

and Technology 13, 139–153.<br />

3. Morin L, Auld BA, Brown JF (1993) Host‐range of Puccinia xanthii<br />

and post‐penetration development on Xanthium occidentale.<br />

Canadian Journal of Botany 71, 959–965.<br />

184 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


12 Helicotylenchus nematode contributing to turf decline in Australia<br />

L. Nambiar{ XE "Nambiar, L." } A , J M. Nobbs B and M. Quader A<br />

A Department of Primary Industries, Private Bag 15, Ferntree Gully Delivery Centre Vic 3156, Australia<br />

B South Australian Research and Development Institute, <strong>Plant</strong> and Soil Health, <strong>Plant</strong> Research Centre, GPO Box 397, Glen Osmond, SA<br />

5064, Australia<br />

Posters<br />

INTRODUCTION<br />

Helicotylenchus species (spiral nematodes) are the most<br />

commonly detected plant‐parasitic nematodes found in turf<br />

samples. These nematodes are generally considered as<br />

migratory ectoparasitic and/or migratory semi‐endoparasitic<br />

feeders, that is, they feed from outside the roots or partially<br />

embedded inside the roots. Damage caused by these nematodes<br />

is light to dark brown or reddish brown necrotic lesions. The aim<br />

of this work was to identify the spiral nematode species causing<br />

turf decline in Australia.<br />

Turf samples were submitted to the Crop Health Services, DPI<br />

Victoria by green keepers and turf consultant companies for the<br />

diagnosis and assessment of plant‐parasitic nematodes<br />

associated with turf decline. These samples were from all<br />

mainland states as well as Tasmania and collected during 1996–<br />

2008.<br />

Helicotylenchus species can be difficult to identify and require<br />

comparison of characters such as tail shape, body on death,<br />

position of vulva and shape of lip region. Morphometrics are also<br />

used which include body length, stylet length, tail length and<br />

position of phasmids in relation to the anus.<br />

METHOD<br />

Extraction and Fixing. Nematodes were extracted from 200ml<br />

soil samples using the Whitehead tray technique, incubated at<br />

room temperature about 25°C for 48 hours. The specimens were<br />

collected in a 38µm sieve, fixed in 4:1 formalin‐acetic acid<br />

fixative and then processed through an alcohol/glycerol series to<br />

be mounted in glycerol on permanent slides. The slides were<br />

deposited in the Victorian <strong>Plant</strong> <strong>Pathology</strong> Herbarium, DPI<br />

Victoria.<br />

RESULTS<br />

Identification. The main morphological characters on which the<br />

identifications are based are presented in Table 1. 3,812 samples<br />

were received from bowling and golf greens for counts and<br />

identification of plant nematodes. Figure 1 shows New South<br />

Wales had the most samples from identified locations compared<br />

to other states (Fig.1).<br />

average of 393–6,345 (Table 2).The specimens found in turf<br />

samples were identified as Helicotylenchus pseudorobustus, H.<br />

dihystera and H. erythrinae. The most commonly found species<br />

was H. pseudorobustus and also it was recorded for the first time<br />

on turf in Australia. There were occasions when more than one<br />

species of Helicotylenchus was identified in a sample.<br />

Table 1. Morphological characters used to identify three species of<br />

Helicotylenchus from turf samples.<br />

Helicoytlenchus<br />

species<br />

Body<br />

length<br />

(μm)<br />

H. pseudorobustus 706–<br />

822<br />

H. dihystera 610–<br />

860<br />

H. erythrinae 480–<br />

610<br />

Tail shape<br />

Tail with distinct<br />

smooth ventral<br />

projection<br />

Tail with<br />

indistinct smooth<br />

ventral projection<br />

Tail with distinct<br />

pointed ventral<br />

projection<br />

Lip<br />

Annule<br />

Position of<br />

phasmid<br />

4–5 5–8 annules<br />

anterior to<br />

anus<br />

4–5 6–12 annules<br />

anterior to<br />

anus<br />

6–12 2 annules<br />

posterior, 4<br />

annules<br />

anterior to<br />

anus<br />

Table 2. Number of identified locations, samples and average of<br />

Helicotylenchus spp detected from various states of Australia.<br />

States<br />

No. bowling<br />

and golf<br />

greens<br />

Number of<br />

samples<br />

Average number of<br />

Helicotylenchus per<br />

sample (200 ml soil)<br />

SA 23 139 6,345<br />

VIC 90 289 1,234<br />

NSW 146 854 870<br />

QLD 34 116 1,662<br />

WA 10 25 613<br />

TAS 7 11 393<br />

DISCUSSION<br />

The threshold damage caused by spiral nematodes on turf has<br />

been calculated as 600 nematodes per 200ml soil. Of the total<br />

1955 Helicotylenchus detected samples, 30% of these were<br />

above the damage threshold. Helicotylenchus spp were found to<br />

be associated with declining couch grass (Cynodon dactylon) and<br />

for the first time in bent grass (Agrostis tenuis). Further survey<br />

work is required to investigate the distribution of each identified<br />

species and their role in turf decline in Australia.<br />

Figure 1. Distribution of Helicotylenchus in Australia<br />

The highest numbers per 200ml soil were detected from South<br />

Australia and lowest from Tasmania. The numbers of<br />

Helicotylenchus species present in each sample varied from an<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 185


Posters<br />

78 The effect of phosphonate on the accumulation of camalexin following challenge<br />

of Arabidopsis by Phytophthora palmivora<br />

Zoe‐Joy Newby{ XE "Newby, Z.J." }, Rosalie Daniel and David Guest<br />

Faculty of Agriculture, Food and Natural Resources, The University of Sydney, 2006, NSW, Australia<br />

INTRODUCTION<br />

Potassium phosphonate (phosphonate) is commonly used to<br />

prevent and treat disease caused by various species of<br />

Phytophthora. Phosphonate acts by enhancing the innate<br />

defence response of the host plant (1), although the precise<br />

mechanisms remain unknown. Arabidopsis thaliana provides an<br />

opportune model for the study of plant‐pathogen interactions<br />

and when inoculated with P. palmivora, as phosphonate<br />

application induces an incompatable defence response (2). Part<br />

of this interaction may include the accumulation of the<br />

phytoalexin camalexin, however A. thaliana mutant studies<br />

suggest that camalexin is not always required for resistance. By<br />

assessing the effect of phosphonate at individual steps of the<br />

camalexin biosynthesis pathway, we aim to learn more about<br />

the phosphonate‐induced defence response and it’s role in<br />

pathogen restriction.<br />

METHODS<br />

<strong>Plant</strong> Material and Inoculations. Approximately 0.2mg of A.<br />

thaliana seed was added to each 125mL conical flask containing<br />

half strength MS supplemented with 20% sucrose with or<br />

without 100 mg/L phosphonate,pH6.5 (Agri‐Fos 600; Agrichem).<br />

Each flask was stoppered then shaken (120 rpm) under 12h<br />

light/dark at 24°C. After 21 days seedlings were inoculated with<br />

zoospores of P. palmivora (2).<br />

Extraction and analysis. Camalexin was sampled at 0, 12, 24 and<br />

48h post‐inoculation (pi). Approximately 0.3g of macerated<br />

tissue was boiled in 80% methanol (20 min, 65°C) then stored<br />

overnight at 4°C. Samples were centrifuged and the liquid<br />

fraction removed and evaporated. The residue was extracted<br />

three times with chloroform and the pooled chloroform fraction<br />

was evaporated to dryness. The pellet was dissolved in 500uL of<br />

50% menthol then samples were filtered. Camalexin was<br />

quantified utilising a C18 column on a Dionex HPLC system and<br />

verified with synthetic camalexin (kindly provide by Jane<br />

Glazebrook, University of Minnesota). Data were transformed if<br />

required and analysed using ANOVA.<br />

RESULTS AND DISCUSSION<br />

Camalexin accumulation in all treatments rapidly increased<br />

between 12 and 24h pi, and then decreased between 24 and 48h<br />

(Figure 1). There were no significant differences between any of<br />

the treatments (p=0.058), however the greatest increase in<br />

camalexin accumulation followed inoculation, whether plants<br />

had been treated with phosphonate or not.<br />

The ‘peak’ in camalexin accumulation at 24h has been reported<br />

in several ecotypes of Arabidopsis and can result from either<br />

biotic or abiotic factors (3). Of all treatments, we found that<br />

inoculated seedlings had the highest level of camalexin<br />

accumulation between 24–48 h pi. This supports the<br />

identification of camalexin as a phytoalexin, produced in<br />

response to pathogen attack.<br />

Figure 1. Camalexin Production in Arabidopsis treated with<br />

phosphonate and infected with P. palmivora; n=6; bars indicate<br />

standard error<br />

Accumulation as a result of abiotic factors reported previously<br />

(3), may explain why camalexin was routinely detected in<br />

uninoculated, untreated seedlings indicating that experimental<br />

conditions were enough to cause a small amount of camalexin to<br />

accumulate.<br />

Phosphonate reduced camalexin accumulation in uninoculated<br />

seedlings at 24 and 48 h pi to levels below that of controls.<br />

Inoculation of phosphonate‐treated seedlings restored<br />

camalexin accumulation however differences were not<br />

significant. This reduction in camalexin accumulation following<br />

phosphonate application was unexpected but has been<br />

confirmed in subsequent experiments.<br />

Although phosphonate has been shown to elicit an incompatible<br />

interaction between Arabidopsis and P. palmivora (2), camalexin<br />

does not appear to play a significant role in phosphonateinduced<br />

resistance. This indicates that the interaction between<br />

phosphonate‐treated Arabidopsis and P. palmivora involves the<br />

activation of multiple lines of defence. The signalling pathways<br />

activated by phosphonate in this pathosystem still remain<br />

unclear, but the availability of Arabidopsis defence‐related<br />

mutants will aid elucidation of the interaction.<br />

REFERENCES<br />

1. Guest & Grant (1991) the complex mode of action of phosphonates<br />

as antifungal agents. Biological Review 66, 159–187<br />

2. Daniel R & Guest D (2006) Defence responses induced by<br />

potassium phosphonate in Phytophthora palmivora‐challenged<br />

Arabidopsis thaliana. Physiological and Molecular <strong>Plant</strong> <strong>Pathology</strong><br />

67, 194–201<br />

3. Schuhegger R, Rauhut T & Glawischnig E (2007) Regulatory<br />

variability of camalexin biosynthesis. Journal of <strong>Plant</strong> Physiology<br />

164, 636–644<br />

186 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


79 Characterisation of Phytophthora capsici isolates from black pepper in Vietnam<br />

N.V. Truong{ XE "Truong, N.V." } A , L.W. Burgess B and E.C.Y. Liew C<br />

A School of Agriculture and Forestry, Hue University, 102 Phung Hung Street, Hue city, Vietnam<br />

B Faculty of Agriculture, Food and Natural Resources, The University of Sydney, NSW 2006, Australia<br />

C Royal Botanic Gardens Sydney, Botanic Gardens Trust, Mrs Macquaries Road, Sydney, NSW 2000, Australia<br />

Posters<br />

INTRODUCTION<br />

Since the establishment of Phytophthora capsici as the causal<br />

agent of black pepper foot rot in Vietnam (Truong et al. 2008),<br />

there is an urgent need to understand the population structure<br />

of this pathogen. The role of genetic diversity and geographic<br />

structuring of P. capsici foot rot epidemics of black pepper is not<br />

known. It is assumed that environmental effects, host<br />

susceptibility and cultivation practices facilitate selection<br />

pressures, which in turn affects the changes in the pathogen<br />

population structure, and subsequently the pattern of disease<br />

incidence. In order to make decisions regarding the direction of<br />

disease management strategies, the population structure of this<br />

pathogen needs to be explored. We begin to address these<br />

issues by testing two hypotheses. The first is that only one<br />

mating type exists in the P. capsici population from black pepper<br />

in Vietnam. The second is that the P. capsici population is<br />

genetically undifferentiated in two different climatic regions.<br />

consists of isolates obtained from all provinces in both regions.<br />

The results also indicate that isolates are not genetically<br />

correlated with mating type.<br />

MATERIALS AND METHODS<br />

Isolates origin. Phytophthora capsici was isolated from soil and<br />

plant samples obtained from provinces in the Southeast region<br />

(SE) and the North Central Coast region (NCC) (Truong et al.<br />

2008).<br />

Mating type analysis. Each isolate was paired on V8 Agar with<br />

known A1 and A2 testers. Test isolates producing oospores with<br />

both A1 and A2 were scored as A1A2. The test was replicated 3<br />

times.<br />

DNA extraction, RAMS and REP‐PCR protocol and genetic<br />

analysis. Fungal mycelium was grown in liquid medium,<br />

incubated at 25°Cin the dark for 5–6 days and DNA was<br />

extracted. The RAMS and REP‐PCR fingerprinting protocols were<br />

performed as previously described by Hantula et al. (1996) and<br />

Rademaker & Bruijn (1997). Cluster analysis was performed<br />

using the DICE similarity coefficient and UPGMA agglomeration<br />

in the program NTSYSpc.<br />

RESULTS<br />

Mating type analysis. Both A1 and A2 mating types were found to<br />

co‐exist within the same farm in 13 cases in Dong Nai (SE) and<br />

Quang Tri (NCC) provinces. In addition, A1 and A2 mating types<br />

were also observed to co‐exist on the same plant in one case in<br />

Quang Tri province.<br />

RAM and REP analysis. In order to assess the overall genetic<br />

diversity of the whole population and relationship between the<br />

two regional subpopulations, the combined data from RAMS and<br />

REP analyses were used to construct an UPGMA dendrogram (Fig<br />

1). The P. capsici isolates from black pepper are distributed in<br />

two main groups, I and II, which are differentiated at DICE<br />

similarity of 53%. Group I comprises 114 isolates with 108<br />

belonging to one large clonal group, two isolates in a small clonal<br />

group and four with unique phenotypes. Group II comprises four<br />

isolates, all with unique phenotypes. The genetic similarity<br />

analysis showed that more than 91% of all isolates were<br />

genetically identical and the whole population was nearly<br />

homogeneous. The clustering of isolates in the dendrogram does<br />

not correlate with geographic origin. The large clonal group<br />

Figure 1. UPGMA dendrograms of 118 isolates of Phytophthora capsici<br />

based on combined RAMS and REP data<br />

DISCUSSION<br />

The analysis of P. capsici isolates obtained from various growing<br />

regions revealed the presence of both mating types in two<br />

different climatic regions, with the A2 type detected at higher<br />

frequency than the A1 type. Overall the level of genetic diversity<br />

detected among the P. capsici isolates from black pepper was<br />

relatively low. One hundred and eight isolates were found to be<br />

identical in their RAMS and REP phenotypes. The genetic pattern<br />

of the P. capsici population was not found to be associated with<br />

geographic origin, and hence the two regional subpopulations<br />

were undifferentiated. The current report provides preliminary<br />

but useful information concerning the extent of genetic diversity<br />

and geographic distribution of P. capsici from black pepper in<br />

Vietnam.<br />

ACKNOWLEDGEMENTS<br />

The authors wish to thank ACIAR for sponsoring this research.<br />

REFERENCES<br />

Truong NV, Burgess LW, Liew ECY, 2008. Prevalence and aetiology of<br />

Phytophthora foot rot of black pepper in Vietnam. <strong>Australasian</strong><br />

<strong>Plant</strong> <strong>Pathology</strong> 37, 431–442.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 187


Posters<br />

80 Characterisation of Phytophthora capsici Isolates from chilli in Vietnam<br />

N.V. Truong{ XE "Truong, N.V." } A , L.W. Burgess B and E.C.Y. Liew C<br />

A School of Agriculture and Forestry, Hue University, 102 Phung Hung Street, Hue city, Vietnam<br />

B Faculty of Agriculture, Food and Natural Resources, The University of Sydney, NSW 2006, Australia<br />

C Royal Botanic Gardens Sydney, Botanic Gardens Trust, Mrs Macquaries Road, Sydney, NSW 2000, Australia<br />

INTRODUCTION<br />

Phytophthora capsici, a pathogen of a range of tropical crops,<br />

such as black pepper, betel, cacao and rubber, is not managed<br />

effectively in developing countries due to lack of knowledge on<br />

the interactions between the pathogen and its many hosts<br />

(Drenth and Guest 2004). Although diseases of crops caused by<br />

P. capsici in temperate regions have been investigated in great<br />

detail, a lot less is known about this species in the South East<br />

Asian region. P. capsici is the causal agent of black pepper foot<br />

rot. With the need for integrated disease management in mind,<br />

questions arise as to the cross‐infectivity of the pathogen from<br />

different hosts and whether chilli could be an alternative host<br />

harbouring the black pepper pathogen. This study described the<br />

testing of three hypotheses. The first was that P. capsici isolates<br />

from chilli were pathogenic on black pepper and conversely<br />

black pepper isolates on chilli. The second was that chilli isolates<br />

were genetically different from black pepper isolates. The third<br />

was that there was genetic diversity among P. capsici isolates<br />

from chilli.<br />

MATERIALS AND METHODS<br />

Phytophthora capsici isolates. Twenty‐two isolates of P. capsici<br />

from chilli pepper (Capsicum frutescens L.) recovered from soil<br />

and diseased plants in Da Lat and Quang Nam provinces were<br />

used in this study. Twenty‐four isolates of P. capsici from black<br />

pepper were also used to compare with isolates from chilli<br />

pepper in the genetic analysis. In addition, an isolate each<br />

representing Phytophthora species P. palmivora and P.<br />

nicotianae was also included as species comparison in the RAMS<br />

and REP‐PCR analysis.<br />

Pathogenicity. Six leaves of black pepper or chilli were placed in<br />

a metal tray in 3 rows on layers of moist tissue. Leaves of black<br />

pepper or chilli were pricked with a sterile needle and inoculated<br />

with 5 µL of zoospore suspension. Uninoculated samples were<br />

similarly inoculated with sterilised deionised water.<br />

DNA extraction, RAMS and REP‐PCR protocol and genetic<br />

analysis. Fungal mycelium was grown in liquid medium,<br />

incubated at 25℃ in the dark for 5–6 days and DNA was<br />

extracted. The RAMS and REP‐PCR fingerprinting protocols were<br />

performed as previously described by Hantula et al. (1996) and<br />

Rademaker & Bruijn (1997). Cluster analysis was performed<br />

using the DICE similarity coefficient and UPGMA agglomeration<br />

in the program NTSYSpc.<br />

RESULTS<br />

Pathogenicity test. Chilli leaves inoculated with isolates from<br />

black pepper developed ‘water‐soaked lesions’, whereas the<br />

isolates from chilli tested on black pepper leaves showed lesions<br />

with fimbriate margins.<br />

RAM and REP analysis. The data from RAMS and REP‐PCR were<br />

combined to construct an UPGMA dendrogram (Figure. 1). A<br />

total of 7 phenotypes were detected, of which 4 were unique<br />

and 3 represented clonal groups. The first clonal group<br />

comprised 4 chilli isolates from Dalat. The second contained 18<br />

chilli isolates from Quang Nam. The third was 20 isolates from<br />

black pepper. The four unique isolates were from black pepper.<br />

The isolates from black pepper were separated from the chilli<br />

isolates at a similarity level of


53 The inhibitory effect of sumac stem extract on some fungal plant pathogens<br />

N. Panjehkeh{ XE "Panjehkeh, N." } A , M. Abdolmaleki A , M. Salari A , S. Bahraminejad B<br />

A University of Zabol, Sistan and Baluchestan, Zabol, Iran<br />

B University of Razi, Kermanshah, Iran<br />

Posters<br />

INTRODUCTION<br />

<strong>Plant</strong>s produce more than 10000 secondary metabolites with<br />

low molecular weights, many of which prevent plant infections<br />

to diseases and pests (2). Methanolic extract of sumac rich in<br />

hydrolysable tannins and proanthocyanidins (5) has<br />

demonstrated high inhibitory activity to Bacillus, Staphylococcus<br />

aureus and Enterobacter phlei (6).<br />

MATERIALS AND METHODS<br />

<strong>Plant</strong> material. Sumac (Rhus coriaria) stems at the final stage of<br />

growth were collected from their natural habitat in Hamedan,<br />

Iran, in September 2007. The stems were dried at ambient<br />

temperature in shade at a laboratory. They were ground using a<br />

mill, and passed through a 1‐mm mesh metal sieve.<br />

Extraction method. Soluble compounds were extracted in<br />

methanol (1). Five grams of powdered stem was placed into a<br />

bottle containing 100 ml absolute methanol. The bottles were<br />

capped and shaken in a rotary shaker at 350 rpm at 20 ° C for 24<br />

hours. Then, 75 ml of the solution from each bottle was removed<br />

and poured separately into another bottle. To each bottle was<br />

added 25 ml of sterilised distilled water and 100 ml hexane. The<br />

bottles were capped and shaken at 350 rpm for 2 hours.<br />

Subsequently, the aqueous phase was separated. The<br />

methanolic extracts were placed into a hood to evaporate the<br />

methanol.<br />

Pathogens. Phytophthora drechsleri and Rhizoctonia solani were<br />

isolated from beet roots, and Fusarium oxysporum and Bipolaris<br />

sorokiniana were isolated from chickpea and wheat roots,<br />

respectively. The pathogenicity of the pathogens confirmed on<br />

their hosts.<br />

Poison food technique. Based on the method of Hagerman and<br />

Butler (4), the inhibitory effect of 500, 1000, 1500, and 2000<br />

ppm from the extract was examined using a completely<br />

randomised design with four replicates. The zero level (just<br />

solvent) was used as control. The calculated extract for each<br />

concentration required to poison 100 ml PDA was dissolved in<br />

1.5 ml of methanol, and added to the autoclaved PDA when<br />

cooled down to 40 ° C. The poisoned PDA was dispensed in 25 ml<br />

aliquots into 9 cm Petri plates and allowed to cool. The plates<br />

were inoculated with 6 mm diameter discs from cultures of the<br />

fungi. The inoculated plates were incubated at 25 ° C. Colony<br />

diameters were measured frequently until the fungal growth in<br />

the control plates completely covered the plates. The<br />

experiment was repeated twice.<br />

Calculation of inhibitory percentage. The inhibitory effect of<br />

different extract concentrations was calculated using the<br />

following formula (3):<br />

IP = 100(C‐T)/C<br />

RESULTS AND DISCUSSION<br />

The inhibitory effects of extracts increased as the extract<br />

concentration increased. The effectiveness of the extract against<br />

R. solani and B. sorokiniana was more than against F. oxysporum<br />

and P. drechsleri (Table 1). Complete inhibition of growth of R.<br />

solani took place at 1500 ppm, while 2000 ppm of extract had<br />

81% inhibitory effect on B. sorokiniana. It is likely that by<br />

increasing extract concentration against this and the other two<br />

fungal species, the inhibitory effect could be increased to 100%.<br />

Table 1. Growth inhibition (%) of four fungi by different concentrations<br />

(ppm) of methanolic extract from sumac.<br />

Extract<br />

conc F. oxysporm R. solani P. drechsleri B. sorokiniana<br />

500<br />

1000<br />

1500<br />

2000<br />

28.12<br />

48.04<br />

51.01<br />

59.08<br />

84.46<br />

89.64<br />

100<br />

100<br />

1.08<br />

1.90<br />

8.15<br />

59.51<br />

68.56<br />

72.15<br />

74.26<br />

81.22<br />

The crude methanolic extract and isolated constituents of<br />

another member of this genus, Rhus glabra, was effective<br />

against 11 gram negative and positive bacteria (7) and nine<br />

pathogenic fungi (6). The research result was in conformity with<br />

previous findings regarding antifungal activity of sumac extract.<br />

In conclusion, the crude methanolic extract of sumac can be<br />

used as an antifungal agent against the examined fungi,<br />

particularly R. solani and B. sorokiniana.<br />

REFERNCES<br />

1. Bahraminejad S, Asenntorfer RE, Riley IT, Zwer P, Schultz CJ,<br />

Schmidt O 2006 Genetic variation of flavonoid defence compound<br />

concentration in Oat (Avena sativa L.) entries and testing of their<br />

biological activity. Proceedings of the 13th <strong>Australasian</strong> <strong>Plant</strong><br />

Breeding Conference pp. 1127–1132.<br />

2. Cowan MM 1999 <strong>Plant</strong> products as antimicrobial agents. Clinical<br />

Microbioligy Reviews 12, 564–582.<br />

3. Hadian JS, Fakhr Tabatabaei M, Salehi P, Hajieghrari B, Ghorban<br />

Pour M 2006 A phytochemical study of Cymbopogon parkeri<br />

essential oil, and its biological activity against some<br />

phytopathogenic fungi. Iranian Journal of Agricultural Sciences 37,<br />

425–431.<br />

4. Hagerman AE, Butler LG 1989 Choosing appropriate methods and<br />

standards for assaying tannin. Journal of Chemical Ecology 15,<br />

1795–1810.<br />

5. Kosar K, Bozan B, Temelli F, Baser KHC 2007 Antioxidant activity<br />

and phenolic composition of sumac (Rhus coriaria L.) extracts. Food<br />

Chemistry 103, 952–959.<br />

6. McCutcheon AR, Ellis, SM, Hancock, REW, Towers, GHN 1994<br />

Antifungal screening of medicinal plants of British Columbian native<br />

peoples. Journal of Ethnopharmacology 44, 157–169<br />

7. Saxena G, McCutcheon AR, Farmer S, Towers GHN, Hancock REW<br />

1994 Antimicrobial constituents of Rhus glabra. Journal of<br />

Ethnopharmacology 42, 95–99.<br />

where IP is inhibitory percentage; C is the colony diameter of the<br />

control; and T is the colony diameter of the treatment.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 189


Posters<br />

33 Aerial photography—a tool to monitor Mallee onion stunt<br />

S.J. Pederick{ XE "Pederick, S.J." }, J.W. Heap, T.J. Wicks, S. Anstis and G.E. Walker<br />

South Australian Research and Development Institute, GPO Box 397, Adelaide, 5001 South Australia<br />

INTRODUCTION<br />

Mallee onion stunt (MOS) is a widespread disease that was<br />

identified in the Murray Mallee of South Australia in 2005 (1). It<br />

is associated with specific strains of Rhizoctonia solani which<br />

cause stunted patches of economic significance. Rhizoctonia<br />

bare patch (AG 8) is a common disease of both cereals and<br />

onions. Cereal crops are often rotated with onions. Monitoring<br />

this disease over large areas is time consuming and complex. In<br />

this research aerial imagery was evaluated as a tool for MOS<br />

mapping and monitoring.<br />

MATERIALS AND METHODS<br />

High resolution aerial imagery was captured using two digital<br />

cameras mounted on an unmanned aerial vehicle (UAV).<br />

Normalised Difference Vegetation Index (NDVI = Red‐<br />

NIR/Red+NIR) was used to estimate relative biomass. Colour<br />

images of complete onion pivots were constructed by mosaicing<br />

multiple images. Images and maps were geo‐referenced using<br />

DGPS ground control points. Low NDVI areas were identified as<br />

putative disease patches, and paired soil samples were collected<br />

from within the disease patches and in adjacent normal areas.<br />

DNA was extracted from soil and assayed for Rhizoctonia using<br />

PCR (2).<br />

RESULTS AND DISCUSSION<br />

MOS patches are shown in an example colour image of an onion<br />

pivot (Fig. 1), and an NDVI (relative biomass) map from the same<br />

pivot is shown in Fig. 2. An example of a mosaiced complete<br />

pivot image is shown in Fig. 3. Levels of Rhizoctonia DNA (AG 8)<br />

in soil from stunted patches were significantly higher (396.4 pg<br />

DNA/g soil) than those from adjacent normal patches (26.1).<br />

Aerial imagery using a UAV is a rapid, cheap and effective tool<br />

for monitoring and mapping onion diseases. Raw images were<br />

viewed on a field computer within 15 minutes of<br />

commencement of the UAV flight. These were used to identify<br />

general areas of interest for immediate disease scouting, and<br />

geo‐referenced NDVI maps allowed sampling points to be<br />

located within 1m on the ground. The imagery was particularly<br />

useful and interesting for growers, who were often able to<br />

contribute valuable interpretation of biomass patterns. Further<br />

research to collect more data from sampling points identified<br />

from biomass maps is planned.<br />

Figure 2. NDVI (relative biomass) map. MOS patches (low biomass) are<br />

light tone patches.<br />

Figure 3. Complete mosaiced onion pivot image.<br />

ACKNOWLEDGEMENTS<br />

This project was facilitated by Horticulture Australia Limited in<br />

partnership with AUSVEG and was funded by the Onion Levy.<br />

The Australian Government provides matched funding for all<br />

HAL Research and Development activities, and onion growers in<br />

the Mallee region are thanked for their support. Thank you to<br />

the SARDI Diagnostic laboratory for PCR assays, and thank you to<br />

Angela Lush for laboratory support.<br />

REFERENCES<br />

1. Pederick SJ, Wicks TJ, Walker GE, Hall BH, Walter A (2007) Studies<br />

on the cause of Mallee Onion Stunt in South Australia. In:<br />

‘Proceedings of the 16th Biennial <strong>Australasian</strong> <strong>Plant</strong> <strong>Pathology</strong><br />

