Abstract
Maintaining genetic diversity in cultured shellfish can be challenging due to high variance in individual reproductive success, founder effects, and rapid genetic drift, but is important to retain adaptive potential and avoid inbreeding depression. To support broodstock management and selective breeding in cultured Pacific oysters (Crassostrea (Magallana) gigas), we developed an amplicon panel targeting 592 genomic regions and SNP variants with an average of 50 amplicons per chromosome. Target SNPs were selected based on elevated observed heterozygosity or differentiation in Pacific oyster populations in British Columbia, Canada. The use of the panel for parentage applications was evaluated using multiple generations of oysters from a breeding program on Vancouver Island, Canada (n = 181) and families selected for Ostreid herpesvirus-1 resistance from the Molluscan Broodstock Program in Oregon, USA (n = 136). Population characterization was evaluated using wild, naturalized, farmed, or hatchery oysters sampled throughout the Northern Hemisphere (n = 189). Technical replicates showed high genotype concordance (97.5%; n = 68 replicates). Parentage analysis found suspected pedigree and sample handling errors, demonstrating the panel's value for quality control in breeding programs. Suspected null alleles were identified and found to be largely population dependent, suggesting population-specific variation impacting target amplification. Null alleles were identified using existing data without the need for pedigree information, and once they were removed, assignment rates increased to 93.0 and 86.0% of possible assignments in the two breeding program datasets. A pipeline for analyzing the amplicon sequence data from sequencer output, amplitools, is also provided.
Keywords: amplicon panel, aquaculture, Crassostrea (Magallana) gigas, genetic diversity, Pacific oyster, parentage assignment
Introduction
Sustainable aquaculture depends on the effective characterization, preservation, exchange and use of genetic resources (Guo 2009). High-resolution genomic tools (e.g. high-density genetic maps, SNP chips, chromosome-level reference genomes) enable genome-wide association studies to identify markers that can be used for marker-assisted or genomic selection (Boudry et al. 2021). Low resolution genetic tools (e.g. amplicon panels) enable high-throughput genotyping at low cost (Meek and Larson 2019). Low-cost applications can allow scalable use, such as in fisheries management through parentage-based tagging (e.g. Beacham et al. 2018), or in aquaculture to assign individuals back to putative family when mixed without physical tagging (Allen et al. 2020). Amplicon panels can provide rapid turnaround time (Arbelaez et al. 2019), and increased flexibility to use lower quality input DNA (Csernák et al. 2017); in general, these panels have been described as reliable, scalable, moderately easy to analyze, and cost-effective (Meek and Larson 2019). Cost-effectiveness can bring genetic advances to small or medium sized operations not able to afford high-density genotyping tools (Boudry et al. 2021; Delomas et al. 2023).
Pacific oyster Crassostrea (Magallana) gigas (Thunberg, 1793; Salvi and Mariottini 2017), is one of the most valuable aquaculture species globally (Botta et al. 2020). Oyster aquaculture is a significant economic driver and food source for many developed and developing nations (Martínez-García et al. 2022). However, production and growth of the industry outside of China has stagnated due to disease, regulatory issues, and other factors (Botta et al. 2020). Large-scale mortality events have been a significant burden on the industry, where disease may be caused by specific or combined protozoan, bacterial, or viral pathogens, as well as by multifactorial or unexplained causes (King et al. 2019). Abiotic stressors such as salinity shifts, elevated temperatures, or heat waves can interact with these biotic stressors, increasing negative impacts (Green et al. 2019; King et al. 2019).
Environmental changes more generally are also challenging cultured and wild oysters. Ocean acidification negatively impacts shell formation and maintenance (Watson et al. 2009; Barton et al. 2012; Barros et al. 2013), and extreme weather has led to massive mortality events in recent years. For example, an atmospheric river and freshwater influx into the San Francisco Bay (California, USA) coincided with a near 100% mortality event of the highly abundant population of wild Olympia oyster Ostrea lurida (Cheng et al. 2016). An unprecedented extreme heatwave in the Pacific Northwest during summer 2021 resulted in large-scale negative impacts to wild and naturalized shellfish in the area (Raymond et al. 2022). With a multitude of stressors and impediments to industry stability and growth, shellfish farmers in the Pacific Northwest consider oyster health and selective breeding to be important scientific adaptation strategies to increase crop resilience (Green et al. 2023; Thompson 2023). As such, improvements to selective breeding infrastructure and genomic resources are valuable to Pacific oyster breeders, as recently advocated for eastern oyster C. virginica (Allen et al. 2020).
Maintenance of genetic diversity is essential to shellfish breeding. Although standing polymorphism levels are very high in shellfish (Plough 2016), where a SNP is expected every 40 bp in C. gigas (Hedgecock et al. 2005; Sauvage et al. 2007), rapid reductions in genetic diversity can occur in cultured lineages (Hedgecock and Sly 1990; Evans et al. 2004; Xiao et al. 2011; Gurney-Smith et al. 2017), with losses compounding over time, leading to significant concerns for breeders (e.g. Li et al. 2007). High genetic load occurs through high fecundity and likely high mutation rates, which can produce severe inbreeding depression affecting growth and survival (Evans et al. 2004). Although inbreeding depression may be counteracted by crossing divergent lines to achieve heterosis (Hedgecock et al. 1995; Hedgecock and Davis 2007), long-term retention of genetic diversity, and maintenance of hatchery lineages is important for broodstock programs (Carlsson et al. 2006; Gurney-Smith et al. 2017). Furthermore, retaining standing genetic diversity is vital for rapid adaptation (Barrett and Schluter 2008) and selective breeding (Guo 2009).
A rapid, inexpensive, automatable, and easily analyzed genotyping tool is needed for the Pacific oyster. Existing genomic resources for the Pacific oyster include high quality reference genome assemblies (Peñaloza et al. 2021; Qi et al. 2021) and genetic maps (Gutierrez et al. 2018; Li et al. 2018; Yin et al. 2020), as well as SNP arrays of high (134 K targets; Qi et al. 2017), medium (41 K; Gutierrez et al. 2017), and low density (384; Lapègue et al. 2014). However, due to significant costs and required expertise, smaller or medium-scale breeding operations may not be able to benefit from these resources (Boudry et al. 2021). Until present, there has been no low-density amplicon panel for the Pacific oyster. In addition to lower-cost per individual, amplicon panels can also be used to genotype non-target variants within an amplicon window, potentially expanding its utility across populations and time, as long as primer sites remain intact and functional. This longevity and malleability may be particularly advantageous to Pacific oysters given their high genetic diversity (Hedgecock et al. 2005; Sauvage et al. 2007), global scale of culture (Boudry et al. 2021; Martínez-García et al. 2022), including genetically diverse large-scale commercial hatcheries (Sutherland et al. 2020), and temporal fluctuations in allele frequencies that occur due to sweepstakes reproductive success (Hedgecock 1994; Hedgecock and Pudovkin 2011; Sun and Hedgecock 2017).
Here we describe the development of a 592 target SNP panel designed using previously identified high heterozygosity and high differentiation markers from naturalized Pacific oysters in British Columbia (BC), Canada, which are similar to other global populations derived from the Japan translocation lineage (Sutherland et al. 2020). We tested the panel on samples from Canada, France, Japan, and China, as well as cultured populations from the United Kingdom, the United States, and Canada. We further evaluate and demonstrated the panel's effectiveness for genetic parentage assignment using single SNP targets through analyzing three generations of the Vancouver Island University (VIU) breeding program, and a set of families bred for viral resistance from Oregon State University's Molluscan Broodstock Program (MBP). All design amplicons and target variant identities and sites are provided here, and the panel is available via the commercial provider (ThermoFisher Scientific). All analytic code is available in amplitools, a repository designed to move from sequencer output to parentage analysis (see Data Availability). Details on the latest version of the panel's target file is available in the accompanying repository amplitargets (see Data Availability), expected to be updated over time. Collectively, with these outputs we aim to facilitate uptake and therefore bring rapid gains in Pacific oyster breeding and aquaculture.
Methods
Data sources and marker selection
Plink and Variant Call Format (VCF) outputs from Sutherland et al. (2020), a restriction-site associated DNA sequencing (RAD-seq) study of Pacific oysters from the Northern Hemisphere, were obtained from the authors and uploaded to FigShare (see Data Availability). Genotypes were read into R (R Core Team 2024) using the read.PLINK function of adegenet (v.2.1.5; Jombart and Ahmed 2011) and data were converted from genlight format to genind, genepop, and hierfstat formats for analysis using adegenet and hierfstat (v.0.5–10; Goudet 2005).
The pre-filtered dataset was limited to only include naturalized collections from British Columbia (BC), the focal region of the study, which includes the locations Hisnit (HIS), Pendrell (PEN), Pipestem (PIP), and Serpentine (SER); for details on collection sites see Sutherland et al. (2020). Within the BC dataset, per locus global minor allele frequency (MAF) was recalculated, and SNP variants with MAF < 0.01 were removed. Per locus observed heterozygosity (HOBS) and average FST (Weir and Cockerham 1984) were calculated using adegenet and pegas (v.0.12; Paradis 2010), respectively. SNPs were excluded from the RAD-seq dataset when HOBS > 0.5 in an effort to avoid duplicated genomic regions, and remaining SNPs with the highest HOBS (n = 300), or the highest FST (n = 300) were selected for panel design. Several additional SNP variants of interest were included based on previous detection as private alleles in Deep Bay, BC (n = 15) or Guernsey, UK (n = 5) farmed populations (Sutherland et al. 2020). Nine SNPs were redundant between the FST and HOBS selected lists; 611 unique SNPs were identified for panel design (Supplementary Table 1).
