Next Article in Journal
Characterization of Lebanese Germplasm of Snake Melon (Cucumis melo subsp. melo var. flexuosus) Using Morphological Traits and SSR Markers
Next Article in Special Issue
High Resolution Melting and Insertion Site-Based Polymorphism Markers for Wheat Variability Analysis and Candidate Genes Selection at Drought and Heat MQTL Loci
Previous Article in Journal
Plant Functional Diversity, Climate and Grazer Type Regulate Soil Activity in Natural Grasslands
Previous Article in Special Issue
QTL Mapping for Seedling and Adult Plant Resistance to Leaf and Stem Rusts in Pamyati Azieva × Paragon Mapping Population of Bread Wheat
 
 
Erratum published on 24 February 2022, see Agronomy 2022, 12(3), 561.
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Cost-Effective and Time-Efficient Molecular Assisted Selection for PPV Resistance in Apricot Based on ParPMC2 Allele-Specific PCR

1
Citriculture and Plant Production Center, Instituto Valenciano de Investigaciones Agrarias (IVIA), CV-315, km 10.7, 46113 Moncada, Valencia, Spain
2
Instituto de Biología Molecular y Celular de Plantas (IBMCP), Consejo Superior de Investigaciones Científicas (CSIC)—Universidad Politécnica de Valencia (UPV), Ingeniero Fausto Elio s/n, 46022 Valencia, Spain
*
Author to whom correspondence should be addressed.
Agronomy 2020, 10(9), 1292; https://doi.org/10.3390/agronomy10091292
Submission received: 31 July 2020 / Revised: 25 August 2020 / Accepted: 25 August 2020 / Published: 31 August 2020
(This article belongs to the Special Issue Molecular Marker Technology for Crop Improvement)

Abstract

:
Plum pox virus (PPV) is the most important limiting factor for apricot (Prunus armeniaca L.) production worldwide, and development of resistant cultivars has been proven to be the best solution in the long-term. However, just like in other woody species, apricot breeding is highly time and space demanding, and this is particularly true for PPV resistance phenotyping. Therefore, marker-assisted selection (MAS) may be very helpful to speed up breeding programs. Tightly linked ParPMC1 and ParPMC2, meprin and TRAF-C homology (MATH)-domain-containing genes have been proposed as host susceptibility genes required for PPV infection. Contribution of additional genes to PPV resistance cannot be discarded, but all available studies undoubtedly show a strong correlation between ParPMC2-resistant alleles (ParPMC2res) and PPV resistance. The ParPMC2res allele was shown to carry a 5-bp deletion (ParPMC2-del) within the second exon that has been characterized as a molecular marker suitable for MAS (PMC2). Based on this finding, we propose here a method for PPV resistance selection in apricot by combining high-throughput DNA extraction of 384 samples in 2 working days and the allele-specific genotyping of PMC2 on agarose gel. Moreover, the PMC2 genotype has been determined by PCR or by using whole-genome sequences (WGS) in 175 apricot accessions. These results were complemented with phenotypic and/or genotypic data available in the literature to reach a total of 325 apricot accessions. As a whole, we conclude that this is a time-efficient, cost-effective and straightforward method for PPV resistance screening that can be highly useful for apricot breeding programs.

1. Introduction

Most cultivated apricots belong to the Prunus armeniaca L. species, a member of the Rosaceae family, Prunus genus and section Armeniaca (Lam.) Koch [1]. World apricot production reached 3.84 million tonnes in 2018, with Turkey, Uzbekistan and Iran as the main producers (http://www.fao.org/faostat/). This means an increase of about 45% since 1998 mainly due to Asian countries. By contrast, European production in this period has just increased slightly while the cultivated area declined up to 19%. Despite its wide geographical spread, apricot has very specific ecological requirements. Consequently, each region usually grows locally adapted cultivars. For this reason, significant breeding efforts have been undertaken since the first apricot breeding program started in 1925 at the Nikita Botanical Garden in Yalta (Crimea, Ukraine) [2]. However, apricot breeding based on biparental controlled crosses and subsequent selection of the best new allelic combinations is hardly limited by the capacity to evaluate trees in the field [3]. On one side, fruit trees show high space requirements to be grown. On the other, their juvenile phase is quite long and reliable pomological phenotyping requires several cropping seasons, which means that at least ten years are needed to release a new variety. Therefore, the implementation of marker-assisted selection (MAS) has a great potential to improve breeding efficiency in fruit trees, including apricot.
Sharka disease, caused by Plum pox virus (PPV), is currently the most important viral disease affecting stone fruit trees (Prunus spp.) [4]. To date, nine PPV strains (D, M, C, EA, W, Rec, T, CR and An) are identified [5]. However, PPV genetic diversity may be even bigger, as observed by Chirkov et al. [6], who recently described the new Tat isolates affecting sour cherry (Prunus cerasus). PPV-D and M are the most widespread and economically important strains [5,7]. A clear host preference is observed: PPV-D/plum/apricot and PPV-M/peach. However, underlying genetic determinants are still unknown [8].
Particularly in apricot, PPV-D has severely hindered production in the last three decades, especially in endemic areas. In this context, development of PPV-resistant varieties is the main objective of apricot breeding programs. However, resistant sources are scarce. Just a handful of North American PPV-resistant cultivars have been identified to date, and they are commonly used as donors in all apricot resistance breeding programs currently in progress [9]. Several independent works aimed at dissecting the genetic control of PPV resistance in apricot have identified the major dominant PPVres locus in the upper part of linkage group 1 [10,11,12,13,14,15,16,17]. According to the pedigree and fine mapping data, a single common ancestor carrying PPVres has been suggested for all PPV-resistant cultivars [16,18,19,20]. Moreover, other minor loci contributing to PPV resistance have been suggested [13,14,15,16], but their role has not yet been well defined. More recently, transcriptomic and genomic analyses of PPVres locus have pointed out ParPMC1 and ParPMC2, two members of a cluster of meprin and TRAF-C homology domain (MATHd)-containing genes, as host susceptibility paralogous genes required for PPV infection [21]. The ParPMC2 allele linked in coupling with PPV resistance (ParPMC2res) accumulates 15 variants, including a 5 nt deletion (ParPMC2-del) that results in a premature stop codon. Moreover, cultivars carrying the ParPMC2res allele show that ParPMC2 and especially ParPMC1 genes are downregulated. As a result, this ParPMC2res was proposed to be a pseudogene that confers PPV resistance by silencing functional homologs, the non-mutated ParPMC2 allele and/or ParPMC1. Another plausible scenario involves epigenetic modifications to explain ParPMC silencing in the resistant cultivars [22].
In spite of evidence supporting linkage with the PPVres locus, some genotype-phenotype incongruencies (GPIs) have been detected in biparental populations segregating for PPV resistance [17,23,24]. In other words, some phenotypically susceptible individuals carrying ParPMC2res were classified as genetically resistant. Possible causes underlying these discrepancies, including other loci contributing to PPV resistance, are still unresolved. However, the potential benefit of using a ParPMC2 allele-specific marker (PMC2) for MAS is still very high since sharka resistance phenotyping is a major bottleneck in apricot breeding programs. The most reliable method for apricot PPV resistance phenotyping is based on a biological test that uses GF-305 peach rootstocks as woody indicators and graft-inoculation with PPV [25]. This procedure is time-consuming and requires visual inspection during two to four growing seasons in several replicates per genotype followed by ELISA [26] and RT-PCR tests [27]. It should be noted that the plant to be tested must be of a significant size in order to have enough buds for grafting replicates, so it takes a couple of years from the time of crossing. As a result of a genetic mapping approach, Soriano et al. [18] reported the first successful MAS application for PPV resistance using 3 SSRs within the PPVres locus resolved by capillary electrophoresis. Afterwards, these SSRs were combined with a single sequence length polymorphism marker (ZP002) interrogating the ParPMC2-del resolved by capillary or acrylamide electrophoresis [24] and by high resolution melting [28]. However, specialized DNA testing services are needed to adopt these MAS approaches, and together with the economic costs, this could be a challenge [29].
Here, we report a method combining high-throughput DNA extraction of 384 samples in 2 days and PMC2 genotyping by allele-specific PCR amplification and agarose gel electrophoresis. This method is proven to be an easily implemented tool for MAS of PPV-resistant seedlings in almost any apricot breeding program. Therefore, bioassays for PPV resistance evaluation will be needed to confirm the phenotype in selected materials. Moreover, PMC2 genotype has been determined and/or revised for 325 worldwide cultivated apricot accessions providing useful information for breeders to select parental genotypes.

2. Materials and Methods

2.1. High-Throughput DNA Isolation in 96-Well Plate

The genomic DNA extraction protocol was optimized from the original Doyle and Doyle method [30] to manage 384 samples per isolation using 8-well 1.2-mL strip tubes (VWR International). For each accession, 2 leaf discs were collected and placed into a tube with 3 glass beads (VWR International). The strips were frozen in liquid N2 and stored at −20 °C before DNA isolation. Frozen tissue was ground for 1 min with a frequency of 26/s using a Qiagen TissueLyser 85210 (Qiagen, Hilden, Germany). Then, 340 μL of preheated CTAB isolation buffer (with 0.2% 2-mercaptoethanol) was added to the ground tissue and incubated at 65 °C for 40 min, shaking gently every 10 min. After a short spin, 340 μL of chloroform-isoamyl alcohol (24:1) was added and mixed inverting the plates. Tubes were centrifuged for 10 min at 3000 rpm and 4 °C. The clean aqueous phase was transferred to new strip tubes, and 1.5 vol of 100% ethanol and 15 mM ammonium acetate were added and mixed gently. After overnight incubation at −20 °C, tubes were centrifuged for 10 min at 3000 rpm at 4 °C. The supernatant was discarded inverting the tubes, and 300 μL of 70% ethanol was added. After centrifugation for 10 min at 3000 rpm at 4 °C, the supernatant was discarded and finally 75 μL of TE was added. DNA at 1:10 dilution was used for PCR. Some random DNA samples from each plate were subjected to quality control. DNA integrity was checked on an agarose gel, and quantification was performed using a Nanodrop ND-1000 spectrophotometer (Nanodrop Technologies, Wilmington, DE, USA).