<strong>Society</strong> Conference’ South Australia, page 59.<br />

2. Ophel Keller K et al. (2006) DNA monitoring tools for soilborne<br />

diseases of potato. In: ‘Proceedings of the 4th <strong>Australasian</strong><br />

soilborne diseases symposium’, New Zealand, page 66.<br />

Figure 1. Colour image showing MOS bare patches.<br />

190 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


81 Survey of viruses infecting sweet potato crops in New Zealand<br />

Z.C. Perez‐Egusquiza{ XE "Perez‐Egusquiza, Z.C." } 3 , L.I. Ward 1 , J.D. Fletcher 2 and G.R.G. Clover 1<br />

1 <strong>Plant</strong> Health and Environment Lab, MAF Biosecurity New Zealand<br />

2 <strong>Plant</strong> and Food Research, Christchurch<br />

Posters<br />

INTRODUCTION<br />

Sweet potato or kumara (Ipomoea batatas) is important<br />

culturally and as a food source in New Zealand. The annual<br />

production of about 20,000 tonnes is grown mainly in the<br />

districts of Auckland, Bay of Plenty and Kaipara in the North<br />

Island. In January 2008, virus symptoms that included chlorotic<br />

spots, ring spots, vein clearing and mottling were observed on<br />

the leaves of commercial sweet potato crops (mainly cvs.<br />

Beauregard, Owairaka Red and Toka Toka Gold) growing in the<br />

three main production areas. A survey was done to determine<br />

the extent of virus infection affecting these crops.<br />

MATERIALS AND METHODS<br />

Samples. Fifty to 100 leaves were collected randomly from each<br />

of 26 different fields, five in Pukekohe (Auckland), nine in<br />

Gisborne (Bay of Plenty) and twelve in Dargaville (Kaipara).<br />

Leaves from each field were bulked into groups of 10, giving a<br />

total of 173 composite samples.<br />

Real‐time PCR. Total nucleic acid was extracted from all 173<br />

composite samples, and used in real‐time PCR assays specific for<br />

Sweet potato leaf curl virus (SPLCV) and real‐time reverse<br />

transcription (RT)‐PCR assays specific for Sweet potato chlorotic<br />

stunt virus (SPCSV), Sweet potato feathery mottle virus (SPFMV),<br />

Sweet potato virus G (SPVG), and Sweet potato virus 2 (SPV2;<br />

synonym Sweet potato virus Y).<br />

Graft inoculation. Representative plants from each field were<br />

grafted onto 3–4 week old Ipomoea setosa plants. Symptoms<br />

were monitored for 3–5 weeks and leaves were collected for<br />

testing.<br />

NCM‐ELISA. Nitrocellulose membrane enzyme‐linked<br />

immunosorbent assays (International Potato Center‐CIP, Lima,<br />

Peru) were done on the original sweet potato samples and<br />

grafted I. setosa. The following viruses were tested for:<br />

Cucumber mosaic virus (CMV), C‐6 virus, Sweet potato caulimolike<br />

virus (SPCaLV), Sweet potato chlorotic fleck virus (SPCFV),<br />

SPCSV, Sweet potato latent virus (SwPLV) and Sweet potato mild<br />

specking virus (SPMSV).<br />

RESULTS AND DISCUSION<br />

Real‐time RT‐PCR detected the potyviruses SPVG, SPFMV and<br />

SPV2 in many of the samples. Table 1 presents results showing<br />

these three viruses are common in the three areas surveyed.<br />

Figure 1 shows symptoms in two different cultivars infected with<br />

the three viruses. SPFMV and SPVG have been reported in New<br />

Zealand (1, 2) but SPV2 had not been reported previously (3).<br />

Samples infected with SPV2, were found as single infections, in<br />

co‐infection with SPVG and SPFMV, or with SPVG but not SPFMV,<br />

but no samples were co‐infected with SPV2 and SPFMV when<br />

SPVG was absent. No samples were infected with SPLCV.<br />

None of the original kumara samples were positive for CMV, C‐6<br />

virus, SPCSV, SPCaLV, SPCFV, SwPLV or SPMSV by NCM‐ELISA. No<br />

samples tested positive for SPCSV by real‐time PCR or NCM‐<br />

ELISA.<br />

Symptoms on grafted I. setosa were typical of potyvirus infection<br />

and no additional viruses were detected when they were tested<br />

by NCM‐ELISA.<br />

Table 1. Viruses in sweet potato in New Zealand.<br />

No. of fields<br />

tested<br />

No. of infected fields<br />

(No. of infected samples)<br />

(No. of<br />

samples<br />

tested) SPV2 SPFMV SPVG<br />

Dargaville 12 (103) 6 (35) 12 (70) 11(82)<br />

Pukekohe 5 (40) 4 (17) 2 (12) 5 (27)<br />

Gisborne 9 (30) 7 (18) 9 (30) 9 (30)<br />

Total 26 (173) 17(70) 23(112) 25(139)<br />

A<br />

Figure 1. Two sweet potato cultivars showing A) chlorotic spots and B)<br />

ring spots. Both samples were infected with SPVG, SPFMV and SPV2.<br />

Single potyvirus infections cause mild or no symptoms in sweet<br />

potato, and consequently yield is not significantly reduced.<br />

However, co‐infection with other viruses such as SPCSV<br />

produces a synergistic effect and more severe disease symptoms<br />

(4).<br />

SPCaLV, SPCFV, SwPLV have been reported in New Zealand<br />

previously (1) but they were not detected during this survey.<br />

ACKNOWLEDGMENT<br />

We would like to thank the Kumara growers who supported this<br />

work by allowing us to sample their crops, and without whose<br />

co‐operation this survey would not have been possible.<br />

REFERENCES<br />

1. Pearson MN, Clover GRG, Guy PL, Fletcher JD, Beever RE (2006) A<br />

review of the plant virus, viroid and mollicute records for New<br />

Zealand. <strong>Australasian</strong> <strong>Plant</strong> <strong>Pathology</strong> 35: 217–252.<br />

2. Rännäli M, Czekaj V (2008) Molecular genetic characterization of<br />

Sweet potato virus G (SPVG) isolates from areas of the Pacific<br />

Ocean and Southern Africa. <strong>Plant</strong> Disease 92: 1313–1320.<br />

3. Perez‐Egusquiza Z, Ward LI, Fletcher JD, Clover GRG (2009)<br />

Detection of Sweet potato virus 2 in sweet potato in New Zealand.<br />

<strong>Plant</strong> Disease 93: 427.<br />

4. Untiveros M, Fuentes S, Salazar LF (2007) Synergistic interaction of<br />

Sweet potato chlorotic stunt virus (Crinivirus) with carla‐, cucumo‐,<br />

ipomo‐, and potyviruses infecting sweet potato <strong>Plant</strong> Disease 91:<br />

669–676.<br />

B<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 191


Posters<br />

54 Evaluation of commercial cultivars for control of white blister rust in Brassica rapa<br />

and Brassica oleracea vegetables<br />

J.E. Petkowski{ XE "Petkowski, J.E." } A , F. Thomson B , E.J.Minchinton A and C. Akem C<br />

A Biosciences Research Division, Department of Primary Industries, Private Bag 15, Ferntree Gully DC 3156, Victoria, Australia<br />

B Biometrics Unit, Department of Primary Industries, Private Bag 15, Ferntree Gully DC 3156, Victoria, Australia<br />

C Ayr Research Station, Department of Primary Industries and Fisheries, PO Box 5, Ayr 4807, Queensland, Australia<br />

INTRODUCTION<br />

White blister rust of Brassicaceae is caused by an oomycete<br />

Albugo candida. The disease affects many economically<br />

important crops including Brassica oleracea and B. rapa<br />

vegetables in Australia and around the world (Petkowski 2008).<br />

White or creamy pustules filled with zoosporangia on leaves,<br />

stems and heads as well as hypertrophy and hyperplasia of plant<br />

organs downgrade the aesthetics and marketability of the<br />

produce. In vegetable production, white blister rust is commonly<br />

controlled with fungicides. <strong>Plant</strong>ing resistant cultivars, however,<br />

is the most cost‐effective and desirable disease control method<br />

(Li et al 2007). A number of commercial cultivars of B. oleracea<br />

and B. rapa are sold as “tolerant” to A. candida pathotypes<br />

occurring on these species in Australia.<br />

We report on the evaluation of commercial cvs of B. oleracea<br />

and B. rapa in glasshouse screening trials.<br />

MATERIALS AND METHODS<br />

In two glasshouse experiments, seedlings of 10 cvs of B. rapa<br />

and 12 cvs of B. oleracea were tested for resistance to A.<br />

candida collections from Chinese cabbage and broccoli,<br />

respectively. Each experiment included hosts with reported<br />

susceptibility or resistance as controls. In each experiment,<br />

seeds were sown in seed‐raising mix in plastic multipot trays of<br />

144 pots per tray. Seedlings were irrigated twice a day for 1<br />

minute by overhead sprinklers. <strong>Plant</strong>s in both experiments were<br />

fertilised weekly with Aquasol.<br />

Inocula were prepared by suspending zoosporangia in sterile<br />

distilled water at the concentration of 1x10 5 zoosporangia per<br />

mL. Prior to inoculation, the suspensions were incubated for four<br />

hours at 13 ºC to induce zoospore release. Inocula were applied<br />

twice in each of the experiments using a trigger atomiser on<br />

seedlings previously misted with sterile distilled water. The first<br />

inoculation was at the fully developed cotyledon stage and the<br />

second at the first true leaf growth stage. Seedlings were<br />

covered with plastic sheets after each inoculation for 24 hours to<br />

ensure sufficient leaf wetness for infection. White blister<br />

incidence and severity on hosts was assessed on 4 week‐old and<br />

on 5 week‐old seedlings in experiments with B. rapa and B.<br />

oleracea, respectively. Incidence was expressed as a percentage<br />

of the seedlings with symptoms. Disease severity was rated on<br />

seedlings with sori using a 0–4 scale. Experiments were designed<br />

as randomised complete blocks of treatments (12 hosts) with 8<br />

replications. Data were analysed using ANOVA.<br />

RESULTS AND DISCUSSION<br />

All varieties of B. rapa vegetables tested were either very of<br />

moderately susceptible to white blister (Table 1). Miyako and<br />

Walz were significantly more susceptible than other cultivars<br />

tested. Pak choi cv Seven Gates had smaller blisters surrounded<br />

by discoloured rings of leaf tissues, indicating some level of<br />

resistance to the A. candida collection tested.<br />

glasshouse conditions. Means followed by different letters are<br />

significantly different at the 5% level.<br />

B. rapa<br />

Cultivar<br />

n=96<br />

Mean<br />

Incidence<br />

(%)<br />

Mean<br />

Severity<br />

Scale<br />

(0–4)<br />

B. oleracea<br />

Cultivar<br />

n=96<br />

Mean<br />

Incidence<br />

(%)<br />

Mean<br />

Severity<br />

Scale<br />

(0–4)<br />

Miyako 91.3 a 2.2 a Greenbelt* 79.8 a 1.6 a<br />

Walz 80.1 a 1.3 b Forte* 76.6 a 1.7 a<br />

Mei Qoing 72.9 b 1.3 b Highfield 53.6 b 1.1 b<br />

Cv 001 70.6 b 1.1 b Brittany 44.5 bc 0.9 b<br />

Matilda 66.9 b 1.0 b Millenium 35.7 cd 0.6 bc<br />

Seven Gates 60.5 b 0.9 b Summer Love 35.6 cd 0.8 bc<br />

Manoko 60.4 b 0.8 b Sting 25.3 d 0.3 c<br />

Reward* 58.0 b 1.0 b Cyrus 4.2 e 0.1 c<br />

Torch* 55.3 b 1.4 b Avalanche 0.0 0.0<br />

Joi Choi 56.3 b 0.7 b Abacus 0.0 0.0<br />

Kailaan* 5.6 c 0.0 c Romulus 0.0 0.0<br />

Regent* 0.0 c 0.0 c NIZ17‐1091 0.0 0.0<br />

*) Cultivars incorporated either as positive or negative controls. Cultivars Regent (B.<br />

napus) and Kailaan (B. oleracea) are negative controls.<br />

Brussels sprouts cultivars Abacus and Romulus had no white<br />

blister, cv Cyrus had a very low incidence (4.2%) of seedlings<br />

with blisters and cv Millenium was susceptible (36% of seedling<br />

affected). All but one cauliflower cultivar, Avalanche, were<br />

susceptible (Table 1). Cabbage cv Sting, which is susceptible to<br />

European A, candida, was moderately susceptible to the<br />

Australian A. candida. The moderately susceptible cv NIZ17‐<br />

1091, however, was resistant to the collection tested. White<br />

blister rust was the most severe on cauliflower cv Forte followed<br />

by broccoli Greenbelt. White blister rust control can be<br />

improved by planting less susceptible cvs combined with<br />

fungicide sprays timed by using the Brassica Spot disease risk<br />

predictive model (Petkowski 2008).<br />

AKNOWLEDGEMENTS<br />

The authors thank seed companies for supplying seeds and HAL,<br />

AusVeg, the State Government of Victoria and the Federal<br />

Government for financial support.<br />

REFERENCES<br />

1. Li CX, Sivasithamparam K, Walton G, Salisbury P, Burton W, Banga<br />

SS, Banga S, Chattopadhyay C, Kumar A, Singh R, Singh D, Agnohotri<br />

A, Liu YC, Fu TD, Wang YF and Barbetti MJ (2007) Expression and<br />

relationships of resistance to white rust (Albugo candida) at<br />

cotyledonary, seedling, and flowering stages in Brassica juncea<br />

germplasm from Australia, China, and India. Australian Journal of<br />

Agricultural Research, 58: 259–264.<br />

2. Petkowski JE (2008) Biology and control of white blister rust of<br />

Brassicaceae vegetables. PhD Thesis, Deakin University, Australia.<br />

Table 1. Incidence and severity of white bister rust on seedlings of<br />

commercial cultivars of B. rapa and B. oleracea vegetables inoculated<br />

with A. candida collections from Chinese cabbage and broccoli in<br />

192 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


34 Diatrypaceae species associated with grapevines and other hosts in New<br />

South Wales<br />

Posters<br />

W.M. Pitt{ XE "Pitt, W.M." }, R. Huang, C.C. Steel and S. Savocchia<br />

National Wine and Grape Industry Centre, School of Agricultural and Wine Sciences, Charles Sturt University, Locked Bag 588, Wagga<br />

Wagga, NSW 2678, Australia<br />

INTRODUCTION<br />

The fungus Eutypa lata is responsible for the canker disease of<br />

grapevines known as Eutypa dieback. Entry of the fungus via<br />

pruning wounds and the formation of cankers in the vascular<br />

tissue of grapevines results in a slow decline and dieback that<br />

reduces vigour, yield and vineyard productivity (1). Symptoms<br />

may include stunted shoots with shortened internodes, and a<br />

characteristic wedge shaped lesion in the trunks and canes of<br />

infected vines. In addition to E. lata, a number of other species<br />

belonging to the Diatrypaceae can be found on grapevines and<br />

other hosts in and around vineyards. To date, the role of many of<br />

the diatrypaceous species in grapevine decline is unknown. The<br />

incidence and distribution of species in the Diatrypaceae was<br />

documented during a recent survey of vineyards throughout<br />

New South Wales (NSW).<br />

MATERIALS AND METHODS<br />

Surveys were conducted from 73 vineyards throughout NSW<br />

between November 2006 and April 2008 to study fungi<br />

associated with grapevine trunk diseases. Wood samples were<br />

taken from 1789 grapevines displaying foliar symptoms typical of<br />

Eutypa dieback or evidence of dead spurs, cankers, or bleached<br />

and discoloured tissue. Isolations of fungi were made by directly<br />

plating out pieces of diseased tissue onto PDA after surface<br />

sterilisation in 1% sodium hypochlorite. Samples were incubated<br />

at 25ºC and monitored for the appearance of fungi.<br />

Diatrypaceous species were tentatively identified based on gross<br />

cultural morphology. In December 2008, additional surveys and<br />

collections were conducted from grapevines and other hosts,<br />

both in the Hunter Valley and Tumbarumba. Isolations of<br />

diatrypaceous fungi from these samples were made directly<br />

from ascospores, with species tentatively identified based on<br />

morphology of the teleomorph and confirmed via sequencing<br />

and analysis of ribosomal DNA regions.<br />

RESULTS AND DISCUSSION<br />

Morphological and molecular analyses of fungi isolated from<br />

diseased wood and cultured from fruiting bodies revealed the<br />

presence of five species belonging to the Diatrypaceae. In<br />

addition to E. lata, Eutypa leptoplaca, Cryptovalsa ampelina and<br />

species of both Diatrypella and Eutypella were isolated. In total,<br />

79 isolates were collected, 12 of E. lata, 23 of C. ampelina, 1<br />

isolate of E. leptoplaca, 19 of Eutypella and 24 of Diatrypella<br />

(Table 1), although isolates from the latter genera could not be<br />

identified to species.<br />

These surveys have shown that Eutypa dieback is more<br />

widespread than first thought and may be increasing in<br />

prominence, especially in cooler climates where lower<br />

temperatures and high rainfall favour the growth of E. lata (2).<br />

Additionally, several other species within the Diatrypaceae are<br />

present not only on grapevines, but on a number of other hosts<br />

in and around vineyards. Many of these species are widely<br />

distributed in NSW and have been shown to have wider<br />

geographic ranges and occur in greater numbers than E. lata.<br />

species that are commonly found both on grapevines and other<br />

hosts. Additional studies are required to determine the<br />

pathogenicity of the diatrypaceous species on grapevines. The<br />

impact of these species and their role in grapevine decline is<br />

under investigation.<br />

Table 1. Incidence and distribution of Diatrypaceous fungi associated<br />

with grapevines and other hosts in New South Wales.<br />

Region Host Fungi<br />

Big Rivers Vitis vinifera Diatrypella<br />

Eutypa lata<br />

Cryptovalsa ampelina<br />

Central Ranges Vitis vinifera Eutypella<br />

Diatrypella<br />

E. lata<br />

C. ampelina<br />

Southern NSW Vitis vinifera Eutypella<br />

Diatrypella<br />

E. lata<br />

C. ampelina<br />

Populus nigra‐italica Eutypa leptoplaca<br />

Ulmus procera<br />

C. ampelina<br />

Northern Rivers Vitis vinifera C. ampelina<br />

Northern Slopes Vitis vinifera Eutypella<br />

Diatrypella<br />

Hunter Valley Vitis vinifera Eutypella<br />

C. ampelina<br />

Citrus paradise<br />

Eutypella<br />

Diatrypella<br />

Citrus sinensis<br />

Eutypella<br />

Ficus carica<br />

Eutypella<br />

Fraxinus excelsior Diatrypella<br />

ACKNOWLEDGEMENTS<br />

This work was supported by the Winegrowing Futures Program,<br />

a joint initiative of the Grape and Wine Research and<br />

Development Corporation and the National Wine and Grape<br />

Industry Centre. The authors wish to thank Florent Trouillas (UC<br />

Davis) and Mark Sosnowski (SARDI) for technical assistance.<br />

REFERENCES<br />

1. Sosnowski MR, Creaser ML, Wicks TJ, Lardner R, Scott E (2008)<br />

Protection of grapevines pruning wounds from infection by Eutypa<br />

lata. Australian Journal of Grape and Wine Research 14, 134–142.<br />

2. Pitt WM, Huang R, Savocchia S, Steel CC (2008) Increasing evidence<br />

that Eutypa dieback of grapevines is widespread in New South<br />

Wales. 6th International Workshop on Grapevine Trunk Diseases,<br />

Florence, Italy, 1–3 September.<br />

3. Mostert L, Halleen F, Creaser ML, Crous PW (2004) Cryptovalsa<br />

ampelina, a forgotten shoot and cane pathogen of grapevines.<br />

<strong>Australasian</strong> <strong>Plant</strong> <strong>Pathology</strong> 33, 295–299.<br />

While E. lata, and the lesser‐known C. ampelina are recognised<br />

pathogens of grapevines (3), little is known about the other<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 193


Posters<br />

13 Biological control of Uncinula necator by mycophagous mites<br />

N. Panjehkeh A , S. Ramroodi{ XE "Ramroodi, S." } A and A.R. Arjmandinezhad B<br />

A University of Zabol,, Zabol, Sistan and Baluchestan, Iran<br />

B Agricultural and Natural Resources Research Centre of Sistan<br />

INTRODUCTION<br />

A number of abiotic and biotic factors influence the incidence,<br />

severity, and spatial scale of diseases in natural and managed<br />

plant systems (2, 3). Although the function of such factors as<br />

temperature, humidity, and host plant resistance traits have<br />

been relatively well‐studied, the impact of natural enemies on<br />

interactions between plants and pathogens has received less<br />

attention. <strong>Plant</strong> pathologists have investigated various<br />

microorganisms as potential biological control agents of<br />

microbial plant pathogens (1, 8). The ability of arthropods to<br />

regulate plant pathogens is poorly understood. Leaf‐inhabiting<br />

mites that are thought to feed on fungi (mycophagous mites)<br />

and other microbes are extremely common on many woody<br />

perennial plant species (9, 10) and represent a potentially<br />

significant group of natural enemies of fungal plant pathogens<br />

that attack leaves and fruit. Erysiphe necator, the causal agent of<br />

grape powdery mildew, is a worldwide pathogen (6). The disease<br />

imposes considerable damages to grapes by reduction of the<br />

quality and quantity of the product in Sistan region. The<br />

powdery mildew agent has a great capacity to develop<br />

resistance to synthetic fungicides (4). Therefore, it seems that<br />

biological control is the best control approach of the disease.<br />

The objective of the project was to identify mycophagous mites<br />

that inhabit grape leaves infected and uninfected by Erysiphe<br />

necator.<br />

METHODS AND RESULTS<br />

Plenty of mites were collected from grape leaves infected and<br />

uninfected by Erysiphe necator. A faunistic survey was<br />

conducted in August 2007 to collect and identify mites living on<br />

the grapevine cultivar Yaghoti in Sistan region, Iran. Six species<br />

belonging to six families were collected and identified from<br />

infected leaves (Table 1). Two species were collected and<br />

identified from uninfected leaves (Table 1).<br />

population dynamics (8). English‐Loeb et al. (5) have suggested<br />

that Orthotydeus lambi (Acari: Tydeidae) could be useful as a<br />

biological control agent of powdery mildew on cultivated grapes<br />

under vineyard conditions. Erysiphe necator is obligately<br />

parasitic, and with the exception of absorptive haustoria within<br />

epidermal cells of the host, the body of the pathogen resides on<br />

the plant surface (6). This characteristic makes it vulnerable to<br />

grazing by mycophagous mites.<br />

REFERENCES<br />

1. Baker KF, Cook RJ 1974 Biological control of plant pathogens.<br />

Freeman, San Francisco.<br />

2. Burdon JJ 1987 Diseases and plant population biology. Cambridge<br />

University Press, New York.<br />

3. Burdon JJ 1993 The structure of pathogen populations in natural<br />

plant communities. Annual Review of Phytopathology 31, 305–324.<br />

4. Erickson EO, Wilcox WF 1997 Distributions of sensitivities to three<br />

sterol demethylation inhibitor fungicides among populations of<br />

Uncinula necator sensitive and resistant to triadimefon.<br />

Phytopathology 87, 784–791.<br />

5. English‐Loeb G, Norton AP, Gadoury DM, Seem RC, Wilcox WF 1999<br />

Control of powdery mildew in wild and cultivated grapes by tydeid<br />

mite. Biological Control 14, 97–103.<br />

6. Pearson RC, Gadoury DM 1991 Powdery mildew of grape. In `<strong>Plant</strong><br />

diseases of international importance: diseases of fruit crops`. (Eds J<br />

Kumar, HS Chaube, US Singh, AN Mukhopadhyay) pp. 129–146.<br />

(Prentice Hall, Englewood Cliffs, NJ.)<br />

7. Shaw PJA. 1992 Fungi, fungivores, and fungal food webs. In `The<br />

Fungal Community: its organisation and role in the ecosystem`.<br />

(Eds GC Carroll, DT Wicklow). pp. 295–310.(Dekker, New York).<br />

8. Sutton JC, Peng G 1993 Manipulation and vectoring of biocontrol<br />

organisms to manage foliage and fruit diseases in cropping<br />

systems. Annual Review of Phytopathology 31, 473–494.<br />

9. Walter DE 1996 Living on leaves: mites, tomenta, and leaf domatia.<br />

Annual Review of Entomology 41, 101–114.<br />

10. Walter DE, O’Dowd DJ 1995 Life on the forest phylloplane: hairs,<br />

little houses, and myriad mites. In `The forest canopy: aspects of<br />

research on this biological frontier`. (Eds ME Lowman, N Nadkarni)<br />

pp 325–352. (Academic Press, New York).<br />

Presence on leaves<br />

Mite species Mite diet Infected Uninfected<br />

Lasioseius mcgregori<br />

(Ascidae)<br />

Typhlodromus sp.<br />

(Phytoseiidae)<br />

Tyrophagus<br />

putrescenteae Schrank<br />

(Acaridae)<br />

Rhizoglyphus robini<br />

(Acaridae)<br />

Anystis baccarum<br />

(Anystidae)<br />

Tydeus sp. (Tydeidae)<br />

Predatory/<br />

mycophagous<br />

Predatory/<br />

mycophagous<br />

+<br />

+ +<br />

Mycophagous + +<br />

Mycophagous +<br />

Predatory/<br />

mycophagous<br />

Predatory/<br />

mycophagous<br />

+<br />

+<br />

DISCUSSION<br />

The experiment revealed that the population of mycophagous<br />

mites in grape leaves infected with E. necator was significantly<br />

higher than uninfected leaves. It is well recognised that some<br />

arthropods including mites use fungi as a food resource,<br />

although it is less understood how this influences fungal<br />

194 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


35 First report of a eucalypt yellowing disease in Syria and its similarity to<br />

Mundulla yellows<br />

Posters<br />

D. Hanold A , B. Kawas B and J.W. Randles{ XE "Randles, J.W." } A<br />

A University of Adelaide, Waite Campus, Glen Osmond, 5064, South Australia<br />

B ICARDA, PO Box 5466, Aleppo, Syria<br />

INTRODUCTION<br />

Eucalyptus camaldulensis is widely grown in Syria for land<br />

rehabilitation, amenity, and as roadside trees. Seed was<br />

imported repeatedly on a large scale from different locations in<br />

Australia up to the 1970s (I. Nahal, L. Makki pers comm). This has<br />

resulted in a national population of E. camaldulensis including<br />

both subspecies camaldulensis and obtusa, as well as local<br />

hybrids of the two. Around 500 000 eucalypt seedlings are being<br />

produced annually from Syrian seed by 11 government nurseries<br />

for distribution around the different regions.<br />

Syria has developed a large, isolated population of E.<br />

camaldulensis over the last century. Little insect or fungal<br />

damage has been observed suggesting that many pathogens<br />

affecting the species in Australia are absent. Nevertheless MYlike<br />

symptoms virtually identical to the syndrome in Australia are<br />

present. Further epidemiological and phenotypic, as well as<br />

molecular comparison of Syrian and Australian symptomatic E.<br />

camaldulensis is needed and may provide information on<br />

aetiology and possibly origin of the disease.<br />

During a visit to ICARDA by D. Hanold in October 2008, yellowing<br />

symptoms similar to the Mundulla Yellows (MY) syndrome (1)<br />

were noted. The MY‐like symptoms appeared to be widespread<br />

in some locations, and preliminary surveys of several sites were<br />

carried out to obtain further information about its distribution.<br />

While a witches' broom disease without associated yellowing<br />

symptoms had been described earlier from Syrian eucalypts (2),<br />

this is the first report of a eucalypt yellowing disease from that<br />

country.<br />

MATERIALS AND METHODS<br />

Based on the description of MY distribution in Australia (1), trees<br />

not adjacent to roads as well as roadside trees were<br />

characterised. Survey sites in the regions around Aleppo, Tel<br />

Hadya, Ein Dara, Idleb, Tabaqah, along the road to the coast<br />

west of Homs, and along the Damascus‐Aleppo highway were<br />

mapped (Fig. 1).<br />

Photographs were taken of trees at each site and additional<br />

information recorded as available. Individual trees on the<br />

ICARDA station (Tel Hadya) were also numbered for future<br />

reference and leaf samples were taken to Adelaide and stored<br />

frozen for molecular analysis. Symptomatic seedlings identified<br />

in a nursery were planted at ICARDA together with<br />

asymptomatic controls to observe disease progress.<br />

RESULTS<br />

Eucalypts with symptoms resembling the descriptors of MY<br />

(Table 1) and including interveinal chlorosis, epicormic shoots<br />

and asymmetric yellowing of the crown were observed in all<br />

areas except west of Homs where only asymptomatic eucalypts<br />

were observed. E. camaldulensis both in single and mixed<br />

species plantings on roadsides, in paddocks, soil rehabilitation<br />

areas, seed gardens and nurseries were affected. Symptoms<br />

occurred on single trees adjacent to asymptomatic ones, in<br />

clusters, and as disease gradients similar to the distribution of<br />

MY in Australia (1).<br />

DISCUSSION<br />

MY is defined by characteristic descriptors distinguishing it from<br />

yellowing disorders due to environmental factors. Its cause is<br />

unknown, but a biotic, contagious agent has been implicated (1).<br />

Yellowing diseases of eucalypts closely resembling the Australian<br />

MY have also been reported from Spain and South America (1).<br />

This is the first report of a similar syndrome from Syria and the<br />

Middle East.<br />

Figure 1. Map of Syria.<br />

Table 1. Comparison of the Syrian eucalypt yellowing disease (SEY) with<br />

descriptors of MY (1).<br />

Descriptors SEY MY<br />

Leaf symptoms:<br />

interveinal chlorosis<br />

distortion of leaf margins<br />

necrotic pin spots<br />

no dead leaves on twigs<br />

Epicormic shoots + +<br />

Twig dieback + +<br />

Disease stages:<br />

early (yellow patches in foliage)<br />

medium (yellow epicormic shoots)<br />

late (overall tree dieback)<br />

Affected trees next to healthy ones<br />

Trees of all ages affected<br />

No recovery observed<br />

Not cured by pruning<br />

Symptoms in paddocks and roadsides<br />

ACKNOWLEDGEMENTS<br />

We gratefully acknowledge the support of I. Nahal and B. Bayaa,<br />

Aleppo University; L. Makki, Syrian Department of Agriculture; A.<br />

El‐Ahmed, R. Brettell, S. Kumari and the staff and management<br />

of ICARDA.<br />

REFERENCES<br />

1. Hanold D, Gowanlock D, Stukely MJC, Habili N, Randles JW (2006)<br />

Mundulla Yellows disease of eucalypts: descriptors and preliminary<br />

studies on distribution and etiology. <strong>Australasian</strong> <strong>Plant</strong> <strong>Pathology</strong><br />