Panel design and chromosomal locations
Flanking 200 bp sequences on either side of selected SNPs were obtained from the contig-level reference genome (GCA_000297895.1; Zhang et al. 2012) that was used in marker discovery (Sutherland et al. 2020) using a bed file derived from the marker discovery VCF and the bedtools function getfasta (Quinlan and Hall 2010). All required code and instructions for this process are provided (see Data Availability, ms_oyster_panel). Target amplicon window sequences (401 bp), variant positions, and reference and alternate alleles (Supplementary File 1) were submitted to the AgriSeq primer design team for the Ion Torrent system (ThermoFisher Scientific). Targets were passed through the AgriSeq quality control process (ThermoFisher Scientific), using the GCA_000297895.1 genome. Those passing QC were then submitted to the primer design phase. In the design phase, oligo candidates were generated, scored, filtered, and the optimal oligo pair for each target was selected to be included in the final panel. The selected oligos were checked in silico for specificity and sensitivity of the intended target regions using the C. gigas reference genome GCA_000297895.1. The resultant amplicon panel is available commercially through ThermoFisher Scientific (SKU A58237 AGRISEQ PACIFIC OYSTER PANEL).
As the marker discovery RAD-seq genotyping (Sutherland et al. 2020) used a contig-level genome assembly (Zhang et al. 2012), the chromosomal positions of the targeted 401 bp sequences were subsequently determined by mapping them against a chromosome assembly (GCF_902806645.1; Peñaloza et al. 2021) using bowtie2 (Langmead and Salzberg 2012) in end-to-end mode allowing multiple alignments (-k 6). The target sequences were also aligned with bwa mem (Li 2013). Amplicons were categorized as mapping to a single or multiple positions, and those mapping to multiple positions were removed during analysis. Positions of amplicon targets along chromosomes were visualized using ggplot2 (Wickham 2016).
Panel testing
The panel was tested for use in characterizing different populations by genotyping oyster collections from 2017 and 2018, as described by Sutherland et al. (2020) and shown in Table 1, including from a farm (Deep Bay, BC, DPB), commercial hatcheries (Guernsey, UK, GUR; and China, QDC), and wild (Japan, JPN; China, CHN) or naturalized sources (Pendrell Sound, BC, PEN; France, FRA). Although these same populations and collection years were used for marker discovery (Sutherland et al. 2020), the panel testing was done on alternate individuals not used for discovery, except for JPN and CHN since additional samples were not available. Notably, JPN and CHN were not used for marker selection, as described above.
Table 1.
Collection | Type | Country of origin | Country collected | Genotyped (n) | Filtered (n) | Retained (%) |
---|---|---|---|---|---|---|
Pendrell (PEN) | Naturalized | Canada | Canada | 31 | 29 | 93.5% |
France (FRA) | Naturalized | France | France | 32 | 31 | 96.9% |
Japan (JPN) | Wild | Japan | Japan | 22 | 22 | 100.0% |
China (CHN) | Wild | China | China | 33 | 33 | 100.0% |
China (QDC) | Commercial hatchery | China | China | 8 | 8 | 100.0% |
Guernsey (GUR) | Farm/commercial hatchery | United Kingdom | BC, Canada | 31 | 15 | 48.4% |
Deep Bay (DPB) | Farm/commercial hatchery | unknown | BC, Canada | 32 | 32 | 100% |
VIU G0 (F0) | University hatchery | Canada | BC, Canada | 16 | 11 | 68.8% |
VIU OFR5 parents (F1) | University hatchery | Canada | BC, Canada | 53 | 40 | 75.5% |
VIU OFR5 offspring (F2) | University hatchery | Canada | BC, Canada | 112 | 91 | 81.3% |
MBP CHR8 F0 | University hatchery | Japan and Canada | USA | 25 | 24 | 96% |
MBP CHR8 F1 | University hatchery | Japan and Canada | USA | 111 | 111 | 100% |
Totals | 506 | 447 | 88.3% |
OFR1 and OFR2 are the broodstock for OFR5, and G0 are the broodstock for OFR1 and OFR2. Acronyms: VIU = Vancouver Island University; MBP CHR8 = Molluscan Broodstock Program chromosome 8 families; BC = British Columbia.
The panel was tested for utility in parentage analysis by genotyping samples from three generations of the VIU breeding program, and oyster families from two MBP generations that were selected based on the presence or absence of a SNP marker on chromosome 8 (CHR8) for resistance to the Ostreid herpesvirus-1 (OsHV-1) (Divilov et al. 2023). The breeding program samples were provided with information about their respective generation, but pedigree information was not used until after assignments were created and the initial analysis reviewed. These analyses are described below.
DNA from the VIU and MBP breeding program samples was extracted using the Monarch Genomic DNA Purification kit (NEB) using the enzymatic cleanup tissue extraction protocol with proteinase K and RNase A and quantified using a BioSpectrometer (Eppendorf). All other samples were extracted using the BioSprint (QIAGEN) method and diluted to 25 ng/μl and stored at −20°C. Samples were further normalized to 10 ng/μl in 25 μl volumes and submitted to ThermoFisher Scientific laboratories in Austin, TX for AmpliSeq library preparation using the AgriSeq™ HTS Library kit (ThermoFisher Scientific) with 16 cycles of amplification and IonCode barcode labeling. Libraries were sequenced using Ion 540 chips on an Ion Torrent S5 (ThermoFisher Scientific). Some samples had low sequence yields, and so a portion of the poorly amplifying libraries were synthesized again from undiluted source samples, then sequenced at both diluted and supplied sample concentrations. Replicate samples with sufficient genotyping rates per sample were used to evaluate repeatability of library preparation, sequencing, and genotyping (more details provided below). The MBP samples were submitted for genotyping as described above but conducted at the Institute de Biologie Intégrative et des Systèmes (IBIS) at Université Laval. At both facilities, target variants were scored using the Torrent Suite software VariantCaller plugin (ThermoFisher Scientific) using the original designed target SNP file (i.e. Cgig v.1.0 hotspots; see Data Availability). The custom pipeline amplitools (see Data Availability) was used to filter multilocus genotypes files for target variants and convert to genepop format. Below, the population genetic samples and the VIU samples are referred to as the “pilot dataset”, and the MBP CHR8 samples are referred to as the “MBP CHR8 dataset”.
Repeatability, filtering, and population genetic analyses
The pilot dataset genepop file was imported into R using adegenet and analyzed using the custom pipeline simple_pop_stats using instructions provided in ms_oyster_panel (see Data Availability). Technical replicate pairs with at least a 50% genotyping rate per individual were used to determine the proportion of concordant genotypes. Downstream analyses used the replicate individual with the highest genotyping rate. Samples, then loci, were filtered to keep those with less than 30% missing genotypes. Monomorphic loci, and loci that were previously determined to be align to more than one position in the genome (see above) were also removed. Although markers were characterized for deviation from Hardy–Weinberg proportions (HWPs) in each population using the hw.test of pegas, deviating markers were not removed, only flagged.
The dataset was converted to a genlight object using dartR (v.2.0.4; Gruber et al. 2018). Inter-individual relatedness between individuals was estimated using the Ritland statistic (Ritland 1996) as implemented in related (v.1.0; Pew et al. 2015). First-degree relatedness values were estimated using known full-sib pairs in the VIU dataset (see below). Pairs of individuals within a population that had relatedness values greater than the lowest value for known first-degree relatives were subjected to a process of elimination where one individual per pair was removed until no pairs remained above the threshold. The purged first-degree relative dataset was then analyzed by principal components analysis (PCA) using the function glPca in adegenet, retaining PC1-4, and population averaged FST (Weir and Cockerham 1984) was calculated for any population with more than 15 retained individuals using hierfstat (v.0.5-11; Goudet 2005). The full dataset with putative first-degree relatives included was used to identify private alleles in each genetic grouping using poppr (v.2.9.3; Kamvar et al. 2014), where Japan, France, and Canada (Pendrell) were pooled together, and the three VIU generations were all pooled together. The hatchery or farm collections were kept as separate groupings in this analysis. The full dataset was used to calculate per locus mean FST using the Fst function of pegas, and per locus heterozygosity using the summary function of adegenet. The data for the MBP CHR8 families were used as a test dataset and filtered as above, removing individuals then loci with more than 30% missing genotypes, and removing monomorphic loci. This dataset was used for parentage analysis to compare with the VIU parentage analysis in order to assess the utility of the panel in another breeding program (see below).
Parentage assignment
The pilot dataset was reduced to only the VIU breeding program samples (now referred to as the VIU dataset), filtered again to keep loci with MAF > 0.01 and to remove monomorphic loci, then converted from genind format to rubias format (Moran and Anderson 2019) using the function genepop_to_rubias of simple_pop_stats (see Data Availability). Data preparation and analysis for parentage closely followed vignettes of close-kin mark recapture-sim (CKMR-sim; Anderson 2024). Using allele frequencies estimated from the VIU dataset, a simulated set of parent–offspring, full-sib, half-sib, and unrelated relationships were created to determine the power for the amplicon panel genotypes to resolve different relationship types, and to estimate expected log likelihood values for the different relationship categories. The dataset was checked for closely matching potential duplicate samples. Parent–offspring relationships and full-sibling relationships in offspring and parents were evaluated using VIU samples (i.e. VIU_F2 vs VIU_F1). Siblings of true parents were included in the analysis to test the power of the amplicon panel to resolve relationships in the present of closely related possible parents (e.g. “aunts” or “uncles”).
