Skip to main content

A novel SNP assay reveals increased genetic variability and abundance following translocations to a remnant Allegheny woodrat population

Abstract

Background

Allegheny woodrats (Neotoma magister) are found in metapopulations distributed throughout the Interior Highlands and Appalachia. Historically these metapopulations persisted as relatively fluid networks, enabling gene flow between subpopulations and recolonization of formerly extirpated regions. However, over the past 45 years, the abundance of Allegheny woodrats has declined throughout the species’ range due to a combination of habitat destruction, declining hard mast availability, and roundworm parasitism. In an effort to initiate genetic rescue of a small, genetically depauperate subpopulation in New Jersey, woodrats were translocated from a genetically robust population in Pennsylvania (PA) in 2015, 2016 and 2017. Herein, we assess the efficacy of these translocations to restore genetic diversity within the recipient population.

Results

We designed a novel 134 single nucleotide polymorphism panel, which was used to genotype the six woodrats translocated from PA and 82 individuals from the NJ population captured before and after the translocation events. These data indicated that a minimum of two translocated individuals successfully produced at least 13 offspring, who reproduced as well. Further, population-wide observed heterozygosity rose substantially following the first set of translocations, reached levels comparable to that of populations in Indiana and Ohio, and remained elevated over the subsequent years. Abundance also increased during the monitoring period, suggesting Pennsylvania translocations initiated genetic rescue of the New Jersey population.

Conclusions

Our results indicate, encouragingly, that very small numbers of translocated individuals can successfully restore the genetic diversity of a threatened population. Our work also highlights the challenges of managing very small populations, such as when translocated individuals have greater reproductive success relative to residents. Finally, we note that ongoing work with Allegheny woodrats may broadly shape our understanding of genetic rescue within metapopulations and across heterogeneous landscapes.

Peer Review reports

Background

Landscape-scale anthropogenic disturbance can cause habitat loss and fragmentation, thereby spatially isolating local wildlife populations and impeding functional connectivity [1,2,3,4]. Species that typically structure as metapopulations may be particularly threatened by spatial isolation of subpopulations. Local extirpations at patches can be common, and persistence of the metapopulation is dependent on ongoing recolonization events [5,6,7]. As such, interrupted dispersal and gene flow among habitat sites can decrease population-wide genetic variability and fitness, promote extirpation of naturally small subpopulations, prevent recolonization events and threaten metapopulation persistence [4, 5, 8, 9]. Allegheny woodrats (Neotoma magister) require rock habitats (e.g., cliff faces, talus slopes, boulder fields), located primarily in high elevation areas throughout the Appalachian Mountains and Interior Highlands [10, 11]. These habitat sites are naturally disjunct. As a result, woodrats typically form small subpopulations (defined by suitable rock habitat) that are connected by dispersal within a larger metapopulation [12, 13]. If movement amongst habitat patches is interrupted, subpopulations become isolated, gene flow is inhibited, genetic diversity is lost through drift, inbreeding depression occurs, population numbers decrease, and recolonization of extirpated sites declines [14,15,16].

New Jersey (NJ) is home to a single remnant population of Allegheny woodrats, located ~ 240 km from the nearest extant population in Pennsylvania. Individuals sampled in 2009, 2011, and 2012 and genotyped at 11 microsatellite loci had relatively low genetic variability, as indicated by allelic diversity and observed heterozygosity (Additional file 1). In response to conservation concerns associated with declining genetic diversity, New Jersey DEP Fish and Wildlife introduced six individuals from a genetically robust population in Pennsylvania in 2015, 2016, and 2017 under the assumption that, if translocated individuals reproduced, population numbers and genetic variability would increase (i.e., genetic rescue; [17,18,19,20]). However, identifying evidence of reproductive success and quantifying genetic variability are dependent on identifying a marker panel with suitable statistical power [21,22,23].

Panels of single-nucleotide polymorphisms (SNPs) can yield a low probability of identity (PID) (i.e., the likelihood that two randomly chosen individuals in a population will present seemingly identical genotypes; [21]), thus aiding accurate reconstruction of familial relationships [24,25,26], even when populations are inbred [23]. Even relatively small SNP panels (e.g., 58–109 markers) can ultimately perform as well or better than small suites of microsatellites [23, 27,28,29,30]. To this end, we sequenced the Allegheny woodrat genome and annotated a draft genome assembly. We subsequently designed a 134 SNP panel incorporating both gene-associated and putatively neutral markers. We conducted preliminary analyses to explore whether the SNP assay provides greater statistical power for individual identification than a commonly used panel of microsatellite markers. The SNP loci were then used to evaluate changes in genetic diversity following translocations to New Jersey’s remnant population and identify offspring of translocated individuals.

Results

Nuclear genome sequencing and SNPtype assay development

We generated 137.6 gigabases (Gb) of raw sequence data from N. magister, including 119.8 Gb from the paired-end (PE) library and 17.8 Gb from the mate-paired (MP) library (Additional File 2). Our draft nuclear genome assembly includes 60,789 scaffolds greater than 2000 basepairs (bp) in length. We used BUSCO v5 to evaluate completeness of the genome by identifying mammalia_odb10 orthologs, finding 77.9% of universal single-copy orthologs were complete (77.2% single copy, 0.7% duplicated), 8.7% fragmented and 13.4% missing.

We initially identified 627,421 high-quality SNPs. Of these, we selected 192 SNPs to include in a Fluidigm SNPtype assay. We subsequently excluded 58 loci for reasons outlined in the methods (e.g., data did not cluster into distinct homozygous and heterozygous states). Of the remaining 134 loci, at least 128 loci amplified for each of the 318 woodrats genotyped (Additional File 3). These loci were roughly divided between gene-associated (n = 72) and neutral markers (n = 62).

Probability of identity using microsatellite and SNP markers

Genotyping 50 woodrats captured in 2017 and 2019 in Adams County, Ohio (OH) at 11 microsatellite markers resulted in a PID of 4.0 × 10–5 and a probability of identity among siblings (PIDsib) of 9.8 × 10–3. By contrast, 134 SNP loci generate values of 5.0 × 10–27 (PID) and 3.1 × 10–14 (PIDsib). If the more conservative data set of 70 loci is used, the PID is 1.9 × 10–13 and the PIDsib is 3.1 × 10–7 across all 50 individuals. Furthermore, our results indicate that a much smaller panel of SNPs might be utilized in subsequent studies to achieve a PID < 0.0001 (Additional File 4; a PID < 0.0001 is considered low enough to distinguish between even closely related individuals in most natural populations [3133]). Given these results, all other samples were genotyped using just the SNP panel. Notably, there was a significant, positive relationship between the number of heterozygous microsatellite loci per individual and the number of heterozygous SNP loci per individual (linear regression: r2 = 0.32, p < 0.0001, Additional File 5).