35, 199–215.<br />

2. Bos LM, Makkouk KM, Bayaa B (1990) Witches' broom and decline<br />

of Eucalyptus, a serious disease in Syria, likely caused by<br />

mycoplasma. Arab Journal of <strong>Plant</strong> Protection 8, 135–133.<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 195


Posters<br />

36 New host records for ‘Candidatus Phytoplasma aurantifolia’ in Australia<br />

J.D. Ray{ XE "Ray, J.D." }<br />

Australian Quarantine and Inspection Service, Marrara, NT, 0812<br />

INTRODUCTION<br />

Phytoplasmas are unculturable wall‐less prokaryotes within the<br />

class Mollicutes that infect plant phloem vessels and are<br />

transmitted by phloem feeding insects (2). They are implicated in<br />

diseases of a wide range of plant species. Symptoms of<br />

phytoplasma infection include reduced leaf size ‘little leaf’,<br />

yellowing of leaves, proliferation of stems leaves and flowers<br />

(witches’ broom) and floral abnormalities (1, 3).<br />

<strong>Plant</strong> health surveys were carried out by the Australian<br />

Quarantine and Inspection Service (AQIS) in northern Western<br />

Australia (northern WA) and the Northern Territory (NT) from<br />

2006 to 2008. This paper reports on the new host records of<br />

phytoplasma collected during these surveys.<br />

MATERIALS AND METHODS<br />

<strong>Plant</strong>s exhibiting symptoms of phytoplasma infection (little leaf<br />

or witches’ broom) during plant health surveys in northern WA<br />

and the NT from 2006 to 2008 were collected for diagnostic<br />

confirmation. <strong>Plant</strong> parts with symptomatic new growth were<br />

collected into plastic bags and kept cool. Within 2 days of<br />

collection leaf petioles and midribs were excised from the<br />

symptomatic material, dehydrated over anhydrous CaCl 2 and<br />

stored at 4°C where possible.<br />

Samples were forwarded to Bioscience North Australia, Charles<br />

Darwin University for molecular analysis. Sample analysis<br />

included polymerase chain reaction (PCR) assays using universal<br />

phytoplasma specific primers. Strain analysis was performed by<br />

restriction fragment length polymorphism (JR153, KN11) or by<br />

sequencing (JR428, JR429).<br />

RESULTS AND DISCUSSION<br />

‘Candidatus Phytoplasma aurantifolia’ is reported for the first<br />

time associated with disease of Ipomea aquatica, Jatropha<br />

gossypifolia, Ocimum basilicum and Canavalia rosea (Table 1).<br />

These new records were all detected in WA whilst no new<br />

records were detected in the NT during this time. These<br />

phytoplasma records represent a diverse host range.<br />

Table 1. Details of new host records for ‘Ca. .P. aurantifolia’ in Australia.<br />

Species, Common name Location Strain Coll. No.<br />

Ipomea aquatica,<br />

Kangkung<br />

Jatropha gossypifolia,<br />

Bellyache bush<br />

Ocimum basilicum,<br />

Lemon basil<br />

Canavalia rosea,<br />

Beach bean<br />

Broome, WA TBB JR153<br />

Cable Beach, WA SPLL‐V4 JR428<br />

Wattle Downs, WA TBB JR429<br />

Lombadina Beach,<br />

WA<br />

SPLL‐V4<br />

KN11<br />

ACKNOWLEDGEMENTS<br />

Thanks to Karen Gibb, Anna Padovan and Claire Streten of<br />

Bioscience North Australia, Charles Darwin University for<br />

diagnostics. Andrew Mitchell and Kirsty Neaylon (AQIS) are<br />

gratefully acknowledged for some collections, and A. Mitchell for<br />

identifying plant specimens.<br />

REFERENCES<br />

1. Davis RI, Jacobson SC, De La Rue SJ, Tran‐Nguyen L, Gunua TG,<br />

Rahamma S (2003) Phytoplasma disease surveys in the extreme<br />

north of Queensland, Australia, and the island of New Guinea.<br />

<strong>Australasian</strong> <strong>Plant</strong> <strong>Pathology</strong>. 32; 269–277.<br />

2. IRPCM Phytoplasma/ Spiroplasma Working Team—Phytoplasma<br />

taxonomy group (2004) ‘Candidatus Phytoplasma’, a taxon for the<br />

wall‐less, non‐helical prokaryotes that colonize plant phloem and<br />

insects. International Journal of Systematic and Evolutionary<br />

Microbiology. 54: 1243‐ 1255.<br />

3. Schneider B, Padovan A, De La Rue S, Eichner R, Davis R, Bernuetz<br />

A, Gibb K (1999) Detection and differentiation of phytoplasmas in<br />

Australia: an update. Australian Journal of Agricultural Research.<br />

50; 33–42.<br />

‘Ca. .P. aurantifolia’ has a wide host range across a diverse group<br />

of plant families and is widespread throughout Australia. The<br />

suggested <strong>Australasian</strong>/Asian origin of this phytoplasma may in<br />

part explain its success in harsh environments and wide host<br />

range (1, 3).<br />

All new records were obtained in northern WA which has a<br />

particularly harsh climate. During surveillance it was observed<br />

that some species of annual plants infected with phytoplasma do<br />

not produce seed and remain green for longer than the same uninfected<br />

hosts (Ray, personal observation). Perhaps the<br />

occurrence of a diverse host range for ‘Ca. .P. aurantifolia’ is a<br />

survival mechanism for both the phytoplasma and the leaf<br />

hopper vectors by providing live host material during periods<br />

when other hosts have died, the creation of a green bridge.<br />

196 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


37 A comparative study of methods for screening chickpea and wheat for resistance<br />

to root‐lesion nematode Pratylenchus thornei<br />

Posters<br />

R.A. Reen{ XE "Reen, R.A." }, J.P. Thompson<br />

DEEDI, Qld Primary Industries and Fisheries, Leslie Research Centre, GPO Box 2282, Toowoomba, 4350, Queensland<br />

© State of Queensland, Department of Employment, Economic Development and Innovation, 2009<br />

INTRODUCTION<br />

Several quantitative methods are available for testing resistance<br />

of crops to root‐lesion nematode (Pratylenchus thornei) under<br />

controlled environments. This study aimed to compare growth<br />

times for wheat and chickpea cultivars and the nematode<br />

extraction procedure of Whitehead tray, (Whitehead and<br />

Hemming 1965) with shake‐elution (Moore et al 1992) and<br />

misting (Hooper 1986) methods. The objective was to optimise<br />

differences between susceptible and resistant chickpea lines for<br />

plant breeding purposes.<br />

P. thornei / kg soil<br />

25000<br />

(a)<br />

20000<br />

15000<br />

10000<br />

5000<br />

0<br />

10 12 14 16 18 20 22<br />

W 7 days<br />

W 4 days<br />

M 7 days<br />

M 4 days<br />

MATERIAL AND METHODS<br />

Five chickpea and two wheat lines covering a range of resistance<br />

to P. thornei were selected for testing. The design consisted of 3<br />

replicates in randomised blocks with 6 harvest times and 3<br />

extraction methods. For each time x variety x replicate<br />

combination there were 2 pots to enable nematode comparison<br />

from one half of each pot using standard Whitehead tray<br />

method, while roots were extracted from the other half for<br />

either shake‐elution or misting. Single plants were grown in pots<br />

of 330 g of steam–sterilised vertosol maintained between 22–<br />

25°C in a glasshouse on a 2 cm tension bottom–watering system.<br />

A 15 ml suspension to provide 10,000 P. thornei/kg soil was<br />

pipetted around the seed at planting. At each harvest of 12, 14,<br />

16, 18 and 20 weeks, soil with roots from pots was sectioned<br />

longitudinally into halves for nematode extractions. For all 3<br />

extraction procedures room temperatures were in the range of<br />

22–26°C and nematodes were collected on a 20 µm sieve.<br />

Whitehead and shake‐elution extractions were assessed at 1, 2,<br />

3, 4 and 7 days while misting extractions were assessed at 4 and<br />

7 days. Nematodes were counted in a 1‐mL Hawksley slide and<br />

expressed as number of P. thornei/kg soil (oven‐dry equivalent)<br />

or P. thornei/g root (fresh weight). A multi–factorial data analysis<br />

was performed using ln(x+c) where x = nematodes/kg soil and c<br />

= constant.<br />

RESULTS AND DISCUSSION.<br />

The Whitehead tray method extracted significantly (P < 0.001)<br />

more P. thornei than either misting or shake‐elution (Fig. 1 a and<br />

b). Growing chickpeas for a longer period (18–20 weeks) than<br />

wheat (16–18 weeks) gave maximum discrimination of<br />

resistance/susceptibility in cultivars (Fig. 2). The extraction<br />

efficiency for 2 days was 70% of that at 7 days when using<br />

Whitehead trays. All the above results showed similar<br />

differences between treatments whether expressed as P.<br />

thornei/kg soil or as P. thornei/g root. The Whitehead trays were<br />

found to be less labour intensive than misting and shake‐elution<br />

procedures, and more practical for assessing large numbers of<br />

plants for resistance.<br />

P. thornei / kg soil<br />

18000<br />

15000<br />

12000<br />

9000<br />

6000<br />

3000<br />

(b)<br />

0<br />

10 12 14 16 18 20 22<br />

<strong>Plant</strong> growth weeks<br />

W 2 days<br />

W 4 days<br />

W 7 days<br />

S-e 2 days<br />

S-e 4 days<br />

S-e 7 days<br />

Figure 1. For each harvest the total number of P. thornei extracted from<br />

wheat and chickpea with Whitehead trays (W) was significantly higher (P<br />

< 0.001) than misting (M; Fig. 1a) and shake‐elution (S‐e; Fig. 1b).<br />

P. thornei / kg soil<br />

13<br />

12<br />

11<br />

10<br />

9<br />

8<br />

7<br />

6<br />

10 12 14 16 18 20 22<br />

<strong>Plant</strong> growth weeks<br />

Batavia (W)<br />

Sona (CP)<br />

Desavic (CP)<br />

QT8343 (W)<br />

Norwin (CP)<br />

Amethyst (CP)<br />

ILWC 246 (WC)<br />

Figure 2. Longer growth periods allowed better discrimination among<br />

chickpea cultivars. Value of l.s.d. at 18 weeks and 20 weeks (P = 0.05) =<br />

0.82 and 1.32 respectively. W=wheat, CP=chickpea, WC=wild chickpea.<br />

ACKNOWLEDGEMENTS<br />

Kerry Bell for statistical analysis and Indooroopilly Research<br />

laboratories for use of their misting facilities.<br />

REFERENCES<br />

Whitehead AG, Hemming JR (1965) A comparison of some quantitative<br />

methods of extracting small vermiform nematodes from soil.<br />

Annals of Applied Biology 55,25–38.<br />

Hooper DJ, (1986) Extraction of nematodes from plant m aterial. In<br />

‘Laboratory methods for work with <strong>Plant</strong> and Soil Nematodes’. (Ed<br />

JF Southey) pp. 51–81. (Her Majesty’s Stationary Office:London)<br />

Moore KJ, Southwell RJ, Schwinghamer MW, Murison RD (1992) A rapid<br />

shake‐elution procedure for quantifiying root lesion nematodes<br />

(Pratylenchus thornei) in chickpea and wheat. <strong>Australasian</strong> <strong>Plant</strong><br />

<strong>Pathology</strong> 21, 70–78.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 197


Posters<br />

82 Multi‐locus sequence typing of isolates of Pseudomonas syringae pv. actinidiae, a<br />

biosecurity risk pathogen<br />

J. Rees‐George{ XE "Rees‐George, J." }, I.P.S. Pushparajah and K.R. Everett<br />

The New Zealand Institute for <strong>Plant</strong> and Food Research Limited, PB 92169, Mt Albert, Auckland, New Zealand<br />

INTRODUCTION<br />

Pseudomonas syringae pv. actinidiae (P.s.a.) is a bacterium that<br />

was first recorded in Japan (1), caus‐ing a trunk canker on<br />

kiwifruit (Actinidia deliciosa ‘Hayward’). This disease has not<br />

been recorded on kiwifruit in New Zealand. PCR primers were<br />

designed to detect this pathogen to prevent importation into<br />

New Zealand on germplasm. To prove specificity of these<br />

primers isolates from different geographic locations were tested.<br />

MATERIALS AND METHODS<br />

Cultures. Cultures for DNA extraction were sourced from the<br />

International Collection of Micro‐organisms from <strong>Plant</strong>s (ICMP),<br />

New Zealand; Korean Agricul‐tural Culture Collection (KACC),<br />

Republic of Korea; National Institute of Agrobiological Sciences<br />

(NIAS), Japan,; National Collection of <strong>Plant</strong> Patho‐genic Bacteria<br />

(NCPPB), UK; and the Culture Collection of <strong>Plant</strong> Pathogenic<br />

Bacteria (PD), The Netherlands.<br />

PCR detection. Primers were designed and tested against 20<br />

strains of P.s.a. from international culture collections. Six strains<br />

were not detected by these primers (Table 1), and multi‐locus<br />

sequence typing (MLST) was conducted to study these strains in<br />

more detail. Sequence was included of three bacteria found on<br />

kiwifruit orchards (2), Pseudomonas sp., P. fluorescens (P.f.) and<br />

P. syringae (P.s.) and four representatives from Group 1 (3), P.s.<br />

maculicola, P.s. tomato, P.s. theae, and P.s.a. P. s. is in Group 2,<br />

and P. f. is an outgroup.<br />

Multi‐locus sequence typing. Five housekeeping genes were<br />

sequenced: those encoding sigma factor 70, aconitate hydratase<br />

B, citrate synthase, phosph‐orglucoisomerase, and gyrase. The<br />

16S–23S rDNA intergenic spacer (IGS) region was also<br />

sequenced.<br />

Phylogenetic analysis. Analyses were performed on individual<br />

gene sequences as well as on the concatenated data set using<br />

maximum parsimony.<br />

RESULTS<br />

By BLAST analysis, the sequence of the IGS region of KAAC 10660<br />

was most similar to Rahnella aquatilis, an unrelated saprotroph.<br />

Isolates KAAC 10582, ISPAVE‐B‐020, ISPAVE‐B‐019, PD2766 and<br />

PD2774 were most similar to P. syringae, and the type strain Kw‐<br />

11 to P.s.a. DNA from KAAC 10660 was not amplified by any<br />

MLST primers. All MLST gene phylogenies yielded similar results:<br />

PD2766 was most similar to P. f., KAAC 10582 to Pseudomonas<br />

sp., ISPAVE‐B‐020, ISPAVE‐B‐019 and PD2774 to P.s. syringae.<br />

None of the isolates not detected by PCR primers F1/R2 and<br />

F3/R4 was closely associated with the type strain Kw‐11, or other<br />

Group 1 pathovars.<br />

Table 1. Isolates of Pseudomonas syringae pv. actinidiae detected by<br />

primer sets F1/R2 and F3/R4<br />

Isolate name Collection Origin Detected<br />

Kw‐11 ICMP Japan +<br />

Kw‐1 Japan +<br />

Kw‐30 Japan +<br />

Kw‐41 Japan +<br />

KAAC 10582 KAAC Korea ‐<br />

KAAC 10584 Korea +<br />

KAAC 10594 Korea +<br />

KAAC 10659 Japan +<br />

KAAC 10660 Korea ‐<br />

KAAC 10754 Korea +<br />

FTRS L1 NIAS Japan +<br />

Sar1 Japan +<br />

Sar2 Japan +<br />

Kiw4 Japan +<br />

Wa1 Japan +<br />

Wa2 Japan +<br />

ISPAVE‐B‐020 NCPPB Italy ‐<br />

ISPAVE‐B‐019 Italy ‐<br />

PD 2766 PD USA ‐<br />

PD 2774 USA ‐<br />

ACKNOWLEDGEMENTS<br />

This work was funded by FfRST contract CO2X0501. DNA was<br />

provided by L. Liefting, and advice by R. Newcomb.<br />

REFERENCES<br />

1. Takikawa Y, Serizawa S, Ichikawa T, Tsuyumu S, and Goto M (1989).<br />

Pseudomonas syringae pv. actinidiae pv. nov.: the causal bacterium<br />

of canker of kiwifruit in Japan. Ann. Phytopath. Soc. Japan 55, 437–<br />

444.<br />

2. Everett KR and Henshall WR (1994). Population ecology and<br />

epidemiology of kiwifruit blossom blight. <strong>Plant</strong> <strong>Pathology</strong> 43,<br />

824–830.<br />

3. Sarkar SF and Guttman DS (2004) Evolution of the core<br />

genome of Pseudomonas syringae, a highly clonal, endemic<br />

plant pathogen. App. Env. Microbiology 70, 1999–2012.<br />

DISCUSSION<br />

These results indicate that isolate KAAC 10660 is an unrelated<br />

saprotroph that has been misidentified. Because the other five<br />

sequenced atypical isolates were not genetically related to<br />

Group 1 pathovars of P. syringae, these are also<br />

misidentifications.<br />

198 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


95 DNA barcoding to support biosecurity decisions<br />

K. Pan, G.F. Bills, M.K. Romberg{ XE "Romberg, M.K." }, W.H. Ho, and B.J.R. Alexander<br />

<strong>Plant</strong> Health and Environment Laboratory, MAF Biosecurity New Zealand, PO Box 2095, Auckland 1140, New Zealand<br />

Posters<br />

INTRODUCTION<br />

Precise pathogen identifications in the biosecurity context must<br />

be accurate and timely, as identification delays can affect trade,<br />

as well as response efforts to invasive exotic species. The need<br />

to rapidly identify cryptic fungi by traditional morphological<br />

techniques presents a distinct challenge to plant diagnostic<br />

laboratories. The use of DNA barcoding techniques to identify<br />

morphologically cryptic species addresses this challenge and has<br />

gained momentum recently, with the establishment of the<br />

International Consortium for the Barcode of Life.<br />

The MAFBNZ <strong>Plant</strong> Health and Environment Laboratory (PHEL) is<br />

currently developing a DNA barcoding platform to resolve<br />

diagnostic cases for which other methods are unable to provide<br />

timely and accurate results. An example of the usefulness of this<br />

tool is presented here.<br />

However, the barcoding approach could challenge or conflict<br />

with phytosanitary regulation. For instance some isolates were<br />

found consistently mis‐identified at species and higher<br />

taxonomic levels, and these organisms could be new to the<br />

country.<br />

DNA barcoding will likely be an international standard for plant<br />

pathogen identification, with the caveat that this powerful tool<br />

relies upon sequences generated from well‐characterised<br />

voucher specimens (preferably including the type species) in<br />

order for DNA barcode results to be well supported.<br />

The genus Diaporthe includes plant pathogens, plant endophytes<br />

and species associated with dying and dead vegetation that are<br />

usually observed as their Phomopsis anamorphs. These fungi are<br />

associated with disease symptoms on a wide range of species,<br />

including many economically important plants. Our objective<br />

was to develop capability to utilise barcoding to precisely and<br />

quickly identify fungal species, in order to assist in the<br />

determination of the regulatory status of intercepted organisms,<br />

and to allow more informed biosecurity decisions.<br />

METHODS<br />

Isolates from surveillance of fungi on plants in New Zealand and<br />

all strains deposited as either Diaporthe or Phomopsis from the<br />

International Collection of Microorganisms (ICMP) were<br />

included. A portion of the mycelia for all samples was stored in<br />

20% glycerol at minus 80°C and DNA for sequencing was<br />

extracted from the remaining mycelia. Sequences of the ITS<br />

region were generated using primer pair ITS1/ITS4 (1). The ITS<br />

sequences of authoritatively identified Diaporthe and Phomopsis<br />

strains were selected from recent revisionary studies. Sequences<br />

were stored and analysed with the software program Geneious<br />

(Biomatters Ltd, Auckland, New Zealand)<br />

RESULTS<br />

To date, the Diaporthe/Phomopsis barcode database consists of<br />

22 sequences from PHEL, 68 New Zealand strains from ICMP<br />

culture collection, and 92 from reference strains (Fig. 1).<br />

Continued survey work and addition of strains from other<br />

collections will expand and contribute new branches to the tree.<br />

The barcode database has proven to be a powerful tool to<br />

rapidly and accurately identify unknown strains. For instance,<br />

strains from surveillance (PHEL09‐2009‐2132) and the culture<br />

collection (ICMP2141) were identified to species level (Fig. 1),<br />

providing an initial identification in the first case, and<br />

clarification in the second.<br />

Isolates without a species name in this project (e.g. Groups A‐E)<br />

will be confirmed by further taxonomic work.<br />

Figure 1. Phylogenetic tree of Diaporthe and Phomopsis derived from<br />

Neighbour‐joining analysis. Expanded data set of P. viticola clade is<br />

presented in more detail.<br />

REFERENCE<br />

1. White, T.J., Bruns, T., Lee, S. and Taylor, J. 1990. Amplification and<br />

direct sequencing of fungal ribosomal RNA genes for phylogenetics.<br />

In: PCR Protocols: A Guide to Methods and Applications.<br />

DISCUSSION<br />

Development of a DNA barcoding platform can lead to resolution<br />

of many questions, including rapid determination if certain<br />

species, including undescribed taxa, are present in a country.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 199


Posters<br />

14 In vitro screening of potential antagonists of Xanthomonas translucens infecting<br />

pistachio<br />

A. Salowi{ XE "Salowi, A." }, D. Giblot Ducray and E.S. Scott<br />

School of Agriculture, Food and Wine, University of Adelaide, PMB1, Glen Osmond, 5064, South Australia<br />

INTRODUCTION<br />

X. translucens is the causal agent of dieback disease of pistachio.<br />

The bacterium infects the vascular tissues of the trees, causing<br />

discolouration of the xylem, lesions on the trunk and major<br />

limbs, decline and, in some cases, death (1). Although hygiene<br />

and application of quaternary ammonium disinfectant to pruning<br />

wounds have been recommended to limit the spread of the<br />

disease (2), effective control methods are lacking. Biological<br />

control offers potential in managing this disease. The aim of this<br />

research is to assess the ability of bacteria to antagonise X.<br />

translucens in vitro, and thus explore the potential of biological<br />

control in managing pistachio dieback.<br />

MATERIALS AND METHODS<br />

Pathogen. Isolate KI of X. translucens (Xtp_KI), obtained from a<br />

commercial orchard in Kyalite, NSW was used (3). The isolate<br />

was resuscitated from storage in sucrose peptone broth (SPB)<br />

and glycerol at ‐80°C.<br />

Potential antagonists. Eight isolates of bacteria obtained from<br />

pistachio wood were tested, along with one isolate of Bacillus<br />

subtilis. Seven of the pistachio isolates were courtesy of E. Facelli<br />

and C. Taylor.<br />

Although most of the bacterial isolates inhibited Xtp_KI in dual<br />

culture, there was no evidence of antibiotic activity in culture<br />

filtrates. B. subtilis was expected to produce diffusible<br />

antibiotics, but most reports concern antifungal activity.<br />

Methods for assessing production of diffusible antibiotics are<br />

being refined. Selected isolates are being tested for ability to<br />

colonise pistachio wood and to reduce colonisation of the wood<br />

by Xtp_KI.<br />

REFERENCES<br />

1. Facelli E, Taylor C, Scott E, Emmett B, Fegan M, Sedgley M (2002)<br />

Bacterial dieback of pistachio in Australia. <strong>Australasian</strong> <strong>Plant</strong><br />

<strong>Pathology</strong> 31, 95–96.<br />

2. Sedgley M, Scott E, Facelli E, Anderson N, Emmett B, Taylor C<br />

(2006) Pistachio canker epidemiology. Final Report to Horticulture<br />

Australia Ltd (Project PS04015).<br />

3. Facelli E, Taylor C, Scott E, Fegan M, Huys G, Noble R D, Swings J,<br />

Sedgley M (2005) Identification of the causal agent of pistachio<br />

dieback in Australia. European Journal of <strong>Plant</strong> <strong>Pathology</strong> 112, 155–<br />