2.2. PMC2 Genotype by Allele-Specific PCR Assay

PMC2 marker genotyping was performed using the allele-specific forward primer (PMC2-F-alleleR: 5′-GTCATTTTCATTGATGTCATTCA-3′, or PMC2-F-alleleS: 5′-GTCGTTTTCATTGATGTCCAAAC-3′, respectively) and one common reverse primer (PMC2-R: 5′-GTGCTCTTTCACATTCTTGCTC-3′), as described by Zuriaga et al. [21]. PCRs were performed in a final volume of 20 μl containing 1 × DreamTaq buffer, 0.2 mM of each dNTP, 5 μM of each primer, 1 U of DreamTaq DNA polymerase (Thermo Fisher) and 2 μL of DNA extraction (diluted 1:10). Cycling conditions were as follows: an initial denaturing of 95 °C for 5 min; 35 cycles of 95 °C for 30 s, 55 °C for 45 s and 72 °C for 45 s; and a final extension of 72 °C for 10 min. PCR products were electrophoresed in 1% (w/v) agarose gels.
Available DNA samples from 120 apricot cultivars and accessions were PCR screened in this work. Part of this collection is currently kept at the collection of the Instituto Valenciano de Investigaciones Agrarias (IVIA) in Valencia (Spain), while other samples were provided by the Departamento de Mejora y Patología Vegetal del CEBAS-CSIC in Murcia (Spain), the University of St. Istvan (Budapest, Hungary) or by SharCo project (FP7-KBBE-2007-1) partners.

2.3. WGS Mapping and PMC2 Screening

WGSs of 73 cultivars were used in this study. Twenty-four of these WGSs and the 454 sequenced BAC clones belonging to the “Goldrich” PPVres locus R-haplotype were already screened in our previous works [20,21,22,23,24,25,26,27,28,29,30,31]. The other 49 WGSs were downloaded from the SRA repository (https://www.ncbi.nlm.nih.gov/sra). All raw reads were processed using the “run_trimmomatic_qual_trimming.pl” script from the Trinity software [32]. After removing the low-quality regions as well as vector and adaptor contaminants, cleaned reads were aligned to the peach genome v.2.0.a1 [33] using Bowtie2 v.2.2.4 software [34]. The presence/absence of the ParPMC2-del was visually inspected using IGV v.2.4.16 [35].

3. Results and Discussion

3.1. High-Throughput DNA Extraction and ParPMC2-del Genotyping for MAS

MAS offers great advantages over traditional seedling selection based just on phenotypic evaluations in fruit breeding [36]. DNA tests in segregating populations can improve the cost efficiency and/or the genetic gain for each seedling selection cycle [29], allowing to identify a few seedlings from among many thousands that have the genetic potential for desired performance levels [37]. As a result, agronomical evaluation in field trials is restricted to the promising selected materials. Implementation of MAS is especially valuable for traits that are difficult and/or expensive to phenotype as PPV resistance. As previously explained, the most reliable PPV resistance phenotyping is based on a biological test that uses graft-inoculated GF-305 peach seedlings [25] (Figure 1A). This protocol requires several replicates per genotype and visual symptoms inspection during 2–4 growing seasons, which entails the main bottleneck in apricot breeding programs. For instance, following this method at the IVIA’s greenhouse and cold chamber facilities, we can phenotype no more than 3000 plants per year, which equals 500 seedlings (i.e., 6 replicates are needed for each seedling).
In this work, we present a new strategy to speed up while reducing costs of the current application of MAS for PPV resistance in apricot [18,24,28]. Here, we combine a high-throughput DNA extraction protocol that does not need sophisticated robotic systems and can be implemented in any regular laboratory, with PMC2 allele-specific PCR amplification using previously described primers [21] and agarose electrophoresis (Figure 1B). Both forwards primers differ at the 3’-end, allowing to easily discriminate the presence/absence of the 5-bp ParPMC2-del (Figure 2). With this DNA extraction method, one person can easily process up to 384 samples (four 96-well sample plates) in 2 working days, enabling high throughput sample preparation. This is 4 times more samples than a standard CTAB method using individual tubes, while the cost of reagents and consumables is similar in both cases (around 0.29–0.30 € per sample) (Table S2). DNA obtained has enough quantity and quality to ensure subsequent regular PCRs. A 1:10 dilution of the DNA obtained was directly used for PCR amplification, without any additional purification step. In contrast, commercial kits are much more expensive in terms of reagents and consumables with costs around 4€ per sample. Then, using this DNA, 3 different methods could be applied for PPV MAS in apricot: the fluorescent labelling of PCR fragments that are resolved using capillary electrophoresis [18], the high-resolution melting (HRM) approach [28], and the use of standard PCR resolved by agarose gel electrophoresis [21]. It should be noted that the first two methods require the use of special equipment that could not be available for some laboratories and that also make the protocol more expensive. For instance, just the capillary electrophoresis costs around 1.5–2€ per sample (PCR not included) and the fluorescently labelled primers needed for PCR (136€ 10 nm) are much more expensive than the non-labelled ones (4€ 20 nm). On the other hand, commercial kits for HRM are not very expensive (around 1€ per sample) but requires the use of real-time PCR machines specially calibrated for this type of experiments and the analysis software. As a resume, although prices differ between laboratories or countries, our rough estimate of the cost points to first and second approaches as 13 and 8 times more expensive, respectively, in terms of reagents and consumables than the protocol proposed in this work (Table S2).
Practical advantages of PMC2 genotyping over classical phenotyping may be illustrated by the following example (Figure 1). The estimated time needed for evaluating 1000 samples at the IVIA’s facilities using bioassays is about 16 months (500 samples/8 months), taking into account that plants should be big enough to be ready-to-graft (approximately 2 years old). In contrast, just about 4 weeks are needed to conduct PMC2 genotyping just after seed germination. This estimated time was calculated assuming a 40-h workweek. As 1000 samples could be distributed into 10.4 96-well plates, ideally the DNA extraction would need 5.2 days (4 plates each 2 days), the 2 allele-specific PCRs would need 7.8 days (3 h each plate) and the agarose electrophoresis would last 2.6 days (2 PCR 96-well plates and 2 h per gel). In total, we would need 15.6 working days to genotype 1000 samples. This improvement removes the phenotyping bottleneck since all seedlings obtained from a particular cross can be PCR screened that same year. Hence, this quick and high-throughput method for DNA testing is expected to have an important effect on the cost efficiency of MAS, as suggested by Edge-Garza et al. [37].

3.2. ParPMC2-del Highly Correlates with PPV Resistance in Apricot Germplasm

One of the main pillars of plant breeding relies on skilful parental selection to create new genetic variation by controlled crossing. Usually, breeders just connect the concept of DNA-informed breeding with the use of molecular markers for seedling selection, but it also can be very helpful for parental selection [36]. This is the case in apricot breeding for PPV resistance. Two decades ago, Martínez-Gómez et al. [9] reviewed phenotypic information regarding apricot cultivar behaviour against PPV. Similarly, here, we compile the PMC2 genotype of a wide set of apricot accessions to facilitate parental selection tasks incorporating also their resistance phenotype, pedigree and origin data from the literature when available. The PPV strain used for phenotyping was also included because differences in severity of the induced symptoms have been observed [10,16]. As a result, after screening 120 accessions by PCR and other 49 by WGS and reviewing the available literature, PMC2 genotype was determined in a total of 325 apricot cultivars or accessions that represent a wide range of geographic origins (Figure 3). A significant part of the materials come from European countries directly involved in PPV resistance research during the last decades, such as Italy (20.9%), Spain (15.7%) or France (14.8%) [38,39,40,41,42]. Regarding viral strain, PPV-M was more frequently used for phenotyping except for PPV-D in Spain and PPV-T in Turkey (Figure 3), in agreement with the prevalence of these two strains in every country [5,43].
In total, 110 accessions were considered phenotypically resistant (Table 1), 108 were susceptible (Table 2) and 11 showed uncertain phenotype against the same or different PPV strains (Table 3). ParPMC2-del highly correlates with PPV resistance, as evidenced by its presence in 92.8% of the resistant accessions (Table 1) and its absence in 92.6% of the susceptible accessions (Table 2). Only 16 out of 219 (7.3%) accessions phenotypically classified as resistant or susceptible showed genotype-phenotype incongruences (GPIs). GPIs were previously reported mainly when using segregating populations [18,23,24,28,44], but clarifying reasons underlying GPIs was found difficult, as quite different factors may be involved. These factors include complex phenotyping protocols, loci other than PPVres contributing to PPV resistance, environmental conditions and/or gene–environment interactions. Additionally, putative misclassifications could also explain some genotypic discrepancies observed in this work. For instance, Sunglo, the resistant donor parent of Goldrich, has been phenotyped as resistant by several authors using PPV-M [15,45,46] and PPV-D [47] and genotypically showed the SSR-resistant alleles targeting the PPVres locus [18]. However, WGS data (SRR2153157) supposedly corresponding to this accession do not have the ParPMC2-del. Something similar occurs with Mirlo Naranja, classified as resistant [48], that was found to carry one copy of the ParPMC2-del by PCR in this work but not in that of Passaro [49]. Detailed accession documentation may be helpful to resolve these discrepancies, but 13 of the 16 identified GPIs have no pedigree data available. This information would be very valuable to increase the efficiency of apricot breeding programs and germplasm management.
Accurate evaluation of PPV resistance is a complex process, and results obtained by different researchers sometimes are contradictory, as exemplified by Farbaly and Pieve (Table 3), which may lead to GPIs. This problem is also observed in well-known accessions. For instance, Goldrich, usually classified as resistant against both PPV-D and M strains, has also been classified as uncertain or even as susceptible at least once (Table 3). Moreover, the effect of the PPV strain used [9,24] has also been observed, as at least 5 accessions showed different behaviour against PPV-M, D or T infection (Table 3). In addition, the environmental effect on symptoms and the different PPV detection techniques employed could also been involved in GPIs [9].
On the other hand, PPV resistance has been related with the downregulation of both ParPMC2 and, especially, ParPMC1, putatively due to an RNA silencing mechanism triggered by the pseudogenization of ParPMC2res [21]. Notwithstanding, the presence of epigenetic changes has also been suggested as a possible cause [22]. In any case, resistant cultivars show residual expression levels that could somehow be influenced by environmental conditions. This might explain sporadic symptoms that eventually lead to GPI classification. Moreover, the role of additional PPV resistance loci or genes may also contribute to GPIs. In this sense, Gallois et al. [105] pointed out that a large part of a resistant phenotype conferred by a given QTL depends on the genetic background due to frequent epistatic effects between resistance genes. In fact, other minor loci, linked or not to PPVres, have been suggested to underlie PPV resistance in apricot [13,14,15,16]. Altogether, the identification and/or confirmation of GPIs in this work pave the way for future studies to unravel the PPV resistance mechanism.
The handful of North American cultivars originally described as PPV resistant [9] have been extensively used as donors in all breeding programs currently in progress. As a result, the PPVres locus has been introduced in different genetic backgrounds. In order to complete our survey, genotypic information was compiled from other 96 accessions without available PPV phenotype data (Table S1, [107,108,109,110,111,112,113]). In summary, 152 accessions (46.8%) have at least one copy of the ParPMC2-del (Figure 3) and 15 out of them are homozygous for ParPMC2-del, including the North American PPV-resistant cultivar Stella [114]. Those materials derived from crosses with North American PPV-resistant cultivars represent an opportunity to accelerate the development of new varieties better adapted to the Mediterranean basin conditions [9]. In this context, it should be highlighted that MAS allows to improve cost efficiency and/or genetic gain in apricot breeding programs aimed to select PPV-resistant seedlings. This improvement is highly significant even if some PPV susceptible individuals among those with ParPMC2-del are dragged, since they will be later identified by PPV phenotyping. Similarly, Tartarini et al. [115] underlined the advantage of the identification of homozygous Rvi6 scab-resistant plants using MAS, despite segregating progenies showing at least 5% of GPIs.