Genetic variability and reproductive success following translocations to the Palisades population

Parentage analysis revealed that the six woodrats translocated from Pennsylvania to New Jersey produced a minimum of thirteen offspring (Table 1). The female translocated in 2015 produced at least four offspring between 2016 and 2019. The male translocated in 2015 produced at least nine offspring, predominantly in 2016 (Table 1). The offspring of the 2015 translocated female and male produced at least thirteen offspring of their own between 2017 and 2019 (Table 1). We found no evidence from trapping and subsequent genotyping that the other four translocated individuals reproduced. The males translocated in 2016 and 2017 were confirmed dead within 11 and 1 weeks of release, respectively. A camera captured footage of the female translocated in 2016 with a pup. It is unclear whether this pup died before reaching adulthood or simply avoided trapping, as the location in which it was photographed was outside of the regular trapping area. The female translocated in 2017 also settled outside of the regular trapping area and was not detected again.

Table 1 Offspring of individuals translocated from Pennsylvania in 2015, 2016 and 2017 and their offspring

Of the 82 tissue samples collected from the Palisades population, five were collected in 2009, thirteen in 2011, nine in 2015, eighteen in 2016, sixteen in 2017, eight in 2018 and thirteen in 2019. Once a Bonferroni correction was applied, a single locus was found to be out of Hardy–Weinberg equilibrium, and only in 2019 (exhibiting evidence of heterozygote excess). STRUCTURE analysis of resident individuals and those translocated to NJ provided evidence of two genetically distinct clusters when the most conservative data set (70 loci) was utilized. The population-wide genetic composition changed following translocation events in 2015, as illustrated by a shift from the blue cluster associated with the resident population prior to human-mediated gene flow, to an increase in the orange cluster associated with the genetic profiles of the PA individuals (Fig. 1). Despite this, all alleles historically present at the loci considered in this study were retained following translocations (data not shown).

Fig. 1
figure 1

STRUCTURE results for 82 woodrats trapped in the Palisades, NJ between 2009 and 2019, as well as six individuals translocated from PA to NJ. PA individuals are labeled with the years in which they were released in NJ (i.e., 2015, 2016 and 2017). All individuals are labeled as being sampled before translocations occurred (“pre-translocation”), during translocations or after translocations occurred (“post-translocation”). STRUCTURE results were CLUMPP-averaged across 10 runs when K is assumed to be equal to two. Admixture is indicated by a shift from the blue cluster associated with the resident population prior to human-mediated gene flow, to an increase in the orange cluster associated with the genetic profiles of the PA individuals

Prior to translocations, observed heterozygosity (HO) and expected heterozygosity (HE) were substantially lower in New Jersey than at sites in Indiana (IN) and Ohio (Table 2). However, genetic variability in the New Jersey population increased notably in the years following translocation (Table 2, Table 3, Fig. 2). For example, observed heterozygosity increased from 0.08 ± 0.02 in 2009 to 0.30 ± 0.02 in 2019 (Table 3, Fig. 2). Average HO and HE were comparable in Indiana, Ohio and (post-translocation) New Jersey (Tables 2, 3, Fig. 2).

Table 2 Mean observed heterozygosity (HO) ± SE, mean expected heterozygosity (HE) ± SE for Allegheny woodrats (Neotoma magister) genotyped at 134 SNPs
Table 3 Mean observed heterozygosity (HO) ± SE, mean expected heterozygosity (HE) ± SE and mean number of alleles (A) for Allegheny woodrats (Neotoma magister) captured in 2009 (n = 5), 2011 (n = 13), 2015 (n = 9), 2016 (n = 18), 2017 (n = 16), 2018 (n = 8) and 2019 (n = 13) in the Palisades, NJ and genotyped at 134 SNP loci
Fig. 2
figure 2

Capture index and mean observed heterozygosity (HO) ± SE for Allegheny woodrats (Neotoma magister) captured in 2009 (n = 5), 2011 (n = 13), 2015 (n = 9), 2016 (n = 18), 2017 (n = 16), 2018 (n = 8) and 2019 (n = 13) in the Palisades, NJ and genotyped at 134 SNP loci. Observed heterozygosity increased following translocations of six woodrats from Pennsylvania in 2015, 2016 and 2017

We identified 20 publications for which a Fluidigm SNPtype assay was used to genotype individuals at relatively few loci (38–192 SNPs) and HO and/or HE were reported (Table 4). Species described were members of Actinopterygii, Aves, Bivalvia, and Mammalia and predominantly considered of “Least concern” by the IUCN. Across studies, HO and HE ranged from 0.13 to 0.45 and 0.14 to 0.42; respectively (Table 4). The vast majority of HO and HE estimates for species characterized as “Least concern” (all but two) fell between 0.25 and 0.37. Median HO and HE were 0.32 and 0.31, respectively.

Table 4 Metrics of genetic variability, sample size and IUCN status for species genotyped using the Fluidigm® BioMark, HD™ Genotyping System

Discussion

Genome sequencing and SNP assay development

The data described herein represent only the second time a member of the genus Neotoma has undergone whole-genome sequencing [34]. Genetic resources for Neotoma magister are particularly limited [35], yet even low coverage sequencing can be used to generate tools that inform management of threatened species. For example, two lanes of paired-end sequencing and one lane of mate-paired sequencing enabled assembly of the complete mitochondrial genome [35] and identification of the 134 SNP loci described in this manuscript. Studies have shown that in some cases SNP genotyping can better reveal fine-scale population structure, provide evidence of differential selection amongst populations and estimate genome wide heterozygosity than other marker panels [36,37,38,39,40,44]. Genotyping 50 woodrats using both microsatellite and SNP loci indicates that our SNP assay provides increased statistical power for analyses. Furthermore, PIDsib estimates are also very low, indicating the panel can be used to distinguish between woodrats even when related individuals are present in the population [3133]. Given the spatial isolation of many woodrat populations (e.g., the remnant NJ Palisades population), the presence of closely related individuals should be assumed. Finally, DNA extracted from naturally shed feathers, hair and fecal samples and subsequent Fluidigm SNP genotyping has been used to identify individual golden eagles [24], wolves, wildcats, and bears [32, 33]. Given the low PID estimates associated with this assay, we anticipate a similar approach could be used to non-invasively monitor Allegheny woodrat populations from hair or fecal samples.

Temporal shifts in genetic variability following translocations to New Jersey’s remnant population

Conservation managers have long worried that translocations across extended geographic distances would result in relatively greater genetic distance amongst introduced and resident individuals, increasing the risk of outbreeding depression [45]. Recent studies, however, indicate that outbreeding depression rarely has negative impacts on the success of translocation programs [19, 46]. Furthermore, factors such as the genetic diversity of translocated individuals may be better predictors of fitness following introduction to a novel population than genetic distance [47]. Despite the relatively great geographic distance between source and resident populations inherent in this study, successful reproduction by translocated individuals clearly drove increases in genetic variability in subsequent years. Parentage analysis provides evidence that at least two woodrats translocated from Pennsylvania to New Jersey in 2015 went on to reproduce, as did their offspring. Increases in HO and HE were apparent as soon as 2016, making observed levels comparable to those among woodrat populations in IN and OH, and persisted through the end of the monitoring period in 2019. We also compared genetic variability in the NJ population to that of other species and determined that observed heterozygosity of New Jersey’s woodrats caught before 2015 was notably lower than any species listed as “Least concern” by the IUCN. Following translocations, HO and HE for the NJ population fall within the range of estimates generated across species. Increased abundance since 2015 provides additional evidence of potential genetic rescue. As such, this study joins relatively few in providing evidence of an increase in population size or growth rate following assisted gene flow (reviewed in [19]).