165.<br />

4. Parente E, Brienza C, Moles M, Ricciardi A (1994) A comparison of<br />

methods for the measurement of bacteriocin activity. Journal of<br />

Microbiological Methods 22, 95–108.<br />

5. Mari M, Guizzardi M, Pratella G C (1996) Biological control of gray<br />

mold in pears by antagonistic bacteria. Biological Control 7, 30–37.<br />

Preliminary screening. A protocol to assess the antagonistic<br />

ability of the above bacteria was modified from Parente et al.<br />

(4). One hundred µl of a suspension of 10 6 CFU/ml of Xtp_KI was<br />

spread on either sucrose peptone agar (SPA) or nutrient agar<br />

(NA). After drying, a 6‐mm diameter well was punched into the<br />

agar, aseptically, and filled with 20 µl of 3 day‐old culture of the<br />

chosen antagonist. At this stage, the concentration of the<br />

antagonist suspension was not determined. Each antagonist<br />

treatment was replicated five times and sterile distilled water<br />

was used for controls. Antagonism was assessed by measuring<br />

the inhibition zone around the wells 48 hours after treatment.<br />

Antibiotic activity of the antagonists. Production of diffusible<br />

antibiotics inhibitory to Xtp_KI was investigated using a<br />

procedure modified from Mari et al. (5). The bacterial isolates<br />

were cultured in SPB, at 28°C on a rotary shaker. After 48 hours,<br />

the suspensions were centrifuged at 10,000g for 20 minutes and<br />

the supernatants filtered through 0.45 µm membrane filters.<br />

Following the protocol described above, 20 µl of culture filtrate<br />

was placed into each well. Antibiotic activity was assessed by<br />

measuring inhibition zones around wells 48 hours after<br />

treatment.<br />

RESULTS AND DISCUSSION<br />

One isolate (64161‐7) efficiently inhibited the growth of Xtp_KI,<br />

both on SPA and NA, with clear inhibition zones of >10mm<br />

radius. Four isolates (SUPP, PC397, PC506 and PC507) caused<br />

moderate inhibition of growth of Xtp_KI on SPA and NA<br />

(inhibition zones


83 Uniform distribution of powdery mildew conidia using an improved spore settling<br />

tower<br />

Posters<br />

Z. Sapak{ XE "Sapak, Z." } 1 , V. Galea 1 , D. Joyce 1 and E. Minchinton 2<br />

1 School of Land, Crop and Food Sciences, The University of Queensland, Gatton, 4343, QLD<br />

2 Department of Primary Industries, Biosciences Research Division, Knoxfield Centre, Knoxfield Victoria<br />

INTRODUCTION<br />

Powdery mildew of cucurbits (Podosphaera fusca (Fr.) S. Blumer)<br />

is an obligate parasite, unable to be cultured and maintained on<br />

agar media. Thereby, it provides a challenge to researchers<br />

studying the infection processes in the laboratory. More<br />

problematic is that the pathogen is sensitive to free water.<br />

Water damages conidia by negatively affecting viability and<br />

infectivity (Sivapalan, 1993). Thus, conidia cannot be applied to<br />

plants using water suspensions. Use of dry conidia applied by<br />

dusting or blowing from infected leaves onto test plants is<br />

commonly used in artificially inoculating plants. This approach<br />

allows for deposition of conidia onto plant surfaces, but with<br />

little uniformity of distribution (Reifschneider and Boiteux,<br />

1988). Techniques such as the use of paint brushes and cotton<br />

swabs provide improved uniformity of conidial distribution, but<br />

are lacking in convenience and accuracy. A spore‐settling tower<br />

applying Stoke’s law of sedimentation as proposed by<br />

Reifschneider and Boiteux (1988) has provided a more<br />

convenient and repeatable method for inoculation of powdery<br />

mildews. These authors constructed a plywood tower and<br />

demonstrated successful use of a low vacuum induced air inrush<br />

to dislodge conidia from infected leaves, and to disperse them<br />

effectively over test plants. This design was improved by the<br />

authors to enhance useability for inoculation of powdery<br />

mildews onto test plants.<br />

MATERIALS AND METHODS<br />

The spore settling tower was constructed from a flanged<br />

Perspex cylinder (100 cm in height, 50 cm diameter, 20 mm<br />

thick). The complete tower comprised of the cylinder, a top<br />

cover, and a base plate. The top cover incorporates a vacuum<br />

line and air valve, a vacuum gauge, a small removable lid with a<br />

25mm inlet valve leading to an internal inoculum platform<br />

suspended below. The vacuum line and air valve is used to<br />

connect the tower to a vacuum pump, and the internal vacuum<br />

applied is measured by the attached gauge. The inlet valve in the<br />

removable lid is used to sharply break the vacuum inside the<br />

tower dislodging conidia from the inoculum source (freshly<br />

harvested infected leaves). The internal inoculum platform (12<br />

cm diameter) was constructed under the top cover with three<br />

holes, each with a 9.0 cm diameter in its side walls (20 cm deep)<br />

to allow inoculum to be expelled and shower the test plants<br />

placed on the base plate at the bottom of the tower. The whole<br />

construction was designed and fabricated to maintain an<br />

internal vacuum by the use of rubber gaskets and sealants.<br />

The distribution of conidia dispersed through the operation of<br />

the spore settling tower was examined using water agar (2%)<br />

trap plates. The base of the trap plates were marked with a 1.0<br />

cm grid using a marker pen. Three open trap plates were used as<br />

replicates per inoculation run, and arranged on the base plate of<br />

the spore settling tower. Powdery mildew (P. fusca) maintained<br />

on cucurbit plants in the glasshouse, was used as the inoculum<br />

source. Leaves were cut into disks 2.0 cm 2 in diameter. In each<br />

inoculation run, 20 leaf disks (~ 2.5 g f.w.) were placed in an<br />

open Petri plate and covered with a layer of tape‐fastened open<br />

plastic mesh (1 cm 2 pore size). The inoculum was placed on the<br />

inoculum platform, the water agar trap plates placed on the<br />

base plate and the unit sealed. A vacuum of 20 kPa (taking ~ 10<br />

sec) was applied followed by closing of the air valve. Sudden<br />

opening of the inlet valve in the removable lid resulted in a sharp<br />

break in the vacuum causing a sudden inrush air onto the<br />

inoculum source and dislodgment of conidia. The trap plates at<br />

the base of the unit plate were then exposed to the resulting<br />

shower of conidia. The unit was left undisturbed for a minimum<br />

of 120 seconds after breaking of vacuum to optimise conidial<br />

distribution (Reifschneider and Boiteux, 1988). After each<br />

inoculation run, the media in the water agar trap plates was cut<br />

into 1 cm squares and fifteen randomly selected sections placed<br />

on glass microscope slides, stained with lactophenol cotton blue<br />

and observed under a light microscope to evaluate the<br />

distribution of conidia by counting. The data were analysed using<br />

a statistical analysis system (SAS).<br />

RESULTS AND DISCUSSION<br />

The spore settling tower developed in this study is relatively<br />

easy to operate for inoculation of plants with powdery mildew<br />

pathogens. The transparent Perspex construction allows users to<br />

monitor the process inside the tower. Provision of a vacuum<br />

gauge allows for more consistent conditions to be achieved,<br />

resulting in higher repeatability among inoculation runs. The<br />

incorporation of an air inlet valve (not present in the unit<br />

designed by Reifschneider and Boiteux, 1988) allows for easy<br />

breaking of vacuum, while modifications to the inoculum<br />

platform further improved useability.<br />

The distribution of conidia observed in this study was uniform<br />

with no significant difference between the numbers observed<br />

within and between trap plates P


Posters<br />

84 The effect of high nutrient loads on disease severity due to Phytophthora<br />

cinnamomi in urban bushland<br />

Kelly Scarlett{ XE "Scarlett, K." }, Zoe‐Joy Newby, David Guest and Rosalie Daniel<br />

Faculty of Agriculture, Food and Natural Resources, University of Sydney, 2006 NSW<br />

INTRODUCTION<br />

The soilborne pathogen Phytophthora cinnamomi has become<br />

infamous for its destruction of native Australian vegetation<br />

communities, particularly in Western Australia and Victoria.<br />

More recently, the pathogen has been reported to occur widely<br />

around NSW. Phytophthora has been indicted with dieback of<br />

iconic species including Angophora costata, Eucalyptus<br />

botryoides, E. piperita and Corymbia gummifera around Sydney<br />

Harbour. However, bushland reserves in which these trees occur<br />

are also often subject to high nutrient loads from urban runoff.<br />

Environmental abiotic factors can play a major role in either<br />

enhancing or suppressing the disease severity on a susceptible<br />

host. Seedlings of E. maculata and E. pilularis grown with lower<br />

amounts of nitrogen and phosphorus express more severe<br />

disease symptoms (1). Conversely, disease severity increased<br />

with increasing levels of inorganic, but not organic, nitrogen<br />

application in durian and papaya inoculated with P. palmivora<br />

(2).<br />

This study investigates the relationship between soil nutrient<br />

loads and the severity of disease due to Phytophthora in four<br />

tree species found in bushland and parks around Sydney<br />

Harbour.<br />

MATERIALS AND METHODS<br />

Site selection. Four 20 x 20 m quadrats were set up in the Lawry<br />

Plunkett Reserve in Mosman, NSW. Sites were identified based<br />

on drainage runoff and the level of disturbance (invasive weeds)<br />

(Table 1). Ten soil samples were taken from each of the four<br />

sites for analysis. Samples were collected from the root zone of<br />

A. costata, E. botryoides, E. piperita and C. gummifera.<br />

Table 1. Sites in the Lawry Plunkett Reserve from which soil was sampled<br />

for analysis of nutrient levels, microbial activity and Phytophthora.<br />

Site<br />

Description<br />

1 Highly disturbed, Phytophthora isolated<br />

2 Highly disturbed, Phytophthora isolated<br />

3 Moderately disturbed, Phytophthora isolated<br />

4 Moderately disturbed, Phytophthora isolated<br />

Soil analysis. All 40 soil samples were analysed for the presence<br />

of Phytophthora by lupin baiting (3). The concentration of<br />

nitrogen (N), phosphorus (P) and other minerals were<br />

determined for three of the 10 samples taken from each site.<br />

Microbial activity was measured for all soil samples using the<br />

FDA assay (4).<br />

Glasshouse trials. The effect of N on the severity of disease due<br />

to P. cinnamomi is being assessed in a glasshouse trial. A.<br />

costata, E. piperita, E. botryoides, C. gummifera and Pinus<br />

radiata (susceptible control) seedlings were inoculated with P.<br />

cinnamomi are treated with 0 mg/ml and 100 mg/mL<br />

ammonium nitrate every two weeks. <strong>Plant</strong> health is being<br />

assessed over time, and root and shoot weight will be measured<br />

at the end of the experiment.<br />

RESULTS AND DISCUSSION<br />

The occurrence of Phytophthora. P. cinnamomi was isolated<br />

from all four sites sampled. A second species of Phytophthora<br />

was isolated from Sites 1 and 2 and is being identified.<br />

Nutrient analysis. The nutrient levels varied significantly within<br />

and between sites. Site 2 had higher levels of total soil N (due to<br />

higher levels of NH 4 + ) than the other three sites (Fig 1). A<br />

drainage channel may account for the high N load observed in<br />

soil sampled from Site 2. Soil P levels were not significantly<br />

different between sites. Magnesium levels were highest in soil<br />

samples from Site 4. Aluminium was present at extreme levels at<br />

Site 3. The effect of excess nitrogen on disease severity in the<br />

four tree species is currently being assessed in a glasshouse<br />

study.<br />

Concentration<br />

(mg/kg)<br />

100<br />

50<br />

0<br />

1 2 3 4<br />

Sit e<br />

N (mg/kg)<br />

P (mg/kg)<br />

Figure 1. Levels of soil nitrogen and phosphorus at the four sites in the<br />

Lawry Plunkett Reserve.<br />

Microbial activity. At 3.5 µg FDA g ‐1 min ‐1 , microbial activity was<br />

greatest in Site 4, although this was not statistically significant (p<br />

= 0.11). Site 3 had the lowest microbial activity at 2.2 µg FDA g ‐<br />

1 min ‐1 (Figure 2).<br />

Microbial Activity<br />

(ug FDA/g soil/min)<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

1 2 3 4<br />

Sit e<br />

Figure 2. Microbial activity at the four sites in the Lawry Plunkett Reserve<br />

as determined by the FDA assay.<br />

In urban areas, nutrients available in the soil can fluctuate due to<br />

run off. High amounts of run off resulting in soil nutrient loading<br />

may play a significant role in either the severity or suppression of<br />

dieback disease associated with P. cinnamomi. A greater<br />

understanding of the interaction between dieback and nutrient<br />

loads will enable effective management decisions for urban<br />

areas where nutrient loads are usually prominent and<br />

Phytophthora is present.<br />

ACKNOWLEDGEMENTS<br />

We gratefully acknowledge the support of Mosman Council in<br />

the supply of seedlings and the soil analysis in this study.<br />

REFERENCES<br />

1. Halsall et al. (1983) Australian Journal of Botany 31 341–355<br />

2. Chee & Newhook (1965) New Zealand Journal of Agricultural<br />

Research 8, 88–95.<br />

3. Schnurer & Rosswall (1982) Applied Environmental Microbiology<br />

43, 1256–1261.<br />

202 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


15 Characterisation of the causal agent of pistachio dieback as a new pathovar of<br />

Xanthomonas translucens, x. Translucens pv. pistaciae pv. nov.<br />

Posters<br />

D. Giblot Ducray A , A. Marefat A# , N.M. Parkinson B , J.P. Bowman C , K. Ophel‐Keller D , E.S. Scott{ XE "Scott, E.S." } A<br />

A School of Agriculture Food and Wine, The University of Adelaide, Waite Campus, Glen Osmond, 5064 SA<br />

B Central Science Laboratory, Sand Hutton, York YO41 1LZ, UK<br />

C University of Tasmania, School of Agricultural Science, Hobart, 7001 TAS<br />

D South Australian Research and Development Institute, Adelaide, 5000 SA<br />

# Current address: Department of <strong>Plant</strong> Protection, Zanjan University, Zanjan, Iran<br />

INTRODUCTION<br />

Pistachio dieback has limited the expansion of the pistachio<br />

industry in Australia over the past 15 years. It is caused by two<br />

genetically distinct groups of strains of X. translucens (1, 2).<br />

However, the pistachio pathogen is atypical because it has a<br />

dicotyledonous woody host, in contrast to typical X. translucens<br />

that cause disease in monocotyledonous hosts in the Poaceae.<br />

Also, the characterisation of integrons, which are known to have<br />

played a key role in genetic diversification of Xanthomonas,<br />

suggested that the pistachio pathogen represented a new<br />

pathovar of the species (3). Here, we report use of DNA‐DNA<br />

hybridisation and gyrB gene sequencing to further clarify the<br />

taxonomic position of the pistachio pathogen and establish its<br />

phylogeny among other Xanthomonas.<br />

MATERIALS AND METHODS<br />

DNA‐DNA hybridisation. DNA/DNA hybridisation analyses were<br />

conducted to determine the DNA similarity of the reference<br />

strains of the two groups of the pistachio pathogen (ICMP 16316<br />

and ICMP 16317) (2) to the pathotype strains of three X.<br />

translucens pathovars, namely X. translucens pv. translucens<br />

(LMG 876), X. translucens pv. poae (LMG 728) and X. translucens<br />

pv. graminis (LMG 726). The type strains of X. theicola (LMG<br />

8764) and X. hyacinthi (LMG 739) were included as outgroups.<br />

Each hybridisation was conducted two‐four times.<br />

gyrB phylogeny. DNA extraction, PCR and sequence analysis of<br />

the gyrB gene were performed as described by Parkinson et al.<br />

(4). Sequences were determined for the reference strains ICMP<br />

16316 and ICMP 16317 of the pistachio pathogen, and compared<br />

to that of other Xanthomonas available in the database (4).<br />

RESULTS<br />

DNA‐DNA hybridisation. DNA‐DNA hybridisation between the<br />

two strains of the pistachio pathogen was 84%. When compared<br />

to X. translucens pathovars, one group had the highest<br />

homology with X. translucens pv. poae, with 84% hybridisation,<br />

whereas the other group had the highest homology with X.<br />

translucens pv. graminis, with 90% hybridisation. Hybridisation<br />

values of both pistachio strains with X. theicola and X. hyacinthi<br />

averaged no more than 54%.<br />

DISCUSSION<br />

DNA.‐DNA hybridisation is considered to provide definitive<br />

species‐level identification: strains from the same species have<br />

above 70% homology, whereas strains from different species<br />

show homology values averaging 40–50%. Therefore, our results<br />

confirmed the classification of the pistachio pathogen as an X.<br />

translucens.<br />

The clustering of the pistachio pathogen among X. translucens<br />

pathovars in the gyrB phylogeny further confirms the DNA‐DNA<br />

hybridisation results and suggests that the pistachio pathogen<br />

has originated through host switching of one of the<br />

Xanthomonas ancestors that had a monocotyledonous host.<br />

These results, together with the distinct pathogenicity to<br />

pistachio and the consistent discrimination of the pistachio<br />

pathogen from other X. translucens (2, 3), support its<br />

designation as a new pathovar of the species, for which we<br />

propose the name Xanthomonas translucens pv. pistaciae pv.<br />

nov.<br />

REFERENCES<br />

1. Facelli E, Taylor C, Scott E, Fegan M, Huys G, Emmett R, Noble D,<br />

Swings J, Sedgley M (2005) Identification of the causal agent of<br />

pistachio dieback in Australia. European Journal of <strong>Plant</strong> <strong>Pathology</strong><br />

112, 155–165.<br />

2. Marefat A, Scott E, Ophel‐Keller K, Sedgley M (2006) Genetic,<br />

phenotypic and pathogenic diversity among Xanthomonads<br />

pathogenic on pistachio (Pistacia vera) in Australia. <strong>Plant</strong> <strong>Pathology</strong><br />

55, 639–49.<br />

3. Giblot Ducray D, Gillings MR, Marefat A, Ophel‐Keller K, Scott ES<br />

(2007) Characterisation of integrons in Xanthomonas translucens<br />

infecting pistachio in Australia. Proceedings of the 16th<br />

<strong>Australasian</strong> <strong>Plant</strong> <strong>Pathology</strong> <strong>Society</strong> Conference, Adelaide. P 37.<br />

4. Parkinson NM, Cowie C, Heeney J, Stead D (2009) Phylogenetic<br />

structure of Xanthomonas determined by comparison of gyrB<br />

sequences. International Journal of Systematic and Evolutionary<br />

Microbiology 59, 264–274.<br />

gyrB phylogeny. In sequence alignments, the two groups of the<br />

pistachio pathogen showed 96% homology. When compared<br />

with other Xanthomonas, the highest similarity was to pathovars<br />

of X. translucens, with percentages ranging from 95 to 99%,<br />

whereas the similarity to other Xanthomonas species and their<br />

pathovars ranged from 81 to 86%. The two exceptions were X.<br />

theicola and X. hyacinthi, which showed similarity values of 92<br />

and 93%, respectively, both with strains of the two groups. In<br />

the phylogenetic tree, both groups clustered among X.<br />

translucens pathovars, but as distinct lineages.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 203


Posters<br />

38 Interactions between Leptosphaeria maculans and fungi associated with canola<br />

stubble<br />

B. Naseri A,C , J.A. Davidson B and E.S. Scott{ XE "Scott, E.S." } A<br />

A School of Agriculture, Food and Wine, The University of Adelaide, PMB1, Glen Osmond, South Australia, 5064<br />

B South Australian Research and Development Institute, GPO Box 397, Adelaide, South Australia, 5001<br />

C Department of <strong>Plant</strong> Protection, Agricultural Research Institute, PO Box 45195‐1474, Zanjan, Iran<br />

INTRODUCTION<br />

Blackleg, caused by Leptosphaeria maculans, is of major<br />

economic importance in the canola‐growing areas of Australia.<br />

The pathogen survives on stubble and releases ascospores from<br />

pseudothecia.<br />

Antagonistic activity of fungi against L. maculans has been<br />

documented (1, 2). Antagonism, competition, stimulation of<br />

growth and decomposition of the stubble may affect survival<br />

and sporulation of L. maculans. A better understanding of the<br />

interactions between L. maculans and potentially antagonistic<br />

fungi on canola stubble could facilitate the development of<br />

strategies to reduce inoculum of the pathogen and contribute to<br />

the control of blackleg.<br />

MATERIALS AND METHODS<br />

In total, 35 species of fungi, including L. maculans, were isolated<br />

from canola stubble or associated soil collected in South<br />

Australia.<br />

Dual culture on agar plates. L. maculans and potential<br />

antagonists were co‐inoculated in triplicate on potato dextrose<br />

agar (PDA) in Petri dishes and incubated at room temperature<br />

for up to 3 weeks. Changes in colony appearance, hyphal growth<br />

and sporulation of L. maculans were examined.<br />

Dual culture on agar‐coated slides. Three replicate water agarcoated<br />

slides were inoculated for each test species and<br />

incubated at room temperature for up to 2 weeks. Interactions<br />

between hyphae of L. maculans and each test fungus were<br />

examined microscopically at 4‐day intervals.<br />

Inoculation of blackleg‐affected stubble. Blackleg‐affected<br />

stubble, three segments on moist sand in each of four Petri<br />

dishes, was inoculated with each test species and incubated at<br />

15ºC in a 12 h photoperiod. After 6 weeks, the density of<br />

pseudothecia (number in a 0.5 × 1 cm area) was assessed for<br />

each stubble segment. Germination of ascospores collected from<br />

stubble inoculated with each test fungus was assessed.<br />

Effect of fungi on decay of canola stubble. Disease‐free stubble<br />

inoculated with Stachybotrys chartarum and Coprinus sp. was<br />

incubated at 20ºC in a 12 h photoperiod for 2 months and<br />

weight loss determined.<br />

Statistical analysis. Data were subjected to ANOVA.<br />

RESULTS<br />

Macroscopic and microscopic observations of plates and agarcoated<br />

slides, respectively, showed lysis, deformation, overgrowth<br />

of hyphae and inhibition of growth and sporulation of L.<br />

maculans by Alternaria spp., Arthrobotrys sp., Aspergillus sp.,<br />

Fusarium equiseti, Gliocladium roseum, Myrothecium sp.,<br />

Trichoderma aureoviride, Sordaria sp., S. chartarum and an<br />

unknown Coelomycete.<br />

reduced by over 67% compared with pathogen‐alone controls<br />

following inoculation with F. equiseti, G. roseum, T. aureoviride,<br />

Sordaria sp. or the Coelomycete. Over 70% of L. maculans<br />

ascospores obtained from the stubble germinated after 24 h at<br />

20ºC, irrespective of the co‐inoculated fungus.<br />

The mass of canola stubble inoculated with S. chartarum and<br />

Coprinus sp. was reduced almost 2‐fold compared with that of<br />

uninoculated controls (Table 1).<br />

Table 1. The effect of Coprinus sp. and Stachybotrys chartarum on<br />

weight of canola stubble after 2 months at 20ºC.<br />

Mean decrease of stubble weight (mg) 1<br />

Test fungus Fresh weight Air‐dry weight<br />

Control 2 5.1 12.1<br />

Coprinus sp. 9.6 20.1<br />

Stachybotrys chartarum 13.4 26.3<br />

1<br />

Mean of the difference of weight before inoculation and after incubation.<br />

2<br />

Each control stubble segment received 2 ml sterile distilled water, whereas each<br />

treated stubble segment was inoculated with two plugs of PDA culture of Coprinus<br />

sp. (plus 2 ml of sterile distilled water) or 2 ml of spore suspension of S. chartarum.<br />

DISCUSSION<br />

This study provides information on interactions between L.<br />

maculans and a wide range of fungal species, isolated from<br />

canola stubble or associated soil, on agar media and on canola<br />

stubble. Several stubble‐associated fungi both antagonised the<br />

pathogen in vitro and reduced pseudothecium formation on<br />

canola stubble. In particular, Coprinus sp. and S. chartarum, as<br />

antagonists of L. maculans and effective decomposers of canola<br />

stubble in this study, have potential to reduce pathogen<br />

inoculum on stubble in the field. This warrants further<br />

investigation, including longer‐term field studies, to assess the<br />

suitability of fungal antagonists as a biological means of<br />

controlling blackleg.<br />

ACKNOWLEDGEMENTS<br />

We thank T. Potter, SARDI, for supplying stubble.<br />

REFERENCES<br />

1. Hysek J, Vach M, Brozova J, Sychrova E, Civinova M, Nedelnik J,<br />

Hruby J (2002) The influence of the application of mineral fertilizers<br />

with the biopreparation supresivit (Trichoderma harzianum) on the<br />

health and the yield of different crops. Archives of Phytopathology<br />

and <strong>Plant</strong> Protection 35, 115–24.<br />

2. Maksymiak MS, Hall AM (2000) Biological control of Leptosphaeria<br />

maculans (anamorph Phoma lingam) causal agent of blackleg<br />

canker on oilseed rape by Cyathus striatus, a bird's nest fungus.<br />

Proceedings of the British Crop Protection Council Conference: Pest<br />

and Diseases. Brighton, UK, 13–6 November.<br />

None of the 21 species tested eliminated L. maculans from<br />

stubble. Density of pseudothecia of L. maculans on stubble was<br />

204 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


86 First report of tomato yellow leaf curl virus in pepper (Capsicum annum) fields<br />

in Iran<br />

Posters<br />

M. Shirazi{ XE "Shirazi, M." } 1,2 , J. Mozafari 1 , F. Rakhshandehroo 2 and M. Shams‐Bakhsh 3<br />

1 Department of Genetics and National <strong>Plant</strong> Gene‐ Bank, Seed and <strong>Plant</strong> Improvement Institute, Mahdasht RD,Karaj, Iran<br />

2 <strong>Plant</strong> <strong>Pathology</strong> Department, Faculty of Agriculture, Islamic Azad University Science and Research Campus, Tehran, Iran<br />

3 <strong>Plant</strong> <strong>Pathology</strong> Department, Faculty of Agriculture, Tarbiat Modares University, Tehran, Iran<br />

mhishirazi@yahoo.com<br />

INTRODUCTION<br />

Tomato yellow leaf curl virus (TYLCV) is one of the most<br />

devastating pathogens affecting tomato (Solanum lycopersicum)<br />

worldwide and is very important in Iran. TYLCV is transmitted by<br />

Bemisia tabaci in persistent and circulative manner. Tomato is<br />

the most important host for TYLCV. Symptoms on hosts except<br />

pepper include stunting, yellowing, leaf curl and flower<br />

senescence, whereas no distinct symptoms has been reported<br />

on pepper and those which occasionally observed, are caused by<br />

vector feeding (4). There are conflicting reports regarding the<br />

susceptibility of peppers (Capsicum spp.) to TYLCV and in Spain,<br />

C. annuum was reported as a host of an uncharacterised strain<br />

of TYLCV (5).<br />

MATERIALS AND METHODS<br />

In order to detection of TYLCV in Southern Iran, samples were<br />

collected from tomato and pepper fields in Bandar Abbas region<br />

including Rezvan and Sarkhun areas and jiroft region in 2007.<br />

Attention to disease symptoms including stunting, leaf curling,<br />

yellowing and deformation of stem end, 38 samples from<br />

Sarkhun, 80 samples from Rezvan and 10 samples from jiroft<br />

were collected. DNA extraction was performed with Dellaporta<br />

protocol (2). Virus detection was conducted by PCR with specific<br />

primers (3) and also by DAS‐ELISA (1).<br />

RESULTS AND DISCUSSION<br />

After DNA extraction and specific PCR, 2 samples of pepper and<br />

6 samples of tomato from 11 samples of pepper and 27 samples<br />

of tomato in Sarkhun, 68 samples from 80 samples of tomato in<br />

Rezvan and 4 samples of tomato and 2 samples of pepper from 4<br />

samples of tomato and 6 samples of pepper in jiroft were<br />

infected by PCR (Table 1), while DAS‐ELISA couldn’t detect any<br />

infection. Indeed a viral DNA fragment of 670 bp including a part<br />

of coat protein and movement protein genes was amplified in<br />

positive samples (Figure 1). Infection of 85% of Rezvan samples<br />

and about 21% of Sarkhun samples and 60% of jiroft samples<br />

indicate the most incidence of TYLCV in Rezvan area and also<br />

detection of virus in pepper samples indicate presence of TYLCV<br />

in pepper fields in southern Iran. This is the first report of pepper<br />

plants infected by Tomato yellow leaf curl virus in Iran.<br />

750 bp<br />

500 bp<br />

Figure 1. A viral DNA fragment of 670 bp was amplified in positive<br />

samples.<br />

(M): 1 Kb ladder‐ (1,2,3): Tomato positive samples‐(C‐): Negative control‐ (4):<br />

Pepper positive sample and (C+): positive control<br />

REFERENCES<br />

–M–—1–— 2 –— 3 –— C–—4 — C+<br />

670 bp<br />

1. Clark, M. R. and Adams, A. N. I977. Characteristics of the<br />

microplate method of enzyme‐linked immunosorbent assay for the<br />

detection of plant viruses. Journal of General Virology, 34: 475–<br />

483.<br />

2. Dellaporta, S.L.; Wood, J. y Hicks, J.B. 1983. A plant DNA<br />

minipreparation: version II. <strong>Plant</strong>. Mol.Biol. Rep. 1(4): 19–21<br />

3. Pico, B., Diez, M.J., Nuez, F. 1998. Evaluation of whitefly‐mediated<br />

inoculation techniques to screen Lycopersicon esculentum and wild<br />

relatives for resistance to tomato yellow leaf curl virus. Euphytica.<br />

101: 259–271<br />

4. Polston, J. E., Cohen, L., Sherwood, T. A., Ben‐Joseph, R., and<br />

Lapidot, M. 2006. Capsicum species: Symptomless hosts and<br />

reservoirs of Tomato yellow leaf curl virus. Phytopathology 96:447–<br />

452<br />

5. Reina, J., Morilla, G., Bejarano, E.R., Rodríguez,M.D., Janssen, D. Y.<br />

Cuadrado, I.M. 1999. First Report of Capsicum annuum <strong>Plant</strong>s<br />

Infected by Tomato Yellow Leaf Curl Virus. <strong>Plant</strong> Dis. 83: 1176<br />

Table 1. sampling regions and number of collected and positive samples<br />

Regions<br />

Collected<br />

samples<br />

(Tomato)<br />

Collected<br />

samples<br />

(Pepper)<br />

Positive<br />

samples<br />

(Tomato)<br />

Rezvan 80 0 69 0<br />

Sarkhun 25 13 5 4<br />

Jiroft 6 4 4 2<br />

Positive<br />

samples<br />

(Pepper)<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 205


Posters<br />

39 Seed‐borne concerns with wheat streak mosaic virus in 2008<br />

S. Simpfendorfer{ XE "Simpfendorfer, S." } A and D. Nehl B<br />

A NSW Department of Primary Industries, 4 Marsden Park Rd, Tamworth, 2340, NSW<br />