4. Conclusions

Here, we present a high-throughput method to quickly perform DNA testing for PPV resistance that may greatly improve the efficiency of apricot breeding programs. The long-lasting PPV phenotyping process will only be performed with those advanced selections showing promising agronomic behaviour in advanced stages to guarantee the selection of PPV-resistant individuals. Additionally, a wide survey over 300 accessions has been made to identify PPV-resistant sources that could also be useful in apricot breeding programs.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4395/10/9/1292/s1, Table S1. PMC2 genotyped apricot accessions without phenotypic data against PPV infection; Table S2. Estimation cost of DNA extraction and PMC genotyping for PPV MAS in apricot.

Author Contributions

Conceptualization: C.R., M.L.B. and E.Z.; experimental procedures: Á.P.-O., I.L. and E.Z.; bioinformatics: E.Z.; funding acquisition: M.L.B.; writing—original draft, E.Z.; writing—review and editing, Á.P.-O., C.R., I.L., M.L.B. and E.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Instituto Nacional de Investigación y Tecnología Agraria y Alimentaria (INIA)-FEDER (RTA2017-00011-C03-01). Á.P.O. was funded by a fellowship cofinanced by the Generalitat Valenciana and European Social Fund (2014–2020) (DOCV 8426/19.11.2018).

Acknowledgments

The authors would like to express their gratitude to Bassi (University of Milan, Italy) for providing pedigree information from their apricot breeding program.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Rehder, A. Manual of Cultivated Trees and Shrubs Hardy in North America, 2nd ed.; The Macmillan Company: New York, NY, USA, 1940. [Google Scholar]
  2. Zhebentyayeva, T.N.; Ledbetter, C.; Burgos, L.; Llácer, G. Apricots. In Fruit Breeding, 1st ed.; Badenes, M.L., Byrne, D.H., Eds.; Springer: New York, NY, USA, 2012; Volume 3, pp. 415–458. [Google Scholar] [CrossRef]
  3. Moreno, M.A. Breeding and selection of Prunus rootstocks at the Aula Dei experimental station, Zaragoza, Spain. Acta Hortic. 2004, 658, 519–528. [Google Scholar] [CrossRef]
  4. García, J.A.; Cambra, M. Plum pox virus and sharka disease. Plant. Viruses 2007, 1, 69–79. [Google Scholar]
  5. García, J.A.; Glasa, M.; Cambra, M.; Candresse, T. Plum pox virus and sharka: A model potyvirus and a major disease. Mol. Plant. Pathol. 2014, 15, 226–241. [Google Scholar] [CrossRef]
  6. Chirkov, S.; Ivanov, P.; Sheveleva, A.; Zakubanskiy, A.; Osipov, G. New highly divergent Plum pox virus isolates infecting sour cherry in Russia. Virology 2017, 502, 56–62. [Google Scholar] [CrossRef]
  7. James, D.; Varga, A.; Sanderson, D. Genetic diversity of Plum pox virus: Strains, disease and related challenges for control. Can. J. Plant. Pathol. 2013, 35, 431–441. [Google Scholar] [CrossRef]
  8. Sihelská, N.; Glasa, M.; Šubr, Z.W. Host preference of the major strains of Plum pox virus—Opinions based on regional and world-wide sequence data. J. Integr. Agric. 2017, 16, 510–515. [Google Scholar] [CrossRef]
  9. Martínez-Gómez, P.; Dicenta, F.; Audergon, J.M. Behaviour of apricot (Prunus armeniaca L.) cultivars in the presence of Sharka (Plum pox potyvirus): A review. Agronomie 2000, 20, 407–422. [Google Scholar] [CrossRef] [Green Version]
  10. Dondini, L.; Lain, O.; Vendramin, V.; Rizzo, M.; Vivoli, D.; Adami, M.; Guidarelli, M.; Gaiotti, F.; Palmisano, F.; Bazzoni, A.; et al. Identification of QTL for resistance to Plum pox virus strains M and D in Lito and Harcot apricot cultivars. Mol. Breed. 2011, 27, 289–299. [Google Scholar] [CrossRef]
  11. Hurtado, M.A.; Romero, C.; Vilanova, S.; Abbott, A.G.; Llácer, G.; Badenes, M.L. Genetic linkage maps of two apricot cultivars (Prunus armeniaca L.) and mapping of PPV (sharka) resistance. Theor. Appl. Genet. 2002, 105, 182–191. [Google Scholar] [CrossRef]
  12. Lalli, D.A.; Abbott, A.G.; Zhebentyayeva, T.N.; Badenes, M.L.; Damsteegt, V.; Polák, J.; Krška, B.; Salava, J. A genetic linkage map for an apricot (Prunus armeniaca L.) BC1 population mapping Plum pox virus resistance. Tree Genet. Genomes 2008, 4, 481–493. [Google Scholar] [CrossRef]
  13. Lambert, P.; Dicenta, F.; Rubio, M.; Audergon, J.M. QTL analysis of resistance to sharka disease in the apricot (Prunus armeniaca L.) ‘Polonais’ x ‘Stark Early Orange’ F1 progeny. Tree Genet. Genomes 2007, 3, 299–309. [Google Scholar] [CrossRef]
  14. Marandel, G.; Pascal, T.; Candresse, T.; Decroocq, V. Quantitative resistance to Plum pox virus in Prunus davidiana P1908 linked to components of the eukaryotic translation initiation complex. Plant. Pathol. 2009, 58, 425–435. [Google Scholar] [CrossRef]
  15. Mariette, S.; Wong Jun Tai, F.; Roch, G.; Barre, A.; Chague, A.; Decroocq, S.; Groppi, A.; Laizet, Y.; Lambert, P.; Tricon, D.; et al. Genome-wide association links candidate genes to resistance to Plum pox virus in apricot (Prunus armeniaca). New Phytol. 2016, 209, 773–784. [Google Scholar] [CrossRef] [Green Version]
  16. Pilarova, P.; Marandel, G.; Decroocq, V.; Salava, J.; Krška, B.; Abbott, A.G. Quantitative trait analysis of resistance to Plum pox virus in the apricot F1 progeny ‘Harlayne’ x ‘Vestar’. Tree Genet. Genomes 2010, 6, 467–475. [Google Scholar] [CrossRef]
  17. Soriano, J.M.; Vera-Ruiz, E.; Vilanova, S.; Martínez-Calvo, J.; Llácer, G.; Badenes, M.L.; Romero, C. Identification and mapping of a locus conferring Plum pox virus resistance in two apricot-improved linkage maps. Tree Genet. Genomes 2008, 4, 391–402. [Google Scholar] [CrossRef] [Green Version]
  18. Soriano, J.M.; Domingo, M.L.; Zuriaga, E.; Romero, C.; Zhebentyayeva, T.; Abbott, A.G.; Badenes, M.L. Identification of simple sequence repeat markers tightly linked to Plum pox virus resistance in apricot. Mol. Breed. 2012, 30, 1017–1026. [Google Scholar] [CrossRef]
  19. Zhebentyayeva, T.N.; Reighard, G.L.; Lalli, D.; Gorina, V.M.; Krška, B.; Abbott, A.G. Origin of resistance to Plum pox virus in apricot: What new AFLP and targeted SSR data analyses tell. Tree Genet. Genomes 2008, 4, 403–417. [Google Scholar] [CrossRef]
  20. Zuriaga, E.; Soriano, J.M.; Zhebentyayeva, T.; Romero, C.; Dardick, C.; Cañizares, J.; Badenes, M.L. Genomic analysis reveals MATH gene(s) as candidate(s) for Plum pox virus (PPV) resistance in apricot (Prunus armeniaca L.). Mol. Plant Pathol. 2013, 14, 663–677. [Google Scholar] [CrossRef]
  21. Zuriaga, E.; Romero, C.; Blanca, J.M.; Badenes, M.L. Resistance to Plum pox virus (PPV) in apricot (Prunus armeniaca L.) is associated with down-regulation of two MATHd genes. BMC Plant Biol. 2018, 18, 25. [Google Scholar] [CrossRef] [Green Version]
  22. Rodamilans, B.; Valli, A.; García, J.A. Molecular plant-Plum pox virus interactions. Mol. Plant. Microbe Interact. 2020, 33, 6–17. [Google Scholar] [CrossRef]
  23. Rubio, M.; Ruiz, D.; Egea, J.; Martínez-Gómez, P.; Dicenta, F. Opportunities of marker assisted selection for Plum pox virus resistance in apricot breeding programs. Tree Genet. Genomes 2014, 10, 513–525. [Google Scholar] [CrossRef]
  24. Decroocq, S.; Chague, A.; Lambert, P.; Roch, G.; Audergon, J.M.; Geuna, F.; Chiozzotto, R.; Bassi, D.; Dondini, L.; Tartarini, S.; et al. Selecting with markers linked to the PPVres major QTL is not sufficient to predict resistance to Plum pox virus (PPV) in apricot. Tree Genet. Genomes 2014, 10, 1161–1170. [Google Scholar] [CrossRef]
  25. Moustafa, T.A.; Badenes, M.L.; Martínez-Calvo, J.; Llácer, G. Determination of resistance to sharka (plum pox) virus in apricot. Sci. Hort. 2001, 91, 59–70. [Google Scholar] [CrossRef]