Following analysis and grouping of individuals into putative families, pedigree information was provided and compared against the genetic parentage analysis relationships. Offspring that failed to assign to genotyped parents were tallied to assess false negatives, and incorrect assignments were tallied as false positives. The number of putative parents was not limited to two, but rather all assignments over the log likelihood cutoff (logl > 5) were retained in the results. This cutoff was based on the total number of comparisons and expected false positive rates as per the CKMR-sim tutorials. The percentage of the total possible assignments that could have been made was calculated for each family in both parentage analyses, and the percentage of the total assignments made that were false positives was also calculated. Suspected pedigree or handling errors were inspected alongside genetic sibship analyses and breeding records to reconcile conflicting relationship information. These errors, when resolvable, were corrected.
Following the initial analysis described above, empirically identified trios (i.e. offspring and two parents) with strong support were identified (i.e. both parents assigned with log likelihood values ≥ 10, only two parental assignments with log likelihood scores above the set cutoff). When a set of parents were supported by more than one trio, the trios were kept as the empirical trio results to be used to identify unexpected genotypes potentially indicating null alleles. The empirical trios were used to infer expected genotypes of offspring per locus and family, and the observed offspring genotypes were compared to expected offspring genotypes to tally the number of unexpected genotypes per locus. Finally, the parentage analysis with all samples was re-done as described above but with the following filtered datasets: (1) all loci passing filters and (2) loci with fewer than four offspring with incompatible genotypes. Percent completeness of assignments and percent of assignments that were false positives were calculated for each of the above two parentage approaches. Relationship networks were plotted using igraph (Csárdi and Nepusz 2006; Csárdi et al. 2024) in R. This same approach was also carried out independently with the MBP CHR8 dataset, and the problematic loci identified with each dataset were compared against each other to determine whether the VIU and the MBP CHR8 dataset had the same loci that were exhibiting potential null alleles or otherwise unexpected offspring genotypes.
Results
SNP selection for marker design and chromosomal positions
The input ddRAD-seq dataset (Sutherland et al. 2020) contained 366 Pacific oyster samples genotyped at 16,492 SNPs (single SNP retained per RAD-tag). It had been filtered to retain SNPs that were present in at least 70% of individuals in all 15 populations, with global MAF ≥ 0.01, and not significantly out of Hardy-Weinberg equilibrium. The dataset was then limited to naturalized British Columbia (BC) samples, retaining 122 individuals from four populations. The BC-only dataset was filtered to keep SNPs with MAF ≥ 0.01, retaining 13,438 SNPs. Marker design candidates were selected for elevated heterozygosity (n = 300; HOBS = 0.41–0.5) or elevated FST between naturalized populations within BC (n = 300; FST = 0.046–0.178), where nine markers were common to both lists. An additional 20 markers were selected due to being previously identified as private alleles (see Methods; Supplementary Fig. 1; Supplementary File 2). Of the 611 submitted SNPs for design, 19 were excluded, and 592 were designed into the panel (i.e. Cgig_v.1.0; see Supplementary File 3 for marker names, contig identifiers, positions and alleles).
Of the 611 submitted design sequences, 499 (81.7%) aligned using bowtie2 to a single location in the chromosome assembly, 55 aligned to multiple locations, 12 aligned singly to non-chromosome scaffolds, and 45 did not map. However, when aligned with bwa mem, 25 of the design sequences aligned to multiple locations, and these were different than the bowtie2 putative multi-mappers (see locus identifiers in Supplementary File 4). Due to this uncertainty, the bowtie2 putative multi-mappers were removed from the present analysis, but remain in the targets file. Each chromosome had an average (± s.d.) of 49.9 ± 26.3 single-mapping amplicons per chromosome (range: 15–83; Fig. 1). Several significant gaps in chromosome coverage with single-mapping markers were observed (Fig. 1).
Panel genotyping and repeatability
The pilot dataset comprised 543 samples including technical replicates and controls, or 370 unique samples and 32 negative controls. Samples were genotyped at 592 amplicons across two sequencing runs (i.e. two sequencing chips). Seven amplicons were uncalled in both chips, leaving 585 genotyped amplicons remaining. The median of the average amplicon depth per individual was 172.5× and 180.1× for the two chips, whereas negative controls were 0.90× and 5.3×, respectively (Supplementary Fig. 2). On the second chip, three of the 19 negative control wells had mean marker coverage greater than 40×, suggesting some contamination in these wells.
Technical replicate sample pairs (n = 68 pairs) showed average (± s.d.) genotype concordance of 97.5 ± 4.2% for SNPs genotyped in both samples (mean = 460 SNPs genotyped in both samples of the pair, with a total of 31,303 genotypes). The most frequent discordant genotypes between technical replicates were homozygous alternate and heterozygous genotypes (n = 552; 1.76%), and less frequent were homozygous reference and heterozygous genotypes (n = 153; 0.49%) or least frequent as homozygous reference and homozygous alternate genotypes (n = 4; 0.01%). Ten replicate sample pairs had more than 5% of the markers discordantly called; on average, these pairs had 359.5 loci typed in both samples, whereas the remainder pairs on average had 477.7 loci. One replicate pair had genotype discordance above 10%, and this pair was only genotyped at 342 loci. Therefore, samples with lower genotyping rates also had lower concordance.
Filtering and population genetics
The 370 oysters in the pilot dataset were comprised of 181 oysters from the VIU breeding program, 71 from aquaculture programs (i.e. hatcheries or farms), and 63 and 55 from naturalized and wild sources, respectively (Table 1). After filtering samples for genotyping rate (GR > 70%), 312 samples were retained (Table 1; Supplementary Fig. 3). Filtering loci for genotyping rate (GR > 70%) removed 111 loci (Supplementary Fig. 4) leaving 474 loci remaining. Removing monomorphic loci removed an additional 43 loci, resulting in the retention of 431 loci. Any remaining of the 55 multi-mapping loci as determined by the bowtie2 alignment were also removed at this stage, dropping 22 additional loci and retaining 409 loci. Retained markers showed a bimodal HOBS distribution (Fig. 2a). The upper HOBS grouping (HOBS > 0.25; n = 246) has 221 loci that were included in the panel due to high HOBS. The lower HOBS grouping (HOBS < 0.25; n = 163) is mainly comprised of loci included due to high FST (i.e. n = 132), as well as those included based on presence as private alleles (n = 11). Per locus HOBS and FST values for the 409 retained loci are shown in Supplementary File 5. Loci HWPs were inspected for wild (JPN, CHN) and naturalized populations (PEN, FRA) finding 15, 24, 26, and 26 variants out of HWP, respectively. This includes 11 loci that were out of HWP in two populations, and one marker in three populations. These markers were not removed from the analysis but can be referenced for future consideration in Supplementary File 6.
The filtered dataset (n = 312 individuals, 409 loci) was then inspected for highly related pairs of individuals within each population (Supplementary Fig. 5). Outlier pairs with elevated relatedness were observed in most populations. The known full-sibs in the largest cross event in the VIU population used (i.e. oyster family run 5; OFR5) were used to calculate average relatedness values within each family, and the 14 families had an average pairwise relatedness value with the Ritland statistic of 0.43, and the average minimum relatedness value per family was 0.26, which was used as a cutoff to remove highly related individuals specifically to reduce family effects on the PCA and between population FST. This resulted in the removal of most of samples from the hatchery or farm populations (i.e. VIU, DPB, and GUR) for PCA and FST analyses. Populations with fewer than 15 individuals were not used for FST calculation.
A PCA with putative first-degree relatives removed found general overlap of all populations except CHN, which were slightly separated from the rest of the samples across PC1 (3.2% variance explained; Supplementary Fig. 6). Similarly, the Japan translocation lineage (i.e. JPN, PEN, FRA) had low FST, with the lower range of the 95% confidence intervals overlapping zero (i.e. no differentiation) and the upper range reaching FST = 0.012 (Supplementary Table 2). By contrast, CHN had higher differentiation, for example JPN-CHN 95% C.I. FST = 0.023–0.048.
Private alleles were empirically identified in the genotyped dataset (with putative first-degree relatives included) at the regional level (i.e. JPN lineage, VIU, DPB, GUR, CHN). In total, 20 private alleles were observed, including some with high frequency. Three markers were included in the panel because they contained private alleles in DPB (Sutherland et al. 2020); although different samples were used here, two of these loci remained after filters and also were observed to be private alleles in this dataset, including locus 39139 (eight occurrences), and locus 538043 (six occurrences). Similarly, locus 336452, included in the panel as a private allele within GUR in Sutherland et al. (2020) was also private in this dataset (nine occurrences). One high frequency private allele was observed in the VIU samples with eight observations (locus 590965). However, this marker would have been designed based on presence in other BC samples, and so is likely just a rare allele that has been enriched in the VIU broodstock program. A private allele was also observed for each of CHN and JPN with five observations (see full list in Supplementary Table 3).
Parentage assignment in the VIU dataset
The panel was tested on families from the oyster family run 5 (OFR5) spawning event of the VIU Pacific oyster breeding program. OFR5 involved 15 crosses using 18 unique parents that originated from cross events OFR1 and OFR2, including nine dams and nine sires, all of which were genotyped. One OFR5 broodstock parent, BR22, was missing. In total, 91 OFR5 offspring (81.3%) were retained after filters (Table 1, Table 2).
Table 2.