B NSW Department of Primary Industries, PMB 8, Camden, 2570, NSW<br />

INTRODUCTION<br />

Wheat streak mosaic virus (WSMV), is a sap‐borne viral disease<br />

transmitted by the wheat curl mite (WCM). Western Australian<br />

research reported globally for the first time in 2005 that low<br />

levels of seed transmission (


87 Effect of white rust infection, bion and phosphonate on glucosinolates in brassica<br />

crops<br />

Posters<br />

Astha Singh{ XE "Singh, A." }, Les Copeland and David Guest<br />

University of Sydney<br />

INTRODUCTION<br />

Broccoli, Rocket and Indian mustard are Brassica crops<br />

consumed throughout the world. The major class of secondary<br />

metabolites formed in these crops are glucosinolates, many of<br />

which affect human health positively or negatively. White rust,<br />

caused by Albugo candida, is the most common disease affecting<br />

brassicas, although there is no information on the effect of<br />

disease on levels of beneficial and harmful glucosinolates, or<br />

their impacts on human health. In this research we describe the<br />

effect of white rust, on glucosinolate levels in leaves of these<br />

brassica crops. In addition, we investigate the effects of two<br />

activators of plant defence, Bion and phosphonate, on disease<br />

and glucosinolate levels.<br />

MATERIALS AND METHODS<br />

<strong>Plant</strong>s. Seeds (10 per pot) of Broccoli (Brassica olearacea) cv.<br />

‘Greenbelt’ from Terranova Seed Company and Rocket (Eruca<br />

sativa) obtained from Yates Seed Company were sown in 10 cm<br />

diameter pots filled with Standard UC Mix + 10 g/kg Osmocote<br />

(Scott’s Australia Pty. ltd).<br />

<strong>Plant</strong>s were grown in Growth Cabinets, with relative humidity of<br />

90% daytime and 70% night, light intensity of 710 µmoles,<br />

temperature of 15°C with daily irrigation for 1 min for broccoli<br />

and Glasshouse with 20°C temperature and daily irrigation with<br />

overhead sprinklers twice for rocket. Chemical treatments were<br />

sprayed 10 days before inoculation. <strong>Plant</strong>s were inoculated 30<br />

days after sowing.<br />

Pathogen. Albugo candida obtained from DPI Victoria (Elizabeth<br />

Minchinton) from broccoli as sporangia on fresh leaves, and<br />

from rocket from a domestic garden. The pathogens were<br />

transferred onto healthy broccoli or rocket plants every 14 days.<br />

Chemical treatments. Twenty day old plants were sprayed with<br />

Bion (10, 25 or 100 mg/L acibenzolar‐S‐methyl; Syngenta) or<br />

phosphonate (0.5, 1.0 or 2.0 g/L a.i. Agrifos Supa 600; Agrichem<br />

Manufacturing Industries) until runoff, requiring approximately<br />

25 mL/pot.<br />

Inoculation technique. Spores were scraped off leaf pustules<br />

into sterile distilled water. After vortexing and centrifuging<br />

@2000 rpm for 5 minutes the pellet was discarded. The<br />

supernatant was incubated at 16°C for 3 h to induce zoospore<br />

release and adjusted to 10 4 zoospores/mL, and sprayed onto all<br />

leaves of 30 day old broccoli and 20 day old rocket plants.<br />

Inoculated plants were held at 16°C and 90% RH in the growth<br />

cabinet. White powdery blisters appeared on lower surfaces of<br />

inoculated leaves 7–10 days after inoculation.<br />

HPLC Analysis. Leaves and stems were detached and frozen in<br />

liquid nitrogen then freeze dried. The sample was then ground<br />

and 0.2 g was heated in a 90°C water bath for 10 min, then 10<br />

mL boiling water was added to the tube and heated in the water<br />

bath for another 10 min. The samples were centrifuged for 10<br />

min @ 3000 rpm. The supernatant was collected and the pellet<br />

was resuspended in 10 mL water, vortexed and centrifuged for<br />

another 10 minutes. The supernatants were pooled and 1 mL<br />

was filtered through 0.45 µm nylon filters into auto sampler<br />

vials.<br />

Samples were analysed using reverse‐phase HPLC (West et al.<br />

2002, modified by R. Jones, DPI Victoria, pers. comm.). Sinigrin<br />

was used as the internal standard and other glucosinolates<br />

(progoitrin, glucoiberin, glucoraphanin, gliucobrassinin and<br />

neoglucobrassin) were identified according to their relative<br />

retention times.<br />

RESULTS<br />

Table 1. Disease development on rocket<br />

Treatment<br />

Conc.<br />

Days for first<br />

symptom<br />

Uninoculated, untreated 11 Nil<br />

Inoculated + Bion 10 mg/L 10 Nil<br />

Inoculated +<br />

Phosphonate<br />

Inoculated + Bion +<br />

Phosphonate<br />

25 mg/L 12 Nil<br />

100 mg/L Nil 11<br />

0.5 g/L 15 Nil<br />

1.0 g/L 12 Nil<br />

2.0 g/L Nil Nil<br />

25 mg/L +<br />

1.0 g/L<br />

11 Nil<br />

Uninoculated + Bion 10 mg/L Nil Nil<br />

Uninoculated +<br />

Phosphonate<br />

Uninoculated + Bion +<br />

Phosphonate<br />

25 mg/L Nil Nil<br />

100 mg/L Nil 11<br />

0.5 g/L Nil Nil<br />

1.0 g/L Nil Nil<br />

2.0 g/L Nil Nil<br />

25 mg/L +<br />

1.0 g/L<br />

Yellowing<br />

The results implied that the optimum concentrations to be used<br />

in further experiments were 25 mg/L Bion and 1.0 g/L<br />

phosphonate.<br />

GLUCOSINOLATE STUDY IN BROCCOLI<br />

Infected leaves and roots were collected at 5 day intervals<br />

before spraying, after spraying, pre and post infection. <strong>Plant</strong>s<br />

were sprayed with Bion (25 mg/L), phosphonate (1.0 g/L) or<br />

combined Bion (25 mg/L) plus phosphonate (1.0 g/L). Results<br />

that indicate differences in levels of different glucosinolates at<br />

different stages of disease and chemical applications. There<br />

were no significant differences between glucosinolates in<br />

infected and uninfected leaves on inoculated plants.<br />

DISCUSSION<br />

Results clearly show that there is effect of the defence activators<br />

on the symptom development. Forthcoming results will describe<br />

the effect of disease as well as the defence activators on the<br />

levels of glucosinolates.<br />

ACKNOWLEDGEMENTS<br />

Rodney Jones, Michael Imsic, Liz Minchinton (DPI Vic).<br />

REFERENCE<br />

1. West L, Tsui I, Haas G. (2002). Single column approach for the liquid<br />

chromatographic separation of polar and non‐polar glucosinolates<br />

from broccoli sprouts and seeds. Journal of Chromatography 966,<br />

227–232.<br />

Nil<br />

Nil<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 207


Posters<br />

55 Fertilisation with N, P and K above critical values required for adequate plant<br />

growth influences plant establishment of cotton varieties in fusarium infested soil<br />

L.J. Smith{ XE "Smith, L.J." } A and J.K. Lehane B<br />

A Queensland Primary Industries and Fisheries, 80 Meiers Road, Indooroopilly, 4068, Qld<br />

B Queensland Primary Industries and Fisheries, PO Box 102, Toowoomba, 4350, Qld<br />

INTRODUCTION<br />

Fertiliser recommendations are developed to optimise nutrient<br />

uptake and provide the crop with adequate nutrients for normal<br />

growth and yield. Once critical levels of nutrients are met, no<br />

response to yield is expected from further nutrient application,<br />

but there may be other benefits. In some instances, nutrient<br />

applications higher than those needed for optimum growth may<br />

result in improved disease resistance. The overall aim of this<br />

work is to determine the effect of N, P and K fertilisation on<br />

nutrient uptake, plant establishment, disease severity and yield<br />

on two cotton varieties grown in Fusarium infested soil. In this<br />

paper the effect of nutrient application on plant establishment<br />

of two varieties differing in Fusarium wilt resistance will be<br />

discussed.<br />

MATERIALS AND METHODS<br />

A field experiment was conducted from November 2008 to May<br />

2009 near Cecil Plains, QLD, in soil naturally infested with the<br />

Fusarium wilt pathogen. Soil cores were collected and analysed<br />

for nutrient availability. Two cotton varieties differing in<br />

Fusarium wilt resistance were investigated: Sicala 45 BRF (F‐rank<br />

126) and Sicala 60 BRF (F‐rank 102). The experimental design<br />

was factorial with 16 treatments in randomised blocks, 6 blocks<br />

per treatment. Triple Superphosphate was applied at 0, 20, 40<br />

and 80 kg/ha; Urea with Entec at 0 and 150 kg/ha; and Muriate<br />

of Potash at 0 and 100 kg/ha. Calcium sulphate (200 kg/ha) was<br />

applied to every plot. Fertiliser treatments were applied by<br />

hand, broadcast to each plot. Hills were reformed following<br />

application. Seeds were sown at a depth of 10 cm. The<br />

experiment was irrigated and managed commercially.<br />

RESULTS AND DISCUSSION<br />

Nutrient availability. Availability of N, P and K of field soil<br />

exceeded the critical values required for adequate plant growth<br />

(Table 1). Therefore, application of N, P and K was considered<br />

above that required for adequate plant growth and optimal<br />

yields.<br />

Table 1. Nutrient and pH analysis of field soil<br />

Measurement 0–15 cm<br />

15–60<br />

cm 60–120 cm<br />

Critical<br />

value<br />

pH 8.01 8.28 8.45 ‐<br />

N mg/kg 22 40 30 20–30<br />

P mg/kg‐Bicarb 29 9 7 6<br />

K mg/kg 331 199 276 100–150<br />

Varietal difference. Significantly more plants established for<br />

Sicala 45 BRF than for Sicala 60 BRF, highlighting the importance<br />

of planting varieties with higher F‐rank in Fusarium infested soils.<br />

N, P and K effects. Application of N at 150 kg/ha, significantly<br />

increased the number of plants established for both varieties<br />

(Table 2). N application has been associated with reduced<br />

Fusarium wilt disease in cotton. This may be due to an effect of<br />

form of N on the pathogen population in the soil.<br />

Table 2. The effect of nitrogen (N) fertilisation on establishment of<br />

varieties Sicala 45 BRF (V1) and Sicala 60 BRF (V2)<br />

N kg/ha V1 F‐rank 126 V2 F‐rank 102<br />

0 67 a 51 a<br />

150 73 b 59 b<br />

Data followed by different letters are significantly different from one another<br />

Application of P at the highest rate, significantly increased<br />

seedling death of Sicala 60 BRF due to Fusarium wilt compared<br />

to plants that had 0 and 40 kg/ha of P applied (Table 3). There<br />

are reports in the literature of increasing P levels both increasing<br />

and decreasing severity of Fusarium wilt. Unfortunately little is<br />

understood about how P influences disease severity.<br />

Table 3. The effect of P fertilisation on emergence and establishment of<br />

varieties Sicala 45 BRF and Sicala 60 BRF<br />

P Kg/ha V1 F‐rank 126 V2 F‐rank 102<br />

0 69 a 58 b<br />

20 73 a 54 ab<br />

40 68 a 56 b<br />

80 69 a 51 a<br />

Data followed by different letters are significantly different from one another<br />

Cotton wilt is commonly found to be more destructive to the<br />

crop when grown on potassium deficient soils, and the<br />

application of high potash fertilisers has real value in the<br />

reduction of wilt, particularly when used on resistant varieties. In<br />

this trial K was abundant, however despite this, addition at 100<br />

kg/ha significantly increased plant establishment of Sicala 45 BRF<br />

(Table 4).<br />

Table 4. The effect of K fertilisation on establishment of varieties Sicala<br />

45 BRF (V1) and Sicala 60 BRF (V2)<br />

K kg/ha V1 F‐rank 126 V2 F‐rank 102<br />

0 68 a 54 a<br />

100 72 b 56 a<br />

Data followed by different letters are significantly different from one another<br />

Interactive effects. When neither N nor K fertiliser was applied,<br />

the number of plants established was significantly reduced<br />

compared to other NK treatment combinations. For Sicala 45<br />

BRF, when N was applied at 150 kg/ha and P at 20 kg/ha, plant<br />

establishment was significantly increased compared to all other<br />

NP treatments (data not shown). These results highlight the<br />

importance of a balanced nutrition which is required for the<br />

functioning of inherent pathogen defence mechanisms.<br />

In conclusion, even when nutrient availability is adequate, plant<br />

establishment in Fusarium infested soil can be influenced by the<br />

application of N, P and K.<br />

ACKNOWLEDGEMENTS<br />

We thank the Cotton Research and Development Corporation<br />

for funding this research.<br />

208 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


56 Eradication of Elsinoe ampelina by burning infected grapevine material<br />

M.R. Sosnowski{ XE "Sosnowski, M.R." } A,D , R.W. Emmett B,D , T.A. Vu Thanh C , T.J. Wicks A,D and E.S. Scott C,D<br />

A South Australian Research and Development Institute, GPO Box 397, Adelaide, South Australia 5001<br />

B Department of Primary Industries Victoria, PO Box 905, Mildura, Victoria 3502<br />

C School of Agriculture, Food and Wine, The University of Adelaide, Waite Campus, Glen Osmond, South Australia 5064<br />

D Cooperative Research Centre for National <strong>Plant</strong> Biosecurity, LPO Box 5012, Bruce ACT 2617<br />

Posters<br />

INTRODUCTION<br />

Burning infected plant material is widely used in the control and<br />

eradication of endemic and exotic pathogens. However, there is<br />

little scientific evidence to confirm that pathogens are<br />

eliminated during this process (1). We report experiments to<br />

assess the efficacy of burning as a means of eradicating<br />

pathogens from woody plants.<br />

Black spot (anthracnose), caused by Elsinoe ampelina, is an<br />

important disease of grapevines worldwide (2). The fungus<br />

infects leaves, stems, petioles and berries. This pathogen was<br />

chosen as a model to develop an eradication strategy for the<br />

exotic disease black rot, caused by Guignardia bidwellii. Black rot<br />

has similar biology and epidemiology to black spot and could<br />

have a severe economic impact on the wine industry if it became<br />

endemic (3).<br />

MATERIALS AND METHODS<br />

An experiment was conducted in the Sunraysia district of<br />

Victoria. Vines (cvs Red Globe, Christmas Rose, Blush Seedless<br />

and Fantasy Seedless) were inoculated in spring 2007 by<br />

spraying a suspension of E. ampelina conidia on new shoots with<br />

2–4 unfolded leaves. The shoots were covered with polyethylene<br />

bags overnight to provide high humidity to promote spore<br />

germination and infection.<br />

In July 2008, vines were drastically pruned in an experiment to<br />

eradicate the disease. On treated vines, all plant material above<br />

the crown was removed and placed in a pit (5 x 3.5 x 0.5 m). In<br />

August 2008, six steel poles were placed upright at random<br />

within the pit. Steel mesh bags, containing infected vine canes<br />

(approx 30 g each) and temperature crayons (Tempilstik) in glass<br />

Petri dishes were attached to the poles at 20 and 50 cm above<br />

the pit floor. Another set of mesh bags was buried 5 cm below<br />

the soil surface on the floor of the pit. After the vine material<br />

was burnt, the mesh bags were collected and the ash was<br />

transferred to plastic tubes. Unburnt canes from untreated<br />

control material and buried samples were grated using a cheese<br />

grater. All samples were stored at 3–4°C until they were used.<br />

A bioassay was conducted in a glasshouse at 22–28°C in<br />

December 2008 using potted grapevines (cv. Thompson<br />

Seedless). The three youngest expanded leaves on each shoot<br />

were sprayed with deionised water and dusted with the ash or<br />

grated vine material. Each treatment was applied to 3–4 shoots<br />

per vine and the inoculated shoots were covered with<br />

polyethylene bags. After 48 hours, the bags were removed, the<br />

leaves were sprayed again with deionised water and the bags<br />

were replaced and left overnight. The experiment was arranged<br />

as a completely randomised design with two replicate vines per<br />

treatment.<br />

Twelve days after inoculation, the vines were assessed for<br />

symptoms of black spot. A Mann‐Whitney U‐test was used to<br />

analyse data.<br />

RESULTS<br />

The temperature crayons indicated that the fire reached in<br />

excess of 250°C and variable temperatures up to 60°C occurred 5<br />

cm below the soil surface. No leaf symptoms developed on<br />

plants inoculated with ash whereas significant symptoms were<br />

observed on plants inoculated with grated material from the<br />

controls and less severe symptoms occurred on plants<br />

inoculated with buried cane material (Fig. 1).<br />

Mean disease score<br />

3<br />

2<br />

1<br />

0<br />

Leaf 1<br />

Leaf 2<br />

Leaf 3<br />

control -5cm 20cm 50cm<br />

Figure 1. Mean disease score on the three newest leaves on grapevine<br />

shoots (cv. Thompson Seedless) inoculated with ash from burnt vine<br />

material positioned 20 and 50 cm above the floor of the bonfire, buried 5<br />

cm below the soil surface under the fire or untreated (control). Samples<br />

of grapevine canes infected with E. ampelina in steel mesh bags were<br />

positioned at 20 and 50 cm above the ground or buried 5 cm below the<br />

surface. Data are presented for each leaf individually, with Leaf 1 being<br />

the oldest.<br />

DISCUSSION<br />

These results confirm the efficacy of burning infected vine<br />

material, as temperatures exceeded those that are lethal to the<br />

fungus and the bioassay verified that this was the case.<br />

However, any pathogen on debris which penetrates the soil may<br />

not be eliminated. Further research is under way to evaluate<br />

burning for eradication of the bacterium Xanthomonas<br />

translucens from pistachio trees.<br />

ACKNOWLEDGEMENTS<br />

We thank Chris Dyson (SARDI) for statistical support, members<br />

of the Department of Sustainability and Environment Victoria for<br />

assisting with the fire and the CRC for National <strong>Plant</strong> Biosecurity<br />

for funding this research.<br />

REFERENCES<br />

1. Sosnowski MR, Fletcher JD, Daly AM, Rodoni BC and Viljanen‐<br />

Rollinson SLH (2009) Techniques for the treatment, removal and<br />

disposal of host material during programmes for plant pathogen<br />

eradication. <strong>Plant</strong> <strong>Pathology</strong>, Early view online DOI:<br />

10.1111/j.1365‐3059.2009.02042.x<br />

2. Magarey RD, Coffey BE and Emmett RW (1993) Anthracnose of<br />

grapevines, a review. <strong>Plant</strong> Protection Quarterly 8, 106–110.<br />

3. Wilcox W (2003) Grapes: Black rot (Guignardia bidwelli (Ellis) Viala<br />

and Ravaz.). Cornell Cooperative Extension Disease Identification<br />

Sheet No. 102GFSG‐D4, Cornell University.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 209


Posters<br />

88 Recent plant virus incursions into Australia<br />

J.E. Thomas A , V. Steele{ XE "Steele, V." } A , A.D.W. Geering A , D.M. Persley A , C.F. Gambley A , and B.H. Hall B<br />

A Queensland Primary Industries and Fisheries, DEEDI, 80 Meiers Rd, Indooroopilly, 4068, Queensland<br />

B SARDI,GPO Box 397, Adelaide, 5001,SA<br />

INTRODUCTION<br />

The continuing improvement in diagnostic methods combined<br />

with the increased frequency of international travel and reduced<br />

trade barriers have all probably contributed to the upsurge in<br />

new virus records. A number of plant viruses have recently been<br />

detected in Australia for the first time. These include; Colombian<br />

datura virus from Brugmansia sp., High plains virus from wheat<br />

(Triticum aestivum), Panicum mosaic virus from buffalo grass<br />

(Stenotaphrum secundatum), Tomato torrado virus from tomato<br />

(Lycopersicon esculentum) and Tomato yellow leafcurl virus from<br />

tomato.<br />

METHODS AND RESULTS<br />

Colombian datura virus (CDV). A sample of Brugmansia sp.<br />

(angel’s trumpets) with leaf mosaic symptoms was obtained<br />

from Bomaderry, NSW, in August 2007. The plant contained<br />

flexuous filamentous particles about 840 nm long and gave a<br />

positive reaction in ELISA with AGDIA potyvirus group<br />

antibodies. The C‐terminal region of the coat protein and the 3’<br />

UTR were amplified by RT‐PCR from an RNA extract (Qiagen<br />

RNeasy) and the products cloned and sequenced [1]. By<br />

comparison with CDV sequences on the GenBank database, the<br />

Australian sequence was 99.6–100% identical in the 3’ UTR, and<br />

the amino acid sequence of the 3’ portion of the coat protein<br />

was 100% identical. The consensus sequence of the Australian<br />

CDV isolate 2079 has been deposited in the GenBank database<br />

under accession FJ821796.<br />

Panicum mosaic virus (PMV). In June, 2008, a sample of<br />

Stenotaphrum secundatum (buffalo grass) showing mosaic<br />

symptoms was obtained from the Sydney basin, New South<br />

Wales. The plant contained isometric virions ca. 25–30 nm<br />

diameter which were trapped by immunosorbent electron<br />

microscopy and decorated, using an antiserum to the St<br />

Augustine decline strain of Panicum mosaic virus (PMV). Specific<br />

PCR primers were designed to amplify the complete coat protein<br />

(CP) gene sequence of PMV. The CP of the Australian sample<br />

(PMV isolate 2349) was 90% and 85% identical to PMV GenBank<br />

Accession PMU55002 at the amino acid and nucleotide levels,<br />

respectively.<br />

Tomato yellow leafcurl virus (TYLCV). Cherry tomato crops<br />

displaying symptoms of leaf curling, chlorosis and stunting were<br />

first reported from Pallara, Brisbane in March, 2006. Subsequent<br />

surveys indicated that the disease was common in the periurban<br />

areas of Brisbane, and incidences of nearly 100% were not<br />

uncommon. Infected plants gave a positive ELISA reaction with<br />

antibodies to African cassava mosaic virus (AGDIA), indicating<br />

the presence of a begomovirus. ELISA‐positive samples were also<br />

obtained from the Lockyer Valley, Caboolture and Bundaberg<br />

areas of south‐east Queensland during March to May, 2006. The<br />

TYLCV‐specific PCR primers, TYLCV‐F1 and TYLCV‐R1 were<br />

designed to produce a 336 bp amplicon from the Rep gene. The<br />

nucleotide sequences of these amplicons from three<br />

representative isolates were ca 99% identical to those of TYLCV<br />

accessions from GenBank.<br />

and occasionally necrotic lesions on leaves, stunting and leaf<br />

distortion and were associated with high greenhouse whitefly<br />

populations. A low concentration of isometric virions was<br />

observed in sap preparations by electron microscopy. PCR using<br />

the TToV specific primer pairs TR1F/R and TR2F/R, designed to<br />

RNA‐1 and RNA‐2 respectively [2], gave amplicons of the<br />

expected size from isolate 1883 (collected in 2006) and isolate<br />

2136 (collected in 2008). Both Australian isolates shared ca 99%<br />

nucleotide sequence identity with each other and with overseas<br />

isolates on both RNA components.<br />

High Plains virus (HPV). HPV is a presently‐unclassified, mitetransmitted<br />

virus which often occurs as mixed infections with<br />

Wheat streak mosaic virus (WSMV). The latter virus was<br />

recorded for the first time in Australia in 2003 [3]. Total RNA<br />

extracts of samples from the 2003 WSMV surveys were used to<br />

test for the presence of HPV. RT‐PCR primers were designed to<br />

amplify part of the putative nucleocapsid gene of HPV (GenBank<br />

accession U60141). RT‐PCR was done using a one‐step RT‐PCR kit<br />

(Qiagen). Amplicons of the expected size (483 bp) were<br />

amplified from some but not all WSMV‐infected plants,<br />

indicating dual infection of some plants with HPV. The<br />

nucleotide sequences of these products were 100% identical to<br />

the published HPV sequence. HPV‐infected wheat samples were<br />

obtained from experimental field plots at Roseworthy and<br />

Adelaide in South Australia and Horsham in Victoria and from a<br />

commercial crop near Moonie in Queensland.<br />

Isolates of all the above viruses have been deposited in the<br />

Queensland Primary Industries and Fisheries <strong>Plant</strong> Virus<br />

Collection.<br />

ACKNOWLEDGEMENTS<br />

We thank G Ellis and E Colson for WSMV survey samples,<br />

Biosecuity Queensland staff for TYLCV survey samples and P<br />

Pezzaniti for assistance with collecting samples of TToV.<br />

REFERENCES<br />

1. Schwinghamer MW, Thomas JE, Parry JN Schilg MA, Dann EK (2007)<br />

First record of natural infection of chickpea by Turnip mosaic virus.<br />

<strong>Australasian</strong> <strong>Plant</strong> Disease Notes 2, 41–43.<br />

2. Pospieszny H, Borodynko N, Obrepalska‐Steplowska A, Hasiow B<br />

(2007) The First Report of Tomato torrado virus in Poland. <strong>Plant</strong><br />

Disease 91, 1364–1364.<br />

3. Ellis MH, Rebetzke GJ, Chu P (2003) First report of Wheat streak<br />

mosaic virus in Australia. <strong>Plant</strong> <strong>Pathology</strong> 52, 808.<br />

Tomato torrado virus (TToV). Since 2005, a new disease of<br />

greenhouse tomatoes has been present in the Northern<br />

Adelaide Plains, South Australia. Symptoms included chlorosis<br />

210 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


16 Investigation of the effect of three essential oils, alone and in combination, on the<br />

in vitro growth of Botrytis cinerea<br />

Posters<br />

S.M. Stewart‐Wade{ XE "Stewart‐Wade, S.M." }<br />

Melbourne School of Land and Environment, The University of Melbourne, Burnley Campus, Richmond, 3121, Victoria<br />

INTRODUCTION<br />

Many essential oils have been screened individually for their<br />

antifungal properties, but little work has been done on the<br />

combination of these oils and their potential synergistic effect<br />

on fungal plant pathogens (1). The fungus Botrytis cinerea Pers.<br />

ex. Fr. causes the disease grey mould worldwide on more than<br />

100 plant species including fruit, vegetables, ornamentals and<br />

field crops, under field, glasshouse and postharvest storage<br />

conditions (2). Three essential oils, derived from clove buds,<br />

cinnamon leaves and red thyme leaves, showed the greatest<br />

inhibition of in vitro growth of B. cinerea in previous studies (2,<br />

3). The aim of this research is to examine the effect of the<br />

essential oils clove, cinnamon and red thyme, alone and in<br />

combination, at varying concentrations, on in vitro growth of<br />

Botrytis cinerea.<br />

Mean Colony Diameter (mm)<br />

30<br />

Cv<br />

Cn<br />

25<br />

T<br />

CvCn<br />

20<br />

CvT<br />

CnT<br />

15<br />

CvCnT<br />

10<br />

5<br />

0<br />

0 125 250 500 1000<br />

Oil Concentration (ppm)<br />

MATERIALS AND METHODS<br />

Stocks of three pure essential oils, namely clove bud (Syzygium<br />

aromaticum (L.) Merr. & Perry) oil, cinnamon leaf (Cinnamomum<br />

zeylanicum J. Presl.) oil and red thyme leaf (Thymus vulgaris L.)<br />

oil (Auroma Pty Ltd, Hallam, Vic.) were prepared with 0.1% v/v<br />

(final) Tween 80 used as an emulsifier. Essential oils and Tween<br />

80 were filter sterilised into cooled molten sterilised PDA. Each<br />

oil was added, either alone or in combination in equal amounts<br />

with each other oil (two and three oil combinations), at the<br />

following concentrations: 0 (nil control), 125, 250, 500 or 1000<br />

ppm. Mycelial plugs (6mm diameter) from the margins of<br />

actively growing, 3–4 day old B. cinerea cultures were inoculated<br />

centrally onto PDA plates containing the essential oils and the<br />

plates were sealed with Parafilm. Seven replicate plates were<br />

used per treatment and cultures were incubated at room<br />

temperature (~18–22°C) under natural light conditions. Mean<br />

colony diameter (mm) was measured 24 and 48 h after<br />

inoculation. The experiment was repeated. Data were analysed<br />

with an ANOVA using GenStat and means were separated using<br />

LSDs at P=0.001.<br />

RESULTS<br />

Presence of essential oils was a significant factor (P


Posters<br />

42 A single plant test for resistance in wheat to crown rot and root‐lesion nematode<br />

(Pratylenchus thornei)<br />

J.P. Thompson{ XE "Thompson, J.P." } and R.B. McNamara<br />

A DEEDI, Qld Primary Industries and Fisheries, Leslie Research Centre, PO Box 2282, Toowoomba, 4350, Queensland<br />