  26. Lommel, S.A.; McCain, A.H.; Morris, T.J. Evaluation of indirect-linked immunosorbent assay for the detection of plant viruses. Phytopathology 1982, 72, 1018–1022. [Google Scholar] [CrossRef]
  27. Wetzel, T.; Candresse, T.; Ravelonandro, M.; Dunez, J. A polymerase chain reaction assay adapted to plum pox potyvirus detection. J. Virol. Methods 1991, 33, 355–365. [Google Scholar] [CrossRef]
  28. Passaro, M.; Geuna, F.; Bassi, D.; Cirilli, M. Development of a high-resolution melting approach for reliable and cost-effective genotyping of PPVres locus in apricot (P. armeniaca). Mol. Breed. 2017, 37, 74. [Google Scholar] [CrossRef]
  29. Ru, S.; Main, D.; Evans, K.; Peace, C. Current applications, challenges, and perspectives of marker-assisted seedling selection in Rosaceae tree fruit breeding. Tree Genet. Genomes 2015, 11, 8. [Google Scholar] [CrossRef]
  30. Doyle, J.J.; Doyle, J.L. A rapid isolation procedure for small quantities of fresh leaf tissue. Phytochem. Bull. 1987, 19, 11–15. [Google Scholar]
  31. Muñoz-Sanz, J.V.; Zuriaga, E.; Badenes, M.L.; Romero, C. A disulfide bond A-like oxidoreductase is a strong candidate gene for self-incompatibility in apricot (Prunus armeniaca) pollen. J. Exp. Bot. 2017, 68, 5069–5078. [Google Scholar] [CrossRef] [Green Version]
  32. Haas, B.J.; Papanicolaou, A.; Yassour, M.; Grabherr, M.; Blood, P.D.; Bowden, J.; Couger, M.B.; Eccles, D.; Li, B.; Lieber, M.; et al. De novo transcript sequence reconstruction from RNA-seq using the trinity platform for reference generation and analysis. Nat. Protoc. 2013, 8, 1494–1512. [Google Scholar] [CrossRef]
  33. Verde, I.; Jenkins, J.; Dondini, L.; Micali, S.; Pagliarani, G.; Vendramin, E.; Paris, R.; Aramini, V.; Gazza, L.; Rossini, L.; et al. The Peach v2.0 release: High-resolution linkage mapping and deep resequencing improve chromosome-scale assembly and contiguity. BMC Genom. 2017, 18, 225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Langmead, B.; Salzberg, S.L. Fast gapped-read alignment with bowtie 2. Nat. Meth. 2012, 9, 357–359. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Thorvaldsdóttir, H.; Robinson, J.T.; Mesirov, J.P. Integrative Genomics Viewer (IGV): High-performance genomics data visualization and exploration. Brief. Bioinform. 2013, 14, 178–192. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Peace, C. DNA-informed breeding of rosaceous crops: Promises, progress and prospects. Hortic. Res. 2017, 4, 17006. [Google Scholar] [CrossRef] [Green Version]
  37. Edge-Garza, D.A.; Luby, J.J.; Peace, C. Decision support for cost-efficient and logistically feasible marker-assisted seedling selection in fruit breeding. Mol. Breed. 2015, 35, 223. [Google Scholar] [CrossRef]
  38. Audergon, J.M.; Blanc, A.; Gilles, F.; Broquaire, J.M.; Clauzel, G.; Gouble, B.; Grotte, M.; Reich, M.; Bureau, S.; Pitiot, C. New recent selections issued from INRA’s apricot breeding programme. Acta Hortic. 2010, 862, 179–182. [Google Scholar] [CrossRef]
  39. Bassi, D.; Audergon, J.M. Apricot breeding: Update and perspectives. Acta Hortic. 2006, 701, 279–294. [Google Scholar] [CrossRef]
  40. Bassi, D.; Bellini, E.; Guerriero, R.; Monastra, F.; Pennone, F. Apricot breeding in Italy. Acta Hortic. 1995, 384, 47–54. [Google Scholar] [CrossRef]
  41. Egea, J.; Dicenta, F.; Burgos, L.; Martínez-Gómez, P.; Rubio, M.; Campoy, J.A.; Ortega, E.; Patiño, J.L.; Nortes, L.; Molina, A.; et al. New apricot cultivars from CEBAS-CSIC (Murcia, Spain) breeding programme. Acta Hortic. 2010, 862, 113–118. [Google Scholar] [CrossRef]
  42. Martínez-Calvo, J.; Font, A.; Llácer, G.; Badenes, M.L. Apricot and Peach breeding programs from the IVIA. Acta Hortic. 2009, 814, 185–188. [Google Scholar] [CrossRef]
  43. Gürcan, K.; Ceylan, A. Strain identification and sequence variability of Plum pox virus in Turkey. Turkish J. Agric. 2016, 40, 746–760. [Google Scholar] [CrossRef]
  44. Decroocq, S.; Cornille, A.; Tricon, D.; Babayeva, S.; Chague, A.; Eyquard, J.P.; Karychev, R.; Dolgikh, S.; Kostritsyna, T.; Li, S.; et al. New insights into the history of domesticated and wild apricots and its contribution to Plum pox virus resistance. Mol. Ecol. 2016, 25, 4712–4729. [Google Scholar] [CrossRef] [PubMed]
  45. Dosba, F.; Orliac, S.; Dutranoy, F.; Maison, P.; Massonie, G.; Audergon, J.M. Evaluation of resistance to Plum pox virus in apricot trees. Acta Hortic. 1992, 309, 211–219. [Google Scholar] [CrossRef]
  46. Karayiannis, I.; Mainou, A. Resistance to Plum pox virus in apricots. EPPO Bull. 1994, 24, 761–766. [Google Scholar] [CrossRef]
  47. Martínez-Gómez, P.; Rubio, M.; Dicenta, F. Evaluation of resistance to Plum pox virus of North American and European apricot cultivars. HortScience 2003, 38, 568–569. [Google Scholar] [CrossRef] [Green Version]
  48. Egea, J.; Rubio, M.; Campoy, J.A.; Dicenta, F.; Ortega, E.; Nortes, M.D.; Martínez-Gómez, P.; Molina, A.; Molina, A., Jr.; Ruiz, D. ‘Mirlo Blanco’, ‘Mirlo Anaranjado’, and ‘Mirlo Rojo’: Three new very early-season apricots for the fresh market. HortScience 2010, 45, 1893–1894. [Google Scholar] [CrossRef] [Green Version]
  49. Passaro, M. Cost-Effective Use of Molecular Markers in the Practical Resolution of Common Horticultural Challenges. Ph.D. Thesis, Agriculture, Environment and Bioenergy in Universitá Degli Studi Di Milano, Milan, Italy, 2016. [Google Scholar]
  50. Krška, B.; Salava, J.; Polák, J. Breeding for resistance: Breeding for Plum pox virus resistant apricots (Prunus armeniaca L.) in the Czech Republic. EPPO Bull. 2006, 36, 330–331. [Google Scholar] [CrossRef]
  51. Krška, B.; Salava, J.; Polák, J.; Komínek, P. Genetics of resistance to Plum pox virus in apricot. Plant Protect. Sci. 2002, 38, 180–182. [Google Scholar] [CrossRef] [Green Version]
  52. Krška, B.; Vachun, Z.; Nečas, T.; Ondrásek, I. New Sharka resistant apricots at the Horticultural Faculty in Lednice. Acta Hortic. 2015, 1063, 105–110. [Google Scholar] [CrossRef]
  53. Rankovic, M.; Duli-Markovic, I.; Paunovic, S. Sharka virus in apricot and its diagnosis. Acta Hortic. 1999, 488, 783–786. [Google Scholar] [CrossRef]
  54. CEP INNOVATION Website. Available online: https://cepinnovation-novadi.com/variete/anegat/ (accessed on 29 July 2020).
  55. Milatović, D.; Nikolić, D.; Krška, B. Testing of self-(in)compatibility in apricot cultivars from European breeding programmes. HortScience 2013, 40, 65–71. [Google Scholar] [CrossRef] [Green Version]
  56. Brans, Y. Evaluation de la sensibilité de cultivars d’abricotier à la Sharka en zone confinée. Présentation de l’essai Ctifl 2012–2015. In Proceedings of the Rencontres Phytosanitaires Ctifl/DGAL—SDQPV Fruits à Noyau Ctifl Balandran, Bellegarde, France, 16 October 2016. [Google Scholar]
  57. Babini, A.R.; Vicchi, V.; Missere, D. L’esame delle cultivar tolleranti alla sharka. Ermes Agricoltura 2010. Available online: http://www.crpv.it/doc/549738/DLFE-9612.pdf (accessed on 29 August 2020).
  58. Finn, C.E.; Clark, J.R. Register of New Fruit and Nut Cultivars List 44. HortScience 2008, 43, 1321–1343. [Google Scholar] [CrossRef] [Green Version]
  59. Krška, B.; Vachůn, Z. Apricot Breeding at the Faculty of Horticulture in Lednice. Agronomy 2016, 6, 27. [Google Scholar] [CrossRef] [Green Version]
  60. Krška, B. Genetic resources of apricot for adaptability improvement and breeding. Acta Hortic. 2010, 862, 203–208. [Google Scholar] [CrossRef]