OFR5 family | Dam [fam.] (loci) | Sire [fam.] (loci) | Offspring | Two-parent assign. | One-parent assign. | % ID | Incorr. assign. |
---|---|---|---|---|---|---|---|
OFR5_2 | BR19 [2–6] (323) | BR11 [1–14] (326) | 6 | 6 | - | 100.0 | |
OFR5_3 | BR14 [2–10] (306) | BR2 [1–12] (322) | 9 | 8 | 1 | 94.4 | |
OFR5_4 | BR14 [2–10] (306) | BR11 [1–14] (326) | 6 | 6 | - | 100.0 | |
OFR5_5 | BR3 [1–4] (317) | BR10 [1–18] (325) | 6 | 5 | 1 | 91.7 | 1 × pat. sib |
OFR5_7 | BR17 [2–11] (323) | BR4 [1–10] (317) | 7 | 7 | - | 100.0 | |
OFR5_8 | BR3 [1–4] (317) | BR7 [2–11] (316) | 8 | 7 | 1 | 93.8 | |
OFR5_9 | BR9 [1–16] (318) | BR23 [2–6] (324) | 8 | 8 | - | 100.0 | |
OFR5_10 | BR25 [1–19] (322) | BR23 [2–6] (324) | 7 | 5 | 2 | 85.7 | 3×mat. sib |
OFR5_12 | BR14 [2–10] (306) | BR20 [1–19] (326) | 7 | 7 | - | 100.0 | 4×pat. sib |
OFR5_13 | BR22 [2–14] (0) | BR26 [1–9] (325) | 3 | na | 3 | 100.0 | |
OFR5_14 | BR24 [1–13] (318) | BR2 [1–12] (322) | 8 | 8 | - | 100.0 | |
OFR5_15 | BR18 [2–13] (320) | BR10 [1–18] (325) | 2 | 2 | - | 100.0 | 1×pat. sib |
OFR5_16 | BR22 [2–14] (0) | BR2 [1–12] (322) | 7 | na | 4 | 57.1 | |
OFR5_17 | BR21 [1–1] (322) | BR5 [2–21] (319) | 8 | 5 | 2 | 75.0 |
In total, 160 of 172 potential assignments were made (93.0% complete), and eight false positives occurred (4.8% of calls). Per family, the individual identifier (with “BR” prefix) for dam and sire are shown alongside their originating family identifier (brackets) and their number of retained genotyped markers (parentheses). Per family, the number of genotyped offspring is shown with the number of correct two- or one-parent assignments, the percent completeness of the assignments (% ID), and any incorrect assignments. All assignments above the applied cutoff were retained (see Methods). Acronyms: fam. = family; pat. = paternal; mat. = maternal.
An initial analysis of the data found likely pedigree or handling errors. Four samples were identified that were suspected handling errors in which the offspring had the incorrect family label. These were evident using parent–offspring, or sibship analyses. Second, family OFR5-15 was found to be a mixed family, with two separate sets of parents (i.e. the expected parents BR18 and BR10, or the parents of OFR5-14, BR24 and BR2; Table 2). Third, a pedigree error was identified where the sire of family OFR5-2 was BR11 and not BR2 (family 1–12). These were all corrected before subsequent analyses.
To identify potential null alleles in the amplicon loci, reliable two-parent assignments were identified and used to identify incompatible offspring genotypes based on the observed parental genotypes (see Methods). Reliable assignments were defined as those trios with exactly two parents assigned above an elevated cutoff. Putative families were estimated from the reliable assignments, keeping parental pairs as a putative family when more than one observation of a parental pair occurred. These identified trios included 37 offspring from 11 families. Within these trios, 36 loci were observed with at least four offspring with incompatible genotypes, and 95 loci with at least two offspring with incompatible genotypes (Supplementary Fig. 7). Parentage datasets were constructed to evaluate the impact of these loci exhibiting potential null alleles: (i) keeping all filtered loci (n = 364 loci retained) or (ii) removing loci with four or more incompatible genotypes (n = 328 loci retained). An improvement in assignments occurred and a reduction of false positives by removing loci with at least four incompatible offspring genotypes, and so this approach was taken (Table 3). In this dataset, the OFR5 broodstock was genotyped with an average of 313.2 loci per individual (Table 2), and the offspring at an average of 307.0 loci per individual (Supplementary File 7).
Table 3.
Cross event/families | Locus filter | Num. loci | Num. assign. | Num. possible assign. | Percent assigned | Num. false positive (FP) | Percent assigns as FP |
---|---|---|---|---|---|---|---|
OFR5 | all filtered loci | 364 | 134 | 172 | 77.9% | 11 | 7.6% |
OFR5 | Remove loci with ≥ 4 incompatibilities | 328 | 160 | 172 | 93.0% | 8 | 4.8% |
MBP CHR8 | all filtered loci | 324 | 163 | 215 | 75.8% | 7 | 4.1% |
MBP CHR8 | Remove loci with ≥ 4 incompatibilities | 289 | 185 | 215 | 86.0% | 17 | 8.4% |
The number of loci present in the dataset, the number of assignments and percentage of possible assignments, as well as the number of false positives and percentage of assignments made as false positives are shown. The datasets with loci having more than four incompatibilities removed were used for downstream analyses.
Within the 328-loci dataset, 160 of 172 potential assignments were made (93.0% complete), and eight false positives occurred (4.8% of assignments made; Table 3; Table 2). The greatest number of false positives occurred in families F10 and F12, where maternal or paternal siblings were assigned in addition to the correct assignment. All parent–offspring relationships are shown in Fig. 3, and assignments with accompanying log likelihood ratios are shown in Supplementary File 8.
Parentage assignment in the MBP CHR8 dataset
An independent dataset comprised of families from the MBP (CHR8) was genotyped to further evaluate the panel performance for parentage assignment. This dataset was expected to be more challenging for the panel given the longer captive breeding history and higher expected relatedness among broodstock. The MBP spawn was comprised of 16 families from MBP cohort 27 (seventh generation of selective breeding) including 25 unique broodstock (nine dams and 16 sires). All parents were genotyped, but one parent (65_8F) was removed due to low genotyping rate. In total, 111 offspring (100.0%) were retained after filtering (Table 1, Table 4). The dataset was comprised of 576 scored loci, but after filtering the excess missing data (missing ≥ 30%), 519 loci were retained. An additional 174 monomorphic loci were removed, leaving 345 loci. Removing the multi-mapping loci from the bowtie2 analysis resulted in an additional 21 loci being removed, leaving 324 loci remaining. The MBP CHR8 family distributions of the likelihoods of simulated relationships were similar to those of the VIU OFR5 families (Supplementary Fig. 8). Further, the observed heterozygosity per locus of the VIU OFR5 family and MBP CHR8 shows many loci with similarly elevated HOBS in both datasets (Fig. 2b).
Table 4.
CHR8 family | Dam [fam.] (loci) | Sire [fam.] (loci) | Offspring | Two-parent assign. | One-parent assign. | % ID | Incorr. assign. |
---|---|---|---|---|---|---|---|
101 | 55_3F [55] (281) | 55_16 M [55] (284) | 7 | 7 | - | 100.0 | 2×[55] aunts |
102 | 55_3F [55] (281) | 55_14 M [55] (284) | 7 | 7 | - | 100.0 | |
103 | 55_45F [55] (283) | 55_17 M [55] (285) | 7 | 7 | - | 100.0 | 6×[55] uncles |
105 | 55_27F [55] (286) | 58_29 M [58] (287) | 6 | 6 | - | 100.0 | 1×[58] uncle, 1×[55] aunt |
106 | 55_27F [55] (286) | 58_20 M [58] (284) | 7 | 7 | - | 100.0 | |
107 | 55_45F [55] (283) | 58_6 M [58] (287) | 7 | 1 | 6 | 57.1 | 3×[58] uncle, 1×[58] aunt |
108 | 55_5F [55] (286) | 65_5 M [65] (281) | 7 | 2 | 5 | 64.3 | |
109 | 55_5F [55] (286) | 65_3 M [65] (286) | 7 | 4 | 3 | 78.6 | |
110 | 55_45F [55] (283) | 65_11 M [65] (286) | 7 | 4 | 3 | 78.6 | |
111 | 55_2F [55] (255) | 79_6 M [79] (286) | 6 | 4 | 2 | 83.3 | 2×[58] unkn. |
112 | 55_2F [55] (255) | 79_4 M [79] (281) | 7 | 2 | 5 | 64.3 | |
113 | 55_45F [55] (283) | 79_18 M [79] (284) | 7 | 6 | 1 | 92.9 | |
114 | 65_4F [65] (286) | 58_33 M [58] (287) | 7 | 7 | - | 100.0 | |
115 | 65_8F [65] (0) | 79_13 M [79] (282) | 7 | na | 7 | 100.0 | 1×[65] uncle |
116 | 55_41F [55] (287) | 65_19 M [65] (282) | 7 | 4 | 3 | 78.6 | |
117 | 58_9F [58] (287) | 79_1 M [79] (284) | 7 | 7 | - | 100.0 | 1×[58] uncle |
In total, 185 of 215 potential assignments were made (86.0% complete), and 17 false positives occurred (8.4% of calls). See VIU OFR5 table for full table explanations.
An initial analysis found suspected handling errors, including three offspring with incorrect family identifiers. Offspring were labeled as family/offspring number. Incorrect offspring identified were as follows: 102/8 was from family 106 and not 102; 106/1 was from family 113 and not 106; and 113/8 was from family 102 and not 113. These were sequential in DNA extraction and library preparation, and so the series of unexpected assignments likely reflects a handling error in DNA extraction or library preparation and not in sampling. These errors were corrected by providing new identifiers to the unexpected offspring to match their correct families, and then the same approach to identify putative null alleles was taken as described above for OFR5 (see above and Methods). Empirically determined trios with strong evidence were identified in the MBP CHR8 families, including 38 offspring from nine different families. Using these trios, 73 loci were identified that had at least two offspring with incompatible genotypes based on expected genotypes inferred from parents, and 35 loci with at least four offspring with incompatible genotypes. The two datasets (i.e. all filtered loci: 324 loci; remove when ≥ 4 incompatibilities: 289 loci), were used for parentage analysis. The dataset with 289 loci resulted in the most assignments, but also the most false positive assignments (Table 3, Table 4). In this dataset, the parent generation (potential broodstock) were genotyped on average at 283.4 loci per individual (Table 4), and the offspring on average at 284.5 loci per individual (Supplementary File 7). This dataset had 185 of 215 possible assignments (86.0% complete), and 17 false positives occurred (8.4% of assignments). All identified parent–offspring relationships are shown in Fig. 4, and assignment details in Supplementary File 8.