INTRODUCTION<br />

Fusarium pseudograminearum, the cause of crown rot, and rootlesion<br />

nematode (Pratylenchus thornei) are the most serious soilborne<br />

pathogens of wheat in the Australian northern grain<br />

region. Only one cultivar (EGA Wylie) is both tolerant to P.<br />

thornei and moderately resistant to crown rot. Glasshouse<br />

methods have been developed to test wheat for resistance to<br />

crown rot (Wildermuth and McNamara 1994) and root‐lesion<br />

nematode (Thompson 2008) separately. This paper reports an<br />

experiment aimed to develop a single‐plant method for<br />

assessing resistance to both crown rot and P. thornei, which<br />

would be very valuable for accelerated wheat breeding.<br />

MATERIALS AND METHODS<br />

Eleven reference wheat cultivars for crown rot (susceptible<br />

Puseas and Vasco, moderately susceptible Hartog, and partially<br />

resistant Gala and 2–49), and for P. thornei (susceptible Gatcher,<br />

Batavia and Cunningham, and partially resistant GS50a, QT9048<br />

and Yallaroi) were tested. Five replicate 67 mm square pots<br />

received 430 g of steam‐sterilised clay‐loam soil moistened to<br />

37.5% moisture (‐0.1 bar), and 10 seeds of each cultivar were<br />

placed on top. Seed was covered with 100 g dry soil (5%<br />

moisture), then 0.3 g of ground barley/wheat seed colonised<br />

with F. pseudograminearum was added followed by 30 g dry soil.<br />

After 7 days in a glasshouse at 25°C, top watering of the pots to<br />

37.5% moisture was commenced. After 3 weeks, soil was<br />

washed away from the seedlings and the first three leaf sheaths<br />

were rated for crown rot symptoms on a 1 to 4 scale and<br />

summed (max score = 12).<br />

After the crown rot test, four plants from each pot, with roots<br />

trimmed to 3 cm, were planted individually in pots of 330 g<br />

steamed vertosolic soil (Irving Series), watered and inoculated<br />

with a suspension of P. thornei to provide 10,000/kg soil. The<br />

plants were placed in a growth room for 4 days after which<br />

permanently wilted leaf tissue was cut off. <strong>Plant</strong>s were then<br />

grown in a glasshouse with temperature at 22°C and constant 2<br />

cm soil water tension (85% moisture). The pots received three<br />

drenches with 0.1% (w/v) benlate over 6 weeks to prevent<br />

crown rot developing further. After 16 weeks, a 150 g subsample<br />

of soil and roots from the bottom half of the pots was extracted<br />

for nematodes in Whitehead trays. P. thornei were counted in a<br />

1‐ml Hawksley slide under a compound microscope and<br />

expressed as number/kg soil and transformed by ln(x+c) for<br />

analysis of variance.<br />

RESULTS<br />

The crown rot standard cultivars performed as expected, with 2–<br />

49 and Gala relatively resistant, and Hartog, Vasco and Puseas of<br />

increasing susceptibility (Fig. 1). All P. thornei standard cultivars<br />

were relatively susceptible to crown rot. Results for P. thornei<br />

are given in Fig. 2 in log units. Backtransformed values ranged<br />

from 16,150 P. thornei/kg soil for QT9048 to 136,380 for Puseas.<br />

The standard cultivars for P. thornei performed as expected with<br />

GS50a, QT9048 and Yallaroi being relatively resistant, and<br />

Batavia, Gatcher and Cunningham being relatively susceptible to<br />

P. thornei. Hartog produced intermediate numbers of P. thornei<br />

as expected from previous experiments. All of the other crown<br />

rot standard cultivars were relatively susceptible to P. thornei.<br />

Figure 1. Crown rot ratings of seedlings<br />

Figure 2. Number of Pratylenchus thornei after 16 weeks<br />

DISCUSSION<br />

This initial experiment shows it is possible to screen individual<br />

plants for resistance to both crown rot and P. thornei. The<br />

approach taken was first to test plants for crown rot by a<br />

standard method, then transplant them for a nematode<br />

resistance test. This was not ideal in that the plants suffered<br />

considerable transplanting stress and the method was labour<br />

intensive. Despite this, meaningful results were obtained for<br />

resistance to P. thornei. Modification of the methods should be<br />

possible to obtain an effective single plant test for resistance to<br />

both crown rot and P. thornei without the need to transplant.<br />

ACKNOWLEDGEMENTS<br />

We thank M. Brady, T. Bull, T. Clewett, S. Coverdale, M. Davis, M.<br />

Harris, N. Seymour, J. Sheedy, K. Trackson, G. Wildermuth and J.<br />

Wood for their input to this study.<br />

REFERENCES<br />

Thompson JP (2008) Resistance to root‐lesion nematodes (Pratylenchus<br />

thornei and P. neglectus) in synthetic hexaploid wheats and their<br />

durum and Aegilops tauschii parents. Australian Journal of<br />

Agricultural Research 59, 432–446.<br />

Wildermuth GB, McNamara RB (1994) Testing wheat seedlings for<br />

resistance to crown rot caused by Fusarium graminearum Group I.<br />

<strong>Plant</strong> Disease 78, 949–953.<br />

© State of Queensland, Department of Employment, Economic<br />

Development and Innovation, 2009<br />

212 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


40 A single plant test for resistance to two species of root‐lesion nematodes and<br />

yellow spot in wheat<br />

Posters<br />

J.P. Thompson{ XE "Thompson, J.P." } A , T.G. Clewett A , J.G. Sheedy A, , S.H. Jones A and P.M. Williamson B<br />

A DEEDI, Primary Industries and Fisheries, Leslie Research Centre, GPO Box 2282, Toowoomba, 4350, Queensland<br />

B 30 Rhyde Street, Mount Lofty, Toowoomba, 4350, Queensland<br />

INTRODUCTION<br />

Root‐lesion nematodes (Pratylenchus thornei and P. neglectus)<br />

and the stubble‐borne fungal disease yellow spot (Pyrenophora<br />

tritici‐repentis) cause substantial loss of wheat production in the<br />

Australian northern grain region. While some wheat varieties<br />

have partial resistance or tolerance to some of these diseases<br />

none has resistance to all. If varieties could be produced that<br />

combine resistance to all these diseases then the savings to the<br />

wheat industry would be very great.<br />

rating of wheat lines. It was probably due to the changed growth<br />

conditions for conducting the yellow spot test resulting in less<br />

reproduction compared with the plants kept in the glasshouse.<br />

Phenotyping for multiple diseases on a single plant could be a<br />

valuable method for rapidly breeding multiple disease resistant<br />

wheat varieties. A method has been developed to test for<br />

resistance to P. thornei and P. neglectus simultaneously (Huang<br />

et al. 2005) and in this study we extend it to include yellow spot.<br />

MATERIALS AND METHODS<br />

Forty‐one wheat cultivars were subjected to three inoculation<br />

treatments (i) root‐lesion nematodes (P. thornei and P.<br />

neglectus), (ii) yellow spot, and (iii) root‐lesion nematodes and<br />

yellow spot together. Five replicates were grown as single plants<br />

in individual pots of 330 g of steam‐sterilised vertosolic soil.<br />

Nematode inoculum of the two species was produced<br />

separately, and mixed in suspension prior to inoculating<br />

5,000/kg soil of each species at sowing. The plants were grown<br />

in a glasshouse with soil maintained at 22°C and 2 cm water<br />

tension. At the 2‐leaf stage, the seedlings of the yellow spot<br />

treatment were spray inoculated with field‐collected Pyr. triticirepentis<br />

conidia (0.45 mg conidia/mL). Inoculated seedlings were<br />

held in a mist chamber for 40 h, and then another 4 days in a<br />

growth room with sprinklers operating for 3 mins every 12 hrs,<br />

and temperature at 23.5°C. The plants were rated for combined<br />

chlorosis and necrosis of the leaves on a 1 (susceptible) to 9<br />

(resistant) scale. All pots were returned to the glasshouse and<br />

laid out in a split block design. After 16 weeks from sowing,<br />

nematodes were extracted from the soil and roots by the<br />

Whitehead tray method. Pratylenchus thornei and P. neglectus<br />

were identified on morphology and counted under a compound<br />

microscope. Nematode numbers [after transformation by<br />

ln(x+c)] and yellow spot ratings were analysed by ANOVA. Mean<br />

values of the 41 wheat lines were used in regression analyses.<br />

RESULTS AND DISCUSSION<br />

There was good discrimination between wheat lines for yellow<br />

spot ratings (P < 0.001) and a highly significant regression<br />

relationship (P < 0.001) between yellow spot ratings in the<br />

presence and absence of Pratylenchus inoculum (Fig. 1). This<br />

indicated that a systemic resistance was not induced and that<br />

yellow spot resistant and susceptible wheats could be reliably<br />

identified in the presence of the nematodes.<br />

The wheat cultivars inoculated with yellow spot or not were<br />

ranked similarly for P. thornei resistance (R 2 = 0.7756 , , P < 0.001)<br />

or P. neglectus (R 2 = 0.484, P < 0.001) or for total Pratylenchus<br />

(R 2 = 0.7738, P < 0.001) (Fig. 2). Numbers of P. thornei and P.<br />

neglectus in the treatment also tested for yellow spot resistance<br />

were significantly lower than in the nematode only treatment.<br />

This effect was not correlated with the yellow spot resistance<br />

Figure 1. Highly significant regression relationship between ratings for<br />

yellow spot of 41 wheat lines when tested either without (x axis) or with<br />

Pratylenchus (y axis).<br />

Figure 2. Highly significant regression relationship between number of<br />

Pratylenchus produced by 41 wheat lines when tested either without (x<br />

axis) or with yellow spot (y axis)<br />

These results indicate that simultaneous testing of single plants<br />

for resistance to P. thornei, P. neglectus and yellow spot is<br />

feasible. With further refinement this method could improve the<br />

efficiency of breeding multiple‐disease resistant wheats which<br />

would be of great value to Australia.<br />

ACKNOWLEDGEMENTS<br />

We thank Megan Brady for technical assistance.<br />

REFERENCES<br />

Huang, R, Thompson, JP, Sheedy, JS, Reen, RA and Brady, MJ (2005) Dual<br />

phenotyping of wheat breeding lines for resistance to root‐lesion<br />

nematodes. 15th Biennial <strong>Plant</strong> <strong>Pathology</strong> Conference Handbook’,<br />

Geelong. p. 110.<br />

© State of Queensland, Department of Employment, Economic<br />

Development and Innovation, 2009<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 213


Posters<br />

90 Role of nematodes and zoosporic fungi in poor growth of winter cereals in the<br />

northern grain region<br />

J.P. Thompson{ XE "Thompson, J.P." }, T.G. Clewett , J.G. Sheedy and K.J. Owen<br />

DEEDI, Qld Primary Industries and Fisheries, Leslie Research Centre, PO Box 2282, Toowoomba, 4350, Qld<br />

INTRODUCTION<br />

The endoparasitic root‐lesion nematodes Pratylenchus thornei<br />

and P. neglectus and the ectoparasitic stunt nematode Merlinius<br />

brevidens occur widely in the northern grain region of Australia.<br />

Grain loss in wheat has been well characterised for P. thornei in<br />

the northern region (Thompson et al 2008) and for P. neglectus<br />

in the southern and western regions (Vanstone et al, 2008).<br />

Information on the role of Merlinius brevidens is sparse although<br />

it has been shown to cause yield loss of wheat in the USA (Smiley<br />

et al. (2006), particularly when associated with the zoosporic<br />

fungus Olpidium brassicae (Langdon et al.1961). Following<br />

diagnosis of high populations of M. brevidens associated with<br />

poor crops of winter cereals in 2007 a glasshouse experiment<br />

was conducted in 2008 to explore further the reasons for the<br />

poor growth<br />

MATERIALS AND METHODS<br />

About 50 kg of soil was collected (on a grid of 36 positions within<br />

a 1,920 m 2 area) from each of nine fields in the northern grain<br />

region located from Garah in northern NSW to Wondai in Qld.<br />

These sites were selected on the basis that M. brevidens and/or<br />

Olpidium sp. had been detected in poorly growing cereals in the<br />

field or on the farm previously. The soil from each site was<br />

mixed, sieved and about half was partially sterilised by steam at<br />

70ºC for 45 min. Quantities of each soil (330 g OD equivalent)<br />

were mixed with 1 g of Osmocote® [Native Gardens plus<br />

micronutrients (17–1.6–8.7 NPK)] slow‐release fertiliser and<br />

placed in 5 cm‐square plastic pots suitable for bottom watering.<br />

Eighteen pots of each of sterilised and unsterilised soil were<br />

prepared to allow for growing 3 replicates of 3 cereals (wheat cv.<br />

Strzelecki, barley cv. Grout and oats cv. Coolibah) at 2 moisture<br />

tensions (2 cm and 7 cm) and 2 harvest times (8 and 16 wks).<br />

The pots were placed on strips of capillary matting for each soil<br />

type and on separate benches for sterilised and unsterilised soil<br />

and for the two water tensions. A single plant per pot was<br />

grown, with the glasshouse temperature kept below 25ºC by<br />

evaporative coolers. At each harvest, plant tops were dried at<br />

85ºC for 4 days and weighed. Soil and roots were removed from<br />

the pots, photographed and split longitudinally. Roots were<br />

extracted from one half of the pots, blotted, weighed, and a<br />

subsample stained with trypan blue. Soil and roots from the<br />

other half were broken into pieces


41 Sources of resistance to root‐lesion nematode (Pratylenchus thornei) in wheat<br />

from West Asia and North Africa<br />

Posters<br />

J.P. Thompson{ XE "Thompson, J.P." }, T.G. Clewett and M.M. O’Reilly<br />

DEEDI, Primary Industries and Fisheries, Leslie Research Centre, GPO Box 2282, Toowoomba, 4350, Queensland<br />

INTRODUCTION<br />

The root‐lesion nematode Pratylenchus thornei occurs widely in<br />

the northern grain region (northern NSW and southern and<br />

central Qld) causing considerable economic loss in wheat<br />

production. Current management tools are hygiene with farm<br />

machinery to prevent transfer of infested soil, crop rotation and<br />

growing tolerant wheat varieties (Thompson et al. 2008). More<br />

effective control of the nematode populations could be achieved<br />

if resistant cultivars were available.<br />

Experiment 3. Thirteen WANA bread wheats (Fig. 1) and 10<br />

durum wheats (data not shown) had P. thornei numbers that did<br />

not differ from GS50a in two experiments. All Australian bread<br />

wheats were susceptible whereas the two Australian durums<br />

were as resistant as GS50a.<br />

To obtain novel sources of resistance we tested two collections<br />

of wheat from the West Asia and North Africa (WANA) region.<br />

The A.E. Watkins Collection was made in Cambridge, UK, in the<br />

late 1920s and early 1930s with landrace wheats from many<br />

countries of the world (Miller et al. 2001). The R.A. McIntosh<br />

Collection was made in 1993 at University of Sydney with wheats<br />

from WANA countries for studies on rusts and flag smut<br />

MATERIALS AND METHODS<br />

Wheat Accessions. The WANA wheats tested comprised 148<br />

bread wheat (Triticum aestivum) and 139 durum wheat (Triticum<br />

turgidum spp. durum) accessions from the Watkins Collection<br />

and 59 bread and 43 durum accessions from the McIntosh<br />

Collection.<br />

Resistance Experiments. Initially each of the above collections<br />

was tested for resistance to P. thornei in two separate<br />

glasshouse experiments that included the reference standards<br />

GS50a (a partially resistant bread wheat) and three susceptible<br />

wheat varieties Gatcher, Suneca and Potam. A number of bread<br />

and durum wheat accessions that produced nematode numbers<br />

not significantly different from GS50a were retested for<br />

resistance in a third experiment.<br />

Resistance test methods. The wheat accessions were grown as 3<br />

replicates in pots of steamed vertosolic soil (1 kg soil in<br />

Experiments 1 and 2 and 650 g in Experiment 3), inoculated with<br />

P. thornei at a rate of 2,500/kg soil. The soil was fertilised to<br />

provide N, P, K, S, Ca and Zn and was watered to pF2 (56%<br />

moisture). The experiments were laid out in randomised blocks<br />

in an evaporatively cooled glasshouse. In Experiment 3, the soil<br />

and root temperature was kept at 22°C with pots in glasshouse<br />

waterbaths. After 16 weeks growth, one half of the soil and<br />

roots was removed, broken to < 1 cm manually and 150 g<br />

extracted for nematodes in Whitehead trays. Nematodes were<br />

counted under a microscope and expressed as P. thornei/kg soil<br />

(oven dry equivalent). Data were transformed by ln(x+1) for<br />

ANOVA and calculation of Fl.s.d. Backtransformed means and<br />

reproduction factors (RF = final number of P. thornei/ initial<br />

number) were calculated.<br />

Figure 1. Reproduction factor of WANA bread wheats (black bars) that<br />

did not differ significantly from GS50a in comparison with reference<br />

standard wheats (open bars) from the northern grain region. The letters<br />

B and D after names indicate bread and durum wheats respectively.<br />

The identification of additional sources of resistance in bread<br />

wheat to P. thornei provides greater options for producing<br />

Australian wheat varieties with greater levels of resistance to P.<br />

thornei than in current varieties.<br />

ACKNOWLEDGEMENTS<br />

We thank Michael Mackay and Greg Grimes of Australian Winter<br />

Cereals Collection, and Professor Bob McIntosh University of<br />

Sydney or provision of seed<br />

REFERENCES<br />

1. Thompson JP, Owen KJ, Stirling GR, Bell MJ (2008) Root‐lesion<br />

nematodes (Pratylenchus thornei and P. neglectus): a review of<br />

recent progress in managing a significant pest of grain crops in<br />

northern Australia. <strong>Australasian</strong> <strong>Plant</strong> <strong>Pathology</strong> 37, 235–242.<br />

2. Miller TE, Ambrose MJ, Reader SM (2001) The Watkins Collection.<br />

In ‘Wheat Taxonomy: The Legacy of John Percival’. (Ed. PDS Caligari<br />

and PE Brandham) pp. 113–120. (Academic Press London).<br />

© State of Queensland, Department of Employment, Economic<br />

Development and Innovation, 2009<br />

RESULTS AND DISCUSSION<br />

Experiments 1 and 2. As a group, the bread wheats were<br />

significantly (P < 0.001) more susceptible to P. thornei than the<br />

durum wheats with backtransformed means for P. thornei/kg<br />

soil of 52,051 for bread wheats and 38,560 for durum wheats in<br />

the Watkins Collection, and 34,200 for bread wheats and 21,268<br />

for durum wheats in the McIntosh Collection.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 215


Posters<br />

89 Sugarcane downy mildew: development of molecular diagnostics<br />

N. Thompson{ XE "Thompson, N." } and B.J. Croft<br />

BSES Limited, PO Box 86, Indooroopilly, 4068, QLD<br />

INTRODUCTION<br />

Sugarcane downy mildew (SDM) is currently one of the most<br />

serious diseases at the Ramu sugarcane plantation in Papua New<br />

Guinea. SDM is rated as a high priority pathogen for the<br />

Australian sugar industry.<br />

SDM can be caused by several fungal species in the genus<br />

Peronosclerospora including P. sacchari, P. spontanea and P.<br />

philippinensis (1). P. sacchari was eradicated from Australia<br />

during the 1950s by resistant varieties and roguing of remaining<br />

infected plants (2). It is estimated that 60% of current<br />

commercial sugarcane varieties in Australia are intermediate or<br />

susceptible to SDM.<br />

The species of Peronosclerospora are difficult to distinguish by<br />

traditional taxonomy, so molecular diagnostics would be useful<br />

to conclusively identify the species during a disease incursion.<br />

Peronosclerospora diagnostics using both hybridisation (3) and<br />

PCR (4) have been developed in the USA; however the number<br />

of isolates available was limited in these studies (D. Luster, pers.<br />

comm.). Therefore molecular diagnostics need to be verified on<br />

a larger sample size of Peronosclerospora, including Australian<br />

domestic and exotic species.<br />

MATERIALS AND METHODS<br />

Target genes and isolates. DNA sequences from isolates held at<br />

USDA‐ARS (Ft Detrick, MD) were used to develop primers for<br />

regions that showed variation (Table 1). Species used to design<br />

and test the primers are shown in Table 2.<br />

Table 1. Peronosclerospora genes used as primer targets<br />

Gene<br />

ITS1*<br />

Actin*<br />

EF1*<br />

Beta‐tubulin<br />

COX‐1<br />

Predicted outcome<br />

General primers; should amplify all<br />

Designed to amplify all Oomycetes<br />

Variation within and between species; may not be<br />

good for diagnostics<br />

High variation; unsure as to efficacy of primers.<br />

Size differential between species, may not amplify<br />

some species<br />

* designed by Dr Clint Magill, Texas A&M University<br />

(ITS1: ribosomal DNA first internal transcribed region; EF1: Transcription Elongation<br />

Factor 1 alpha; COX‐1: Cytochrome oxidase‐1)<br />

Table 2. Peronosclerospora species used in this study<br />

Species<br />

Detail<br />

P. sacchari Sequence from up to 3 isolates<br />

P. philippinensis Sequence from up to 3 isolates<br />

P. sorghi Sequence from up to 3 isolates<br />

P. maydis Sequence from up to 6 isolates<br />

P. sacchari 2 herbarium isolates<br />

P. sorghi 1 herbarium isolate.<br />

P. eriachloae 1 herbarium isolate<br />

P. noblei 1 herbarium isolate<br />

P sp.<br />

2 unknown herbarium isolates<br />

ARS. Peronosclerospora isolates from the QDPI&F <strong>Plant</strong><br />

<strong>Pathology</strong> Herbarium were used to test efficacy of primers.<br />

Analysis. DNA was extracted from isolates using QIAGEN DNeasy<br />

kit. PCR was done using primers developed for the genes of<br />

interest, and the results analysed on agarose gels.<br />

RESULTS AND DISCUSSION<br />

Amplification with ITS‐1 was used as a positive control to ensure<br />

that amplifiable DNA was obtained after extraction.<br />

Unfortunately, one of the P. sacchari herbarium isolates was not<br />

amplifiable. Further investigation is needed to test this isolate.<br />

The general Oomycete primers (actin) gave amplified products in<br />

some, but not all species. This is unexpected, because this<br />

primer set was designed to amplify products from all<br />

Oomycetes.<br />

The primers for EF1 were tested and a high degree of variation<br />

observed, with complex banding patterns between all species<br />

tested. This primer set is likely to not be useful for diagnostics<br />

unless further sequencing reveals diagnostic differences.<br />

Beta‐tubulin primers showed differential amplification between<br />

species, however some unexpected complex banding was<br />

observed. Further investigation of this primer set is also<br />

required.<br />

The Cox‐1 primer set was designed to show differences between<br />

species, and it showed promise as a diagnostic, however the<br />

amplification of P. sacchari remains inconsistent.<br />

Future work is planned to obtain more isolates of the causal<br />

agents of SDM to troubleshoot and refine the molecular<br />

diagnostic test.<br />

ACKNOWLEDGEMENTS<br />

Thanks to: Dr Clint Magill, Dr Ram Perumal (Texas A&M<br />

University) and Dr Doug Luster (USDA‐ARS, Ft Detrick) for initial<br />

sequencing information; to Dr Roger Shivas (QDPI&F <strong>Plant</strong><br />

<strong>Pathology</strong> Herbarium) for isolates and collaboration. This project<br />

was partially funded by an SRDC Travel and Learning<br />

Opportunity scholarship.<br />

REFERENCES<br />

1. Suma, S and Magarey, R.C (2000) Downy Mildew in “A guide to<br />

Sugarcane Diseases” (Eds Rott, P., Bailey, R.A., Comstock, J.C.,<br />

Croft, B.J. and Saumtally, A.S.) pp 90–95 (CIRAD and ISSCT)<br />

2. Hughes, C.G. (1957) The eradication of Downy mildew. Cane Pest<br />

and Disease Control Boards’ Conference Minutes and Proceedings,<br />

pp 29.<br />

3. Yao, C‐L, Magill, C.W, Frederiksen, R.A., Bonde, M.R., Wang, Y. and<br />

Wu, P‐S. (1991) Detection and Identification of Peronosclerospora<br />

sacchari in Maize by DNA Hybridization. Phytopathology 81(8) 901–<br />

905.<br />

4. Yao, C‐L, Magill, C.W and Frederiksen, R.A (1992) Length<br />

heterogeneity in ITS 2 and the methylation status of CCGG and<br />

GCGC sites in the rRNA genes of the genus Peronosclerospora.<br />

Current Genetics 22:415–420.<br />

Amplification. Theoretical amplification was done using<br />

Peronosclerospora gene sequences from isolates held at USDA‐<br />

216 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


43 Effect of irrigation method on disease development in a carrot seed crop<br />

R.S. Trivedi{ XE "Trivedi, R.S." } 1 , J.M. Townshend 2 , M.V. Jaspers 3 , H.J. Ridgway 3 , and J.G. Hampton 1<br />

1 Bio‐Protection Research Centre and 3 Department of Ecology, PO Box 84, Lincoln University, Canterbury, New Zealand<br />

2 Midlands Seed Ltd, PO Box 65, Ashburton, Canterbury, New Zealand<br />

Posters<br />

INTRODUCTION<br />

For carrot (Daucus carotae L.) seed production in Canterbury,<br />

New Zealand, overhead irrigation is commonly used from mid‐<br />

November until March to prevent soil moisture deficits.<br />

However this irrigation method may assist the spread of plant<br />

pathogens such as Alternaria radicina/A. daucii/Cercospora<br />

carotae within the crop. We compared the effects of overhead<br />

irrigation and drip (T‐Tape) irrigation on disease level in a carrot<br />

seed crop in Canterbury.<br />

MATERIALS AND METHODS<br />

A field assessment of the effect of irrigation method (overhead<br />

vs. drip) was made in the 2007/08 season. An 8 ha field was<br />

divided into two; 4 ha was irrigated via a conventional overhead<br />

system, and 4 ha was irrigated via a drip (T‐Tape) system. The<br />

drip lines were dug 3–4 cm below the soil surface in early<br />

November. Soil moisture was monitored using a neutron probe<br />

and irrigation applied as required. Irrigation treatments were not<br />

replicated, but within each 4 ha block, twelve 30 m 2 plots were<br />

selected at random and used for three assessment of foliage<br />

disease symptoms using a 1 to 10 rating scale where 1= no<br />

symptoms and 10= dead plant and an assessment of root disease<br />

symptoms using a 0 to 4 rating scale where 0= no evidence of<br />

root infection and 4= shoulder region of root completely girdled<br />

with black rot. At maturity ten primary umbels were hand<br />

harvested from each plot and hand threshed. Seeds were then<br />

dried at 30°C to bring the seed moisture level down to 8%, hand<br />

rubbed to remove spines and thoroughly mixed. Hands were<br />

sterilised with 90% ethanol to prevent contamination while<br />

handling different seed samples. For seed infection, 100 carrot<br />

seeds per plot were plated onto a semi selective agar media for<br />

A. radicina, and incubated at 27°C in 24 h dark for 14 days. For<br />

seed germination, 100 seeds per plot were placed between<br />

moist germination paper towels and incubated at 20°C in 24 h<br />

dark for 14 days before evaluation. The collected data were<br />

statistically analysed using an unpaired t‐ test.<br />

RESULTS<br />

The carrot plants from drip irrigated blocks had significantly less<br />

foliage infected by A. radicina/A. daucii/Cercospora carotae at all<br />

three assessment times (Fig 1, P


Posters<br />

91 Pathogenicity of Radopholus similis on ginger in Fiji<br />

U. Turaganivalu{ XE "Turaganivalu, U." } A , G.R. Stirling B , S. Reddy A and M. K. Smith C<br />

A Fiji Ministry of Agriculture, Box 77, Nausori, Fiji<br />

B Biological Crop Protection Pty Ltd, 3601 Moggill Rd, Moggill,, Queensland<br />

C Department of Primary Industries and Fisheries, PO Box 5083, Nambour, Queensland<br />