  61. Gürcan, K.; Çetinsağ, N.; Pınar, H.; Macit, T. Molecular and biological assessment reveals sources of resistance to Plum pox virus—Turkey strain in Turkish apricot (Prunus armeniaca) germplasm. Sci. Hortic. 2019, 252, 348–353. [Google Scholar] [CrossRef]
  62. Drogoudi, P. ΕΥΠAΘΕΙA ΠOΙΚΙΛΙΩΝ ΒΕΡΙΚOΚΙAΣ ΣΤOΝ ΙO ΤHΣ ΕΥΛOΓΙAΣ ΤHΣ ΔAΜAΣΚHΝΙAΣ (PPV) (Sensitivity of apricot varieties against PPV). Π. Δρογούδη 28o Συνέδριο Ελληνικής Εταιρείας πιστήμης Oπωροκηπευτικών ’50 χρόνια από την ίδρυση της ΕΕΕO’ . In Proceedings of the 28th Conference of the Hellenic Fruit and Vegetable Credit Society ‘50 Years Since the Founding of EEEO’, Thessaloniki, Greece, 16–20 October 2017. [Google Scholar]
  63. Goffreda, J.C.; Voordeckers, A.; Butenis-Vorsa, L.; Cowgill, W.P., Jr.; Maletta, M.H.; Frecon, J.L. NJA53 apricot. HortScience 1995, 30, 389–390. [Google Scholar] [CrossRef]
  64. Faggioli, F.; Barba, M. Screening of stone fruit germplasm for resistance to Plum Pox Potyvirus. Adv. Hortic. Sci. 1996, 10, 91–94. [Google Scholar]
  65. Poggi Pollini, C.; Bianchi, L.; Babini, A.; Vicchi, V.; Liverani, A.; Brandi, F.; Giunchedi, L.; Autonell, C.; Ratti, C. Evaluation of Plum pox virus infection on different stone fruit tree varieties. J. Plant. Pathol. 2008, 90, S27–S31. [Google Scholar]
  66. Brooks, R.M.; Olmo, H.P. The Brooks and Olmo Register of Fruit and Nut Varieties, 3rd ed.; ASHS Press: Alexandria, VA, USA, 1997. [Google Scholar]
  67. Leccese, A.; Bartolini, S.; Viti, R. Genotype harvest season, and cold storage influence on fruit quality and antioxidant properties of apricot. Int. J. Food Prop. 2012, 15, 864–879. [Google Scholar] [CrossRef]
  68. Fuchs, E.; Grünzig, M.; Kegler, H. Investigation on the Plum pox virus resistance in different apricot genotypes. Acta Virol. 1998, 42, 222–225. [Google Scholar]
  69. Polák, J.; Oukropec, I.; Krška, B.; Pívalová, J.; Miller, W. Difference in reactions of apricot and peach cultivars to Plum pox virus: Serological and symptomatological evaluation. HortScience 2003, 30, 129–134. [Google Scholar] [CrossRef] [Green Version]
  70. Audergon, J.M.; Dosba, F.; Karayiannis, I.; Dicenta, F. Amélioration de l’abricotier pour la résistance à la sharka. EPPO Bull. 1994, 24, 741–748. [Google Scholar] [CrossRef]
  71. Mesarović, J.; Trifković, J.; Tosti, T.; Akšić, M.F.; Milatović, D.; Ličina, V.; Milojković-Opsenica, D. Relationship between ripening time and sugar content of apricot (Prunus armeniaca L.) kernels. Acta Physiol. Plant. 2018, 40, 157. [Google Scholar] [CrossRef]
  72. Krška, B.; Oukropec, I.; Polak, J.; Kominek, P. The evaluation of apricot (Prunus armeniaca L.) cultivars and hybrids resistant to Sharka. Acta Hortic. 2000, 538, 143–146. [Google Scholar] [CrossRef]
  73. Salazar, J.A.; Rubio, M.; Ruiz, D.; Tartarini, S.; Martínez-Gómez, P.; Dondini, L. SNP development for genetic diversity analysis in apricot. Tree Genet. Genomes 2015, 11, 15. [Google Scholar] [CrossRef]
  74. Syrgiannidis, G.; Mainou, A. Two new apricot varieties resistant to Sharka (Plum pox virus) disease created by crossing. In Agriculure, Proceedings of the Programme de recherche Agrimed. Deuxiemes Rencontres sur L’abricotier, Avignon, France, 27–31 May 1991; CEC Commission of the European Communities: Luxembourg, 1993. [Google Scholar]
  75. Egea, J.; Ruiz, D.; Dicenta, F.; Burgos, L. Murciana apricot. HortScience 2005, 40, 254–255. [Google Scholar] [CrossRef]
  76. Ledbetter, C.A.; Peterson, S.J. ‘Apache’ and ‘Kettleman’: Two early season apricots for the fresh market. HortScience 2005, 40, 2202–2203. [Google Scholar] [CrossRef]
  77. Trandafirescu, M.; Dumitru, L.M.; Trandafirescu, I. Evaluating the Resistance to the Plum pox virus of Some Apricot Tree Cultivars and Hybrids in South-Eastern Romania. Proc. Latvian Acad. Sci. 2013, 67, 203–206. [Google Scholar] [CrossRef] [Green Version]
  78. Egea, J.; Ruiz, D.; Martínez-Gómez, P. Influence of rootstock on the productive behaviour of ‘Orange Red’ apricot under Mediterranean conditions. Fruits 2004, 59, 367–373. [Google Scholar] [CrossRef] [Green Version]
  79. Karayannis, I.; Di Terlizzi, B.; Audergon, J.M. Susceptibility of apricot cultivars to Plum pox virus. Acta Hortic. 1999, 488, 753–760. [Google Scholar] [CrossRef]
  80. Halász, J. Molecular Background of the S-locus Controlled Self-Incompatibility in Apricot. Ph.D. Thesis, Department of Genetics and Plant Breeding, Corvinus University, Budapest, Hungary, 2007. [Google Scholar]
  81. Ledbetter, C.A.; Ramming, D.W. Apricot cv. Robada. United. States Patent USPP9890P, 13 May 1997. [Google Scholar]
  82. Karayiannis, I.; Ledbetter, C.A. Susceptibility of certain apricot and plumcot cultivars to Plum pox virus infection. Acta Hortic. 2009, 825, 153–156. [Google Scholar] [CrossRef] [Green Version]
  83. Egea, J.; Dicenta, F.; Burgos, L. ‘Rojo Pasión’ Apricot. HortScience 2004, 39, 1490–1491. [Google Scholar] [CrossRef]
  84. Egea, J.; Dicenta, F.; Martínez-Gómez, P.; Burgos, L. Selene apricot. HortScience 2004, 39, 192–1493. [Google Scholar] [CrossRef]
  85. Syrgiannidis, G. Selection of two apricot varieties resistant to Sharka virus. Acta Phytopathol. Acad. Sci. Hung. 1980, 15, 85–87. [Google Scholar] [CrossRef]
  86. Adascalului, M.; Hoza, D.; Ion, L. Behaviour study for pollination a Romanian apricot varieties using different source of resistance to Sharka. J. Hortic. For. Biotechnol. 2014, 18, 13–17. [Google Scholar]
  87. Ion, L.; Asănică, A.; Moale, C. Studies of resistance to Sharka in several Romanian apricot progenies. In Proceedings of the International Conference on Chemical, Agricultural and Biological Sciences (CABS-2015), Istanbul, Turkey, 4–5 September 2015. [Google Scholar]
  88. Elibüyük, S.; Erdiller, G. The susceptibility of some apricot and plum varieties to Plum pox (sharka) virus. Acta Hortic. 1995, 384, 549–552. [Google Scholar] [CrossRef]
  89. Rodríguez, J.; Andrés, V.; Gil, L.; Martínez, J.; Hita, I. Sensibilidad a Sharka en Variedades de Albaricoquero de Murcia; Frutales Hueso Fund. La Caixa: Barcelona, Spain, 1995; pp. 56–64. [Google Scholar]
  90. Dosba, F.; Lansac, M.; Maison, P.; Massonie, G.; Audergon, J.M. Tolerance to Plum pox virus in apricot. Acta Hort. 1988, 235, 275–281. [Google Scholar] [CrossRef]
  91. Dondini, L.; Lain, O.; Geuna, F.; Banfi, R.; Gaiotti, F.; Tartarini, S.; Bassi, D.; Testolin, R. Development of a new SSR-based linkage map in apricot and analysis of synteny with existing Prunus maps. Tree Genet. Genomes 2007, 3, 239–249. [Google Scholar] [CrossRef]
  92. Martínez-Gómez, P.; Dicenta, F. Evaluation of resistance of apricot cultivars to a Spanish isolate of plum pox potyvirus (PPV). Plant. Breed. 2000, 119, 179–181. [Google Scholar] [CrossRef]
  93. Karayiannis, I. Susceptibility of apricots cultivars to Plum pox virus in Greece. Acta Hortic. 1989, 235, 271–274. [Google Scholar] [CrossRef]
  94. Balan, V.; Stoian, E. Susceptibility of certain apricot-tree to the Plum pox virus pathogenic action. Acta Hortic. 1995, 384, 565–569. [Google Scholar] [CrossRef]
  95. Avinent, L.; Hermoso de Mendoza, A.; Llácer, G.; García, S. Transmisión del virus de la sharka y sensibilidad varietal en albaricoquero. In Proceedings of the II Congreso Ibérico Ciencias Hortícolas, Zaragoza, Spain, 27–30 April 1993. [Google Scholar]
  96. Maghuly, F.; Borroto-Fernandez, E.; Ruthner, S.; Pedryc, A.; Laimer, M. Microsatellite variability in apricots (Prunus armeniaca L.) reflects their geographic origin and breeding history. Tree Genet. Genomes 2005, 1, 151–165. [Google Scholar] [CrossRef]