To determine whether a common set of loci were exhibiting genetic incompatibilities (i.e. putative null alleles) in the VIU OFR5 and the MBP CHR8 families, overlap was inspected between the incompatible loci. Of the 36 loci that were incompatible in VIU (≥4 incompatibilities) and the 35 incompatibility loci in MBP, only seven were found common in both datasets (i.e. 19% or 20% of the total VIU or MBP loci, respectively). This low overlap occurred even though 35 of 36 of the incompatible VIU loci were present in the MBP CHR8 dataset (before incompatibles were removed), and 28 of 35 of the incompatible MBP CHR8 loci were present in the VIU dataset (before incompatibles were removed; Supplementary File 9). This suggests that putative null alleles identified are largely population-specific.
Discussion
This work demonstrates the effectiveness of the Pacific oyster Cgig_v.1.0 amplicon panel for characterizing genetic variation and conducting parentage assignment, as well as provides the analysis framework to support additional use of the panel, as well as future development of the panel in subsequent design iterations. The unique characteristics of the Pacific oyster, including high levels of genetic polymorphism, likely led to challenges that were faced during the analysis, such as the presence of numerous, population-specific null alleles. Initial results were sufficient to detect suspected handling or pedigree errors, including in breeding program samples from populations not part of the panel design, and with careful filtering steps, parentage assignment success was able to be improved. We expect that the tools provided here will bring advances to Pacific oyster breeding programs for both small and large operations.
Maintaining genetic diversity and supporting breeding program development
Several genetic approaches may be used to increase the resiliency of cultured Pacific oysters. First, maintaining overall genetic diversity retains variation required for rapid adaptation (Barrett and Schluter 2008) and avoids significant inbreeding depression that can impact growth and survival (Wada and Komaru 1994; Evans et al. 2004). Second, artificial selection methods including family-based selection (i.e. breeding families that show most favorable phenotypes as selection targets), marker-assisted selection (i.e. selecting broodstock based on genotypes at loci associated with a phenotype), and genomic selection (e.g. breeding based on multilocus genotype relationships derived from a training set of phenotyped individuals), can be used to improve phenotypic targets of breeding populations (Boudry et al. 2021). Genomic or marker-assisted selection cannot be undertaken without genotyping. Similarly, maintaining diversity or separation of distinct breeding lines benefits from the ability to empirically measure diversity and confirm family relationships.
Maintaining genetic diversity in Pacific oyster breeding populations is important as rapid reductions in genetic diversity can occur (Hedgecock and Sly 1990), as can small sustained reductions that accumulate over time (Carlsson et al. 2006; Xiao et al. 2011; Varney and Wilbur 2020). Using genetic markers, the effective population size and other measures of diversity such as allelic richness can be directly monitored in the breeding population (Hedgecock and Sly 1990; Gurney-Smith et al. 2017). Allelic richness may be a more appropriate measure to monitor diversity than observed heterozygosity or per-individual inbreeding coefficients (Yu and Guo 2004; Varney and Wilbur 2020; Sutherland et al. 2023), in part because these other metrics can be skewed by crossing divergent genetic lines for heterosis, an approach sometimes applied within the bivalve aquaculture industry. When reduced diversity is detected, solutions may be needed to replenish diversity (Li et al. 2007). Some countries, including Canada, have naturalized populations of conspecifics that can be incorporated within breeding programs (Sutherland et al. 2020) to increase diversity and avoid inbreeding. Having tools to empirically measure genetic diversity is important for breeding programs (Gurney-Smith et al. 2017) for a range of industry scales (Boudry et al. 2021).
Family-based selection or crossing of divergent lines both require maintenance of separate genetic groups, and benefit from confirming crosses and pedigree information to correct errors that may occur during animal husbandry or data recording. This has been a subject of concern for many decades (Hedgecock and Davis 2007), and has led to the development of high-resolution melt curve genotyping panels (Zhong et al. 2013; Sun et al. 2015). Pedigree errors have been identified in the present work and elsewhere (Hedgecock and Pan 2021) that would not have been identified without the use of a low-cost, effective genotyping panel, and if or when left unchecked, could slow progress in family-based breeding programs (Hedgecock and Pan 2021). The amplicon panel presented here provides a rapid and lower-cost tool to evaluate the parentage of crosses and correct pedigree errors, thereby increasing the precision of family-based selection protocols.
There are many potential sources of error that can disrupt maintenance of separate genetic groups and directional family-based selection. They may occur during spawning, larval rearing, or growth phases, and can be caused by incorrect labeling, contaminating families (i.e. inadvertent mixing), as well as handling and record keeping errors. Many of these errors disrupt pedigree relationships, and if left uncorrected, can compound over time and impede directional selection progress (Hedgecock and Davis 2007). As presented here, we identified several different sources of error: (1) mislabeled single offspring; (2) a full-sib family contaminated with a second full-sib family (i.e. the preceding family); (3) handling or recording errors for the sire of two families within one spawn event; (4) handling errors after sampling during laboratory preparation. Importantly, as discussed by Hedgecock and Pan (2021), these errors could impede selective breeding gains were they not detected, and may increase the mismatch between the genetic breeding values and the phenotypic outcomes of trials.
The amplicon panel may bring additional benefits to aquaculture programs. Currently, the application of genomic selection in shellfish aquaculture is limited by high genotyping costs per individual (Delomas et al. 2023). One framework, described and recently evaluated in silico for Pacific oyster by Delomas et al. (2023), is to use whole-genome resequencing of parents, lower-density amplicon panel genotyping of offspring, and imputation to enable more cost-effective application of modern genomic selection methods. Furthermore, field deployments for phenotyping trials can be conducted in common garden replicates, rather than using separate nets that confound spatial and family effects, by using genetic parentage analysis enabled by lower cost genetic marker panels (Allen et al. 2020). Finally, future versions of the panels can be designed to include putatively functional markers for marker-assisted selection, for example for shell colur or body size, or traits which can increase survival from thermal stress (Raymond et al. 2022), or for field survival likely linked to disease resistance (e.g. OsHV-1; Divilov et al. 2023).
Future directions
The current panel performed well on the populations included in this analysis. Future work should continue to evaluate the panel in other populations to determine any ascertainment bias and to assess how broadly applicable the tool can be. As an example, in the present study, moving to different populations that were not involved in design, such as MBP's CHR8 families, resulted in a specific loss of SNPs selected due to high FST in BC, whereas the SNPs selected for high heterozygosity were more likely to pass filters. The high FST markers are the most likely candidates to remove in future panel iterations. In this project, most genotyped individuals originated from populations that were included in marker discovery and comprised oysters from the Japan translocation lineage, China, and several commercial aquaculture hatcheries (Sutherland et al. 2020). In new populations, other variants within the panel's amplified sequences (i.e. not the target designed SNP) may prove more powerful for parentage, and these could be characterized via adjustments to the data analysis post sequencing, and not requiring any physical changes to the current panel. This is beneficial because physical changes to the primer composition would be more laborious and costly. This is a particular advantage for amplicon-based sequencing that is not possible in array or high-resolution melt curve genotyping.
Using non-target variants from the existing amplicons may provide improvements to the panel for parentage purposes without changes to the primer pool. This is apparent given that some markers within Cgig_v.1.0 in the pilot study had low heterozygosity and MAF, including those selected due to high differentiation potential or presence as private alleles. For technical replicates, low genotyping rate pairs had the lowest percent concordance, suggesting that poor genotyping rate is not only an issue for missing data, but the data that is present has a higher chance of miscalls. Therefore, emphasis should be placed on retaining high quality throughout all phases of tissue collection, library preparation, and sequencing to minimize the loss of data or the generation of poor quality data. End-users of the panel should determine levels of missing data that are allowable for their uses. As missing data was more prevalent in specific collections, it is likely that this is driven by DNA quality or quantity; other sample collections with similar genetic backgrounds to the high failure collections had more consistent and complete genotyping, suggesting that the missing data is not largely due to primer failures due to variants at binding sites but rather due to DNA quality or quantity. Including higher value markers (i.e. higher MAF) may improve resolution of secondary assignments to full-sibs of the true parent (i.e. false positives) without a significant increase in false negatives (i.e. missing assignments). Increased statistical power to assign genetic relationships may also be realized through a microhaplotype approach (Baetscher et al. 2018), which is an active area of research for the current panel.
Other advances to the panel are possible but would require modifying the primer suite by adding new or removing existing amplicons. The uniformity of coverage across chromosomes could be improved, filling in gaps observed in the current work, and thereby improving the potential for genomic selection (Delomas et al. 2023). As discussed above, when additional variants with putatively functional roles (or close linkage with causal variants) are identified, such as the recently identified SNP on chromosome eight associated with increased field survival associated with OsHV-1 outbreaks (e.g. Divilov et al. 2023), these will likely be valuable to add to the panel for breeding purposes. Although rigorous documentation and standardization of locus names is required, the malleability of the amplicon panel is a major strength, as loci can be added or dropped by changing the primer composition of the primer pool. Once it can be more conclusively determined which loci are true multi-mapping loci, for example through detecting more than two alleles in diploid individuals with microhaplotypes, these loci will be permanently removed from the panel. Furthermore, removing loci that exhibit null alleles would strengthen the use of the panel, however, the observation of the two different breeding programs largely showing different loci that had putative null alleles negates the ability to screen for null allele loci across a broad geographic scale without including this in an initial panel development protocol. In addition, given the high mutation rate of Pacific oysters, new mutations are likely to regularly emerge that could disrupt primer binding sites, producing new null alleles occurrences. Thus, users of the current panel are strongly encouraged to screen for null alleles within the target population of interest as removing these loci improves genetic parentage assignment success, as was demonstrated in the present work.
The amplicon panel provides a new tool for emerging Pacific oyster selection programs, but it is also important to also consider the ease of which the data can be analyzed. Although the tab-delimited genotype data produced by the genotyping platform may be easier to handle for those with less bioinformatics experience than high volume sequence data, moving from this output to the formats required for parentage or population genetic analysis does require moderate skills in data manipulation, quality control, and data management. With this in mind, we developed amplitools, a repository that is designed to make the use of this panel more accessible, including bash and R scripts for converting from raw variant calling output from standard sequencer software to a standard format used by numerous genetic analysis programs with a wide range of analysis types being available.
Some aspects of documentation within breeding programs will support the use of amplitools, such as not including third generations when running parent–offspring analyses, given the potential for grandparents to be miss-assigned as parents. Including parental sex information may also be beneficial for analysis, although this has not yet been implemented in amplitools. Additionally, we developed the associated repository amplitargets to keep and version control target files for genotyping, as well as any notes regarding changes between versions of SNP target files. Although the primer panel, amplitools and amplitargets are currently only implemented for the AgriSeq platform (ThermoFisher Scientific), the source marker sequences and target variants are all provided here and so they could be used for future panel development on other sequencing platforms, and the code could also be adapted and expanded to other data processing platforms.
Conclusions
The Pacific oyster Cgig_v.1.0 amplicon panel is an effective tool for characterizing genetic variation, genetic similarity among populations, and relatedness between individuals, including for genetic parentage analysis applications. Combined with the amplitools workflow provided here, this amplicon SNP panel could benefit oyster breeding programs by facilitating monitoring of genetic variation and confirming pedigree relationships. The amplicon panel is a useful and lower-cost genotyping tool that has the ability to easily evolve and improve in future iterations by targeting novel variants and microhaplotypes in silico, as well as by more substantial revisions, such as including putatively adaptive loci, and achieving more uniform coverage across the genome.
Supplementary Material
Acknowledgments
Thanks to Chen Yin Walker (Vancouver Island University), Claire Rycroft (University of British Columbia), and the Molecular Genetics Lab of Fisheries and Oceans Canada for DNA extraction for the study. Thanks to those who collected or provided samples for this project, including Dr. Kristi Miller (DFO), Rob Saunders (RKS Labs), Prof. Nicolas Bierne and Dr. Ismaël Bernard (Eurêka Modélisation), Prof. Li Li and Dr. Sheng Liu, David Howell, J.P. Hastey (Nova Harvest Ltd.), Keith Reid (Odyssey Shellfish), Brian Yip (Fanny Bay Oysters), and Bruce Evans. Thanks to Geneviève Morin and Brian Boyle of IBIS (Université Laval) for support in implementing the amplicon panel for genotyping. Thanks to Dr. Thomas Delomas (USDA-ARS) for discussions on identifying null alleles without using pedigree information. Thanks to three anonymous reviewers and the Editor for comments on an earlier version of the manuscript.
Contributor Information
Ben J G Sutherland, Sutherland Bioinformatics, Lantzville, BC V0R 2H0, Canada; Faculty of Science and Technology, Vancouver Island University, Nanaimo, BC V9R 5S5, Canada.
Neil F Thompson, United States Department of Agriculture, Hatfield Marine Science Center, Pacific Shellfish Research Unit, Agricultural Research Service, Newport, OR 97365, USA.
Liam B Surry, Faculty of Science and Technology, Vancouver Island University, Nanaimo, BC V9R 5S5, Canada.
Krishna Reddy Gujjula, ThermoFisher Scientific, 2130 Woodward Street, Austin, TX 78744, USA.
Claudio D Carrasco, ThermoFisher Scientific, 2130 Woodward Street, Austin, TX 78744, USA.
Srinivas Chadaram, ThermoFisher Scientific, 2130 Woodward Street, Austin, TX 78744, USA.
Spencer L Lunda, Department of Microbiology, Oregon State University, 226 Nash Hall, Corvallis, OR 97331, USA.
Christopher J Langdon, Hatfield Marine Science Center, 2030 SE Marine Science Dr., Oregon State University, Coastal Oregon Marine Experiment Station, Newport, OR 97365, USA.
Amy M Chan, Department of Earth, Ocean and Atmospheric Sciences, The University of British Columbia, Vancouver, BC V6T 1Z4, Canada.
Curtis A Suttle, Department of Earth, Ocean and Atmospheric Sciences, The University of British Columbia, Vancouver, BC V6T 1Z4, Canada; Department of Microbiology and Immunology, The University of British Columbia, Vancouver, BC V6T 1Z3, Canada; Department of Botany, The University of British Columbia, Vancouver, BC V6T 1Z4, Canada; Institute for the Oceans and Fisheries, The University of British Columbia, Vancouver, BC V6T 1Z4, Canada.
Timothy J Green, Faculty of Science and Technology, Vancouver Island University, Nanaimo, BC V9R 5S5, Canada.
Data availability
The amplicon marker panel referred to here as Cgig_v.1.0 is available commercially from ThermoFisher Scientific as SKU A58237 AGRISEQ PACIFIC OYSTER PANEL. The input data required for marker selection (single SNP per locus; plink and VCF formats), the VariantCaller multilocus genotypes for the pilot study and for the MBP CHR8 families, alignment results of submitted sequences against the reference genome, names of markers and their reason for inclusion in the panel, and sample population annotation is available on FigShare, DOI: 10.6084/m9.figshare.23646471 . The raw fastq data of the pilot study amplicon data has been uploaded to NCBI SRA under BioProject PRJNA1118744, accessions SAMN41615571-SAMN41615942.
Three repositories support this project:
Manuscript code repository: https://github.com/bensutherland/ms_oyster_panel
amplitools: https://github.com/bensutherland/amplitools
amplitargets: https://github.com/bensutherland/amplitargets
Supplemental material available at G3 online.
Funding
This work was funded by the Gordon and Betty Moore Foundation (GBMF#5600) to CAS. This research used resources provided by the SCINet project of the USDA Agricultural Research Service, ARS project number 0500-00093-001-00-D. Funding for NFT was provided by USDA Agricultural Research Service CRIS project 2076-63000-005-000-D. Production and parentage assignment of the MBP CHR8 families was supported by NOAA Award NA18NMF4720007 through the Pacific States Marine Fisheries Commission. Mention of trade names or commercial products in this publication is solely for the purpose of providing specific information and does not imply recommendation or endorsement by the U.S. Department of Agriculture.
Literature cited
- Allen SK, Rexroad C, Rheault R. 2020. Improving production of the eastern oyster Crassostrea virginica through coordination of genetic improvement programs, research, and technology transfer activities. J Shell Res. 39(2):175–179. doi: 10.2983/035.039.0201. [DOI] [Google Scholar]
- Anderson, E.C., 2024CKMRsim: Inference of pairwise relationships using likelihood ratios. https://github.com/eriqande/CKMRsim.
- Arbelaez JD, Dwiyanti MS, Tandayu E, Llantada K, Jarana A, Ignacio JC, Platten JD, Cobb J, Rutkoski JE, Thomson MJ, et al. 2019. 1k-RiCA (1K-Rice Custom Amplicon) a novel genotyping amplicon-based SNP assay for genetics and breeding applications in rice. Rice. 12(1):55. doi: 10.1186/s12284-019-0311-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Baetscher DS, Clemento AJ, Ng TC, Anderson EC, Garza JC. 2018. Microhaplotypes provide increased power from short-read DNA sequences for relationship inference. Mol Ecol Res. 18(2):296–305. doi: 10.1111/1755-0998.12737. [DOI] [PubMed] [Google Scholar]
- Barrett RDH, Schluter D. 2008. Adaptation from standing genetic variation. Trend Ecol Evol. 23(1):38–44. doi: 10.1016/j.tree.2007.09.008. [DOI] [PubMed] [Google Scholar]
- Barros P, Sobral P, Range P, Chícharo L, Matias D. 2013. Effects of sea-water acidification on fertilization and larval development of the oyster Crassostrea gigas. J Exp Mar Biol Ecol. 440:200–206. doi: 10.1016/j.jembe.2012.12.014. [DOI] [Google Scholar]
- Barton A, Hales B, Waldbusser GG, Langdon C, Feely RA. 2012. The Pacific oyster, Crassostrea gigas, shows negative correlation to naturally elevated carbon dioxide levels: Implications for near-term ocean acidification effects. Limnol Ocean. 57(3):698–710. doi: 10.4319/lo.2012.57.3.0698. [DOI] [Google Scholar]
- Beacham TD, Wallace C, MacConnachie C, Jonsen K, McIntosh B, Candy JR, Withler RE. 2018. Population and individual identification of Chinook salmon in British Columbia through parentage-based tagging and genetic stock identification with single nucleotide polymorphisms. Can J Fish Aquat Sci. 75(7):1096–1105. doi: 10.1139/cjfas-2017-0168. [DOI] [Google Scholar]
- Botta R, Asche F, Borsum JS, Camp EV. 2020. A review of global oyster aquaculture production and consumption. Mar Pol. 117:103952. doi: 10.1016/j.marpol.2020.103952. [DOI] [Google Scholar]
- Boudry P, Allal F, Aslam ML, Bargelloni L, Bean TP, Brard-Fudulea S, Brieuc MSO, Calboli FCF, Gilbey J, Haffray P, et al. 2021. Current status and potential of genomic selection to improve selective breeding in the main aquaculture species of International Council for the Exploration of the Sea (ICES) member countries. Aquacult Rep. 20:100700. doi: 10.1016/j.aqrep.2021.100700. [DOI] [Google Scholar]
- Carlsson J, Morrison CL, Reece KS. 2006. Wild and aquaculture populations of the eastern oyster compared using microsatellites. J Hered. 97(6):595–598. doi: 10.1093/jhered/esl034. [DOI] [PubMed] [Google Scholar]
- Cheng BS, Chang AL, Deck A, Ferner MC. 2016. Atmospheric rivers and the mass mortality of wild oysters: insight into an extreme future? Proc Royal Soc B: Biol Sci. 283(1844):20161462. doi: 10.1098/rspb.2016.1462. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Csárdi G, Nepusz T. 2006. The igraph software package for complex network research. InterJournal, Complex Systems. 1695. [Google Scholar]
- Csárdi G, Nepusz T, Traag V, Horvát S, Zanini F, Noom D, Müller K. 2024. igraph: Network analysis and visualization in R. doi: 10.5281/zenodo.7682609. [DOI]
- Csernák E, Molnár J, Tusnády GE, Tóth E. 2017. Application of targeted next-generation sequencing, TruSeq custom amplicon assay for molecular pathology diagnostics on formalin-fixed and paraffin-embedded samples. App Immun Mol Morphol. 25(7):460–466. doi: 10.1097/PAI.0000000000000325. [DOI] [PubMed] [Google Scholar]
- Delomas TA, Hollenbeck CM, Matt JL, Thompson NF. 2023. Evaluating cost-effective genotyping strategies for genomic selection in oysters. Aquaculture. 562:738844. doi: 10.1016/j.aquaculture.2022.738844. [DOI] [Google Scholar]
- Divilov K, Merz N, Schoolfield B, Green TJ, Langdon C. 2023. Marker-assisted selection in a Pacific oyster population for an antiviral QTL conferring increased survival to OsHV-1 mortality events in Tomales Bay. Aquaculture. 567:739291. doi: 10.1016/j.aquaculture.2023.739291. [DOI] [Google Scholar]
- Evans F, Matson S, Brake J, Langdon C. 2004. The effects of inbreeding on performance traits of adult Pacific oysters (Crassostrea gigas). Aquaculture. 230(1–4):89–98. doi: 10.1016/j.aquaculture.2003.09.023. [DOI] [Google Scholar]
- Goudet J. 2005. hierfstat, a package for R to compute and test hierarchical F-statistics. Mol Ecol Notes. 5:184–186. 10.1111/j.1471-8286.2004.00828.x. [DOI] [Google Scholar]
- Green TJ, Siboni N, King WL, Labbate M, Seymour JR, Raftos D. 2019. Simulated marine heat wave alters abundance and structure of Vibrio populations associated with the Pacific oyster resulting in a mass mortality event. Microb Ecol. 77(3):736–747. doi: 10.1007/s00248-018-1242-9. [DOI] [PubMed] [Google Scholar]
- Green KM, Spalding AK, Ward M, Levine A, Wolters EA, Hamilton SL, Rice L. 2023. Oregon shellfish farmers: Perceptions of stressors, adaptive strategies, and policy linkages. Ocean Coastal Management. 234:106475. doi: 10.1016/j.ocecoaman.2022.106475. [DOI] [Google Scholar]
- Gruber B, Unmack PJ, Berry OF, Georges A. 2018. Dartr: an r package to facilitate analysis of SNP data generated from reduced representation genome sequencing. Mol Ecol Res. 18(3):691–699. doi: 10.1111/1755-0998.12745. [DOI] [PubMed] [Google Scholar]
- Guo X. 2009. Use and exchange of genetic resources in molluscan aquaculture. Rev Aquacult. 1(3–4):251–259. doi: 10.1111/j.1753-5131.2009.01014.x. [DOI] [Google Scholar]
- Gurney-Smith HJ, Wade AJ, Abbott CL. 2017. Species composition and genetic diversity of farmed mussels in British Columbia, Canada. Aquaculture. 466:33–40. doi: 10.1016/j.aquaculture.2016.08.038. [DOI] [Google Scholar]
- Gutierrez AP, Bean TP, Hooper C, Stenton CA, Sanders MB, Paley RK, Rastas P, Bryrom M, Matika O, Houston RD. 2018. A genome-wide association study for host resistance to ostreid herpesvirus in Pacific oysters (Crassostrea gigas). G3 (Bethesda). 8(4):1273–1280. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gutierrez AP, Turner F, Gharbi K, Talbot R, Lowe NR, Peñaloza C, McCullough M, Prodöhl PA, Bean TP, Houston RD. 2017. Development of a medium density combined-species SNP array for Pacific and European oysters (Crassostrea gigas and Ostrea edulis). G3 (Bethesda). 7(7):2209–2218. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hedgecock D. 1994. Does variance in reproductive success limit effective population sizes of marine organisms. In: Beaumont AR, editors. in Genetics and Evolution of Aquatic Organisms. London, UK: Chapman and Hall. p. 122–134. [Google Scholar]
- Hedgecock D, Davis JP. 2007. Heterosis for yield and crossbreeding of the Pacific oyster Crassostrea gigas. Aquaculture. 272(Supplement 1):S17–S29. doi: 10.1016/j.aquaculture.2007.07.226. [DOI] [Google Scholar]
- Hedgecock D, Gaffney PM, Goulletquer P, Guo X, Reece K, Warr GW. 2005. The case for sequencing the Pacific oyster genome. J Shell Res. 24(2):429–441. doi: 10.2983/0730-8000(2005)24[429:TCFSTP]2.0.CO;2. [DOI] [Google Scholar]
- Hedgecock D, McGoldrick DJ, Bayne BL. 1995. Hybrid vigor in Pacific oysters: an experimental approach using crosses among inbred lines. Aquaculture. 137(1):285–298. doi: 10.1016/0044-8486(95)01105-6. [DOI] [Google Scholar]
- Hedgecock D, Pan FTC. 2021. Genetic divergence of selected and wild populations of Pacific oysters (Crassostrea gigas) on the West Coast of North America. Aquaculture. 530:735737. doi: 10.1016/j.aquaculture.2020.735737. [DOI] [Google Scholar]
- Hedgecock D, Pudovkin AI. 2011. Sweepstakes reproductive success in highly fecund marine fish and shellfish: a review and commentary. Bull Mar Sci. 87(4):971–1002. doi: 10.5343/bms.2010.1051. [DOI] [Google Scholar]
- Hedgecock D, Sly F. 1990. Genetic drift and effective population sizes of hatchery-propagated stocks of the Pacific oyster, Crassostrea gigas. Aquaculture. 88(1):21–38. doi: 10.1016/0044-8486(90)90316-F. [DOI] [Google Scholar]
- Jombart T, Ahmed I. 2011. Adegenet 1.3-1: new tools for the analysis of genome-wide SNP data. Bioinformatics. 27(21):3070–3071. doi: 10.1093/bioinformatics/btr521. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kamvar ZN, Tabima JF, Grünwald NJ. 2014. Poppr: an R package for genetic analysis of populations with clonal, partially clonal, and/or sexual reproduction. Peer J. 2:e281. doi: 10.7717/peerj.281. [DOI] [PMC free article] [PubMed] [Google Scholar]
- King WL, Jenkins C, Seymour JR, Labbate M. 2019. Oyster disease in a changing environment: decrypting the link between pathogen, microbiome and environment. Mar Environ Res. 143:124–140. doi: 10.1016/j.marenvres.2018.11.007. [DOI] [PubMed] [Google Scholar]
- Langmead B, Salzberg SL. 2012. Fast gapped-read alignment with bowtie 2. Nat Methods. 9(4):357–359. doi: 10.1038/nmeth.1923. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lapègue S, Harrang E, Heurtebise S, Flahauw E, Donnadieu C, Gayral P, Ballenghien M, Genestout L, Barbotte L, Mahla R, et al. 2014. Development of SNP-genotyping arrays in two shellfish species. Mol Ecol Res. 14(4):820–830. doi: 10.1111/1755-0998.12230. [DOI] [PubMed] [Google Scholar]
- Li H. 2013Aligning sequence reads, clone sequences and assembly contigs with BWA-MEM. arXiv:1303.3997, preprint: not peer reviewed.
- Li C, Wang J, Song K, Meng J, Xu F, Li L, Zhang G. 2018. Construction of a high-density genetic map and fine QTL mapping for growth and nutritional traits of Crassostrea gigas. BMC Genomics. 19(1):626. doi: 10.1186/s12864-018-4996-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li Q, Xu K, Yu R. 2007. Genetic variation in Chinese hatchery populations of the Japanese scallop (Patinopecten yessoensis) inferred from microsatellite data. Aquaculture. 269(1):211–219. doi: 10.1016/j.aquaculture.2007.04.017. [DOI] [Google Scholar]
- Martínez-García MF, Ruesink JL, Grijalva-Chon JM, Lodeiros C, Arreola-Lizárraga JA, de la Re-Vega E, Varela-Romero A, Chávez-Villalba J. 2022. Socioecological factors related to aquaculture introductions and production of Pacific oysters (Crassostrea gigas) worldwide. Rev Aquacult. 14(2):613–629. doi: 10.1111/raq.12615. [DOI] [Google Scholar]
- Meek MH, Larson WA. 2019. The future is now: amplicon sequencing and sequence capture usher in the conservation genomics era. Mol Ecol Res. 19(4):795–803. doi: 10.1111/1755-0998.12998. [DOI] [PubMed] [Google Scholar]
- Moran BM, Anderson EC. 2019. Bayesian inference from the conditional genetic stock identification model. Can J Fish Aquat Sci. 76(4):551–560. doi: 10.1139/cjfas-2018-0016. [DOI] [Google Scholar]
- Paradis E. 2010. Pegas: an R package for population genetics with an integrated–modular approach. Bioinformatics. 26(3):419–420. doi: 10.1093/bioinformatics/btp696. [DOI] [PubMed] [Google Scholar]
- Peñaloza C, Gutierrez AP, Eöry L, Wang S, Guo X, Archibald AL, Bean TP, Houston RD. 2021. A chromosome-level genome assembly for the Pacific oyster Crassostrea gigas. GigaScience. 10(3):giab020. doi: 10.1093/gigascience/giab020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pew J, Muir PH, Wang J, Frasier TR. 2015. Related: an R package for analysing pairwise relatedness from codominant molecular markers. Mol Ecol Res. 15(3):557–561. doi: 10.1111/1755-0998.12323. [DOI] [PubMed] [Google Scholar]
- Plough LV. 2016. Genetic load in marine animals: a review. Curr Zool. 62(6):567–579. doi: 10.1093/cz/zow096. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Qi H, Li L, Zhang G. 2021. Construction of a chromosome-level genome and variation map for the Pacific oyster Crassostrea gigas. Mol Ecol Res. 21(5):1670–1685. doi: 10.1111/1755-0998.13368. [DOI] [PubMed] [Google Scholar]
- Qi H, Song K, Li C, Wang W, Li B, Li L, Zhang G. 2017. Construction and evaluation of a high-density SNP array for the Pacific oyster (Crassostrea gigas). PLoS One. 12(3):e0174007. doi: 10.1371/journal.pone.0174007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Quinlan AR, Hall IM. 2010. BEDTools: a flexible suite of utilities for comparing genomic features. Bioinformatics. 26(6):841–842. doi: 10.1093/bioinformatics/btq033. [DOI] [PMC free article] [PubMed] [Google Scholar]
- R Core Team . 2024. R: A language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria. https://www.R-project.org/. [Google Scholar]
- Raymond WW, Barber JS, Dethier MN, Hayford HA, Harley CDG, King TL, Paul B, Speck CA, Tobin ED, Raymond AET, et al. 2022. Assessment of the impacts of an unprecedented heatwave on intertidal shellfish of the Salish Sea. Ecology. 103(10):e3798. doi: 10.1002/ecy.3798. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sauvage C, Bierne N, Lapègue S, Boudry P. 2007. Single nucleotide polymorphisms and their relationship to codon usage bias in the Pacific oyster Crassostrea gigas. Gene. 406(1):13–22. doi: 10.1016/j.gene.2007.05.011. [DOI] [PubMed] [Google Scholar]
- Salvi D, Mariottini P. 2017. Molecular taxonomy in 2D: a novel ITS2 rRNA sequence-structure approach guides the description of the oysters’ subfamily Saccostreinae and the genus Magallana (Bivalvia: Ostreidae). Zool J Linn Soc. 179(2):263–276. doi: 10.1111/zoj.12455 [DOI] [Google Scholar]
- Sauvage C, Bierne N, Lapègue S, Boudry P. 2007. Single nucleotide polymorphisms and their relationship to codon usage bias in the Pacific oyster Crassostrea gigas. Gene. 406(1):13–22. doi: 10.1016/j.gene.2007.05.011. [DOI] [PubMed] [Google Scholar]
- Sun X, Hedgecock D. 2017. Temporal genetic change in North American Pacific oyster populations suggests caution in seascape genetics analyses of high gene-flow species. Mar Ecol Prog Ser. 565:79–93. doi: 10.3354/meps12009. [DOI] [Google Scholar]
- Sun X, Shin G, Hedgecock D. 2015. Inheritance of high-resolution melting profiles in assays targeting single nucleotide polymorphisms in protein-coding sequences of the Pacific oyster Crassostrea gigas: Implications for parentage assignment of experimental and commercial broodstocks. Aquaculture. 437:127–139. doi: 10.1016/j.aquaculture.2014.11.009. [DOI] [Google Scholar]
- Sutherland BJG, Itoh N, Gilchrist K, Boyle B, Roth M, Green TJ. 2023. Genomic diversity of wild and cultured Yesso scallop Mizuhopecten yessoensis from Japan and Canada. G3 (Bethesda). 13(12):jkad242. doi: 10.1093/g3journal/jkad242. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sutherland BJG, Rycroft C, Ferchaud AL, Saunders R, Li L, Liu S, Chan AM, Otto SP, Suttle CA, Miller KM. 2020. Relative genomic impacts of translocation history, hatchery practices, and farm selection in Pacific oyster Crassostrea gigas throughout the Northern Hemisphere. Evol App. 13(6):1380–1399. doi: 10.1111/eva.12965. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Thompson, K., 2023Pacific Coast Shellfish Growers Association (PCSGA) 2023 Research Priorities. https://pcsga.org/wprs/wp-content/uploads/2023/06/FINALPCSGA-2023-Research-Priorities.pdf. [Google Scholar]
- Varney RL, Wilbur AE. 2020. Analysis of genetic variation and inbreeding among three lines of hatchery-reared Crassostrea virginica broodstock. Aquaculture. 527:735452. doi: 10.1016/j.aquaculture.2020.735452. [DOI] [Google Scholar]
- Wada KT, Komaru A. 1994. Effect of selection for shell coloration on growth rate and mortality in the Japanese pearl oyster, Pinctada fucata martensii. Aquaculture. 125(1):59–65. doi: 10.1016/0044-8486(94)90282-8. [DOI] [Google Scholar]
- Watson S-A, Southgate PC, Tyler PA, Peck LS. 2009. Early larval development of the Sydney rock oyster Saccostrea glomerata under near-future predictions of CO2-driven ocean acidification. J Shell Res. 28(3):431–437. doi: 10.2983/035.028.0302. [DOI] [Google Scholar]
- Weir BS, Cockerham CC. 1984. Estimating F-statistics for the analysis of population structure. Evolution. 38(6):1358–1370. doi: 10.1111/j.1558-5646.1984.tb05657.x. [DOI] [PubMed] [Google Scholar]
- Wickham H. 2016. ggplot2: Elegant Graphics for Data Analysis. New York (NY): Springer-Verlag. [Google Scholar]
- Xiao J, Cordes JF, Moss JA, Reece KS. 2011. Genetic diversity in U.S. hatchery stocks of Crassostrea ariakensis (Fujita, 1913) and comparison with natural populations in Asia. J Shell Res. 30(3):751–760. doi: 10.2983/035.030.0315. [DOI] [Google Scholar]
- Yin X, Arias-Pérez A, Kitapci TH, Hedgecock D. 2020. High-density linkage maps based on genotyping-by-sequencing (GBS) confirm a chromosome-level genome assembly and reveal variation in recombination rate for the Pacific oyster Crassostrea gigas. G3 (Bethesda). 10(12):4691–4705. doi: 10.1534/g3.120.401728. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yu Z, Guo X. 2004. Genetic analysis of selected strains of eastern oyster (Crassostrea virginica Gmelin) using AFLP and microsatellite markers. Mar Biotechnol. 6(6):575–586. doi: 10.1007/s10126-004-3600-5. [DOI] [PubMed] [Google Scholar]
- Zhang G, Fang X, Guo X, Li L, Luo R, Xu F, Yang P, Zhang L, Wang X, Qi H, et al. 2012. The oyster genome reveals stress adaptation and complexity of shell formation. Nature. 490(7418):49–54. doi: 10.1038/nature11413. [DOI] [PubMed] [Google Scholar]
- Zhong X, Li Q, Yu H, Kong L. 2013. Development and validation of single-nucleotide polymorphism markers in the Pacific oyster, Crassostrea gigas, using High-resolution Melting analysis. J World Aquacult Soc. 44(3):455–465. doi: 10.1111/jwas.12044. [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Data Citations
- Csárdi G, Nepusz T, Traag V, Horvát S, Zanini F, Noom D, Müller K. 2024. igraph: Network analysis and visualization in R. doi: 10.5281/zenodo.7682609. [DOI]
Supplementary Materials
Data Availability Statement
The amplicon marker panel referred to here as Cgig_v.1.0 is available commercially from ThermoFisher Scientific as SKU A58237 AGRISEQ PACIFIC OYSTER PANEL. The input data required for marker selection (single SNP per locus; plink and VCF formats), the VariantCaller multilocus genotypes for the pilot study and for the MBP CHR8 families, alignment results of submitted sequences against the reference genome, names of markers and their reason for inclusion in the panel, and sample population annotation is available on FigShare, DOI: 10.6084/m9.figshare.23646471 . The raw fastq data of the pilot study amplicon data has been uploaded to NCBI SRA under BioProject PRJNA1118744, accessions SAMN41615571-SAMN41615942.
Three repositories support this project:
Manuscript code repository: https://github.com/bensutherland/ms_oyster_panel
amplitools: https://github.com/bensutherland/amplitools
amplitargets: https://github.com/bensutherland/amplitargets
Supplemental material available at G3 online.