INTRODUCTION<br />

Radopholus similis was first associated with a disease of ginger<br />

(Zingiber officinale) in the early 1970s, when stunted, chlorotic,<br />

low yielding crops in Fiji were found to be infested with the<br />

nematode (1). Nematodes were observed in small, shallow,<br />

water soaked lesions on the rhizome surface, and these lesions<br />

eventually enlarged until the rhizome was destroyed.<br />

Although R.similis was considered primarily responsible for the<br />

symptoms observed, secondary organisms were also thought to<br />

be involved (1). This study sought to confirm this by examining<br />

the pathogenicity of the nematode on ginger in a more<br />

controlled environment.<br />

MATERIALS AND METHODS<br />

Twenty 4 L pots were filled with autoclaved potting mix and<br />

planted with a Radopholus‐free ‘seed piece’ of ginger (a section<br />

of rhizome used as planting material). Pots were then<br />

transferred to a glasshouse and 6 weeks later, half the pots were<br />

inoculated with 1,500 R. similis. The nematode was obtained<br />

from a ginger farm at Veikoba, Fiji and had been multiplied in<br />

the laboratory on sterile carrot tissue. Fifteen and 20 weeks after<br />

pots were inoculated, the number of yellowing or dead shoots in<br />

each pot was recorded, above‐ground biomass in five inoculated<br />

and five control pots was measured and symptoms on seed<br />

pieces and newly‐developing rhizomes were assessed.<br />

Nematodes were extracted by spreading macerated rhizomes or<br />

200 mL samples of soil and roots on an extraction tray and<br />

recovering them on a 38µm sieve.<br />

Portions of seed pieces and rhizomes showing symptoms<br />

possibly caused by R. similis were assessed by removing small<br />

pieces of tissue from affected areas, macerating them in water<br />

and checking for nematodes after 24 hours. Discoloured tissue<br />

was also checked for fungal pathogens known to cause<br />

discoloration or rotting of ginger (i.e. Fusarium and Pythium) by<br />

placing small pieces of tissue onto potato dextrose agar (PDA) or<br />

corn meal agar with carbendazim, ampicillin, rifampicin,<br />

pentachloronitrobenzene and pimaricin (CARPP), and observing<br />

plates after 24 and 48 hours.<br />

RESULTS<br />

Non‐inoculated plants grew normally and after 15 weeks they<br />

had several healthy green shoots up to 90 cm long. In contrast,<br />

the stem bases and lower leaves of 6 of the 10 inoculated plants<br />

were yellow and in some cases the affected shoots had died.<br />

Yellowing was first observed about 12 weeks after inoculation<br />

and at 15 weeks, 40% of shoots were chlorotic or dead. By 20<br />

weeks, three inoculated plants had died, shoots on the<br />

remaining plants were dying back, rhizomes and seed pieces<br />

were discoloured or badly rotted and plant biomass was<br />

significantly reduced relative to the control (Table 1).<br />

into the rhizome, and from seed pieces. In some cases, the<br />

nematode population in parts of a seed piece or rhizome was as<br />

high as 500 R. similis/g. Fusarium, Pythium or other fungi were<br />

not isolated from discoloured tissues. Estimates of the number<br />

of R. similis recovered after 20 weeks indicated that significant<br />

nematode multiplication had occurred in roots, seed pieces and<br />

rhizomes (Table 2).<br />

Table 1. Effect of Radopholus similis on ginger 20 weeks after plants<br />

growing in potting mix were either inoculated with the nematode or left<br />

uninoculated<br />

Treatment Dry wt. shoots (g) Fresh wt.<br />

seed piece (g)<br />

Fresh wt.<br />

rhizome (g)<br />

Control 79.2 a 52.8 a 54.4 a<br />

R. similis 7.5 b 23.7 b 24.5 b<br />

Numbers in the same column followed by different letters are significantly different<br />

(P= 0.05)<br />

Table 2. Numbers of Radopholus similis recovered 20 weeks after ginger<br />

plants were inoculated with 1,500 nematodes or left uninoculated<br />

No. R. similis females<br />

/seed<br />

piece<br />

/rhizome<br />

Control 0 0 0<br />

/pot<br />

(soil + roots)<br />

R. similis 397 884 15,628<br />

DISCUSSION<br />

Our results suggest that R. similis multiplies initially on seed<br />

pieces and roots and then invades newly‐developing rhizomes.<br />

The nematode first seems to feed on outer parts of the rhizome<br />

and the resulting damage leads to yellowing at the base of<br />

shoots and of the lower leaf sheath. As more tissues are<br />

destroyed, older leaves turn yellow, shoots eventually collapse<br />

and discoloration extends further into the rhizome. The end<br />

result is that plants eventually die and the rhizome is totally<br />

destroyed.<br />

Given the results of our experiment, we suggest that R. similis is<br />

a pathogen of ginger in its own right. The nematode is capable of<br />

killing plants and destroying rhizomes, with secondary organisms<br />

playing little role in symptom development.<br />

ACKNOWLEDGEMENTS<br />

Funding from ACIAR is gratefully acknowledged.<br />

REFERENCES<br />

1. Vilsoni F, McClure MA, Butler LD (1976). Occurrence, host range<br />

and histopathology of Radopholus similis in ginger (Zingiber<br />

officinale). <strong>Plant</strong> Disease Reporter 60: 417–420.<br />

Observations on tissue collected from affected plants showed<br />

that R. similis was present in the lowest leaf sheaths, in the collar<br />

at the base of the shoot, and in rhizome tissue at the point<br />

where shoots emerged from the rhizome. The nematode was<br />

also recovered from sunken lesions and blackened tissue on the<br />

rhizome surface, from discoloured tissue that extended 1–3 mm<br />

218 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


17 Survival of the pistachio dieback bacterium in buried wood<br />

T.A. Vu Thanh{ XE "Vu Thanh, T.A." } A , D. Giblot‐Ducray A , M.R. Sosnowski B,C and E.S. Scott A,C<br />

A School of Agriculture, Food and Wine, The University of Adelaide, Waite Campus, PMB 1, Glen Osmond, 5064 SA<br />

B South Australian Research and Development Institute, GPO Box 397, Adelaide, 5001 SA<br />

C Cooperative Research Centre for National <strong>Plant</strong> Biosecurity, LPO Box 5012, Bruce, 2617 ACT<br />

Posters<br />

INTRODUCTION<br />

Pistachio dieback is a bacterial disease causing internal staining,<br />

trunk and limb lesions, decline, dieback and, in some instances,<br />

death of pistachio trees (1). Apparently endemic in Australia, the<br />

causal agent is a strain of Xanthomonas translucens. It is a<br />

vascular pathogen that provides a local model to assess the<br />

effectiveness of existing eradication strategies for systemic<br />

bacterial pathogens of woody perennials which are considered<br />

high‐priority emergency plant pests in Australia. Burial and<br />

burning are two accepted means of disposal of diseased plant<br />

material. However, there is little or no information on the<br />

survival of bacterial pathogens following burial or burning of<br />

infected wood. The efficacy of burial and burning as means of<br />

safe disposal of diseased wood is being evaluated. This paper<br />

reports the survival of the pistachio dieback bacterium in buried<br />

wood to date.<br />

MATERIALS AND METHODS<br />

Infected pistachio wood was collected from the Waite Campus<br />

orchard in August 2008. Staining was assessed and the presence<br />

of X. translucens was confirmed by culturing on antibiotic<br />

benlate sucrose peptone agar (ABSPA, 1) and by polymerase<br />

chain reaction (PCR) using strain‐specific primers (2).<br />

Methods adapted from those described by Naseri et al. (3) were<br />

used. Plastic mesh bags containing pieces of infected wood or<br />

mulched pistachio wood (approx 20 g each) were buried 10 cm<br />

deep in pots filled with orchard soil or left on the soil surface.<br />

Pots were placed outdoors in August 2008 and retrieved<br />

monthly to assess survival of X. translucens. Upon retrieval, each<br />

piece of wood was cut into half. One half was surface‐sterilised<br />

and bacterial suspensions obtained by soaking in sterile distilled<br />

water overnight at room temperature. Suspensions were<br />

streaked on several culture media amended with various<br />

antibiotics and incubated at 28°C for 2–8 days. Suspensions<br />

prepared for culturing were also assayed by PCR using strainand<br />

species‐specific primers.<br />

The remaining wood samples were sliced into pieces (1–1.5 mm<br />

thick), surface sterilised and placed directly onto V8 juice agar<br />

amended with streptomycin for fungal isolation and nutrient<br />

agar (NA, Oxoid) amended with benlate for bacterial isolation.<br />

Plates were incubated at 25°C for 3–5 days, then bacteria and<br />

fungi were subcultured onto NA and PDA, respectively. To screen<br />

for antagonism, a suspension of X. translucens was spread on<br />

sucrose peptone agar plates an hour before applying to the<br />

centre of each plate a small amount of bacteria isolated from the<br />

wood. Plates were incubated at 28°C and observed for 3–7 days.<br />

RESULTS<br />

X. translucens was detected by PCR with species‐specific primers<br />

in most wood samples up to 7 months after burying in soil or<br />

placing on the soil surface. X. translucens was rarely detected by<br />

culturing, as plates were often over‐<br />

selected as the optimal culture medium for subsequent isolation<br />

of X. translucens from buried wood.<br />

A number of bacteria isolated from wood, buried or left on the<br />

soil surface, produced clear inhibition zones and some prevented<br />

the growth of X. translucens. Fungi isolated from wood samples<br />

have been stored for future use.<br />

DISCUSSION<br />

Viable X. translucens has been detected in pistachio wood buried<br />

for 7 months. The burial experiment will continue, to examine if<br />

X. translucens can survive in buried infected wood for up to 2<br />

years. In comparison, X. campestris pv. campestris survived for<br />

507 days in cabbage stem residues buried in soil (4).<br />

Inhibition of the growth of X. translucens on agar by bacteria<br />

isolated from the wood may explain the difficulty in isolating X.<br />

translucens from wood. Antibiotic activity of these bacteria is<br />

being assessed. Another explanation for the difficulty in isolating<br />

X. translucens from the wood might be that X. translucens enters<br />

into the viable but non‐culturable (VBNC) state in response to<br />

the environmental conditions during burial. This possibility is<br />

being examined.<br />

Frequently isolated fungi will be assessed for their ability to<br />

decompose pistachio wood.<br />

A preliminary trial conducted in August 2008 to assess burning<br />

for disposal of diseased wood yielded no viable X. tranlucens<br />

from ash or wood buried 5 cm below the pit surface. However,<br />

eradication from debris which penetrates the floor of the pit<br />

may depend on that wood reaching the lethal temperature for<br />

the bacterium. Further experiments are planned for winter 2009<br />

to study the effect of heat and burning on survival of X.<br />

translucens.<br />

REFERENCES<br />

1. Facelli E, Taylor C, Scott ES, Fegan M, Huys G, Emmett R, Noble D,<br />

Sedgley M (2005) Identification of the causal agent of pistachio<br />

dieback in Australia. European Journal of <strong>Plant</strong> <strong>Pathology</strong> 112, 155–<br />

165.<br />

2. Marefat A, Ophel‐Keller K, Scott ES and Sedgley M (2006) The use<br />

of ARMS PCR in detection and identification of xanthomonads<br />

associated with pistachio dieback in Australia. European Journal of<br />

<strong>Plant</strong> <strong>Pathology</strong> 116, 57–68.<br />

3. Naseri B, Davidson JA, Scott ES (2008) Survival of Leptosphaeria<br />

maculans and associated mycobiota on oilseed rape stubble buried<br />

in soil. <strong>Plant</strong> <strong>Pathology</strong> 57, 280–289.<br />

4. Schultz T, Gabrielson RL (1986) Xanthomonas campestris pv.<br />

campestris in Western Washington crucifer seed fields: occurrence<br />

and survival. Phytopathology 76, 1306–1309.<br />

grown by soil and/or wood microorganisms. However, X.<br />

translucens was isolated on NA amended with ampicillin,<br />

cephalexin and gentamycin from some wood and mulch samples<br />

after 8 months buried in soil. Nutrient agar with antibiotics was<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 219


Posters<br />

92 The effect of dryland salinity on the diversity of arbuscular mycorrhizal fungi<br />

B.A. Wilson{ XE "Wilson, B.A." }, G.J. Ash and J.D.I. Harper<br />

EH Graham Centre for Agricultural Innovation (an alliance between New South Wales Department of Primary Industries and Charles Sturt<br />

University), Locked Bag 588, Wagga Wagga, 2678, NSW<br />

INTRODUCTION<br />

Dryland salinity is a significant environmental problem in<br />

Australia. The large‐scale removal of native vegetation and the<br />

sowing of shallow‐rooted pastures and crops is a major<br />

contributor to dryland salinity. Much of the research that aims to<br />

improve land productivity and maintain biodiversity focuses on<br />

above ground biomass. Less research has investigated the<br />

deleterious effects of dryland salinity on soil microorganisms<br />

such as arbuscular mycorrhizal fungi, considered essential for<br />

the establishment and growth of plants. The aim of study was to<br />

investigate the diversity of arbuscular mycorrhizal (AM) fungi<br />

from a saline agricultural field site using denaturing gradient gel<br />

electrophoresis (DGGE).<br />

MATERIALS AND METHODS<br />

Field site and sampling Soil sampling commenced after an<br />

electromagnetic survey (EM) of the saline field site and electrical<br />

conductivity (EC) testing of the soil was complete. The soil<br />

salinity ranged from non‐saline (1600 mS/m) based on plant salinity tolerance<br />

guidelines used in Western Australia (1). The paddock was<br />

divided into 24 plots from which soil cores (5 mm x 40 mm) were<br />

randomly taken. Soil was sampled in the summer and spring of<br />

2007.<br />

Trap cultures Soil removed from each of the 24 plots at both<br />

sampling times was used to set‐up trap cultures using several<br />

salt‐tolerant pasture and grass species. The purpose of the trap<br />

cultures is to provide conditions for maximum sporulation with<br />

the aim to increase diversity compared to that seen in field soil.<br />

Cultures were grown for 5 months before being assessed for<br />

sporulation. Sporulation was so poor for both sets of cultures<br />

that pots were re‐sown and grown for an additional 4 months.<br />

present and to provide a general picture of the diversity of AM<br />

fungi at a dryland salinity‐affected site.<br />

Figure 2. DGGE gel of summer field samples. Lane 1 (M) = ladder, S =<br />

slightly saline, N = non‐saline, V = very saline. One sample is represented<br />

by two lanes except for lane 1 (ladder) and lane 16.<br />

REFERENCES<br />

1.<br />

http://www.agric.wa.gov.au/content/LWE/SALIN/SMEAS/sali<br />

nity_units.htm<br />

2. Schwarzott, D and Schüßler, A (2001) A simple and reliable method<br />

for SSU rRNA gene DNA extraction, amplification, and cloning from<br />

single AM fungal spores. Mycorrhiza 10, 203–207<br />

3. Helgason et al (1998) Ploughing up the wood‐wide web. Nature<br />

394, 431<br />

4. Simon et al (1992) Specific amplification of the 18S fungal<br />

ribosomal genes from vesicular‐arbuscular endomycorrhizal fungi<br />

colonising roots. Applied Environmental Microbiology 58, 291–295.<br />

5. Kowalchuk et al (1997) Detection and characterisation of fungal<br />

infections of Ammophilia arenaria (Marram Grass) roots by<br />

denaturing gradient gel electrophoresis of specifically amplified 18S<br />

rDNA. Applied and Environmental Microbiology 63, 3858–3865.<br />

Spore extraction, DNA extraction and PCR‐DGGE Spores were<br />

extracted from 50 g soil sub‐samples (summer and spring field<br />

soil and trap culture soil) using wet sieving and sucrose<br />

centrifugation and DNA was extracted using the Powersoil DNA<br />

isolation kit (Mo Bio Laboratories Inc, USA). A nested PCR was<br />

used to amplify the partial small subunit gene 18S ribosomal<br />

DNA gene (550 bp). The first reaction employed the universal<br />

fungal primers GeoA2/Geo11 (2) and the second reaction used<br />

the fungal primer AM1 (3) and the universal eukaryotic primer<br />

NS31 (4) with an attached 5’ GC clamp on the NS31 primer<br />

(NS31‐GC) (5) for analysis with DGGE. Gels contained 7% (w/v)<br />

polyacrylamide (37:1 acrylamide/bis‐acrylamide) and the linear<br />

gradient was 30–50% denaturant. Gels were run at a constant<br />

temperature of 60 °C for 16 hrs at 70V using Bio‐rad’s Dcode<br />

Universal Mutation Detection System (Biorad, Australia).<br />

RESULTS AND DISCUSSION<br />

Preliminary results suggest that genetic diversity exists across<br />

the field site irrespective of salinity levels (compare N1‐7). It is<br />

hypothesised that salinity will reduce the genetic diversity of AM<br />

fungi at this field site, given that large salt scalds on the property<br />

have reduced vegetation cover. Further analysis is under way,<br />

which aims to reveal the diversity between soil samples across<br />

the salinity gradient on both a spatial and temporal scale.<br />

Cloning and sequencing will be used to identify the AM fungi<br />

220 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


18 Discovery of a Ceratocystis sp. associated with wilt disease of two native<br />

leguminous tree hosts in Oman and Pakistan<br />

Posters<br />

A.O. Al Adawi A,C* , I. Barnes B , A.A. Al Jahwari C , M.L. Deadman D , B.D. Wingfield B and M.J. Wingfield{ XE "Wingfield, M.J." } A<br />

A Department of Microbiology and <strong>Plant</strong> <strong>Pathology</strong>, Forestry and Agricultural Biotechnology Institute (FABI), University of Pretoria,<br />

Pretoria 0002, South Africa<br />

B Department of Genetics, FABI, University of Pretoria, Pretoria 0002, South Africa<br />

C Ghadafan Agriculture Research Station, Ministry of Agriculture, PO Box 204, Sohar, 311, Sultanate of Oman<br />

D Department of Crop Sciences, PO Box 34, Sultan Qaboos University, Al Khod 123, Sultanate of Oman<br />

INTRODUCTION<br />

Ceratocystis fimbriata Ellis and Halst sensu lato represents a well<br />

recognised group of cryptic species that cause serious canker<br />

stain and vascular wilt diseases on a wide range of mostly woody<br />

hosts. An epidemic wilt disease, devastating thousands of mango<br />

trees, has occurred in Oman and Pakistan since 1998 (1, 2). The<br />

pathogen causing the disease was identified as the new and<br />

cryptic species Ceratocystis manginecans M. van Wyk, A. Al<br />

Adawi and M.J. Wingf., based on morphology and DNA sequence<br />

comparisons (2). Recently, native Ghaf (Prosopis cineraria L.<br />

Druce) trees in Oman began to show symptoms of wilt similar to<br />

those displayed by mangos. Intriguingly, a similar wilt diseases<br />

has been observed on native Shisham (Dalbergia sissoo, Roxb.)<br />

trees in Pakistan. The objective of this study was to identify the<br />

pathogen causing the wilt disease of P. cineraria and D. sissoo in<br />

Oman and Pakistan.<br />

MATERIALS AND METHODS<br />

During 2004–2006, samples from P. cineraria trees showing<br />

recent wilting symptoms were collected from Wilayat Sohar in<br />

the northern region of the Sultanate of Oman. In May 2006,<br />

samples were collected from D. sissoo plantations in Fasilabad,<br />

Shorkot, Chenab negar and Multan, in Pakistan. Wood samples<br />

exhibiting vascular discolouration were placed in moist<br />

chambers and incubated at 25°C for 7 days. Wood samples were<br />

also placed between carrot slices (3) to bait for the possible<br />

presence of Ceratocystis spp.<br />

Morphological observations of the isolated fungi were made<br />

from cultures on 2% Malt Extract Agar (MEA) that had been<br />

incubated for 10 days at 25 ºC. The ITS regions of two isolates<br />

from each of the two host species were amplified and<br />

sequenced. Sequence data were compared to those in the<br />

Genbank database using a blast search<br />

(http://blast.ncbi.nlm.nih.gov/Blast.cgi).<br />

Two isolates (CMW17225 and CMW17570) of the Ceratocystis<br />

sp. isolated from P. cineraria in Oman were tested for<br />

pathogenicity. Five seedlings, approx 2‐yrs‐old, were inoculated<br />

with each isolate, 15 cm above soil level. Five control seedlings<br />

were wounded and inoculated with sterile MEA. Lesion lengths<br />

were measured after 42 days.<br />

RESULTS<br />

Symptoms on diseased P. cineraria and D. sissoo included<br />

streaked discolouration of the vascular tissue and rapid wilting<br />

of the foliage. Cultures produced on MEA resembled those of C.<br />

fimbriata s. l. based on the morphology and size of the<br />

perithecia, ascospores, conidia and conidiophores. Blast<br />

searches of the ITS sequences of two isolates from P. cineraria<br />

showed 100% similarity with C. manginecans, while the two<br />

isolates D. sissoo shared 98% similarity with that species.<br />

All the inoculated P. cineraria seedlings had vascular<br />

discolouration and 40% of the inoculated seedlings wilted and<br />

died after four weeks. Lesions produced by isolates CMW17225<br />

and CMW17570 were 105 and 131 mm long respectively, and<br />

significantly longer than those of the control treatments (5 mm).<br />

Lesion length (mm)<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

Pathogencity test 2006<br />

CMW17225 CMW17570 Control<br />

Isolate<br />

Figure 1. Mean lesion lengths on P. cineraria seedlings 42 days after of<br />

inoculation with two Ceratocystis isolates.<br />

DISCUSSION<br />

The results of this study have shown clearly that a Ceratocystis<br />

sp. is closely associated with the wilt disease of P. cineraria in<br />

Oman and D. sissoo in Pakistan. Based on ITS sequence data, this<br />

fungus is very similar, if not identical, to C. manginecans.<br />

Furthermore, pathogenicity tests have provided good evidence<br />

that the Ceratocystis sp. is the cause of the wilt disease of P.<br />

cineraria and most probably also D. sissoo trees in Oman and<br />

Pakistan.<br />

The relatedness of the pathogen of P. cineraria and D. sissoo to<br />

C. manginecans, which causes a devastating disease of Mango in<br />

Oman and Pakistan, is intriguing. It has been suggested<br />

previously that C. manginecans is most likely an introduced<br />

pathogen in Oman and Pakistan (2). It is thus possible that this<br />

pathogen has subsequently adapted the ability to infect and kill<br />

native trees in those countries. This question deserves further<br />

and urgent study.<br />

ACKNOWLEDGEMENTS<br />

We thank members of the Tree Protection Co‐operative<br />

Programme (TPCP), University of Pretoria, South Africa, and the<br />

Ministry of Agriculture, Oman, for funding. We also thank the<br />

Nuclear Institute for Agriculture and Biology (NIAB) for<br />

facilitating surveys and shisham sample collection in Pakistan.<br />

REFERENCES<br />

1. Al Adawi AO, Deadman ML, Al Rawahi A, Al Maqbali YM, Al Jahwari<br />

AA, Al Saadi BA, Al Amri IS, Wingfield MJ (2006) Aetiology and<br />

causal agents of mango sudden decline disease in the Sultanate of<br />

Oman. European Journal of <strong>Plant</strong> <strong>Pathology</strong> 116, 247–254.<br />

2. Van Wyk M, Al Adawi AO, Khan IA, Deadman ML, Al Jahwari A,<br />

Wingfield BD, Ploetz RC, Wingfield MJ (2007) Ceratocystis<br />

manginecans sp. nov., causal disease of a destructive mango wilt<br />

disease in Oman and Pakistan. Fungal Diversity 27, 213–230.<br />

3. Moller WJ, DeVay JE (1968) Carrot as a species‐selective isolation<br />

medium for Ceratocystis fimbriata. Phytopathology 58, 123–124.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 221


Posters<br />

93 Evaluation of plant extracts for control of sclerotinia pathogens of vegetable crops<br />

D. Wite{ XE "Wite, D." } A , O. Villalta A and I.J. Porter A<br />

A Department of Primary Industries, 621 Burwood Hwy, Knoxfield, Victoria 3180<br />

INTRODUCTION<br />

Sclerotinia, caused by Sclerotinia sclerotiorum and S.minor, is<br />

one of the most damaging soilborne diseases of vegetable crops<br />

in Australia. These pathogens are difficult to control because<br />

they have a wide host range, survive in soil for many years as<br />

melanised sclerotia, and there are limited fungicide and genetic<br />

resistance options available. Beneficial management practices to<br />

reduce disease risk and IPM compatible control measures,<br />

suitable for on‐farm use, are both required to achieve<br />

sustainable control of Sclerotinia. Our research is therefore<br />

evaluating a range of new practices and control methods to<br />

improve the management of Sclerotinia in vegetable production.<br />

One potential control method is the use of plant extracts with<br />

antifungal volatile compounds. For instance, isothiocyanates<br />

(ITCs) released from cruciferous plant residues during hydrolysis<br />

of glucosinolates reduced the viability of S. sclerotiorum sclerotia<br />

(1) and unknown volatile compounds released from fennel oil<br />

inhibited mycelial growth of S. sclerotiorum (2). We report here<br />

on the efficacy of two mixtures of plant extracts and one<br />

essential oil on the viability of mycelium and sclerotia of S.<br />

sclerotiorum and S. minor.<br />

METHODS<br />

Treatments. Voom® (mustard and other essential oils, 15–20%<br />

allyl‐ITCs, Akhil), Dazitol® (mustard oil and capsaicanoids,<br />

unknown % ITCs, Champon) and bitter fennel oil (Foeniculum<br />

vulgare, Essential Oils of Tasmania) were tested at<br />

concentrations from 1–8% v/v. Fluke (20% allyl‐ITCs) was used<br />

for comparison as standard control in soil bioassay.<br />

In vitro tests. Mycelial plugs and sclerotia of S. sclerotiorum and<br />

S. minor isolates from bean and lettuce, respectively, were<br />

plated onto PDA amended with different concentrations of the<br />

treatments (contact diffusion). Inocula were also exposed to<br />

treatments using an inverted Petri dish assay (vapour phase).<br />

Plates were incubated for 7 days at room temperature and<br />

colony diameters recorded until growth reached the edge of the<br />

plates. Inocula that did not grow were transferred to<br />

unamended PDA to test whether the activity was fungicidal or<br />

merely fungistatic.<br />

Soil bioassay (sclerotia). Sclerotia in mesh bags (5/bag) were<br />

exposed to treatments (vapour phase) in non‐sterile sandy soil (1<br />

kg sealed pots) outdoors for 24 hrs (n = 3). After this period,<br />

sclerotia were surface sterilised and plated onto PDA to determine<br />

viability.<br />

PRELIMINARY RESULTS<br />

Effect on mycelial growth in vitro. All concentrations of Voom<br />

(1%, 3%, 5% v/v) were biocidal to mycelium of both Sclerotinia<br />

spp, irrespective of application method. Dazitol (4, 6 and 8% v/v)<br />

was biocidal to mycelium of both spp. when using the vapour<br />

phase method. Using the diffusion method, 6 and 8% were<br />

biocidal but 4% was only fungistatic. Fennel oil (2, 4 and 6% v/v)<br />

significantly suppressed mycelial growth of both pathogens.<br />

Effect on sclerotia viability in vitro. Voom (3% and 5% v/v) was<br />

biocidal to sclerotia of both pathogens (Table 1). At 1%, Voom<br />

was more effective in reducing the viability of sclerotia of S.<br />

sclerotiorum than S. minor. Dazitol (6% and 8% v/v) was biocidal<br />

to S. minor sclerotia and significantly reduced the viability of S.<br />

sclerotiorum sclerotia by 78–89%. Dazitol (4% v/v) also<br />

significantly reduced sclerotia viability by 78%. Fennel oil (2, 4<br />

and 6% v/v) had no effect on sclerotia viability.<br />

Table 1. Mean percentage reduction of sclerotia viability in vitro.<br />

Treatment (v/v) S. minor S. sclerotiorum<br />

untreated 0 a 0 a<br />

Voom 1% 11 b 100 c<br />

Voom 3% 100 d 100 c<br />

Voom 5% 100 d 100 c<br />

Dazitol 4% 78 c 78 b<br />

Dazitol 6% 100 d 78 b<br />

Dazitol 8% 100 d 89 b<br />

Means with the same letters are not significantly different (P6%) significantly reduced the viability of<br />

sclerotia of both spp.<br />

DISCUSSION<br />

The three products tested all released volatile compounds with<br />

antifungal activity against the two Sclerotinia species. Voom and<br />

Dazitol, tested at concentrations considered economic for<br />

disease control, were biocidal to inocula in vitro. In soil,<br />

however, Voom® (15–20% allyl‐ITCs) was more effective than<br />

Dazitol (unknown levels of ITCs) in reducing sclerotial viability of<br />

both pathogens. The standard treatment, Fluke, with similar<br />

levels of ITCs (20%) to that of Voom, caused total sclerotia<br />

mortality. Future studies will determine if differences in<br />

efficacies in soil were due to higher levels of ITCs released from<br />

Voom during the hydrolysis of glucosinolates. Further<br />

investigations to determine concentrations of ITCs required to<br />

achieve high levels of sclerotia mortality in soil are necessary to<br />

optimise biofumigant treatments such as plant extracts and<br />

biofumigant crops. Volatile compounds released from the low<br />

concentrations of fennel oil tested were only inhibitory to<br />

mycelial growth. Further studies of essential oils would be<br />

necessary to determine whether they have useful anti‐fungal<br />

compounds at concentrations that might be biocidal to<br />

Sclerotinia inoculum, and if so, to determine their mode of<br />

action.<br />

ACKNOWLEDGEMENTS<br />

We thank Horticulture Australia Ltd and the Department of<br />

Primary Industries Victoria for financial support.<br />

REFERENCES<br />

1. Smolinska, U., Horbowicz, M. (1999). Journal of Phytopathology<br />

147, 119–124.<br />

2. Soylu S, Yigitbas H, Soylu EM, Kurt S (2007). Journal of Applied<br />

Microbiology 103, 1021–1030<br />

222 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


94 First report of Macrophominia phaseolina on rapeseed stem in some provinces<br />

of Iran<br />

Posters<br />

A. Zaman Mirabadi{ XE "Zaman Mirabadi, A." } A , A. Esmaailifar B , A. Alian C and R. M. Alamdalou A<br />

A Oilseed Research and Development Company, Applied Research Center in North of Country, Sari, Iran<br />

B Department of <strong>Plant</strong> Protection, Faculty of Agriculture and Natural Resources Islamic Azad University, Arak, Iran<br />

C The Central <strong>Plant</strong> Diagnostic Clinic, Babolsar, Iran<br />

INTRODUCTION<br />

Rapeseed (Brassica napus) is one of the most important<br />

oleaginous crops in different areas of Iran. Recently, in some<br />

cases, gray‐black lesions with few microsclerotia were observed<br />

on rapeseed stem.<br />

MATERIALS AND METHODS<br />

In August 2006, rapeseed basal stem and taproot samples were<br />

collected from Azarbayejan sharghi (As), Ardabil (Ar),<br />

Kermanshah (Ke), Khuzestan (Ku), Fars (F) and Hamadan (H)<br />

provinces in Iran. Small pieces (5 mm) of these tissues were<br />

surface sterilised with NaOCL(1%) for 1 min and then positioned<br />

in the center of plates containing potato dextrose agar (PDA).<br />

Plates were maintained at 25°c for 4 days in the dark condition.<br />

Average diameter of 200 microsclerotia was calculated for each<br />

isolate.<br />

RESULTS<br />

Grey‐black color mycelia with spherical and black microsclerotia<br />

observed after 5 days. Yielding fungal colonies identified as<br />

Macrophomina phaseolina (Tassi) Goidanich based on mycelia<br />

and size of the microsclerotia. Size of 6 isolates microsclerotia<br />

was estimated for different areas (five provinces), from 50 to<br />

180 µm in diameter. To our knowledge, this is the first report of<br />

M. phaseolina on rapeseed stem (cultivars: Okapi, Zarfam, Licord<br />

and Hyola401) in mentioned provinces of Iran.<br />

DISCUSSION<br />

Charcoal rot on canola has been reported from Argentina (2)<br />

with the presence of microsclerotia 71–94 µm in diameter and<br />

the United States (1). Average diameter of microsclerotia in Iran<br />

was consisted: Ku‐102(70–150) µm, As1‐ 96.7(60–150) µm, F‐<br />

95(60–150) µm, Ke‐92.81(50–180) µm, Ar‐91.87(60–150) µm<br />

and As2‐ 87.03(60–180) µm. The highest and lowest size of<br />

sclerotia was related to warmest and coldest areas respectively.<br />

It seems to be diversity between Iranian isolates.<br />

REFERENCES<br />

1. Baird RE, Hershman DE, Christmas EP (1994) Occurrence of<br />

Macrophomina phaseolina on Canola in Indiana and Kentucky.<br />

<strong>Plant</strong> Disease, 78:316.<br />

2. Gaetán SA, Fernandez L, Madia M (2006) Occurrence of Charcoal<br />

Rot Caused by Macrophomina phaseolina on Canola in Argentina.<br />

<strong>Plant</strong> Disease, 90:524.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 223


Posters<br />

44 First report of rapeseed blackleg caused by pathogenicity group T (PGT)<br />

of Leptosphaeria maculans in Mazandaran province of Iran<br />

A. Zaman Mirabadi{ XE "Zaman Mirabadi, A." } A , K. Rahnama B , R.M. Alamdalou A and A. Esmaailifar C<br />

A Oilseeds Research and Development Company, Pasdaran Avenue, Sari, Iran<br />

B Department of <strong>Plant</strong> Protections, Gorgan University of Agricultural Sciences and Natural Resources, Gorgan, Iran<br />

C Department of <strong>Plant</strong> Protections, Faculty of Agriculture and Natural Resources Islamic Azad University, Arak, Iran<br />

INTRODUCTION<br />

Rapeseed (Brassica napus L.) is cultivated in a large scale<br />

approximately, 20.000 ha in Mazandaran province in the north<br />

of Iran. Phoma blackleg (Leptosphaeria biglobosa), pathogenicity<br />

group 1 (PG‐1) or non‐aggressive type, has been reported on<br />

rapeseed from Golestan province (1). Recently, a typical<br />

symptom of rapeseed blackleg has been observed in the regions<br />

with a long history of cultivation (Dasht‐E‐ Naz).<br />

MATERIALS AND METHODS<br />

In September 2008, Ascospores of Leptosphaeria maculans was<br />

obtained from infected stubble residues of commercial<br />

Rapeseed Hybrid (Hyola 401)(2). These spores were cultured in<br />

V8 medium. Thirteen isolates of L.maculans were used for<br />

determining Pathogenicity groups (PG) according to the<br />

phenotypic interaction (PI) on 7‐day old rapeseed cultivars;<br />

Glacier, Westar and Quinta. After 10 days, disease severity was<br />

rated on a 0–9 scale (3). Wounded cotyledons were inoculated<br />

with 10µl of conidial suspensions at 2×10 7 spores per ml. All<br />

plants were maintained in a growth chamber at 21°C (light) to<br />

16°C (dark), with a 16‐h photoperiod and relative humidity of<br />

95%. The test was repeated three times.<br />

ACKNOWLEDGEMENTS<br />

Special thanks to the Leibniz Institute of <strong>Plant</strong> Genetics and Crop<br />

<strong>Plant</strong> Research (IPK) for providing the seeds needed for this<br />

testing.<br />

REFERENCES<br />

1. Fernando WGD, Ghanbarnia K, Salati M (2007) First report on the<br />

presence of phoma blackleg pathogenicity group 1(Leptosphaeria<br />

biglobosa) on Brassica napus (canola/ rapeseed) in Iran. <strong>Plant</strong><br />

disease, 91: 465.<br />

2. Mengistu A, Rimmer RS, Williams PH (1993) Protocols for in vitro<br />

sporulation, ascospore release, sexual mating, and fertility in<br />

crosses of Leptosphaeria maculans. <strong>Plant</strong> Disease 77, 538–540.<br />

3. Williams PH (1985) Crucifer Genetics Cooperatives (CrGC) Resource<br />

Book. University of Wisconsin, Madison, Wisc., 160pp.<br />

RESULTS<br />

Two isolates (Es‐3 and ES‐ 12) were classified as belonging PGT, 2<br />

(Es‐5 and ES‐7) as PG2 and 9 isolates as PG1. PGT isolates gave PI<br />

reactions 3 to 4, 7 to 9 and 7 to 9 on Glacier, Quinta and Westar,<br />

respectively. As our survey, this is the first report of the<br />

occurrence of Leptosphaeria maculans PGT in Iran(Figure 1).<br />

DISCUSSION<br />

In some regions in the north of IRAN, the leaf symptoms have<br />

been observed as circle lesions with chlorotic border and gray<br />

color in center, sometimes lesions separate from leaf. The spots<br />

on stems are more elongate and often surrounded by a purple or<br />

black border. With attention to high sensitivity of current<br />

cultivar of Mazandaran province (Hyola 401) to PG2 and PGT of<br />

L.maculans, It is possible that to be epidemic of blackleg disease<br />

in the case of suitable climatic condition, in future.<br />

Figure 1. Phenotypic interaction (PI) of Leptosphaeria maculans isolates<br />

on five rapeseed cultivars after 10 days.<br />

224 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


19 In vitro study on the effect of NanoSilver (Nanosid) on Sclerotinia sclerotiorum<br />

fungi the causal agent of rapeseed white stem rot<br />

Posters<br />

A. Zaman Mirabadi{ XE "Zaman Mirabadi, A." } A , K. Rahnama B , R. Mehdi Alamdarlou A and A. Esmaailifar c<br />

A Oilseeds Research and Development Company, Pasdaran Avenue Sari, Iran<br />

B Department of <strong>Plant</strong> Protection, Gorgan University of Agricultural Sciences and Natural Resources, Grgan, Iran<br />

c Department of <strong>Plant</strong> Protections, Faculty of Agriculture and Natural Resources Islamic Azad University, Arak, Iran<br />

INTRODUCTION<br />

White stem rot of rapeseed (Brassica napus L.) caused by<br />

Sclerotinia sclerotiorum (Lib.) de Bary is a serious disease in<br />

many areas in the world. This fungus attacks over 360 species<br />

of plants (2). It has been reported in Golestan and Mazandaran<br />

provinces in north of Iran (1). Chemical spraying is the simplest<br />

option for controlling sclerotinia disease in rapeseed fields<br />

however many surveys have been done for potential on<br />

bicontrol methods and new fungicide applications.<br />

MATERIALS AND METHODS<br />

In this study, the in vitro effectiveness of nanosilver different<br />

doses consist of 5, 10, 30, 50, 100,120, 130 and 150 ppm were<br />

tested against one isolate of S. sclerotiorum. Nanosid solution<br />

was incorporated into potato dextrose agar (PDA) medium,<br />

after autoclaving. Sclerotia of S. sclerotiorum were positioned<br />

in the center of Petri dishes containing PDA plus nanosid and<br />

were maintained in 25 centigrade. Fungal growth<br />

characteristics were determined after 8 and 14 days at all of<br />

the treatments.<br />

RESULTS<br />

After 8 days, there was no mycelia growth or sclerotia<br />

production in doses more than 10 ppm, while there was<br />

observed a significant difference between the mycelia growth<br />

of control (Sterile water), 5 and 10 ppm doses of nanosilver.<br />

Average per cent of inhabitation effect of nanosilver doses (8<br />

doses) and control treatment on sclerotia were estimated after<br />

14 days respectively 0, 25.33%, 48.17%, 53.33%, 56%, 56%,<br />

57.66%, 61% and also 0 for control (Table 1). After 14 days, a<br />

few numbers of new sclerotia were observed in doses of more<br />

than 10 ppm (2 to 4 sclerotia) in comparison white 5ppm and<br />

control (18–22 sclerotia). Obtained results showed that<br />

Nanosilver has fungistatic action and could be used for<br />

decreasing sclerotia production and mycelia growth (Fig.1)<br />

Table 1. Comparison of means for effect of Nanosilver doses on<br />

sclerotia of S.sclerotiorum after 8 and 14 days.<br />

Nanosilver(ppm)<br />

after8days<br />

Mean of square<br />

Control 75a z 75a z<br />

5 31.37b 75a<br />

10 16.75c 56b<br />

after14days<br />

30 0d 38.87c<br />

50 0d 35cd<br />

100 0d 33cd<br />

120 0d 33cd<br />

130 0d 31.75cd<br />

150 0d 29.25cd<br />

z Within columns, numbers fallowed by a common letter are not significantly<br />

different (P=0.05) according to Duncan’s multiple range test.<br />

DISCUSSION<br />

Practical Tests of Nanosid compounds will be done in field<br />

condition in future.<br />

REFERENCES<br />

1. Hossein B, Hamid Z, Gafar E, Abdoreza F (2001) Distribution of<br />

rapeseed white stem rot in Mazandaran. Proceedings of the 18th<br />

Iranian plant protection congress, 295p.<br />

2. Purdy, LH (1979) Sclerotinia sclerotiorum history, diseases and<br />

symptomatology, host range, geographic distribution, and impact.<br />

Phytopathology 69:875–880.<br />

Figure 1. Effect of Nanosilver doses on radial growth of S.sclerotiorum<br />

after 14 days: a) Control and 5ppm, b) 10 ppm, c)30 ppm and d)50ppm,<br />

after 8 days: e)100ppm, f)120 ppm, g)130 ppm and h)150 ppm.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 225


Posters<br />

20 Study on the effect of number of spraying with fungicides on rapeseed sclerotinia<br />

stem rot control<br />

R. Mehdi Alamdarlou A , A. Zaman Mirabadi{ XE "Zaman Mirabadi, A." } A , A. Esmaailifar B and K. Foroozan C<br />

A Applied Research Center in North of Country_Oilseeds Research and Development Company, Sari, Iran<br />

B Department of <strong>Plant</strong> Protection, Faculty of Agriculture and Natural Resources Islamic Azad University, Arak, Iran<br />

C Oilseeds Research and Development Company,Tehran, Iran<br />

INTRODUCTION<br />

Sclerotinia stem rot disease caused by Sclerotinia sclerotiorum<br />

(Lib.) de Bary fungus, is one of the rapeseed important diseases<br />

in the world and also in the north of Iran (1). Germination of<br />

fungus sclerotia produce apothecia on the soil and the<br />

ascospores that released from apothecia can infect rapeseed<br />

plants at the flowering stage. Spraying with fungicide in this<br />

stage protect the plant from infection.<br />

MATERIALS AND METHODS<br />

In order to study the effect of number of spraying with<br />

fungicides against rapeseed sclerotinia stem rot disease caused<br />

by S. sclerotiorum, an experiment was conducted in complete<br />

randomised block design with 3 replications and 7 treatments in<br />

Mazandaran province for two years (2006–2007). An infected<br />

field (with many sclerotia in the soil) was selected and the<br />

rapeseed hybrid Hyola401 planted in it. For disease control at<br />

the flowering stage of rapeseed one or twice spraying with<br />

fungicides carbendazim (WP60%) and tebuconazole (EC25%) was<br />

done. The first spraying was done at the 20%–30% of flowering<br />

stage and the second one 14 days after that. In both two years<br />

the period of fungi apothecia appearance and ascospores<br />

releasing had been started before spraying. The percentage of<br />

disease infection on leaves and stems was determined. The<br />

disease severity on the plants was scored from 1(lowest<br />

infection) to 9(highest infection) and determined in different<br />

treatments.<br />

DISCUSSION<br />

Chemical control of rapeseed sclerotinia stem rot is an effective<br />

method in the north of Iran. Depending on the weather<br />

condition and field growth, one or twice fungicide application is<br />

needed. For best disease control we recommend the disease<br />

managing by integration of different methods like mechanical,<br />

agronomical, biological and chemical.<br />

REFERENCES<br />

1. Hossein B, Hamid Z, Gafar E, Abdoreza F (2001) Distribution of<br />

rapeseed white stem rot in Mazandaran. Proceedings of the 18th<br />

Iranian plant protection congress, 295p.<br />

2. Saharan GS, Mehta N (2008) Sclerotinia disease of crop plants:<br />

biology, ecology and disease management, Springer science,<br />

531pp.<br />

3. Thomson JR, Thomas PM, Evans JR (1984) Efficacy of aerial<br />

application of benomyl and iprodione for the control of sclerotinia<br />

stem rot of canola (rapeseed) in central Alberta, Canadian journal<br />

of plant pathology, 6:75–77.<br />

RESULTS<br />

Results of variances analysis and comparison between<br />

treatments means by Duncan’s test indicated that in first year<br />

sprayed treatments compared to control had lower infection and<br />

higher yield, but there were not significant differences between<br />

one and twice sprayed treatments. In second year and two years<br />

combined analysis, control had highest infection and lowest yield<br />

and twice sprayed treatments had lowest infection and highest<br />

yield.<br />

Table1. Comparison of leaf and stem infection percentage, disease<br />

severity and yield between different treatments in two years.<br />

Treatments<br />

%leaf<br />

infection<br />

%stem<br />

infection<br />

Disease<br />

severity<br />

Yield Kg/ha<br />

S 55a Z 49a Z 6.6a Z 3208d Z<br />

C 15bc 11bc 3.2c 3793bc<br />

CC 3.5d 0.5d 0.7f 4020a<br />

F 21b 15b 4.1b 3720c<br />

FF 8.5cd 5.5cd 2.4d 3975a<br />

FC 4d 1d 0.7f 4003a<br />

CF 7.5cd 3.5d 1.6e 3921ab<br />

Control(S), Carbendazim(C), Twice carbendazim (CC), folicur(F), Twice folicur (FF),<br />

Folicur at first and carbendazim at second application(FC), Carbendazim at first and<br />

folicur at second application (CF)<br />

z Within columns, numbers fallowed by a common letter are not significantly<br />

different (P=0.01) according to Duncan’s multiple range test.<br />

226 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009


98 The biology and management of chestnut rot in south‐eastern Australia<br />

L. Shuttleworth{ XE "Shuttleworth, L" }, D. Guest, E. Liew<br />

A University of Sydney. Room 403, Level 4 McMillan Building NSW 2006<br />

Posters<br />

INTRODUCTION<br />

The Australian chestnut industry is currently small, but has<br />

potential to expand due to increasing local and export demand.<br />

The European chestnut, Castanea sativa is the main species<br />

grown in Australia today. Internal rot of chestnuts is a significant<br />

problem facing the industry and is what this study will be<br />

investigating. Internal rot affects the kernel, manifesting in light,<br />

medium and/or dark brown lesions occurring on the endosperm,<br />

and embryo. It is often not visible externally and is only realised<br />

when the nut is opened. The incidence of chestnut rot at<br />

Melbourne markets has been recorded as high as 40% (1).<br />

Consumer rejection of the chestnuts is highly likely with rot<br />

incidence at these levels. Without an appropriate control<br />

strategy, the Australian chestnut industry is unlikely to grow.<br />

The main cause of chestnut rot is reported to be endophytic,<br />

caused by the fungal Ascomycete Phomopsis castanea (Sacc.)<br />

Höhn (1,3). Research also shows that infection by ascospores<br />

during flowering is a mode of infection (2).<br />

The aims of this study are to survey the incidence of chestnut rot<br />

in south eastern Australia, New South Wales (NSW) and Victoria<br />

(VIC), identify the pathogen responsible for causing chestnut rot,<br />

and to study the disease cycle.<br />

MATERIALS AND METHODS<br />

Field and market surveys. A hierarchical sampling strategy was<br />

used. Twenty two orchards were sampled across VIC and NSW.<br />

Orchards were classed into 6 regions based on geographical<br />

distance i.e. 100km proximity (see fig 1 for orchard locations).<br />

Four commercial varieties were assessed for rot incidence:<br />

Buffalo Queen, Red Spanish, Purton’s Pride, and Decoppi<br />

Marone (300 nuts/orchard). Nuts were dissected and visually<br />

assessed for rot.<br />

Flemington Markets in NSW were also sampled for chestnut rot.<br />

Nuts sourced from 5 orchards were sampled, 300 nuts/orchard.<br />

Decoppi Marone was surveyed from 2 orchards and Purton’s<br />

Pride from 3 orchards.<br />

Pathogen identification. Diseased chestnuts from 12 of the<br />

surveyed farms in VIC and NSW had the pathogen isolated from<br />

them using standard isolating protocols (1, 3). The pathogen was<br />

then identified morphologically.<br />

Observation of the teleomorph was completed through<br />

microscopic observation of 30 decaying burrs from an orchard in<br />

Mullion Creek, NSW.<br />

Endophyte studies. Samples were collected from healthy floral<br />

and vegetative tissues at Mullion Creek NSW. Seventy‐five<br />

samples per tissue type were tested.<br />

RESULTS AND DISCUSSION<br />

Field Survey<br />

Figure 1. Incidence of chestnut rot from nuts sampled field survey across<br />

6 regions in NSW and VIC. Total number of nuts in survey =6600.<br />

Disease incidence of chestnuts tested from Flemington markets,<br />

NSW was found to be ≤10% for all orchards and varieties tested.<br />

This value is within the acceptable level of rot set by the<br />

Australian chestnut industry.<br />

Pathogen identification. Of the 568 isolates obtained, 62% of<br />

were identified as Phomopsis.castanea. The teleomorph of<br />

Diaporthe castaneti Nitschke. was observed growing on decaying<br />

burr tissue. This indicates that aerosols of ascospores are a<br />

source of infection in the disease cycle.<br />

Endophyte studies. Tissues displaying the highest rates of<br />

pathogen isolation include female flowers (82%), male flowers<br />

(59%), pedicel (28%), leaf margin (33%), and 1 st year growth<br />

(17%). Low isolation levels were recorded in petioles (9%), and<br />

leaf mid veins (9%), second year stems (8%), and third and<br />

fourth year bark (3%). The pathogen was not recorded in third<br />

and fourth year xylem tissue, indicating pathogen travel via<br />

xylem tissue is not significant and likely not systemic.<br />

CONCLUSION<br />

Disease incidence varies greatly between regions. This is likely<br />

due to factors including regional and micro‐climates, rainfall<br />

during flowering, the susceptibility of host variety to the<br />

pathogen, and the virulence of the pathogen.<br />

REFERENCES<br />

1. Anderson CR (1993) A study of Phomopsis castanea (Sacc.) Höhn,<br />

and post harvest decay of chestnuts. Thesis. University of<br />

Melbourne.<br />

2. Smith HC The life cycle, pathology and taxonomy of two different<br />

nut rot fungi in chestnut. Australian Nutgrower. 22:2, 10‐15.<br />

3. Washington WS, Stewart‐Wade S, Hood, V (1999) Phomopsis<br />

castanea, a seed‐borne endophyte in chestnut trees. Australian<br />

Journal of Botany. 47, 77‐84.<br />

APPS 2009 | PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH 227


Index<br />

Abkhoo, J................................ 128, 129, 130<br />

Aftab, M. ................................................131<br />

Agarwal, A. ...............................................65<br />

Akem, C. ...................................................50<br />

Akem, C.N....................................... 132, 133<br />

Akinsanmi, O.A. .......................................22<br />

Alhudaib, K. ............................................134<br />

Amponsah, N.T.........................................70<br />

Anderson, J.M...........................................54<br />

Auer, D.P.F................................................58<br />

Auyong, A.S.M. .......................................135<br />

Baskarathevan, J.......................................52<br />

Beever, R.E. ..............................................77<br />

Beresford, R.M. ........................................44<br />

Bhuiyan, S.A........................................47, 61<br />

Billones, R.G. ......................................69, 79<br />

Bleach, C.M. ...........................................104<br />

Blomley, C....................................... 114, 136<br />

Borines, L.M. ..........................................137<br />

Boyd‐Wilson, K.S.H. ................................138<br />

Braithwaite, M........................................103<br />

Brett, R.W.................................................29<br />

Burgess, L.W...........................................120<br />

Burgess, T.I. .............................. 37, 111, 139<br />

Burgess. L.W.............................................42<br />

Cahill, D.M................................................83<br />

Casonato, S.G. ..........................................99<br />

Chomic, A. ................................................31<br />

Christ, B. .................................................122<br />

Cobon, J.A...............................................140<br />

Collins, S.J.........................................27, 141<br />

Dann, E.K. ...............................................142<br />

Davidson, J.A. .........................................143<br />

Davies, P.A.B................................... 108, 144<br />

Davis, R.I.................................................109<br />

Dean, J.R.................................................145<br />

Deland, L...................................................88<br />

Donald, E.C. ..............................................68<br />

Donovan, N.J. .........................................146<br />

Dore, D.S. .................................................71<br />

Drenth, A. .................................................21<br />

Edwards, J.................................................24<br />

Erwin, E.....................................................38<br />

Evans, K.J. .................................................46<br />

Everett, K.R...............................................45<br />

Fahim, M. .................................................66<br />

Falloon, R.E..................................... 147, 148<br />

Ferguson, K.L. ...........................................89<br />

Ford, R. .....................................................82<br />

Forsyth, L.M............................................150<br />

Fullerton, R.A..........................................151<br />

Gambley, C. ............................................152<br />

Ge, Y. ......................................................153<br />

Glen, M................................... 154, 155, 156<br />

Golzar, H...........................................41, 157<br />

Gouk, C. ..................................................158<br />

Gouk, S.C. .................................................90<br />

Greer, L.A................................................159<br />

Haghighi, S..............................................160<br />

Hall, B.H.................................. 161, 162, 163<br />

Hardham, A.R. ..........................................64<br />

Hidayat, S.H. .............................................49<br />

Hill, G.N. ...................................................57<br />

Hodda, M................................................100<br />

Hohmann, P............................................102<br />

Huang, R. ................................................164<br />

Hüberli, D..................................................85<br />

Jackson, S.L.............................................165<br />

Jayasena, K.W......................... 166, 167, 168<br />

Johnson, G.I. .............................................20<br />

Keane, P.J. ................................................91<br />

Ketabchi, S..............................................169<br />

Khanam, N.N...........................................170<br />

Knight, N.L. .............................................106<br />

Kueh, K.H. ...............................................171<br />

Kumari, S.G. ..............................................33<br />

Lehmensiek, A. .......................................172<br />

Linde, C.C..................................................51<br />

Lomavatu, M.F........................................173<br />

Lovelock, D. ............................................174<br />

Luck, J.E. .................................................118<br />

MacDiarmid, R.M......................................32<br />

Magarey, P.A. .............................59, 78, 175<br />

Magarey, R.C. .....................................56, 98<br />

Malligan, C.D. ...................................96, 176<br />

Mannan, S.........................................80, 177<br />

Maora, J.Y.S. .............................................55<br />

McMahon, P.J...........................................23<br />

Meldrum, R.A. ........................................179<br />

Miles, A.K........................................180, 181<br />

Minchinton, E.J. ........................................60<br />

Mintoff, S.J.L...........................................182<br />

Moran, J.R.................................................26<br />

Morin, L. ................................. 115, 183, 184<br />

Mundy, D.C...............................................86<br />

Namaliu, Y. ...............................................63<br />

Nambiar, L. .............................................185<br />

Newby, Z.J. .............................................186<br />

Oliver, J.E. ...............................................112<br />

Panjehkeh, N. .........................................189<br />

Park, E.W. ...............................................117<br />

Paul, P.K..................................................113<br />

Pederick, S.J............................................190<br />

Pegg, G.S...................................................35<br />

Perez‐Egusquiza, Z.C...............................191<br />

Peterson, S.A. .........................................110<br />

Petkowski, J.E. ....................................... 192<br />

Petrisko, J.E.............................................. 84<br />

Phillips, D. ................................................ 67<br />

Pitt, W.M. .............................................. 193<br />

Porter, I.............................................28, 121<br />

Probst, C.M. ............................................. 25<br />

Ramroodi, S. .......................................... 194<br />

Randles, J.W........................................... 195<br />

Ray, J.D. ................................................. 196<br />

Reen, R.A. .............................................. 197<br />

Rees‐George, J. ...................................... 198<br />

Romberg, M.K........................................ 199<br />

Sakalidis, M.L. .......................................... 36<br />

Salam, M.................................................. 75<br />

Salam, M.U. ......................................74, 119<br />

Salowi, A. ............................................... 200<br />

Sapak, Z.................................................. 201<br />

Scarlett, K............................................... 202<br />

Scoble, C.A. ............................................ 116<br />

Scott, E.S.........................................203, 204<br />

Sdoodee, R............................................... 76<br />

Seem, R.................................................... 92<br />

Shirazi, M............................................... 205<br />

Shuttleworth, L ...................................... 227<br />

Simpfendorfer, S...............................95, 206<br />

Singh, A.................................................. 207<br />

Smith, L.J................................................ 208<br />

Sosnowski, M.R.................................97, 209<br />

Steele, V................................................. 210<br />

Stewart, A. ............................................. 101<br />

Stewart‐Wade, S.M................................ 211<br />

Sutherland, M.W.................................... 105<br />

Sutton, T.B. .............................................. 43<br />

Taylor, P.W.J. ........................................... 53<br />

Tesoriero, L.A........................................... 40<br />

Thompson, J.P................. 212, 213, 214, 215<br />

Thompson, N. ........................................ 216<br />

Trivedi, R.S. .......................................73, 217<br />

Truong, N.V.....................................187, 188<br />

Turaganivalu, U...................................... 218<br />

Vanneste, J.L............................................ 87<br />

Viljanen‐Rollinson, S.L.H. ....................... 107<br />

Vu Thanh, T.A. ....................................... 219<br />

Walter, M............................................48, 62<br />

Wellings, C.R. ......................................93, 94<br />

Wiechel, T.J.........................................30, 39<br />

Wilson, B.A. ........................................... 220<br />

Wingfield, M.J...................................34, 221<br />

Wite, D................................................... 222<br />

Wunderlich, N.......................................... 72<br />

Zaman Mirabadi, A. ........ 223, 224, 225, 226<br />

228 PLANT HEALTH MANAGEMENT: AN INTEGRATED APPROACH | APPS 2009

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!