  97. Audergon, J.M.; Morvan, G.; Dicenta, F.; Chastelliere, G.; Karayiannis, I. A method to determine the susceptibility of apricot to Plum pox virus. Acta Hortic. 1995, 384, 575–579. [Google Scholar] [CrossRef]
  98. Egea, J.; Burgos, L. Detecting cross-incompatibility of three North American apricot cultivars and establishing the first incompatibility group in apricot. J. Am. Soc. Hort. Sci. 1996, 121, 1002–1005. [Google Scholar] [CrossRef] [Green Version]
  99. McLaren, J. Apricot tree, ‘F168 cv’. United. States Patent USPP16071P2, 25 October 2005. [Google Scholar]
  100. Dicenta, F.; Audergon, J.M. Localization of Plum pox virus (PPV) in tissues of susceptible and resistant apricot cultivars. Phytopathol. Med. 1995, 34, 83–87. [Google Scholar]
  101. Lachkar, A.; Mlika, M. New apricot varieties selected from the Tunisian breeding programme. Acta Hort. 2006, 717, 189–192. [Google Scholar] [CrossRef]
  102. Dosba, F.; Denise, F.; Audergon, J.M.; Maison, P.; Massonie, G. Plum pox virus resistance of apricot. Acta Hortic. 1991, 293, 569–580. [Google Scholar] [CrossRef]
  103. Egea, J.; Campoy, J.A.; Dicenta, F.; Burgos, L.; Patiño, J.L.; Ruiz, D. ‘Estrella’ and ‘Sublime’ apricot cultivars. HortScience 2009, 44, 469–470. [Google Scholar] [CrossRef] [Green Version]
  104. Eynard, A.; Roggero, P.; Lenzi, R.; Conti, M.; Milne, R.G. Test for pollen and seed transmission on Plum pox virus (Sharka) in two apricot cultivars. Adv. Hortic. Sci. 1991, 5, 104–106. [Google Scholar]
  105. Gallois, J.L.; Moury, B.; German-Retana, S. Role of the Genetic Background in Resistance to Plant Viruses. Int. J. Mol. Sci. 2018, 19, 2856. [Google Scholar] [CrossRef] [Green Version]
  106. Polák, J.; Kominek, P.; Jokes, M.; Oukropec, I.; Krška, B. The evaluation of resistance of apricots to Plum pox virus by ELISA and ISEM. Acta Hortic. 1995, 386, 285–289. [Google Scholar] [CrossRef]
  107. Layne, R.E.C.; Hunter, D.M. ‘AC Harostar’ apricot. HortScience 2003, 38, 140–141. [Google Scholar] [CrossRef] [Green Version]
  108. Hegedűs, A.; Lénárt, J.; Halász, J. Sexual incompatibility in Rosaceae fruit tree species: Molecular interactions and evolutionary dynamics. Biol. Plant 2012, 56, 201–209. [Google Scholar] [CrossRef]
  109. Austin, P.T. Pollination of Sundrop Apricot. Ph.D. Thesis, Massey University, Auckland, New Zealand, 1995. [Google Scholar]
  110. Egea, J.; Ruiz, D.; Burgos, L. “Dorada” apricot. HortScience 2005, 40, 1919–1920. [Google Scholar] [CrossRef] [Green Version]
  111. Muñoz-Sanz, J.V.; Zuriaga, E.; López, I.; Badenes, M.L.; Romero, C. Self-(in)compatibility in apricot germplasm is controlled by two major loci, S and M. BMC Plant. Biol. 2017, 17, 82. [Google Scholar] [CrossRef] [Green Version]
  112. Corrin, A.A. “Ruby” Apricot Tree. United. States Patent USPP8177, 16 March 1993. [Google Scholar]
  113. Zaiger, C.F. Apricot Tree (Spring Giant). United. States Patent USPP5138, 15 November 1983. [Google Scholar]
  114. Dicenta, F.; Audergon, J.M. Inheritance of resistance to plum pox potyvirus (PPV) in ‘Stella’ apricot seedlings. Plant. Breed. 1998, 117, 579–581. [Google Scholar] [CrossRef]
  115. Tartarini, S.; Sansavini, S.; Vinatzer, B.; Gennari, F.; Domizi, C. Efficiency of marker assisted selection (MAS) for the Vf scab resistance gene. Acta Hortic. 2000, 538, 549–552. [Google Scholar] [CrossRef]
Figure 1. Comparison between traditional Plum pox virus (PPV) resistance phenotyping (A) and high-throughput marker-assisted selection (MAS) based on PMC2 allele-specific PCR (B). (*) Estimated duration based on Instituto Valenciano de Investigaciones Agrarias (IVIA) facilities.
Figure 1. Comparison between traditional Plum pox virus (PPV) resistance phenotyping (A) and high-throughput marker-assisted selection (MAS) based on PMC2 allele-specific PCR (B). (*) Estimated duration based on Instituto Valenciano de Investigaciones Agrarias (IVIA) facilities.
Agronomy 10 01292 g001
Figure 2. PMC2 genotyping by allele-specific PCR using forward primers differing at the 3’-end (A): R-allele (B) and S-allele (C) amplifications in 1% agarose gel electrophoresis for 46 apricot accessions (1: Goldrich, 2: Harlayne, 3: Henderson, 4: Lito, 5: Orange Red, 6: Pandora, 7: SEO, 8: Stella, 9: Veecot, 10: Bebeco, 11: Bergeron, 12: Canino, 13: Currot, 14: Ginesta, 15: Katy, 16: Mitger, 17: Palau, 18: Tyrinthos, 19: Piera, 20: Selene, 21: Colorao, 22: Moixent, 23: Perla, 24: Dama Vermella, 25: Maravilla, 26: Ninfa, 27: Palabras, 28: Sublime, 29: Dorada, 30: Castlebrite, 31: Martinet, 32: Corbató, 33: Gandía, 34: Cristalí, 35: Manri, 36: Gavatxet, 37: Pisana, 38: Xirivello, 39: Velazquez, 40: Mirlo Rojo, 41: Rojo Carlet, 42: Bulida, 43: ASP, 44: Silvercot, 45: Bora and 46: Roxana).
Figure 2. PMC2 genotyping by allele-specific PCR using forward primers differing at the 3’-end (A): R-allele (B) and S-allele (C) amplifications in 1% agarose gel electrophoresis for 46 apricot accessions (1: Goldrich, 2: Harlayne, 3: Henderson, 4: Lito, 5: Orange Red, 6: Pandora, 7: SEO, 8: Stella, 9: Veecot, 10: Bebeco, 11: Bergeron, 12: Canino, 13: Currot, 14: Ginesta, 15: Katy, 16: Mitger, 17: Palau, 18: Tyrinthos, 19: Piera, 20: Selene, 21: Colorao, 22: Moixent, 23: Perla, 24: Dama Vermella, 25: Maravilla, 26: Ninfa, 27: Palabras, 28: Sublime, 29: Dorada, 30: Castlebrite, 31: Martinet, 32: Corbató, 33: Gandía, 34: Cristalí, 35: Manri, 36: Gavatxet, 37: Pisana, 38: Xirivello, 39: Velazquez, 40: Mirlo Rojo, 41: Rojo Carlet, 42: Bulida, 43: ASP, 44: Silvercot, 45: Bora and 46: Roxana).
Agronomy 10 01292 g002
Figure 3. Geographic distribution of apricot accessions: PMC2 genotypes (RR: homozygous for the resistant allele; SS: homozygous for the susceptible allele; and RS: heterozygous) and PPV strain used for phenotyping are also indicated.
Figure 3. Geographic distribution of apricot accessions: PMC2 genotypes (RR: homozygous for the resistant allele; SS: homozygous for the susceptible allele; and RS: heterozygous) and PPV strain used for phenotyping are also indicated.
Agronomy 10 01292 g003
Table 1. Apricot PPV-resistant accessions genotyped for PMC2.
Table 1. Apricot PPV-resistant accessions genotyped for PMC2.
NameCountry aOriginPedigreePPV Resistance Phenotype bPPV Strain UsedFirst Phenotype RefPMC2 Genotype cPMC2 Genotype Ref
A4316IT RM [15]RSWGS
A4804IT RM [15]RSWGS
Adriana (= Le-3241)CRHorticulture Faculty, LedniceVestar × SEO [50]RM [51]RR [24]
Rec [52]
Alfred (= NY345)USAGeneva, NY State Expt Sta, by Robert C. LambOP seedling of selection from (Doty × Geneva)RM [53]RSWGS
AndsweeIR RM [15]RSWGS
AnegatFRINRA, CEP Innovation RM/D [54]RS [49]
Bergarouge (= Avirine A2914)FRINRABergeron × Orange Red [55]RD [23]RS [49]
M [56]
Bergeval (= Aviclo, A3950)FRINRA RM [56]RS [49]
BO03615011IT Goldrich × Harlayne [28]RM* [49]RS [49]
BO03615025IT Goldrich × Harlayne [28]RM* [49]RR [49]
BO03615034IT Goldrich × Harlayne [28]RM* [28]RR [28]
BO03615049IT Goldrich × Harlayne [28]RM* [28]RR [28]
BO03615053IT Goldrich × Harlayne [28]RM* [28]RS [28]
BO03615070IT Goldrich × Harlayne [28]RM* [49]RR [49]
BO04624031IT Portici × Goldrich [28]RM* [28]RS [28]
BO04624039IT Portici × Goldrich [28]RM* [49]SS [49]
BO05636034IT Kyoto × Priscilla [28]RM* [28]RS [28]
BO06609012IT Silvercot × Bora [28]RM* [49]RS [49]
BO06609013IT Silvercot × Bora [28]RM* [49]RS [49]
BO06609024IT Silvercot × Bora [28]RM* [49]RS [49]
BO06609033IT Silvercot × Bora [28]RM* [49]RS [49]
BO06609036IT Silvercot × Bora [28]RM* [49]RS [49]
BO06609037IT Silvercot × Bora [28]RM* [49]RS [49]
BO06609039IT Silvercot × Bora [28]RM* [49]RS [49]