Management implications

Ongoing research on the conservation of Allegheny woodrats may inform best practices in translocating individuals to very small populations. Management guidelines recommend translocating a number of non-resident individuals that represent 20% of the recipient population to minimize the likelihood of swamping out local adaptive genetic variation [17, 20]. Genetic swamping (i.e., the rapid increase in frequency of alleles introduced by gene flow; [48, 49]) can result in the loss of private alleles within the recipient population. This, in turn, can lead to a loss of species-wide allelic diversity [49], even as the resident population’s genetic diversity increases. Efforts to minimize genetic swamping can lead to translocating very few individuals when recipient populations have low abundance. This study joins others in suggesting that successful reproduction by just one to three non-resident individuals can promote increased genetic diversity and abundance [50,51,52,53]. In some cases, however, these few immigrants achieve substantially elevated reproductive success in comparison to resident individuals, contributing to inbreeding in subsequent generations (e.g., arctic foxes, [51]; wolves, [54,55,56]). Even in the absence of direct observations of inbreeding, reproductive skew is known to decrease effective population size and result in the accelerated loss of genetic diversity due to drift [57]. Disproportionate reproductive success by translocated individuals may prove to be common in Allegheny woodrats if sex ratios are skewed in small populations [50], resident individuals prefer to mate with translocated individuals as an inbreeding avoidance mechanism [58], or F1 offspring have increased fitness stemming from heterosis [59,60,61]. Both this study and Davis et al., [50] document observations of translocated male Allegheny woodrats siring 39 and 35% of young trapped in the subsequent season, in their respective populations. A few known instances of inbreeding amongst relatives followed in subsequent generations (Table 2, [50]) but, encouragingly, coincided with stable or increasing population-wide genetic variability and abundance. Conservation managers monitoring small populations might consider genotyping individuals, conducting parentage analyses, and monitoring genetic diversity on a bi-yearly or yearly basis. This would allow for rapid translocation of additional individuals to small populations if elevated reproductive success seems likely to lead to inbreeding events, or, to supplement previous, unsuccessful attempts at genetic rescue (i.e., if non-resident survivorship is low). It is worth noting that even in the absence of inbreeding, truly isolated populations (like that of the Palisades) will ultimately require additional human-mediated gene flow to counteract loss of genetic diversity due to genetic drift.

There are additional ways in which ongoing studies of woodrat translocations have the potential to add depth to our understanding of genetic rescue and restoration. Studies of genetic rescue have typically considered populations as discrete units, uninterrupted by landscape features. The natural tendency for woodrats to exist in metapopulations give scientists the opportunity to study genetic rescue throughout heterogeneous landscapes and, in particular, how alleles introduced at one habitat patch have the potential to move amongst sites. Just as recent studies have proposed choosing specific individuals for their ability to reduce inbreeding depression in genetically depauperate populations [19, 62], specific habitat sites might be targeted for releases if they are connected by natural dispersal corridors to other portions of the metapopulation. Indeed, recent work on the landscape genetics of Virginia’s Allegheny woodrats suggests that low elevation, rather than anthropogenic barriers such as roads, might prevent translocated individuals and/or their offspring from dispersing amongst habitat sites [63].

Conclusions

Herein, we describe a novel SNP assay, which provides increased statistical power to studies of a species commonly found in small and consanguineous populations. Our study has important implications for remnant populations of threatened species that are geographically isolated from the nearest metapopulation. Translocating small numbers of individuals to very small populations may increase the risk of reproductive skew followed by genetic drift and inbreeding, necessitating increased monitoring following introductions. Despite this, human-mediated gene flow is likely to be integral to the persistence of remnant populations. Our results indicate, encouragingly, that small numbers of introduced, genetically variable individuals can successfully reproduce, increase population-wide genetic diversity, and facilitate increased abundance.

Methods

Genome assembly and annotation

We extracted deoxyribonucleic acid (DNA) from a tail clip of a single N. magister individual by pairing commercially available extraction (DNEasy Blood and Tissue, Qiagen, Venlo, the Netherlands) and clean-up (DNA clean & Concentrator, Zymo Research, Irvine, California) kits in accord with the manufacturers’ instructions. We conducted three lanes of paired-end and one lane of mate-paired sequencing using an Illumina HiSeq2500. We used Trimmomatic [64] to remove adaptors, discard short reads and trim poor quality bases from 5′ and 3′ ends of raw sequence reads as described in Schofield et al., [35]. We used ABySS 1.9.0 [65] to conduct several assemblies with kmer lengths ranging from 40 to 85. PE reads were used to generate contigs. MP reads were used to infer the order, orientation, and distance between contigs, linking them together in scaffolds. The assembly with the greatest N50 value and longest scaffold was used for downstream analyses. BUSCO v5 [66], implemented by gVolante [67], was used to evaluate completeness of the genome.

We used the MAKER 2.28 pipeline [68] to annotate all scaffolds greater than 10 kb, following the methods described in Doyle et al. [69] and Doyle et al. [26]. To briefly summarize, we first used Repeat-Masker to identify and mask stretches of repetitive DNA. Second, we downloaded 6762 Mus musculus protein sequences from the UniProtKB database (www.uniprot.org) and used the protein2genome setting in MAKER to generate gene annotations. These annotations were subsequently used to train SNAP [70] and generate ab initio predictions. Third, we aligned protein sequences and 93,400 Mus musculus expressed sequence tag (EST) sequences to the genome using BLAST and used InterProScan to identify putative protein domains. Finally, all ab initio gene predictions supported by protein, EST or InterProScan evidence were promoted to gene annotations.

SNP discovery and assay design

We identified SNPs as in Doyle et al. [26]. Briefly, we aligned paired-end reads back to the draft genome assembly using BWA 0.7.12 [71] and used Picard 2.3 (http://broadinstitute.github.io/picard) to sort and identify duplicate reads. We used the GATK 3.6 pipeline [72, 73] to realign reads around indels and identify high quality SNPs with a Phred quality score ≥ 30. We then selected 95 autosomal nuclear markers associated with gene deserts (i.e., “neutral” markers) and 97 autosomal nuclear markers associated with protein-coding genes. We deliberately chose no more than one SNP of each category from a given scaffold to minimize linkage disequilibrium. To identify neutral markers, we quantified the distances between all SNPs and genes using the BEDtools suite [74], ultimately choosing markers at the 95% percentile distance from genes. We used SnpEff 4.3 [75] to find SNPs associated with non-synonymous changes in the exonic regions of genes (i.e., “gene-associated” markers). IGV 2.3 [76] was used to confirm that at least 60 nucleotides of high-quality flanking sequence were present upstream and downstream of the marker, that guanine-cytosine (GC) content was less than 65%, and that no other variable sites were present within 20 nucleotides.

DNA extraction and SNP genotyping

We trapped and subsequently genotyped 82 woodrats sampled from the Palisades, NJ between 2009 and 2019 (Tables 2, 3), including 18 and 64 individuals sampled before and after translocations began, respectively. We followed standard live-trapping protocols [77, 78] and collected a 2-mm ear punch from each individual, which was preserved in 70–100% ethanol. For each year trapping occurred, we calculated a capture index by dividing the number of unique individuals caught by the number of trap nights and multiplying by 10 [79]. Calculating a trap index allows us to control for differences in the number of nights trapping occurred across years [77]. DNA extractions were performed using an ammonium acetate protocol [14] or the Zymo Quick-DNA Miniprep Plus Kit. We used the Fluidigm® BioMark HD™ Genotyping System to genotype these individuals. Additionally, we genotyped the six individuals translocated from Pennsylvania to New Jersey between 2015 and 2017. Finally, we opportunistically genotyped 172 and 58 samples collected from Indiana and Ohio, respectively (Table 2). These samples were collected between 2015 and 2019 as part of long-term monitoring studies.

SNP calls were edited using the Fluidigm® Genotyping Analysis Software. We excluded markers from downstream analysis when data did not cluster into distinct homozygous and heterozygous states and if minor allele frequencies were less than 0.025. We used chi-squared tests implemented by GenAlEx [24, 26] to test for departures from Hardy–Weinberg equilibrium. Following Bonferroni correction, a single locus was found to be consistently out of Hardy–Weinberg equilibrium across years and was omitted, leaving 134 loci. We excluded individuals from analysis if ≥ 7 loci were not successfully genotyped, as call rates tend to be negatively correlated with genotyping errors [24, 32]. Using snpStats [80], we identified a number of markers in linkage disequilibrium but assumed that in many cases this was due to consanguinity, rather than two markers being in close proximity along the genome. However, to meet the assumptions of Cervus 3.0.7 [81] and STRUCTURE 2.3.4 [82, 83], we identified all pair-wise comparisons with r2 > 0.2 and removed one marker in each case, creating a reduced dataset of 70 loci in which all SNPs are in linkage equilibrium.

Probability of identity using microsatellite and SNP markers

Previous studies of N. magister utilized relatively small panels of 11–22 microsatellite markers (e.g., [15, 63, 84, 85]). To evaluate the statistical power associated with each approach, we genotyped 50 woodrats captured in 2017 and 2019 in the Adams County, OH at both 11 microsatellites and 134 SNP loci (Additional File 3 and Additional File 4). We subsequently calculated the probability that two randomly chosen individuals in the population would have identical genotypes (PID), using each marker panel. We additionally calculated PIDsib, which represents a conservative upper bound for the likelihood that two individuals sampled from a population will have the same genotype by chance [32, 33]. This estimate is particularly useful when substructure is present in the population (i.e., related individuals; [32, 33]).

Reproductive success and genetic variability following translocations to the Palisades

We used Cervus 3.0.7 [81] to assign individuals sampled between 2015 and 2019 to dams and sires. For individuals trapped for the first time in each year, all woodrats trapped in that same year and in all previous years were considered candidate parents. Simulations included 100,000 replicate cycles. The proportion of candidate dams and sires sampled was estimated to be 0.80, based on the probability of capture estimated from comparable trapping approaches of other woodrat populations [16]. The proportion of typed loci was 0.99 and the proportion of loci mistyped was set to 0.04 [26]. The minimum confidence level for parentage assignment was 95%.

STRUCTURE 2.3.4 [82, 83], STRUCTURE HARVESTER 0.6.94 [86], and Clumpak [87] were used to visualize admixture in the Palisades, NJ woodrat population across time. We utilized the reduced dataset (i.e., with all 70 loci in linkage equilibrium, see above) for 82 individuals sampled between 2009 and 2019, as well as the six individuals translocated from PA. We considered values of K = 1–8, running each value 10 times with an initial burn-in of 100,000 Markov chain Monte Carlo (MCMC) iterations and 1,000,000 subsequent iterations for each value. We assumed an admixture ancestry model and allowed for correlated allele frequencies [82]. The results were interpreted using mean likelihood values of K and ΔK [86].

We used GenAlEx [88] to calculate allele frequencies and expected and observed heterozygosity in the years before (2009, 2011), during (2015–2017) and after translocations (2018, 2019) to the Palisades, NJ population. To provide context for our interpretation of temporal changes in genetic variability in the Palisades population, we (1) used GenAlEx [88] to calculate allele frequencies and expected and observed heterozygosity in Indiana, New Jersey and Ohio and (2) surveyed the literature for estimates of observed and expected heterozygosity generated using the Fluidigm® BioMark HD™ Genotyping System and relatively small SNP assays (e.g., 96–192 loci). To conduct our literature review, we searched for the phrases “SNP type assay”, “SNPtype assay”, “Fluidigm SNP assay”, “Fluidigm SNP chip” and “Fluidigm Genotyping Analysis Software” in Google Scholar. For all studies of non-human animals for which observed and/or expected heterozygosity were described, we recorded the number of loci in the assay, sample size, metrics of genetic variability and International Union for Conservation of Nature (IUCN) status. If average HO and HE were not provided, we averaged across per-locus values or population-specific values when able. When interpreting these results, an important caveat is that we do not have a robust understanding of how SNP HE, HO and allelic diversity vary with categorizations such as body size, conservation status, habitat, migratory behavior, taxonomic group and trophic class. In contrast, these relationships have been extensively studied utilizing estimates of genetic variation generated with microsatellite loci [89,90,91,92,93].

Availability of data and materials

The datasets generated and/or analyzed during the current study are available in NCBI’s Short Read Archive (BioSample accession # SAMN25554335, BioProject accession # PRJNA802531), Dryad (https://doi.org/10.5061/dryad.rjdfn2zdb) and the supplementary files associated with this manuscript.

Abbreviations

NJ:

New Jersey

SNPs:

Single nucleotide polymorphisms

Gb:

Gigabases

bp:

Basepairs

kb:

Kilobases

PE:

Paired-end

MP:

Mate-paired

OH:

Ohio

PID :

Probability of identity

PIDsib:

Probability of identity among siblings

PA:

Pennsylvania

HO :

Observed heterozygosity

HE :

Expected heterozygosity

IN:

Indiana

DNA:

Deoxyribonucleic acid

EST:

Expressed sequence tag

GC:

Guanine-cytosine

MCMC:

Markov chain Monte Carlo iterations

IUCN:

International Union for Conservation of Nature

References

  1. Wilcox BA, Murphy DD. Conservation strategy: the effects of fragmentation on extinction. Am Nat. 1985;125:879–87.

    Article  Google Scholar 

  2. Young A, Boyle T, Brown T. The population genetic consequences of habitat fragmentation for plants. Trends Ecol Evol. 1996;11:413–8.

    Article  CAS  PubMed  Google Scholar 

  3. Mikoláš M, Tejkal M, Kuemmerle T, Griffiths P, Svoboda M, Hlásny T, et al. Forest management impacts on capercaillie (Tetrao urogallus) habitat distribution and connectivity in the Carpathians. Landsc Ecol. 2017;32:163–79.

    Article  Google Scholar 

  4. Blanton RE, Cashner MF, Thomas MR, Brandt SL, Floyd MA. Increased habitat fragmentation leads to isolation among and low genetic diversity within populations of the imperiled Kentucky Arrow Darter (Etheostoma sagitta spilotum). Conserv Genet. 2019;20:1009–22.

    Article  CAS  Google Scholar 

  5. Saccheri I, Kuussaari M, Kankare M, Vikman P, Fortelius W, Hanski I. Inbreeding and extinction in a butterfly metapopulation. Nature. 1998;392:491–4.

    Article  CAS  Google Scholar 

  6. Hanski I. Metapopulation ecology. Oxford: Oxford University Press; 1999.

    Google Scholar 

  7. Dallas TA, Saastamoinen M, Schulz T, Ovaskainen O. The relative importance of local and regional processes to metapopulation dynamics. J Anim Ecol. 2020;89:884–96.

    Article  PubMed  Google Scholar 

  8. Couvet D. Deleterious effects of restricted gene flow in fragmented populations. Conserv Biol. 2002;16:369–76.

    Article  Google Scholar 

  9. Nonaka E, Sirén J, Somervuo P, Ruokolainen L, Ovaskainen O, Hanski I. Scaling up the effects of inbreeding depression from individuals to metapopulations. J Anim Ecol. 2019;88:1202–14.

    Article  PubMed  Google Scholar 

  10. Poole E. A life history sketch of the Allegheny woodrat. J Mammal. 1940;21:249–70.

    Article  Google Scholar 

  11. Castleberry SB, Mengak MT, Ford WM. Neotoma magister. Mamm Species. 2006;789:1–5.

    Article  Google Scholar 

  12. Hassinger J, Butchkoski C, Diefenbach D. Managing surface rock communities for Neotoma magister. In: Peles J, Wright J, editors. The Allegheny Woodrat: Ecology, conservation, and management of a declining species. New York: Springer; 2008. p. 133–52.

    Chapter  Google Scholar 

  13. Wood P. Woodrat population dynamics and movement patterns. In: Peles J, Wright J, editors. The Allegheny Woodrat: ecology, conservation, and management of a declining species. New York: Spring Publishing Company; 2008. p. 45–62.

    Chapter  Google Scholar 

  14. Smyser TJ, Duchamp JE, Johnson SA, Larkin JL, Rhodes OE. Consequences of metapopulation collapse: comparison of genetic attributes between two Allegheny woodrat metapopulations. Conserv Genet. 2012;13:849–58.

    Article  Google Scholar 

  15. Smyser TJ, Page LK, College W, Eugene O Jr. Synergistic stressors and the dilemma of conservation in a multivariate world: a case study in Allegheny woodrats. Anim Conserv. 2012;15:205–13.

    Article  Google Scholar 

  16. Smyser TJ, Stauffer GE, Johnson SA, Hudson CM, Rhodes OE, Swihart RK. Annual survival of Allegheny woodrats in a nonequilibrium metapopulation. J Mammal. 2016;97:1699–708.

    Article  Google Scholar 

  17. Hedrick P. Gene flow and genetic restoration: the Florida Panther as a case study. Conserv Biol. 1995;9:996–1007.

    Article  PubMed  Google Scholar 

  18. Frankham R. Genetic rescue of small inbred populations: meta-analysis reveals large and consistent benefits of gene flow. Mol Ecol. 2015;24:2610–8.

    Article  PubMed  Google Scholar 

  19. Whiteley AR, Fitzpatrick SW, Funk WC, Tallmon DA. Genetic rescue to the rescue. Trends Ecol Evol. 2015;30:42–9.

    Article  PubMed  Google Scholar 

  20. Frankham R. Genetic rescue benefits persist to at least the F3 generation, based on a meta-analysis. Biol Conserv. 2016;195:33–6.

    Article  Google Scholar 

  21. Paetkau D, Strobeck C. Microsatellite analysis of genetic variation in black bear populations. Mol Ecol. 1994;3:489–95.

    Article  CAS  PubMed  Google Scholar 

  22. Kalinowski S. How many alleles per locus should be used to estimate genetic distances? Heredity. 2002;88:62–5.

    Article  CAS  PubMed  Google Scholar 

  23. Tokarska M, Marshall T, Kowalczyk R, Wójcik JM, Pertoldi C, Kristensen TN, et al. Effectiveness of microsatellite and SNP markers for parentage and identity analysis in species with low genetic diversity: The case of European bison. Heredity. 2009;103:326–32.

    Article  CAS  PubMed  Google Scholar 

  24. Doyle JM, Katzner TE, Roemer GW, Cain JW, Millsap BA, McIntyre CL, et al. Genetic structure and viability selection in the golden eagle (Aquila chrysaetos), a vagile raptor with a Holarctic distribution. Conserv Genet. 2016;17:1307–22.

    Article  Google Scholar 

  25. DeWoody J, Fernandez N, Brüniche-Olsen A, Antonides J, Doyle J, San Miguel P, et al. Characterization of the gray whale (Eschrichtius robustus) genome and a genotyping array based on single nucleotide polymorphisms in candidate genes. Biol Bull. 2017;232:186–97.

    Article  CAS  PubMed  Google Scholar 

  26. Doyle J, Bell D, Bloom P, Emmons G, Katzner T, LePre L, et al. New insights into the phylogenetics and population structure of the prairie falcon (Falco mexicanus). BMC Genomics. 2018;19:1–14.

    Article  Google Scholar 

  27. Hauser L, Baird M, Hilborn R, Seeb LW, Seeb JE. An empirical comparison of SNPs and microsatellites for parentage and kinship assignment in a wild sockeye salmon (Oncorhynchus nerka) population. Mol Ecol Resour. 2011;11(SUPPL. 1):150–61.

    Article  PubMed  Google Scholar 

  28. Buchanan JW, Woronuk GN, Marquess FL, Lang K, James ST, Deobald H, et al. Analysis of validated and population-specific single nucleotide polymorphism parentage panels in pedigreed and commercial beef cattle populations. Can J Anim Sci. 2017;97:231–40.

    CAS  Google Scholar 

  29. Kaiser SA, Taylor SA, Chen N, Sillett TS, Bondra ER, Webster MS. A comparative assessment of SNP and microsatellite markers for assigning parentage in a socially monogamous bird. Mol Ecol Resour. 2017;17:183–93.

    Article  CAS  PubMed  Google Scholar 

  30. Thongda W, Zhao H, Zhang D, Jescovitch LN, Liu M, Guo X, et al. Development of SNP Panels as a new tool to assess the genetic diversity, population structure, and parentage analysis of the Eastern Oyster (Crassostrea virginica). Mar Biotechnol. 2018;20:385–95.

    Article  CAS  Google Scholar 

  31. Waits LP, Luikart G, Taberlet P. Estimating the probability of identity among genotypes in natural populations: cautions and guidelines. Mol Ecol. 2001;10:249–56.

    Article  CAS  PubMed  Google Scholar 

  32. von Thaden A, Cocchiararo B, Jarausch A, Jüngling H. Assessing SNP genotyping of noninvasively collected wildlife samples using microfluidic arrays. Sci Rep. 2017;7:10768.

    Article  Google Scholar 

  33. von Thaden A, Nowak C, Tiesmeyer A, Reiners TE, Alves PC, Lyons LA, et al. Applying genomic data in wildlife monitoring: development guidelines for genotyping degraded samples with reduced single nucleotide polymorphism panels. Mol Ecol Resour. 2020;20:662–80.

    Article  Google Scholar 

  34. Campbell M, Oakeson KF, Yandell M, Halpert JR, Dearing D. The draft genome sequence and annotation of the desert woodrat Neotoma lepida. Genomics Data. 2016;9:58–9.

    Article  PubMed  PubMed Central  Google Scholar 

  35. Schofield M, Duchamp J, Larkin JL, Smyser TJ, Doyle JM, Schofield M, et al. Mitochondrial genome of an Allegheny Woodrat (Neotoma magister). Mitochondrial DNA Part B. 2018;3:256–8.

    Article  PubMed  PubMed Central  Google Scholar 

  36. Ruegg KC, Anderson EC, Paxton KL, Apkenas V, Lao S, Siegel RB, et al. Mapping migration in a songbird using high-resolution genetic markers. Mol Ecol. 2014;23:5726–39.

    Article  PubMed  Google Scholar 

  37. Bekkevold D, Helyar SJ, Limborg MT, Nielsen EE, Hemmer-Hansen J, Clausen LAW, et al. Gene-associated markers can assign origin in a weakly structured fish, Atlantic herring. ICES J Mar Sci. 2015;72:1790–801.

    Article  Google Scholar 

  38. Malenfant RM, Coltman DW, Davis CS. Design of a 9K illumina BeadChip for polar bears (Ursus maritimus) from RAD and transcriptome sequencing. Mol Ecol Resour. 2015;15:587–600.

    Article  CAS  PubMed  Google Scholar 

  39. Ferchaud A-L, Pedersen SH, Bekkevold D, Jian J, Niu Y, Hansen MM. A low-density SNP array for analyzing differential selection in freshwater and marine populations of threespine stickleback (Gasterosteus aculeatus). BMC Genomics. 2014. https://doi.org/10.1186/1471-2164-15-867.

    Article  PubMed  PubMed Central  Google Scholar 

  40. Limborg MT, Helyar SJ, De Bruyn M, Taylor MI, Nielsen EE, Ogden R, et al. Environmental selection on transcriptome-derived SNPs in a high gene flow marine fish, the Atlantic herring (Clupea harengus). Mol Ecol. 2012;21:2355–6.

    Article  Google Scholar 

  41. DeWoody YD, DeWoody JA. On the estimation of genome-wide heterozygosity using molecular markers. J Hered. 2005;96:85–8.

    Article  CAS  PubMed  Google Scholar 

  42. Roffler GH, Amish SJ, Smith S, Cosart TED, Kardos M. SNP discovery in candidate adaptive genes using exon capture in a free-ranging alpine ungulate. Mol Ecol Resour. 2016;16:1147–64.

    Article  CAS  PubMed  Google Scholar 

  43. Hoffman JI, Simpson F, David P, Rijks JM, Kuiken T, Thorne MAS, et al. High-throughput sequencing reveals inbreeding depression in a natural population. Proc Natl Acad Sci USA. 2014;111:3775–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Clemento AJ, Crandall ED, Garza JC, Anderson EC, Garza JC, Anderson EC. Evaluation of a single nucleotide polymorphism baseline for genetic stock identification of Chinook Salmon (Oncorhynchus tshawytscha) in the California current large marine ecosystem. Fish Bull. 2014;112:112–30.

    Article  Google Scholar 

  45. Edmands S. Between a rock and a hard place: evaluating the relative risks of inbreeding and outbreeding for conservation and management. Mol Ecol. 2007;16:463–75.

    Article  PubMed  Google Scholar 

  46. Frankham R, Ballou JD, Eldridge MDB, Lacy RC, Ralls K, Dudash MR, et al. Predicting the probability of outbreeding depression. Conserv Biol. 2011;25:465–75.

    Article  PubMed  Google Scholar 

  47. Scott PA, Allison LJ, Field KJ, Averill-Murray RC, Bradley SH. Individual heterozygosity predicts translocation success in threatened desert tortoises. Science. 2020;370:1086–9.

    Article  PubMed  Google Scholar 

  48. Hufford K, Mazer S. Plant ecotypes: genetic differentiation in the age of ecological restoration. Trends Ecol Evol. 2003;18:147–55.

    Article  Google Scholar 

  49. Lenormand T. Gene flow and the limits to natural selection. Trends Ecol Evol. 2002;17:183–9.

    Article  Google Scholar 

  50. Davis MM, Smyser TJ, Johnson SA, Duchamp J, Larkin JL, Swihart RK, et al. Reproductive success of captive-reared Allegheny Woodrats (Neotoma magister) released into genetically depauperate populations. Conserv Genet. 2021;22:903–12.

    Article  Google Scholar 

  51. Lotsander A, Hasselgren M, Larm M, Wallén J, Angerbjörn A, Norén K. Low persistence of genetic rescue across generations in the arctic fox (Vulpes lagopus). J Hered. 2021;112:276–85.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Gustafson KD, Vickers TW, Boyce WM, Ernest HB. A single migrant enhances the genetic diversity of an inbred puma population. R Soc Open Sci. 2017;4:1–12.

    Article  Google Scholar 

  53. Vilà C, Sundqvist AK, Flagstad Ø, Seddon J, Björnerfeldt S, Kojola I, et al. Rescue of a severely bottlenecked wolf (Canis lupus) population by a single immigrant. Proc R Soc B. 2003;270:91–7.

    Article  PubMed  PubMed Central  Google Scholar 

  54. Hedrick PW, Peterson RO, Vucetich LM, Adams JR, Vucetich JA. Genetic rescue in Isle Royale wolves: genetic analysis and the collapse of the population. Conserv Genet. 2014;15:1111–21.

    Article  Google Scholar 

  55. Hedrick P, Robinson J, Peterson R, Vucetich J. Genetics and extinction and the example of Isle Royale wolves. Anim Conserv. 2019;22:302–9.

    Article  Google Scholar 

  56. Robinson JA, Räikkönen J, Vucetich LM, Vucetich JA, Peterson RO, Lohmueller KE, et al. Genomic signatures of extensive inbreeding in Isle Royale wolves, a population on the threshold of extinction. Sci Adv. 2019. https://doi.org/10.1126/sciadv.aau0757.

    Article  PubMed  PubMed Central  Google Scholar 

  57. Stoffel MA, Humble E, Paijmans AJ, Acevedo-Whitehouse K, Chilvers BL, Dickerson B, et al. Demographic histories and genetic diversity across pinnipeds are shaped by human exploitation, ecology and life-history. Nat Commun. 2018;9:1–12.

    Article  Google Scholar 

  58. Jones C, Noble L, Jones J, Tegelstrom H, Triggs G, Berry R. Differential male genetic success determines gene flow in an experimentally manipulated mouse population. Proc R Soc B. 1995;260:251–6.

    Article  CAS  PubMed  Google Scholar 

  59. Ingvarsson PK, Whitlock MC. Heterosis increases the effective migration rate. Proc R Soc B. 2000;267:1321–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Heinsohn R, Ebert D, Legge S, Peakall R. Genetic evidence for cooperative polyandry in reverse dichromatic Eclectus parrots. Anim Behav. 2007;74:1047–54.

    Article  Google Scholar 

  61. Saccheri IJ, Brakefield PM. Rapid spread of immigrant genomes into inbred populations. Proc R Soc B. 2002;269:1073–8.

    Article  PubMed  PubMed Central  Google Scholar 

  62. Allendorf FW, Hohenlohe P, Luikart G. Genomics and the future of conservation genetics. Nat Rev Genetics. 2010;11:697–709.

    Article  CAS  PubMed  Google Scholar 

  63. Kanine JM, Kierepka EM, Castleberry SB, Mengak MT, Nibbelink NP, Glenn TC. Influence of landscape heterogeneity on the functional connectivity of Allegheny woodrats (Neotoma magister) in Virginia. Conserv Genet. 2018;19:1259–68.

    Article  CAS  Google Scholar 

  64. Bolger A, Lohse M, Usadel B. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics. 2014;30:2114–20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Simpson JT, Wong K, Jackman SD, Schein JE, Jones SJM, Birol I. ABySS: a parallel assembler for short read sequence data. Genome Res. 2009;19:1117–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Simao FA, Waterhouse RM, Ioannidis P, Kriventseva EV, Zdobnov EM. BUSCO: assessing genome assembly and annotation completeness with single-copy orthologs. Bioinformatics. 2015;31:3210–2.

    Article  CAS  PubMed  Google Scholar 

  67. Nishimura O, Hara Y, Kuraku S. gVolante for standardizing completeness assessment of genome and transcriptome assemblies. Bioinformatics. 2017;33:3635–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Cantarel BL, Korf I, Robb SMC, Parra G, Ross E, Moore B, et al. MAKER: an easy-to-use annotation pipeline designed for emerging model organism genomes. Genome Res. 2008;18:188–96.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Doyle JM, Katzner TE, Bloom PH, Ji Y, Wijayawardena BK, DeWoody JA. The genome sequence of a widespread apex predator, the golden eagle (Aquila chrysaetos). PLoS ONE. 2014;9:20–2.

    Article  Google Scholar 

  70. Korf I. Gene finding in novel genomes. BMC Bioinformatics. 2004;5:59.

    Article  PubMed  PubMed Central  Google Scholar 

  71. Li H, Durbin R. Fast and accurate short read alignment with Burrows-Wheeler transform. Bioinformatics. 2009;25:1754–60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. van der Auwera GA, Carneiro MO, Hartl C, Poplin R, del Angel G, Levy-Moonshine A, et al. From fastQ data to high-confidence variant calls: the genome analysis toolkit best practices pipeline. Current Protoc Bioinf. 2013;43:1–33.

    Article  Google Scholar 

  73. DePristo MA, Banks E, Poplin R, Garimella K, Maguire JR, Hartl C, et al. A framework for variation discovery and genotyping using next-generation DNA sequencing data. Nat Genet. 2011;43:491–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Quinlan AR, Hall IM. BEDTools : a flexible suite of utilities for comparing genomic features. Bioinformatics. 2010;26:841–2.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Cingolani P, Platts A, Wang L, Coon M, Nguyen T, Wang L, et al. A program for annotating and predicting the effects of single nucleotide polymorphisms, SnpEff: SNPs in the genome of Drosophila melanogaster strain w1118; iso-2, iso-3. Fly. 2012;6:1–13.

    Article  Google Scholar 

  76. Thorvaldsdóttir H, Robinson J, Mesirov J. Integrative Genomics Viewer (IGV): high-performance genomics data visualization and exploration. Brief Bioinform. 2013;14:178–92.

    Article  PubMed  Google Scholar 

  77. Caughley G. Analysis of vertebrate populations. Hoboken: Wiley; 1977.

    Google Scholar 

  78. Johnson SA. Reassessment of the Allegheny woodrat (Neotoma magister) in Indiana. Proc Indiana Acad Sci. 2002;1:56–66.

    Google Scholar 

  79. Mengak M, Butchkoski C, Feller D, Johnson S. Lessons from long-term monitoring of woodrat populations. In: Peles J, Wright J, editors. The Allegheny Woodrat: ecology, conservation, and management of a declining species. New York: Springer; 2008. p. 109–31.

    Chapter  Google Scholar 

  80. Clayton D. snpStats: SnpMatrix and XSnpMatrix classes and methods. 2014.

  81. Kalinowski ST, Taper ML, Marshall TC. Revising how the computer program CERVUS accommodates genotyping error increases success in paternity assignment. Mol Ecol. 2007;16:1099–106.

    Article  PubMed  Google Scholar 

  82. Pritchard JK, Stephens M, Donnelly P. Inference of population structure using multilocus genotype data. Genetics. 2000;155:945–59.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Falush D, Stephens M, Pritchard JK. Inference of population structure using multilocus genotype data: linked loci and correlated allele frequencies. Genetics. 2003;164:1567–87.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Castleberry SB, King TL, Wood PB, Ford WM. Microsatellite DNA analysis of population structure in Allegheny woodrats (Neotoma magister). J Mammal. 2002;83:1058–70.

    Article  Google Scholar 

  85. Smyser TJ, Johnson SA, Page LK, Hudson CM, Rhodes OE. Use of experimental translocations of allegheny woodrat to decipher causal agents of decline. Conserv Biol. 2013;27:752–62.

    Article  PubMed  Google Scholar 

  86. Earl D, VonHoldt BM. STRUCTURE HARVESTER: a website and program for visualizing STRUCTURE output and implementing the Evanno method. Conserv Genet Resour. 2011;4:359–61.

    Article  Google Scholar 

  87. Kopelman NM, Mayzel J, Jakobsson M, Rosenberg NA, Mayrose I. CLUMPAK: A program for identifying clustering modes and packaging population structure inferences across K. Mol Ecol Resour. 2015;15:1179–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Peakall R, Smouse PE. GenAlEx 6,5: genetic analysis in excel population genetic software for teaching and research–an update. Bioinformatics. 2012;28:2537–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. DeWoody J, Avise JC. Microsatellite variation in marine, freshwater and anadromous fishes compared with other animals. J Fish Biol. 2000;56:461–73.

    Article  CAS  Google Scholar 

  90. Eo S, Doyle J, DeWoody J. Genetic diversity in birds is associated with body mass and habitat type. J Zool. 2011;283:220–6.

    Article  Google Scholar 

  91. Doyle JM, Hacking CC, Willoughby JR, Sundaram M, DeWoody JA. Mammalian genetic diversity as a function of habitat, body size, trophic class, and conservation Status. J Mammal. 2015;96:564–72.

    Article  Google Scholar 

  92. Willoughby JR, Sundaram M, Wijayawardena BK, Kimble SJA, Ji Y, Fernandez NB, et al. The reduction of genetic diversity in threatened vertebrates and new recommendations regarding IUCN conservation rankings. Biol Conserv. 2015;191:495–503.

    Article  Google Scholar 

  93. Willoughby JR, Sundaram M, Wijayawardena BK, Lamb MC, Kimble SJA, Ji Y, et al. Biome and migratory behaviour significantly influence vertebrate genetic diversity. Biol J Lin Soc. 2017;121:446–57.

    Article  Google Scholar 

  94. Schunter C, Garza JC, Macpherson E, Pascual M. SNP development from RNA-seq data in a nonmodel fish: how many individuals are needed for accurate allele frequency prediction? Mol Ecol Resour. 2014;14:157–65.

    Article  CAS  PubMed  Google Scholar 

  95. Baetscher DS, Hasselman DJ, Reid K, Palkovacs EP, Garza JC. Discovery and characterization of single nucleotide polymorphisms in two anadromous alosine fishes of conservation concern. Ecol Evol. 2017;7:6638–48.

    Article  PubMed  PubMed Central  Google Scholar 

  96. Reid K, Palkovacs EP, Hasselman DJ, Baetscher D, Kibele J, Gahagan B, et al. Comprehensive evaluation of genetic population structure for anadromous river herring with single nucleotide polymorphism data. Fish Res. 2018;206:247–58.

    Article  Google Scholar 

  97. Davis CD, Garza JC, Banks MA. Identification of multiple genetically distinct populations of Chinook salmon (Oncorhynchus tshawytscha) in a small coastal watershed. Environ Biol Fishes. 2017;100:923–33.

    Article  Google Scholar 

  98. Ford MJ, Murdoch AR, Hughes MS, Seamons TR, Lahood S. Broodstock history strongly influences natural spawning success in hatchery Steelhead (Oncorhynchus mykiss). PLoS ONE. 2016;11:1–20.

    Article  Google Scholar 

  99. Katzner TE, Nelson DM, Braham MA, Doyle JM, Fernandez NB, Duerr AE, et al. Golden Eagle fatalities and the continental-scale consequences of local wind-energy generation. Conserv Biol. 2017;31:406–15.

    Article  PubMed  Google Scholar 

  100. Mathur S, Tomeček JM, Heniff A, Luna R, DeWoody JA. Evidence of genetic erosion in a peripheral population of a North American game bird: the Montezuma quail (Cyrtonyx montezumae). Conserv Genet. 2019;20:1369–81.

    Article  Google Scholar 

  101. Turley B, Reece K, Shen J, Ho J, Ximing L, Jan G. Multiple drivers of interannual oyster settlement and recruitment in the lower Chesapeake Bay. Conserv Genet. 2019;20:1057–71.

    Article  Google Scholar 

  102. Turley BD. Oyster reef connectivity inferred via population genetic analysis. 2015.

  103. Day GQ, Ng J, Oldt RF, Houghton PW, Smith DG, Kanthaswamy S. DNA-based determination of ancestry in Cynomolgus Macaques (Macaca fascicularis). J Am Assoc Lab Anim Sci. 2018;57:432–42.

    Article  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

This work used the Extreme Science and Engineering Discovery Environment (XSEDE), which is supported by National Science Foundation grant number ACI-1548562. Specifically, this work used Comet at the San Diego Supercomputer Center through allocation TG-BIO170095. The authors thank Jeffrey Larkin for preparing proposals for funding that supported this work; Amanda Menasion, Sharon Petzinger, Dave Kenny, Kathleen LoGuidice, Allegra Mitchell, Kate Hausman for help collecting N. magister samples and tracking the translocated individuals with telemetry and cameras; and Kylie Brandt, Miracle Davis, Heather Farrington, Mandy Schofield and Eli Seeman for assistance with SNP panel development, DNA extraction and quantification and microsatellite genotyping.

Funding

This research was supported by the Indiana Department of Natural Resources, the New Jersey DEP Fish and Wildlife’s Endangered and Nongame Species Program, the USFWS Fish and Wildlife Service Wildlife Restoration Program, the Pennsylvania Game Commission, the Nature Conservancy, Purdue University and Towson University Fisher College of Science and Mathematics.

Author information

Authors and Affiliations

Authors

Contributions

JD, GF, GGT, TJS and JMD conceived and designed the research. JD, GF, CM, BW and GGT collected field data and samples. MMG, SK, TJS and JMD performed DNA extractions, DNA quantification and SNP genotyping. TJS and JMD conducted whole-genome sequencing, assembled the draft genome and designed the SNP panel. MMG, GF, SK and JMD led the writing effort and all authors (MMG, GF, JD, SK, CM, TJS, GGT, BW and JMD) reviewed, revised, commented on and approved the final version of the manuscript.

Corresponding author

Correspondence to Jacqueline M. Doyle.

Ethics declarations

Ethics approval and consent to participate

All methods and protocols were approved by an institutional and/or licensing committee and carried out in accordance with relevant guidelines and regulations. Standard live trapping and sample collection protocols were used in all cases and have been approved by Institutional Animal Care and Use Committees (IACUC) of Indiana University of Pennsylvania, Indiana, PA (protocol #10-1617 and 10-1920) and Purdue University, West Lafayette, IN (protocol #1111000236 and 1201000596). All trapping and tissue sample collection in Indiana, New Jersey and Pennsylvania was completed by the Pennsylvania Game Commission, the New Jersey Division of Fish and Wildlife’s Endangered and Nongame Species Program and Indiana Department of Natural Resources (agencies mandated by state law to review and perform all monitoring and permitting). All trapping and tissue sample collection in Ohio was completed by WildWork with permission from the Ohio Division of Wildlife (scientific collecting permit #21–181). This work is subject to approval by the National Environmental Policy Act, which is administered by the United States Fish and Wildlife Service. This study is reported in accordance with ARRIVE guidelines.

Consent for publication

Not applicable.

Competing interests

The authors declare they have no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Additional file 1:

Microsatellite genotypes of woodrats sampled in 2009, 2011 and 2012 in the Palisades, NJ. The spreadsheets labeled “Raw Data” and “data GenAlEx” include the genotypes for twenty-eight individuals genotyped at 11 microsatellite loci. These individuals had relatively low genetic variability, as indicated by number of alleles (spreadsheets AFP and AGL), observed heterozygosity and expected heterozygosity (spreadsheet HFP).

Additional file 2:

Summary statistics associated with nuclear genome sequencing, assembly and annotation.

Additional file 3:

SNP genotypes of woodrats sampled in Indiana, New Jersey, Ohio and Pennsylvania. The spreadsheet labeled “Raw Data” includes the genotypes for 318 individuals genotyped at 134 SNP loci.

Additional file 4:

Relationship between probability of identity (PID), probability of identity between siblings (PIDsib) and the number of genotyped SNP or microsatellite loci.

Additional file 5:

Comparison of microsatellite- and SNP-based genetic variation.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Muller-Girard, M., Fowles, G., Duchamp, J. et al. A novel SNP assay reveals increased genetic variability and abundance following translocations to a remnant Allegheny woodrat population. BMC Ecol Evo 22, 137 (2022). https://doi.org/10.1186/s12862-022-02083-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12862-022-02083-w

Keywords