BO06609045IT Silvercot × Bora [28]RM* [49]RS [49]
BO06609048IT Silvercot × Bora [28]RM* [28]RS [28]
BO06609055IT Silvercot × Bora [28]RM* [28]RS [28]
BO06609060IT Silvercot × Bora [28]RM* [49]RS [49]
BO06609068IT Silvercot × Bora [28]RM* [49]RS [49]
BO06609074IT Silvercot × Bora [28]RM* [49]RS [49]
BO06609079IT Silvercot × Bora [28]RM* [49]RS [49]
BO06609083IT Silvercot × Bora [28]RM* [49]RS [49]
BO06609087IT Silvercot × Bora [28]RM* [49]RS [49]
BO06609099IT Silvercot × Bora [28]RM* [49]RS [49]
BO06609104IT Silvercot × Bora [28]RM* [49]RS [49]
BO06609113IT Silvercot × Bora [28]RM* [49]RS [49]
BO06609129IT Silvercot × Bora [28]RM* [49]RS [49]
BO06609133IT Silvercot × Bora [28]RM* [49]RS [49]
BO06609136IT Silvercot × Bora [28]RM* [49]RS [49]
BO96621002IT Goldrich × Lito [28]RM [57]RR [28]
BO96621030IT Goldrich × Lito [28]RM [57]RS [28]
Bora (BO90610010)ITUniversity of Bologna and Milan, by D. BassiEarly Blush × PA 7005-2 [58]RM/D [58]RS [21,28]
Candela (= LE-2927)CRHorticulture Faculty, LedniceHungarian Best × SEO [59]RM [60]RR [49]
CebirTU RT [61]RS [61]
CongatFRINRA, CEP Innovation R- [62]RS [49]
Early Blush (= RUTBHART, NJA53, Aurora46)USRutgers Horticultural Research Farm, New Brunswick, N.J.RR17–62 × NJA-13 [63]RD [64]RSPCR; [21,28,61]
M [65]
T [61]
FarlisFRMarie-France BOIS, France (IPS) RM* [28,49]RS [28]
Farmingdale (=NY346)USAGeneva, NY State Expt Sta, by Robert C. LambOP seedling of selection from (Doty × Geneva) [66]RM [53]RS [28]
Flavor cot (=Bayoto)USAWashington State University Research, by Tom Toyama RM [57]RS [28]
FlopriaFRPSB Producción Vegetal S.L. RM* [28]RSPCR; [28]
GG9310SPIVIA, Moncada, ValenciaGoldrich × Ginesta [18]RDIVIARSPCR
GG9318SPIVIA, Moncada, ValenciaGoldrich × Ginesta [18]RDIVIARSPCR
GG937SPIVIA, Moncada, ValenciaGoldrich × Ginesta [18]RDIVIARSPCR
GG941SPIVIA, Moncada, ValenciaGoldrich × Ginesta [18]RDIVIARSPCR
GG979SPIVIA, Moncada, ValenciaGoldrich × Ginesta [18]RDIVIARSPCR
GG9869SPIVIA, Moncada, ValenciaGoldrich × Ginesta [18]RDIVIARSPCR
GilgatFRINRA / CEP INNOVATION RM* [28]; [49]RS [28]
GP9817SPIVIA, Moncada, ValenciaGoldrich × Palau [18]RDIVIARSPCR
Dama Rosa (GG9871)SPIVIA, Moncada, ValenciaGoldrich × Ginesta [18]RDIVIARSPCR; [49]
Dama Taronja (GK988)SPIVIA, Moncada, ValenciaGoldrich × Katy [18]RDIVIARSPCR; [49]
DulcineaITPisa UniversityMoniqui OP [67]RD [64]SSPCR; [49]
FracassoIT TolerantT [61]SS [61]
HarlayneCAgr. Canada, Res. Station, Harrow, Ontario, by REC LayneV51092 ((Reliable × OP) × OP) × Sun Glo [66]RD [68]RSPCR; [20,21,24,28,61]
M [45]
Harval (=HW437)CAgr. Canada, Res. Station, Harrow, Ontario, by REC LayneVeecot × HW435 (Rouge du Roussillon × NJA2 (Morden604 OP)) [66]RM[69]RS [28]
HendersonUSAGeneva, NY, by GW HendersonUnknown [66]RM [46]RSPCR; [21]
D [70]
Kaniş (=M2252)TU RT [61]SS [61]
KarumTU RT [61]RS [61]
Lady cot (=HYB 3-3)FRCOT International RM* [28]RS [28]
LaycotC V51092 ((Reliable o.p.) o.p.) × NJA1 [71]RM [15]RRWGS
LE-2904CRHorticulture Faculty, LedniceVelkopavlovická × SEO [19]RM [72]RS [49]
LE-3205CRHorticulture Faculty, Lednice RM* [49]RR [49]
Le-3246CRHorticulture Faculty, LedniceVestar × SEO [51]RM [51]RS [24]
LE-3662CRHorticulture Faculty, Lednice RM [72]RR [49]
LifosTU RT [61]RS [61]
LillycotFRSDR Fruit Llc (US)Unknown [73]RM* [28]RS [28]
LitoGR SEO × Tirynthos [18]RM [74]RSPCR; [24,28]
DIVIA
Mediabel (=Mediabell)FRNewcot and IPS RM* [28]RS [28]
Mirlo Naranja (= Mirlo anaranjado)SPCEBAS-CSIC, MurciaRojo Pasión × Búlida Precoz [48]RD [48]RSPCR
SS [49]
Mirlo BlancoSPCEBAS-CSIC, MurciaRojo Pasión × Búlida Precoz [48]RD [48]RS [28]
Mirlo RojoSPCEBAS-CSIC, MurciaRojo Pasión × Búlida Precoz [48]RD [48]RSPCR; [49]
MogadorSPPSB Producción Vegetal S.L. RM* [28,49]RSPCR; [28]
Moixent (=GM961)SPIVIA, ValenciaGoldrich × Mitger [18]RDIVIARSPCR; [49]
MurcianaSPCEBAS-CSIC, MurciaOrange Red × Currot [73]RD [75]RSWGS; PCR; [49]
M [15]
NikitskiiUKR RM [15]RSWGS
NJA42USANew JerseyNJA12 × NJA13 [76]R? [77]RSPCR
Orange Red (=Barth; NJA-32)USANew JerseyLasgerdi Mashhad × NJA2 (= Morden 604 OP) [78]RD [68]RSPCR; [21]
M [79]
PandoraGR SEO × Tirynthos [18]RM [74]RSPCR
D [47]
Pelese di GiovannielloIT TolerantD [64]SS [49]
PerlaSPMurcia RD [64]SSPCR
Petra (BO88617102)ITUniversity of Bologna and Milan, Italy, by D BassiGoldrich × Pelese di Giovanniello [73]RM* [28]RS [28]
Precoce d’ImolaIT tolerantD [64]SSWGS
Priboto (=Zebra)FR bud mutation of Goldrich [80]RM [15]RSWGS; [49]
PriciaFRMarie-France BOIS, France (IPS) RM* [28,49]RS [28]
Pseudo RoyalUSA RM [15]RSWGS
Robada (= K106-2)USAParlier, CaliforniaOrange Red × K113-40 (ancestry includes Blenheim, Blush and Perfection) [81]RM [82]RSWGS
Rojo PasiónSPCEBAS-CSICOrange Red × Currot [83]RD [83]RSPCR; [49]
RosaSPCEBAS-CSIC, MurciaOrange Red × Palsteyn [73]RD [41]RS [49]
Tolerant [23]
RubistaFRMarie-France BOIS, France (IPS) RM* [28,49]RS [28]
Sabbatani (= Selezione Sabbatani?)IT RD [64]SS [49]
SeleneSPCEBAS-CSICGoldrich × A2564 (=Screara × SEO) [18]RD [84]RSPCR; [49]
SEOP934SPIVIASEO × Palau [18]RDIVIARSPCR
Spring Blush (= EA3126TH)FREscande EARL RM* [57]RR [49]
Stark Early Orange (= SEO, Earle Orange)USAGrandview, Washington, by WL RobertsUnknown [66]RM [85]RSPCR; [20,21,24,28,61]
D [70]
StellaUSA Unknown [18]RM [85]RRPCR; [21]
D [70]
Sunglo (= Sun Glo)USAColumbia & Okanogan Nursery Co.Unknown [66]RM [45]RSPCR
D [47]SSWGS
Sunnycot (= 97-3-203)USASDR FRUIT LLC – USA R [62]RS [49]
TraianRO RD [86]RSPCR; [87]
Tsunami (= EA 5016)FREscande EARL RM* [28]RS [28]
Wonder Cot (= RM 7)USASDR FRUIT LLC – USA RM* [28]RS [28]
ZardCA RT [61]RS [61]
M *: strain likely used for phenotyping by the Phytosanitary Service, Emilia-Romagna (Italy). a Countries: C: Canada, CA: Central Asia, CR: Czech Republic, FR: France, GR: Greece, IR: Iran, IT: Italy, RO: Romania, SP: Spain, TU: Tunisia, TR: Turkey, UKR: Ukraine, US: United States of America; b Phenotype: R: Resistant, S: Susceptible; c Genotype: RR: homozygous for PMC2 resistant allele, SS: homozygous for PMC2 susceptible allele, RS: heterozygous.
Table 2. Apricot PPV susceptible accessions genotyped for ParPMC2-del.
Table 2. Apricot PPV susceptible accessions genotyped for ParPMC2-del.
CultivarCountry aOriginPedigreePPV Resistance Phenotype bPPV Strain UsedFirst Phenotype RefPMC2 Genotype cPMC2 Genotype Ref
A3521IR SM [15]SSWGS
A3522IR SM [15]SSWGS
Amabile VecchoniITSeedling by Prof. F. ScaramuzziUnknown [67]SM [45]SS [49]
AprikozTR SM [88]SSPCR
ArroganteSPMurcia SD [89]SS [21]
Avikaline FR SM [15] SSWGS
Bebecou (Bebeco)GR Unknown [18]SM/D [90]SSPCR; [21,28]
Bella Di ImolaIT Spontaneous seedling [23]SD [64]SS [28]
BergeronFRSaint-Cyr-au-Mont-d’Or, LyonSpontaneous seedling [23]SM [90]SSPCR; [21]
Big Red (EA4006)FREscande EARL, France SM [57]RS [28]
BO04624042IT Portici × Goldrich [28]SM* [28]SS [28]
BO04624043IT Portici × Goldrich [28]SM* [28]SS [28]
BO06609003IT Silvercot × Bora [28]SM* [49]RS [49]
BO81604311IT San Castrese × Reale di Imola [73]SD [91]SS [24]
BO96621021IT Goldrich × Lito [28]SM* [28]RS [28]
Boucheran BoutardFR SM [15]SSWGS
BúlidaSPMurciaUnknown [73]SD [92]SSPCR; [21]
M [93]
Cafona IT Vesuvian area SM [94]SSWGS
D [64]
CAID AGDZ n2MO SM [15]SSWGS
CaninoSPValenciaUnknown [18]SD [95]SSPCR; [20,21]
M [90]
Castlebrite (=K111-6)USAUSDA, Fresno, CaliforniaOP seedling of B60-12 (= Perfection × Castleton) [66]SM [45]SSPCR
Ceglédi BíborHUCegléd Horticultural Research InstituteChance seedling [96] SM [46]SS [28]
Colorado (Colorao 43-15)SPPSB Producción Vegetal SLUnknown SM* [49]SSPCR; [28]
D [89]
CorbatóSPValencia SD [95]SSPCR
M [46]
CurrotSPValenciaUnknown [18]SD [95]SSPCR
M [46]
EstrellaSPCEBAS-CSICOrange Red × Z211-18 (= Goldrich × Pepito del Rubio) [23]SD [23]SSPCR; [49]
FaraliaFRMarie-France BOIS, IPS SM* [28]SS [28]
FarcloFRMarie-France BOIS, IPS SM [57]SS [28]
FavoritRO SM [94,97]SS [49]
Geç AbligozTR ST [61]SS [61]
GinestaSPValenciaUnknown [18]SD [95]SSPCR
M [46]
Dama Vermella (HG9869)SPIVIAHarcot × Ginesta [18]SDIVIASSPCR; [49]
HacıhaliloğluTR ST [61]SS [61]
Hargrand (= HW410)CRichard EC Layne, Agr. Canada, Res. StationV51092 ((Reliable × OP) × OP) × NJA1 (Phelps × Perfection) [66]SM [45]SS [21]
HasanbeyTR SM [45] ¹SSPCR
Hungarian Best = (Best of Hungary?)HU/RO ST [61]SS [61]
M [94]
KatyUSAZaiger’s Genetics⁴ SD [18]SSPCR; [21]
KrasnoshchekiiUKRAdvanced/improved cultivar SD [20]SS [20,21]
Kyoto (= Kioto)FREscandeUnknown [73]SM* [28]SS [28]
Lambertin-1USAUSDA, Fresno, CaliforniaA95-45 × B69-85 (=Perfection × Royal) [98]SM [45]SS [21]
Larclyd (= F168 cv; Jenny Cot)NZCentral OtagoSundrop × Moorpark [99]SM [15]SSWGS
Le-3218CRFaculty of Horticulture in LedniceVestar × SEO [51]SM [51]SS [24]
Luizet (= Suchet; Hatif du clos; Abricot du Clos)FR Spontaneous seedling [71]SM [93]SSWGS
LunaIT SM* [28]RS [28]
Madarska Narijlepsia SL SM [15] SSWGS
Magic cot (= RM 22) USASDR FRUIT LLC - USAUnknown [23]SD [23]SS [49]
Manicot FR SD [97,100]SSWGS
M [15]
MaravillaSPCEBAS-CSIC, MurciaOrange Red × Z211-18 (= Goldrich × Pepito) [23]SD [23]SSPCR; [49]
Mari de CenadRO UnknownS [86]RSPCR
MarkuleştiTR ST [61]SS [61]
MarlénCRHorticulture Faculty, Ledniceclone of Hungarian Best [59] SRec [14]SSPCR; [24]
Marouch 14 MO Local landraceSM [15]SSWGS
Marouch 4 MO SM [15] SSWGS
Mei Hwang CH Traditional cultivar/landraceSM [15] SSWGS
MektepTR ST [61]SS [61]
Mektep 8TR ST [61]SS [61]
MitgerSPCastellón [30] Unknown [18]SD [95]SSPCR
M [46]
Monaco BelloIT SM [97]SSWGS
MoniquiSPMurciaUnknownSM [90]SSPCR; [21,24]
MonoUSALe Grand, California, by FW AndersonPerfection OP [66] SM [93]SS [49]
Moongold (= Moongola?)USAUniversity of Minessota S- [77]SSPCR
Moorpark (=Moor Park)USA SM [46]SSWGS
Morden 604 CMorden, Manitoba, by Canada Dept. Agr. Res. Sta.Scout × McClure [66]SM [15] SSWGS
Ninfa (BO81602075)ITUniversity of Bologna and Milan, by D. BassiOuardy × Tyrinthos [55]SM* [28]SSPCR; [28,61]
T [61]
OlimpRO SM [45]SSWGS; [49]
Orange Rubis (=Couloumine)FRMallard SM [57]SS [28]
Ordubat B.TR ST [61]SS [61]
OuardiTUINRAT, ArianaCanino × Hamidi [101] SM [46]SS [49]
Palsteyn (Palstein)SA Blenhein × Canino [73]SM [102]SSWGS
PalabrasSP SD [95]SSPCR
M [46]
PalauSP Unknown [18]SD [95]SSPCR
Paviot FR SM [93]SSWGS
Peche De Nancy FR SM [15] SSWGS
PerfectionUSAWaterville, WashingtonUnknown [66]SM [46]SS [21]
Piera SM [65]RSPCR
Poizat FR SM [15] SSWGS
PolonaisFR Spontaneous seedling [23]SM [93]SS [24]
Poppy USAZaiger Genetics, Inc., Modesto, CA78EB575 × 123GD161 [58] SD [23]SS [49]
Portici (= Pertini)ITVesuvian areaUnknown; Local selection [23]SM [46]SSPCR; [28]
D [64]
Precoce AmpuisFR SM [15]SSWGS
Reale d’ImolaIT Luizet OP [23] SM [46]SS [21,24,49]
D [64]
Rojo de CarletSPValencia SD [95]SSPCR
M [46]
Rouge Du Roussillon FR SM [45]SSWGS
Rouge De FournesFR SM [15]SSWGS
SaturnRO SM [45]SSWGS
Screara FR SD [70]SSWGS
M [45]
Şekerpare B.TR ST [61]SS [61]
Shalakh (=Yerevani, Erevani)AR Local selection [23]SM [93]SSWGS; [20,21]
Silistra × Ananas (Marculesti 43/1)RO SM [15] SSWGS
Sucre De HolubHUBohême, by M. Holub SM [15]SSWGS
SublimeSPCEBAS-CSICOrange Red × Z211-18 (= Goldrich × Pepito del Rubio) [103] SD [103]SSPCR; [49]
Super Rouge FR SM [15] SSWGS
Sweet RedFR SM [57]SS [49]
Szegedi mamut (=Szegadti Mamut?)HU Foki István and Kovács ImreHybrid of Cegledi orias, "Giant" group [96] SM [94]SS [49]
TabrizTR S [86]SSPCR
Tadeo (= Taddeo)SPValencia SD [95]SSPCR
M [45]
Tardif De Bordaneil FR Unknown [23] SM [46]SSWGS
D [64]
Tardif De TainFR SM [15]SSWGS
Tonda di costiglioleITPiedmont S [104]SS [49]
TrevattAU SM [45]SSPCR
TyrinthosGR Unknown [18]SD [70]SSPCR; WGS; [49]
M [97]
UleanosSPUlea, Murcia SD [89]SS [49]
VelázquezSPMurcia SD [89]SSPCR; [21]
Venus (= Venus 1414?)RO (Umberto × Ananas) × (Luizet × Umberto) [96] SM [46]SS [49]
VestarCR Hungarian Best × mixture of pollen from Chinese cultivars [55]SM [105]RSWGS; [24]
VivagoldCVineland Station, OntarioVeecot × V49024 (= Geneva × Gibb) [66]SM [15]SSWGS
Xirivello (=Chirivello)SPValenciaUnknownSM [46]SSPCR
Yilbat (=M2243)TR ST [61]RS [61]
M *: strain likely used for phenotyping by the Phytosanitary Service, Emilia-Romagna (Italy). a Countries: AR: Armenia, AU: Australia, C: Canada, CH: China, CR: Czech Republic, FR: France, GR: Greece, HU: Hungary, IR: Iran, IT: Italy, MO: Morocco, NZ: New Zealand, RO: Romania, SA: South Africa, SL: Slovakia, SP: Spain, TR: Turkey, UKR: Ukraine and US: United States of America; b Phenotype: R: Resistant, S: Susceptible; c Genotype: RR: homozygous for PMC2 resistant allele, SS: homozygous for PMC2 susceptible allele and RS: heterozygous.
Table 3. Apricot accessions with uncertain PPV resistance phenotype genotyped for ParPMC2-del.
Table 3. Apricot accessions with uncertain PPV resistance phenotype genotyped for ParPMC2-del.
CultivarCountry aOriginPedigreePPV Resistance Phenotype bPPV Strain UsedFirst Phenotype RefPMC2 GenotypecPMC2 Genotype Ref
Badami IR SM [102]SSWGS
TD
FarbalyFRMarie-France BOIS, IPS SM* [28]RS [28]
RM [56]
GoldrichUSAUSDA and Washington State University, Prosser, WashingtonSun Glo × Perfection [73]RD [68]RSPCR; [20,21,24,28]
M [45]
uncertainM [106]
D [68]
SD [64]
HarcotC Agr. Canada, Res. Station, Harrow, Ontario, by REC Layne(T2 (Geneva × Naramata) × Morden 604 (Scout × McClure)) × NJA1 (Phelps × Perfection) [66]T? - [28]RSPCR; [21,24,28]
RM [90]
D [70]
ST [61]SS [61]
Incomparable de Malissard (= Valssard)FRMalissard, Valence RM [15]RSWGS
S [57]
PisanaIT ICAPI 26/5 OP [55]SM* [28]SSPCR; [28]
RM [65]
RD [64]
SD [28]
Pieve (BO89608015)ITUniversity of Bologna and Milan, by D. BassiHarcot × Reale di Imola [73]SM* [28]SS [28]
RM [65]
San CastreseITNaplesUnknown [73]TD [64]SSWGS; [49]
SM [46]
SulmonaRO (Luizet × Re Umberto) × (Ananas × Ananas) [71]SM [45]SS [49]
R- [77]
VeecotCOntario Dept Agr Res Inst, Vineland Station, Ontario, by OA BradtReliable OP [18]RM [45]RSPCR; [21]
S [105]
TD [47]
Viceroy (=Viceroy_603_G?)RO R- [77]SSPCR
S- [86]
M *: strain likely used for phenotyping by the Phytosanitary Service, Emilia-Romagna (Italy). a Countries: C: Canada, FR: France, IR: Iran, IT: Italy and RO: Romania; b Phenotype: R: Resistant and S: Susceptible; and c Genotype: RR: homozygous for PMC2 resistant allele, SS: homozygous for PMC2 susceptible allele and RS: heterozygous.

Share and Cite

MDPI and ACS Style

Polo-Oltra, Á.; Romero, C.; López, I.; Badenes, M.L.; Zuriaga, E. Cost-Effective and Time-Efficient Molecular Assisted Selection for PPV Resistance in Apricot Based on ParPMC2 Allele-Specific PCR. Agronomy 2020, 10, 1292. https://doi.org/10.3390/agronomy10091292

AMA Style

Polo-Oltra Á, Romero C, López I, Badenes ML, Zuriaga E. Cost-Effective and Time-Efficient Molecular Assisted Selection for PPV Resistance in Apricot Based on ParPMC2 Allele-Specific PCR. Agronomy. 2020; 10(9):1292. https://doi.org/10.3390/agronomy10091292

Chicago/Turabian Style

Polo-Oltra, Ángela, Carlos Romero, Inmaculada López, María Luisa Badenes, and Elena Zuriaga. 2020. "Cost-Effective and Time-Efficient Molecular Assisted Selection for PPV Resistance in Apricot Based on ParPMC2 Allele-Specific PCR" Agronomy 10, no. 9: 1292. https://doi.org/10.3390/agronomy10091292

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop