Open Access
How to translate text using browser tools
16 September 2022 Pronounced Genetic Separation among Varieties of the Primula cusickiana Species Complex, a Great Basin Endemic
Austin Koontz, William D. Pearse, Paul G. Wolf
Author Affiliations +
Abstract

Distinguishing between populations with strong genetic structure and unique species is a common challenge in systematics, especially for taxa occurring in fragmented habitats where allopatric speciation may be widespread and distinct groups may be morphologically similar. Such is often the case with species complexes across sky island environments. In these scenarios, biogeography may help to explain the taxonomic relations between species complex members, and restriction site-associated DNA (RAD) sequencing methods are commonly used to compare closely related taxa across thousands of loci. Here we use RADseq to clarify the boundaries separating the geographically distinct but morphologically similar varieties of the Primula cusickiana species complex, and to contextualize past findings of strong genetic structure among populations within varieties. Our genetic analyses demonstrate pronounced separation between isolated populations of this Great Basin endemic, indicating that the current varietal classification of complex members is inaccurate, and emphasizing their conservation importance. We discuss how these results correspond to recent biogeographical models used to describe the distribution of other sky island taxa in western North America. Our findings also fit into a wider trend observed for alpine Primula species complexes, and we consider how edaphic specialization and heterostylous breeding systems may be contributing to frequent diversification via allopatric speciation in this genus.

A canonical driver of biological diversification is allopatry, whereby geographic barriers lead to population isolation and, eventually, speciation. Sky islands are places where sharp changes in elevation lead to pronounced ecological differences over relatively short distances, providing the types of barriers required for allopatric speciation. Historically, climatic fluctuations have determined the presence and distribution of sky island environments for mountain ranges across the world, and this in turn is reflected by the genetic patterns seen in montane species today (Hewitt 1996). However, in this biogeographic context, distinguishing between closely related species and genetically structured populations may prove challenging (Huang 2020), especially if similar niches across mountain ranges maintain phenotypic similarities (e.g. Yang et al. 2019). Additionally, in the short-term, genetic patterns will be influenced by particular aspects of a species' biology, such as dispersal and breeding systems, which may facilitate or hinder reproductive isolation between genetically distinct entities. Here, we examine the genetic relations between the sky island populations of members of the Primula cusickiana species complex, a group of plants endemic to the Great Basin region of the western United States.

The Primula cusickiana species complex (Holmgren and Kelso 2001) is a group of herbaceous, perennial plants that fall within Primula section Parryi (Wright Smith and Fletcher 1949). When including the phylogenetically nested Dionysia Fenzl., Dodecatheon L., and Cortusa L. (Trift et al. 2002; Schmidt-Lebuhn et al. 2012), the genus is composed of roughly 500 species, distributed worldwide but largely found in temperate and montane habitats in the northern hemisphere, with the Himalayas as the center of diversity (Richards 2003). Primula section Parryi (five species, eight taxa; Kelso et al. 2009) is one of 37 sections of the genus (Richards 2003) and one of three parts of Primula subgenus Auriculastrum Schott (Schott 1852), which also includes Primula sections Auricula Duby and Auneifolia Balfour (Zhang and Kadereit 2004), as well as Dodecatheon L. (Mast and Reveal 2007) and Primula suffrutescens (Gray 1888). Primula subgenus Auriculastrum makes up many of the Primula taxa endemic to western North America, with Primula section Parryi consisting of those found in the Intermountain West, the region between the Sierra Nevada range to the west and the Rocky Mountains to the east. The members of Primula section Parryi are largely found in alpine to subalpine habitats, but while species Primula parryi and P. rusbyi Green are widely distributed within their respective mountain ranges, P. angustifolia Torrey, P. capillaris (Holmgren and Holmgren 1974), and members of the P. cusickiana species complex (P. cusickiana varieties maguirei, domensis, nevadensis, and cusickiana (see Fig. 1)) are more restricted to patches of anomalously cool habitat in otherwise xeric environments. For the P. cusickiana species complex, this is particularly true of P. cusickiana var. maguirei (Williams 1936), found growing on cliffs and small ledges of dolomitic limestone substrates in a 20 km stretch of Logan Canyon within the Bear River Range in northern Utah. Primula cusickiana varieties domensis (Kass and Welsh 1985) and nevadensis (Holmgren 1967) also occur in very small pockets on limestone substrates, but at higher elevations past treeline, with P. cusickiana var. domensis being found in the House Range (west central Utah) and P. cusickiana var. nevadensis having documented populations in the Snake Range (just west of the House Range) and the Grant Range of Nevada. Finally, P. cusickiana var. cusickiana (Gray 1888) is more widespread, found in cool, wet microsites at relatively low elevations along the Snake River Plain in Idaho, but with documented populations as far west as eastern Oregon (on the Owyhee High Desert Plain) and as far south as the Cougar Mountains (near Jarbidge) of northeastern Nevada (Mansfield 2010).

The geographic separation and ecological differences of Primula cusickiana varieties maguirei, nevadensis, and cusickiana contributed to their original classification as unique species (i.e. P. maguirei, P. nevadensis, and P. cusickiana) upon initial description. The discovery and publication of P. domensis in 1985, along with the continued collection of the other varieties, began to cast doubt on the species distinction for each complex member. Morphologically, the differences among the four varieties are subtle: P. cusickiana vars. maguirei and cusickiana are entirely glabrous, and distinguished from one another by relative calyx length, while in P. cusickiana vars. nevadensis and domensis, plants are pubescent, with P. cusickiana var. nevadensis having a shorter corolla tube length than P. cusickiana var. domensis (Holmgren et al. 2005). A review in 2001 considered these morphological distinctions “meager at best,” and concluded that “recognition at the varietal level would be most appropriate” (Holmgren and Kelso 2001).

Fig. 1.

Four members of the Primula cusickiana species complex. A. P. c. var. maguirei, in Right Hand Fork of Logan Canyon. B. P. c. var. cusickiana, near Cougar Point in Jarbidge, Nevada. C. P. c. var. domensis, at Notch Peak in the House Range, Utah. D. P. c. var. nevadensis, on Mount Washington in the Snake Range (Great Basin National Park), in Nevada.

img-z2-1_887.jpg

At the time of this shift, no genetic data was available to justify classification at the variety level. However, a 1997 analysis of P. cusickiana var. maguirei used allozyme marker genes to uncover a significant degree of genetic structure between the relatively proximate (∼10 km) populations (Wolf and Sinclair 1997) within this one taxon. A later analysis of the same populations using amplified fragment length polymorphism (AFLP) loci confirmed this finding, and found similar levels of polymorphism between the upper and lower canyon groups, suggesting this genetic structure is not the result of a past bottleneck event (Bjerregaard and Wolf 2008). A further analysis of AFLP and chloroplast DNA from members of Primula section Parryi showed P. cusickiana var. maguirei and the other P. cusickiana complex members as being monophyletic, but relationships within the complex were incongruent, with only weak support of a clade containing P. cusickiana vars. nevadensis and domensis being sister to a clade made up of P. cusickiana vars. maguirei and cusickiana (Kelso et al. 2009). To better resolve the relationships between varieties, the authors suggested an analysis utilizing more populations from across the range of this species complex. Restriction site-associated DNA sequencing (RADseq) methods available today, with their ability to generate reads over many sequence regions of closely related individuals, are well-suited to provide the data required for such an analysis.

In addition to clarifying the genetic relations between geographically distinct varieties, a more detailed analysis of the P. cusickiana species complex can meaningfully contribute to ongoing conservation efforts. Primula cusickiana var. maguirei was listed as Threatened in 1985, due to its unique habitat in Logan Canyon and threats of habitat loss due to development (US Fish and Wildlife Service 1985). Given the strong genetic structure between P. cusickiana var. maguirei's populations, either population may be more closely related to populations of a different complex variety than the neighboring Logan Canyon population, a finding which would have significant implications for the protection of this variety. More broadly, an understanding of the genetic relations in this species complex will determine whether the varietal classification properly reflects the extent of divergence of each complex member, and thus the extent of unique evolutionary history. This understanding can direct management of the narrow-range endemics included in this species complex, such as P. cusickiana var. maguirei, but also P. cusickiana vars. nevadensis and domensis, and also inform the identification of potential evolutionarily significant units (Coates et al. 2018).

We sought to clarify the relatedness of the P. cusickiana complex members by using a RADseq approach to genotype all four varieties located at distinct populations scattered throughout the Great Basin. In addition to contextualizing the genetic structure between the upper and lower Logan Canyon P. cusickiana var. maguirei populations, this analysis provides insights into the biogeographic history of this species complex, and could have important conservation implications for this rare endemic plant.

Materials and Methods

Sampling—Samples from all four P. cusickiana species complex members were gathered in the field. Additionally, we collected samples from P. parryi, a species outside of the P. cusickiana species complex, in order to compare genetic differences among complex members to those extending outside of the complex. Finally, because past research has shown variable relations between P. capillaris and the P. cusickiana species complex (Kelso et al. 2009), we also tried to collect P. capillaris in the field. However, we were unable to locate any wild P. capillaris individuals in the Ruby Mountains: at one location suggested by past herbaria data, a population of P. parryi was found instead. To compensate, two P. capillaris samples were sourced from herbaria (see Appendix 1). By sampling all four species complex varieties, P. capillaris, and P. parryi, we captured 75% of the taxa in the Primula section Parryi. The number of collected samples for each population of each variety, as well as P. parryi, along with population coordinates, are provided in Table 1. Sampling sites were determined by examining all populations documented through specimens in herbaria across the United States, and selecting the populations of each variety that would maximize the geographic extent of our genetic survey. For P. cusickiana vars. maguirei, domensis, and nevadensis, this meant collecting from previously documented locations within each variety's mountain ranges. For P. cusickiana var. cusickiana, we collected from populations documented at the most westerly (Craters of the Moon National Monument), northerly (Bear, Idaho), southerly (Jarbidge, Nevada), and easterly (Owyhee High Desert, Oregon) extent of the variety's known range.

Table 1.

Number of samples and locations per taxon.

img-z3-13_887.gif

At each population location, an individual plant was removed as completely as possible as a voucher specimen. For DNA samples, two leaves from each of ten plants were removed and placed in labeled paper envelopes, which were stored on silica crystals to keep samples dry. Vouchers were deposited at the Intermountain Herbarium (UTC); P. cusickiana var. nevadensis voucher specimens collected from Mt. Washington were additionally deposited at the Great Basin National Park herbarium.

Leaf tissue from 89 samples (87 silica-dried field collections representing all sample sites, and two herbarium specimens of P. capillaris) were placed into Qiagen Collection Microtubes (catalog number 19560) and sent to the University of Wisconsin-Madison Biotechnology Center, for DNA extraction, library prep, and DNA sequencing (described below). Seven replicate samples were also included to assess the quality of sequencing results, and were distributed across all four P. cusickiana varieties, as well as P. parryi.

DNA Extraction—DNA was extracted using the Qiagen Dneasy mericon 96 QIAcube HT kit. DNA was then quantified using the Quant-iT PicoGreenR© dsDNA kit (Life Technologies, Grand Island, New York).

Library Preparation and Sequencing—Libraries were prepared following Elshire et al. (2011). ApekI (New England Biolabs, Ipswich, Massachusetts) was used to digest 100 ng of DNA. Following digestion, Illumina adapter barcodes were ligated onto DNA fragments using T4 ligase (New England Biolabs, Ipswich, Massachusetts). Size selection was run on a PippinHT (Sage Science, Inc., Beverly, Massachusetts) to subset samples down to 300–450 bp fragments, after which samples were purified using a SPRI bead cleanup. To generate quantities required for sequencing, adapter-ligated samples were pooled and then amplified, and a post-amplification SPRI bead cleanup step was run to remove adapter dimers. Final library qualities were assessed using the Agilent 2100 Bioanalyzer and High Sensitivity Chip (Agilent Technologies, Inc., Santa Clara, California), and concentrations were determined using the Qubit© dsDNA HS Assay Kit (Life Technologies, Grand Island, New York). Libraries were sequenced on an Illumina NovaSeq 6000, generating paired-end data for fragments of approximately 150 bp in length.

Data Processing—Raw FASTQ data files were demultiplexed and processed using steps 1–7 of the ipyrad software, v. 0.9.31 (Eaton and Overcast 2020). Single nucleotide polymorphisms (SNPs) recognized by ipyrad were used as the basis for variation between individuals for downstream analyses, and assemblies were generated de novo. All ipyrad and STRUCTURE parameter files, as well as R scripts used for analysis and data visualization, can be found in the Supplementary Materials available on Dryad (Koontz et al. 2022), as well as on GitHub (github.com/akoontz11/Primula/). Raw, demultiplexed sequencing data can also be accessed on the NCBI Sequence Read Archive (SRA; accession number PRJNA705310).

COMPLEX-WIDE GENETIC SURVEY—For our complex-wide genetic survey, we ran ipyrad twice: we used the results from our initial run to confirm sequencing consistency for replicate samples, and to identify samples with low coverage. For both runs, demultiplexed sequences were paired and merged, and low quality bases, adapters, and primers were filtered prior to SNP calling. Default values were used for the ipyrad parameters in these steps, as well as for the clustering threshold (clust_threshold; 0.85) and minimum sequencing depth (mindepth_statistical; 6) parameters.

For our initial run, we specified a minimum number of samples per locus (min_samples_locus) parameter of 10, in order to obtain loci shared between two to three sample locations for any taxa. Using the results from this run, we used a Python script (vcf2Jaccard.py) to compare samples with replicates by calculating the mean Jaccard similarity coefficients between all samples. We found that all replicates matched highly with their corresponding samples (Fig. S1, Koontz et al. 2022).

After merging replicates and removing low coverage (generally, less than 30 loci in the final assembly) samples from the dataset, 82 of our 87 original samples remained for our complex-wide analysis. We reran ipyrad (steps 1–7) using these 82 samples to select for loci specific to this subset. We used a min_samples_locus parameter of 32 for this second run, to cluster using loci shared across the species complex; ipyrad default values were used otherwise. Because very low numbers of loci were retrieved for both herbarium specimens of P. capillaris (possibly due to the age of these specimens), we were unable to include P. capillaris in downstream clustering analyses.

Variety Specific Clustering—In addition to our complex-wide survey, we were interested in exploring population structure within P. cusickiana var. maguirei which could not be resolved using loci shared across all species complex members. To do so, we ran ipyrad on just the 18 P. cusickiana var. maguirei samples used in our complex-wide survey. Because five samples from each of the upper Logan canyon sampling sites were included in our ipyrad assembly, we specified a min_samples_locus parameter of 5; ipyrad default parameter values were used otherwise.

Population Analyses—STRUCTURE—To visualize relations between complex members across their geographic range, and to determine the number of identifiable genetic clusters within the complex, we used the program STRUCTURE v. 2.3 (Pritchard et al. 2000). STRUCTURE uses Bayesian clustering analysis to probabilistically assign individuals to one or more of K source populations, where the loci within each population are assumed to be in Hardy-Weinberg proportions and linkage equilibrium. While the presence of an outgroup is not required for the STRUCTURE software, our preliminary findings suggested that including P. parryi did not significantly alter our clustering results, and we chose to include this species for this and our other population level analyses (see below). For all STRUCTURE runs, we used a burn-in length of 50,000, and 100,000 MCMC reps after burn-in. For our complex-wide survey, we ran STRUCTURE for K values of 2–16, with 50 replicates per K value. For our P. cusickiana var. maguirei-only analyses, we ran STRUCTURE for K values of 2–6, with 50 replicates per K value. We used the CLUMPAK server (Kopelman et al. 2015) to summarize results across replicates for each K value, and to build STRUCTURE plots.

For all of our STRUCTURE analyses, we used two different methods to determine the “optimal” K value: the method described in the STRUCTURE manual (Pritchard et al. 2000), which identifies the K value with the greatest likelihood (the natural logarithm of the posterior probability of the genetic data at a specific K value); and the method described in Evanno et al. (2005), which measures the second order rate of change in the posterior probability of the genetic data between successive K values (ΔK). While the ΔK method described by Evanno was designed to better accommodate scenarios in which unequal migration is occurring between putative source populations, it is worth noting that both techniques are ad hoc and require interpretation (as described by their authors). Furthermore, there is an inherent difficulty in inferring an unambiguous number of genetic clusters from any given set of populations (Novembre 2016). Therefore, following recommendations in the original STRUCTURE manual (Pritchard et al. 2000; Novembre 2016), we generated STRUCTURE outputs within a range of K values, and examined how relationships changed or remained constant across values. Because robust species boundaries consider genetic data in concert with other types of data (Carstens et al. 2013), we took the geographic distribution of each population into consideration when determining a biologically reasonable value of source populations. We were particularly interested in visualizing the extent of genetic differences between geographically distinct populations within varieties across K values, and how different values of K illustrated the extent of admixture (as indicated by individuals having portions of their genomes assigned to different sources) between adjacent populations. Finally, given evidence for strong genetic structure between the populations of P. cusickiana var. maguirei (Wolf and Sinclair 1997; Bjerregaard and Wolf 2008), we were interested to see if either P. cusickiana var. maguirei population in Logan Canyon was more closely related to another species complex population than it was to the neighboring Logan Canyon population.

SpliTstree—In addition to our STRUCTURE analysis, we used the NeighborNet split network algorithm (Bryant and Moulton 2002) to visualize the taxa sampled in this survey using a phylogenetic network. NeighborNet works by iteratively grouping pairs of taxa together based on similarities, in the same way that a neighbor joining tree is built. However, rather than a tree, the end result is a split network, in which splits in the taxa are represented by parallel lines, and conflicting signals in relations of taxa are represented by boxes. The ability to represent different phylogenetic hypotheses simultaneously allows this method to accommodate processes which undermine traditional phylogenetic analyses, such as scenarios involving recombination and hybridization. This makes the NeighborNet algorithm useful for summarizing RADseq data, where patterns of loci across individuals may imply different genealogies. For our complex-wide survey, we passed the output from our ipyrad de novo assembly to the software SplitsTree v. 4.17.1 (Huson and Bryant 2006), which we used to construct the NeighborNet split network.

Discriminant Analysis of Principal Components—In addition to STRUCTURE and the NeighborNet algorithm, we analyzed the results of our complex-wide survey using Discriminant Analysis of Principal Components (DAPC; Jombart et al. 2010) in the package adegenet in R v. 3.6.3 (R Core Team 2020). DAPC is a statistical technique designed to accommodate the size of genomic data sets and capable of differentiating within-group variation from between-group variation. SNP data is first transformed using a principal components analysis (PCA), and then k-means clustering is run to generate models and likelihoods corresponding to each number of population clusters. The best-fitting model, and so the best-supported number of populations, is assessed using the models' Bayesian Information Criterion (BIC) scores. We chose to utilize DAPC to visualize population clusters in a PCA format, and to determine whether the supported number of clusters was congruent with our STRUCTURE and SplitsTree results, indicating a more robust determination of the number of unique taxa contained within the complex (Carstens et al. 2013).

Fst Estimates—Because we wanted to measure the extent of genetic variance within the groups analyzed, we used the VCFtools software (Danecek et al. 2011) to generate weighted FST estimates (Weir and Cockerham 1984). We generated an FST estimate for our complex-wide analysis (across all populations of all P. cusickiana varieties, excluding P. parryi) as well as for the samples included in our P. cusickiana var. maguirei-only analysis.

Results

Complex-Wide Genetic Survey—We retrieved, on average, 2.04 × 106 reads per sample, and our complex-wide ipyrad run identified 1277 loci that were used in our subsequent STRUCTURE analysis. Using the Evanno et al. (2005) method yielded an optimal K value of K = 5; using the method described in the STRUCTURE manual (Pritchard et al. 2000) identified the K value with the greatest likelihood as K = 14 (Fig. 2). Given the limited differentiation between the makeup of individuals for K values beyond K = 7 (demonstrated by the addition of minor genetic proportions in individuals composed largely of a single genetic group), we felt K = 14 was too large a number of source populations to describe these taxa, and that distinctions at this level of clustering were observable at much lower values of K. While the value recommended by the Evanno method provides a more conservative hypothesis, in that it recommends a lower number of putative source populations, this level of clustering fails to illustrate the genetic distinction between the geographically separated Nevada (Jarbidge) and Oregon (Owyhee) populations of P. cusickiana var. cusickiana, which is illustrated at values greater than K = 5. Similarly, we felt that the clustering at K = 6 failed to properly distinguish the distinct populations of P. cusickiana var. nevadensis in Great Basin National Park (GRBA) and the Grant Range (Troy), a distinction which is quite notable at higher values of K. Therefore, following the recommendations provided in Pritchard et al. (2000) by examining how relationships between our sampled populations change across K values ranging from K = 2 to K = 16 (Figs. S2–S4, Koontz et al. 2022), we determined K = 7 to be the most biologically relevant value to describe this group of plants. This level of source populations best captures the geographic isolation of all sampled populations, with P. cusickiana vars. domensis and maguirei being clearly delineated, P. cusickiana var. nevadensis showing distinctions between its two populations, and P. cusickiana var. cusickiana split into three groups composed of populations from the Snake River Plain in Idaho (SRP), Nevada (Jarbidge), and Oregon (Owyhee). Since higher K values emphasize the divisions seen at this level, and further subdivide isolated populations of P. cusickiana vars. cusickiana and nevadensis, K = 7 is an estimate which reflects the strong divisions within this complex while allowing for more nuanced groupings of unique populations to be made in light of future evidence.

Fig. 2.

Sample STRUCTURE plots at K = 5 (the optimal K value determined by the Evanno ΔK method), K = 7 (the recommended value of K based on all STRUCTURE outputs), and K = 14 (the value of K with the greatest likelihood). Bars and names on the bottom indicate populations grouped based on K = 7 clustering. While groupings remain largely coherent across K values, nevadensis individuals at Troy peak appear more admixed and less unique at K = 5, while separation between cusickiana populations along the Snake River plain is more observable at K = 14.

img-z5-1_887.jpg

Our NeighborNet split network analysis supports this clustering value, and divisions seen in our STRUCTURE analysis are similarly reflected in this split network (Fig. 3). Primula cusickiana var. maguirei is separated from all other varieties, while P. cusickiana var. domensis and the Great Basin National Park population (GRBA) of P. cusickiana var. nevadensis group together, and separately from the Grant Range population of P. cusickiana var. nevadensis (Troy). Finally, P. cusickiana var. cusickiana is split into the same three groups shown in our STRUCTURE analysis: populations from the Snake River Plain in Idaho (SRP), Nevada (Jarbidge), and Oregon (Owyhee), with the last population splitting from a branch shared with P. parryi. Our DAPC analysis, however, revealed that the greatest supported number of clusters (i.e. the value with the lowest BIC score) was eleven (data not shown), a value incongruent with our SplitsTree and STRUCTURE results, suggesting that boundaries within this complex are elaborate. At this level of genetic clusters, several groups were quite small (consisting of only one or two samples), and groupings were incoherent with the spatial distribution of populations. To provide a clearer comparison to our STRUCTURE results, and to examine relations strictly within the species complex, we removed P. parryi outgroup samples from our dataset and ran our DAPC with a specification of six clusters (Fig. S5, Koontz et al. 2022). At this level of clustering, the population of P. cusickiana var. nevadensis in the Snake Range of Great Basin National Park (GRBA) is shown as a unique cluster, while the P. cusickiana var. nevadensis population further south in the Grant Range groups with the P. cusickiana var. cusickiana population sampled from Oregon (Owyhee). Primula cusickiana var. domensis is a unique cluster which groups closely to both of these. Thus, while our population level analyses do not point to an unambiguous number of “true” genetic clusters, some patterns are shared across techniques.

The extreme level of divergence between the sky island populations in this species complex is reflected not only in our population level analyses, but also in our relatively large FST estimate across all complex populations, which was 0.72. Figure 4 illustrates proportions of sample membership to clusters based on our STRUCTURE analysis at K = 7 for all populations in their geographic context across the Great Basin.

Variety Specific Clustering—In our complex-wide analysis, all P. cusickiana var. maguirei samples grouped as a single cluster, distinct from all other populations of all other varieties, indicating that neither Logan Canyon population is more closely related to any populations of another variety. Even at values of K = 16, the upper and lower Logan Canyon populations of P. cusickiana var. maguirei were not resolved from one another.

However, reducing our sample set to only maguirei samples allowed us to retain loci informative to this variety but unshared with other complex member populations. Our P. cusickiana var. maguirei-only ipyrad run generated an assembly with 68,492 loci, indicating a large number of loci specific to P. cusickiana var. maguirei and not shared with the wider species complex. To speed up processing times, we ran STRUCTURE on a 17,988 loci subset of maguirei-specific markers. Using the CLUMPAK server, we found optimal K values of K = 4 (using the Evanno method) and K = 3 (using the likelihood method described in the STRUCTURE manual). Figure 5 shows the STRUCTURE plot at K = 3, which resolves similar groupings of maguirei populations supported in (Bjerregaard and Wolf 2008), and the distinctions between upper and lower canyon populations. We also estimated an FST value of 0.33 among these three populations, which is comparable to previous estimates in Bjerregaard and Wolf (2008).

Fig. 3.

Split network analysis generated using the NeighborNet algorithm. Tips indicate individuals, and groups are labeled by variety and populations, with coloration matching the groupings used in the K = 7 clustering of Fig. 2. Splits between taxa are represented by parallel edges (lines), and the lengths of edges are proportional to the weight of the associated split.

img-z6-1_887.jpg

Discussion

Our analysis of RADseq data from Primula cusickiana complex members demonstrates that the disjunct geographical distribution of populations across the Great Basin is reflected by pronounced genetic divergences. While the results of our clustering analyses largely coincide with the divisions of current varietal classifications, there are notable exceptions. Distinctions between isolated populations within varieties, as well as similarities between neighboring populations of different varieties, can be observed in our STRUCTURE plots of low K values (i.e. ranging from 2–6; see Figs. S2–S4) and in our split network analysis. For instance, we found Great Basin National Park (GRBA) P. cusickiana var. nevadensis populations to be admixed, with genetic contributions coming from P. cusickiana var. domensis to the east and (to a lesser extent) Grant Range P. cusickiana var. nevadensis populations to the south. This is in accordance with past analyses of AFLP and chloroplast DNA from the Primula section Parryi, which found these two varieties to be extremely close (Kelso et al. 2009).

Our results also suggest a more nuanced understanding of P. cusickiana var. cusickiana. Populations of this variety are split into distinct clusters in our analysis, with Jarbidge (Nevada) and Owyhee (Oregon) populations appearing unique from each other and the remaining Snake River Plain (SRP) populations in Idaho. That these distinctions are seen in our STRUCTURE and SplitsTree analyses imply the robustness of this result. Given the relatively wide distribution of this variety (growing in moist soils at lower elevations than other complex members), our findings of genetic divergence between its populations are noteworthy, and support past evidence of phenotypic differences in different portions of its range. For instance, past morphological research of Idaho P. cusickiana var. cusickiana populations has suggested dividing this taxa into three unique species (Mansfield 1993), with Owyhee populations being classified as P. wilcoxiana. Interestingly, our SplitsTree analysis suggests this population has a close and possible reticulated relationship with P. parryi, perhaps belonging outside of the P. cusickiana species complex entirely. While all interpretations of this group implicate its distinction from the rest of P. cusickiana var. cusickiana, a more detailed phylogenetic and morphological description of the Owyhee populations of P. cusickiana var. cusickiana would be required before species recognition is warranted.

Fig. 4.

Map of sample locations with cluster membership. Sampling locations are represented by pie charts indicating percentage of population membership to clusters determined at K = 7 STRUCTURE clustering threshold. With exception to nevadensis, most samples fall almost entirely within a specified cluster.

img-z7-1_887.jpg

Fig. 5.

STRUCTURE plot for maguirei samples at a clustering threshold of K = 3, with a map depicting the lower and upper canyon collection sites. While maguirei clustered together in the complex-wide analysis, maguirei-only analysis was able to resolve divisions suggested in past studies.

img-z8-1_887.jpg

The separation between populations within P. cusickiana var. cusickiana, as well as our support of past findings of significant genetic distances between the proximate populations of P. cusickiana var. maguirei, underscore our discovery of profound genetic divergences between all members of this species complex, despite their distribution over a relatively small geographic area. This trend is reflected not only in our clustering analyses, but also in our weighted FST estimate of 0.72 across complex populations, a high value compared to similar estimates for other plant taxa (for instance, the mean FST for plant taxa in a meta-analysis by Leinonen et al. (2008) was estimated to be 0.24). Our sampling sites were determined by maximizing sampling across the known geographic distribution of all populations of P. cusickiana var. cusickiana, in addition to collecting from all documented populations of the more geographically restricted varieties (P. cusickiana vars. maguirei, nevadensis, and domensis). Therefore, short of the discovery of new populations (e.g. in different sky island habitats in the Great Basin, or outside P. cusickiana var. cusickiana's known locations), we believe this is an exhaustive geographic survey of this complex. Although these results support the historical designation of species for these complex members, rather than variety, we restrain from asserting that designation here, based on our reported population genetic comparisons alone. In accordance with using several types of nongenetic data, in addition to more standard genetic tests, to support species definitions (Carstens et al. 2013), we believe more thorough morphological and phylogenetic analyses (ideally including both nuclear and cpDNA), as well as breeding surveys, are necessary before unambiguous species designations can be made. Strong genetic structure between groups, while typically indicative of separate species, is not sufficient on its own for establishing species identification (Sukumaran and Knowles 2017), especially given the potential for genetically separate groups to generate fertile offspring, thereby violating the central criterion of the biological species concept (Huang 2020). However, regardless of formal classification, these findings of strong genetic separation are worthy of interest simply due to their environmental and biogeographical context. Below, we consider how three phenomena (biogeographical trends in the Great Basin, edaphic endemism, and reproductive traits specific to Primula) may contribute to the significant genetic divergence observed across this species complex.

Great Basin Sky Island Biogeography—Members of the P. cusickiana complex are found at relatively high (2000–3300 m) elevations throughout the Great Basin. Many of these are sky island locations associated with strong ecological shifts as habitat transitions from lower sagebrush steppe to cooler, more forested regions dominated by pinyon and juniper. Now separated by arid basins due to climatic warming in the Holocene, these sky islands are understood to be the fragmented remnants of a continuous region of cool, moist habitat which once extended across the Great Basin (Thompson and Mead 1982). This has led to their characterization as refugia for various taxa, particularly mammals (Brown 1971; Badgley et al. 2014), but also butterflies (Boggs and Murphy 1997) and plants (Harper et al. 1978; Nowak et al. 1994; Charlet 2007).

However, research regarding the populations unique to these sky island habitats has noted that many species distribution patterns among Great Basin mountaintops do not follow a strictly island biogeographical model (Lawlor 1998; Fleishman et al. 2001), in that neither island surface area nor proximity to “mainland” source populations (typically identified as the western Sierra Nevada or eastern Rocky Mountains) is predictive of species abundance (Fleishman et al. 2001). And in some taxa, there is evidence for regular, modern dispersal between Great Basin ranges (Floyd et al. 2005). An alternative scenario is that this complex has followed what has been described as an “expanding-contracting archipelago” (ECA) model, in response to Quaternary glacial cycles (DeChaine and Martin 2005a). The ECA model has been used to describe the divergence between Rocky Mountain sky island plant taxa (Dechaine and Martin 2005b; Hodel et al. 2021), and provides a framework for explaining the genetic structure observed between isolated montane populations on a broad spatial scale. In this model, populations are assumed to become fragmented as they contract up-slope during warmer interglacials; during glacial periods, populations expand down-slope as moist, cool habitat becomes widespread, leading to hybrid zones and possible admixture. Given the degree of fragmentation between P. cusickiana's populations in today's climate (which resembles past interglacial periods), and the admixture between the relatively proximate populations of P. cusickiana vars. domensis and nevadensis revealed in our analysis, this model offers a viable explanation for the trends observed in this species complex. In addition to determining areas of refugia and secondary contact between these varieties, Quaternary glacial cycles may have influenced the colonization of the Intermountain West region more broadly. In the late Tertiary, Primula section Parryi is considered to have derived from an ancestral lineage within the Auriculastrum clade in east Asia (Zhang and Kadereit 2004; Zhang et al. 2004), with Primula parryi being the first to diverge from a common North American ancestor (Kelso et al. 2009). Given the prevalence and diversification of Primula in the Arctic (Kelso 1992), climatic fluctuations have likely played a role in the north-south dispersal of this genus across the North American continent.

Edaphic Endemism—In conjunction with climatic niche preferences, complex P. cusickiana vars. maguirei, domensis, and nevadensis are found on the cliffs and crevices of exclusively limestone substrates. Edaphic endemism is known to play a role in the diversification of plant species, both globally (Rajakaruna 2018; Hulshof and Spasojevic 2020) and within the Great Basin region specifically (De Queiroz et al. 2012; Brown and Mansfield 2017), and we might consider whether the trends uncovered in our genetic survey are a result of edaphic preferences of these populations. Within Primula section Parryi, the only species that is as narrowly restricted to calcareous substrates is P. capillaris, which is endemic to the Ruby Mountains of Nevada and has been shown to be closely related to P. cusickiana complex members (Kelso et al. 2009). P. rusbyi, which has a range from the southern Rocky Mountains to northern Mexico, is also found in limestone soils, but has been associated with granitic habitats as well. Given the wider edaphic niches of members of Primula section Parryi outside of the P. cusickiana species complex, it seems more likely that the edaphic preferences of P. cusickiana vars. maguirei, domensis, and nevadensis are the result of a tendency towards moisture retaining substrates, rather than specific mineral or pH constraints (Kelso et al. 2009). If so, we would expect the available niche space of the populations within this species complex to be contingent on the historical climatic fluctuations of the Great Basin, as described above, and not the edaphic heterogeneity of the region alone. Regardless of the ultimate cause, allopatry across distinctive climatic and edaphic niches seems to contribute to the genetic divergences in P. cusickiana's populations, a trend observed in other sections of Primulaceae, as well (Boucher et al. 2016).

Speciation and Heterostyly in Primula—Recent research has shown several different alpine Primula species complexes to contain previously undescribed cryptic species, in China (Huang et al. 2019; Ren et al. 2020) and in Europe (Schorr et al. 2013; Theodoridis et al. 2019). Our findings on the P. cusickiana species complex resonate with these trends, and raise the question of what unique traits Primula possesses which might cause such frequent diversification via allopatric speciation. In addition to edaphic preferences in this genus, it has been argued that heterostyly, a widespread breeding system in angiosperms to promote outcrossing, may be a driving force leading to speciation in Primula (He et al. 2021). In heterostyly, “pin” and “thrum” floral morphologies prevent self-fertilization via insect pollination (Darwin 1877), and are associated with a sporophytic-incompatibility system which follows a Mendelian pattern of inheritance (Li et al. 2016). Changes in the prevalence of heterostyly across populations could lead to the divergence between distylous and homostylous populations, and ultimately speciation. Within populations, the possible deleterious effects of reduced effective population size (due to exclusive mating with individuals of the opposite morphology) seem to be counterbalanced by the genetic advantages of outcrossing, leading to greater net diversification over evolutionary time (De Vos et al. 2014). While the extent of distyly within populations has been well documented in P. cusickiana var. maguirei (Davidson et al. 2014), observation of distylous morph ratios in other species complex populations would be required to determine if these dynamics are driving the divergences seen at the species complex level.

Conclusion—The results of our genetic survey of Primula cusickiana fit into a wider trend demonstrating abundant allopatric speciation despite little niche divergence in other alpine Primula species complexes. Our findings support the historical classification of each of these complex members as unique species, rather than the varietal classification taken in (Holmgren and Kelso 2001). Furthermore, these results warrant a more detailed understanding of the isolated and genetically unique populations in this complex (such as P. cusickiana var. cusickiana populations in Nevada and Oregon), and of the admixture observed in the populations of P. cusickiana var. nevadensis. Similarly, updated morphological comparisons between varieties, breeding experiments across the groups described in this study, and observations into the levels of heterostyly in disjunct populations, would offer a clearer understanding of the mechanisms of speciation occurring within this complex. Further genetic comparisons including P. capillaris would be able to clarify the currently ambiguous relationship between this species and the P. cusickiana species complex, and may bolster or conflict with the trends of allopatry seen in our results. Finally, the endemic species with narrow niches included in this study, such as P. cusickiana var. maguirei, but also P. cusickiana vars. nevadensis and domensis, and P. capillaris, warrant concern of extinction, and more work needs to be done to better understand the threats faced by each of these taxa in order to ensure their survival in an increasingly arid Great Basin.

Acknowledgments

This research was supported by funds provided by the Margaret Williams Research Grant offered by the Native Nevada Plant Society, the Lawrence Piette Graduate Scholarship offered by the Utah State University (USU) College of Science, the Dr. Ivan J. Palmblad Graduate Research Award offered by the USU Department of Biology, and the Graduate Research Award offered by the USU Ecology Center. WDP and the Pearse Lab are funded by NSF EF-1802605, NSF ABI-1759965, and UKRI-NERC NE/V009710/1.

The authors would like to thank Dr. Carol Rowe for her help with the genetic analyses, her Python script used to calculate Jaccard similarities from SNP data, and her Python scripts for creating the complex-wide and maguirei map. Thanks also to Dr. Barbara Ertter and especially Dr. Don Mansfield at the College of Idaho, for their assistance in collecting P. cusickiana var. cusickiana. Trish Winn and Jennifer Lewihnsohn in the United States Forest Service provided the collection permit for variety maguirei from the Uinta-Wasatch-Cache National Forest; Todd Stefanic and Gretchen Baker with the National Park Service coordinated collection from Craters of the Moon National Monument and Great Basin National Park, respectively. Michael Piep, Elizabeth Makings, and Jerry Tiehm allowed for Primula specimens to be sampled from the Intermountain Herbarium, Arizona State Vascular Plant Herbarium, and University of Nevada, Reno Herbarium, respectively. Thanks also to Kris Valles at the Intermountain Herbarium. Dr. Leila Shultz assisted with identification of the Owyhee P. cusickiana var. cusickiana population. The authors would also like to thank Jean Howerton, Buddy Smith, and Noel and Pat Holmgren. Finally, all collections were made on the ancestral lands of the Western Shoshone, Eastern Shoshone, Shoshone-Bannock, Southern Paiute, Goshute, and Nez Perce Native American tribes.

Author Contributions

AK determined sample locations, performed the majority of sample collection, and ran genetic analyses. WDP contributed to study design and assisted with genetic analyses and manuscript writing. PW guided study design and assisted with genetic analyses, sample collection, and manuscript writing.

Literature Cited

1.

Badgley, C., T. M. Smiley, and J. A. Finarelli. 2014. Great Basin mammal diversity in relation to landscape history. Journal of Mammalogy 95: 1090–1106. Google Scholar

2.

Bjerregaard, L. and P. G. Wolf. 2008. Strong genetic differentiation among neighboring populations of a locally endemic primrose. Western North American Naturalist 68: 66–75. Google Scholar

3.

Boggs, C. L. and D. D. Murphy. 1997. Community composition in mountain ecosystems: Climatic determinants of montane butterfly distributions. Global Ecology and Biogeography Letters 6: 39–48. Google Scholar

4.

Boucher, F. C., N. E. Zimmermann, and E. Conti. 2016. Allopatric speciation with little niche divergence is common among alpine Primulaceae. Journal of Biogeography 43: 591–602. Google Scholar

5.

Brown, J. H. 1971. Mammals on mountaintops: Nonequilibrium insular biogeography. American Naturalist 105: 467–478. Google Scholar

6.

Brown, B. J. and D. H. Mansfield. 2017. Edaphic and geographic origins of varietal differentiation in Eriogonum calcareum . Madro ño 64: 22–31. Google Scholar

7.

Bryant, D. and V. Moulton. 2002. NeighborNet: An agglomerative method for the construction of planar phylogenetic networks. Algorithms in Bioinformatics 375–391. Google Scholar

8.

Carstens, B. C., T. A. Pelletier, N. M. Reid, and J. D. Satler. 2013. How to fail at species delimitation. Molecular Ecology 22: 4369–4383. Google Scholar

9.

Charlet, D. A. 2007. Distribution patterns of Great Basin conifers: Implications of extinction and immigration. Aliso: A Journal of Systematic and Floristic Botany 24: 31–61. Google Scholar

10.

Coates, D. J., M. Byrne, and C. Moritz. 2018. Genetic diversity and conservation units: Dealing with the species-population continuum in the age of genomics. Frontiers in Ecology and Evolution 6: 165. Google Scholar

11.

Danecek, P., A. Auton, G. Abecasis, C. A. Albers, E. Banks, M. A. DePristo, R. E Handsaker, G. Lunter, G. T. Marth, S. T. Sherry, G. McVean, R. Durbin, and 1000 Genomes Project Analysis Group. 2011. The variant call format and VCFtools. Bioinformatics 27: 2156–2158. Google Scholar

12.

Darwin, C. 1877. The Different Forms of Flowers on Plants of the Same Species . London: John Murray. Google Scholar

13.

Davidson, J. B., S. L. Durham, and P. G. Wolf. 2014. Breeding system of the threatened endemic Primula cusickiana var. maguirei (Primulaceae). Plant Species Biology 29: E55–E63. Google Scholar

14.

De Queiroz, T. F., C. Baughman, O. Baughman, M. Gara, and N. Williams. 2012. Species distribution modeling for conservation of rare, edaphic endemic plants in White River Valley, Nevada. Natural Areas Journal 32: 149–158. Google Scholar

15.

De Vos, J. M., C. E. Hughes, G. M. Schneeweiss, B. R. Moore, and E. Conti. 2014. Heterostyly accelerates diversification via reduced extinction in primroses. Proceedings. Biological Sciences 281: 20140075. Google Scholar

16.

DeChaine, E. G. and A. P. Martin. 2005a. Historical biogeography of two alpine butterflies in the Rocky Mountains: Broad-scale concordance and local-scale discordance. Journal of Biogeography 32: 1943–1956. Google Scholar

17.

Dechaine, E. G. and A. P. Martin. 2005b. Marked genetic divergence among sky island populations of Sedum lanceolatum (Crassulaceae) in the Rocky Mountains. American Journal of Botany 92: 477–486. Google Scholar

18.

Eaton, D. A. R. and I. Overcast. 2020. ipyrad: interactive assembly and analysis of RADseq datasets. Bioinformatics 36: 2592–2594. Google Scholar

19.

Elshire, R. J., J. C. Glaubitz, Q. Sun, J. A. Pol, K. Kawamoto, E. S. Buckler, and E. Sharon. 2011. A robust, simple genotyping-by-sequencing (GBS) approach for high diversity species. PLoS One 6: e19379. Google Scholar

20.

Evanno, G., S. Regnaut, and J. Goudet. 2005. Detecting the number of clusters of individuals using the software STRUCTURE: A simulation study. Molecular Ecology 14: 2611–2620. Google Scholar

21.

US Fish and Wildlife Service. 1985. Endangered and Threatened Wildlife and Plants: Final rule to determine Primula maguirei (Maguire's primrose) to be a threatened species (Vol. 50, pp. 49–52). Department of the Interior. Google Scholar

22.

Fleishman, E., G. T. Austin, and D. D. Murphy. 2001. Biogeography of Great Basin butterflies: Revisiting patterns, paradigms, and climate change scenarios. Biological Journal of the Linnean Society. Linnean Society of London 74: 501–515. Google Scholar

23.

Floyd, C. H., D. H. Van Vuren, and B. May. 2005. Marmots on great basin mountaintops: Using genetics to test a biogeographic paradigm. Ecology 86: 2145–2153. Google Scholar

24.

Gray, A. 1888. Synoptical Flora of North America . Ivison, Blakeman, Taylor & Co. Google Scholar

25.

Harper, K. T., C. D. Freeman, K. W. Ostler, and L. G. Klikoff. 1978. The flora of Great Basin mountain ranges: Diversity sources and dispersal ecology. Great Basin Naturalist Memoirs 2: 81–103. Google Scholar

26.

He, X., J. J. Cao, W. Zhang, Y. Q. Li, C. Zhang, X. H. Li, G. H. Xia, and J. W. Shao. 2021. Integrative taxonomy of herbaceous plants with narrow fragmented distributions: A case study on Primula merrilliana species complex. Journal of Systematics and Evolution :  https://doi.org/10.1111/jse.12726Google Scholar

27.

Hewitt, G. M. 1996. Some genetic consequences of ice ages, and their role in divergence and speciation. Biological Journal of the Linnean Society. Linnean Society of London 58: 247–276. Google Scholar

28.

Hodel, R. G. J., R. Massatti, S. G. D. Bishop, and L. L. Knowles. 2021. Testing which axes of species differentiation underlie covariance of phylogeographic similarity among montane sedge species. Evolution 75: 349–364. Google Scholar

29.

Holmgren, N. H. 1967. A new species of primrose from Nevada. Madroño 19: 27–29. Google Scholar

30.

Holmgren, N. H. and A. H. Holmgren. 1974. Three new species from the Great Basin. Brittonia 26: 309–315. Google Scholar

31.

Holmgren, N. H., P. K. Holmgren, and A. Cronquist. 2005. Intermountain Flora: Vascular Plants of the Intermountain West, USA, vol. 2, Part B: Subclass Dilleniidae . New York: New York Botanical Garden. Google Scholar

32.

Holmgren, N. H. and S. Kelso. 2001. Primula cusickiana (Primulaceae) and its varieties. Brittonia 53: 154–156. Google Scholar

33.

Huang, J. 2020. Is population subdivision different from speciation? From phylogeography to species delimitation. Ecology and Evolution 36: 303. Google Scholar

34.

Huang, Y. F., L. N. Dong, and W. B. Xu. 2019. Lysimachia fanii, a new species of Primulaceae from limestone area of Guangxi, China. Phyto-Keys 130: 75–84. Google Scholar

35.

Hulshof, C. M. and M. J. Spasojevic. 2020. The edaphic control of plant diversity. Global Ecology and Biogeography 29: 1634–1650. Google Scholar

36.

Huson, D. H. and D. Bryant. 2006. Application of phylogenetic networks in evolutionary studies. Molecular Biology and Evolution 23: 254–267. Google Scholar

37.

Jombart, T., S. Devillard, and F. Balloux. 2010. Discriminant analysis of principal components: a new method for the analysis of genetically structured populations. BMC Genetics 11: 94. Google Scholar

38.

Kass, R. and S. Welsh. 1985. New species of Primula (Primulaceae) from Utah. The Great Basin Naturalist 45: 548–550. Google Scholar

39.

Kelso, S. 1992. The genus Primula as a model for evolution in the Alaskan flora. Arctic and Alpine Research 24: 82–87. Google Scholar

40.

Kelso, S., P. M. Beardsley, and K. Weitemier. 2009. Phylogeny and Biogeography of Primula sect. Parryi (Primulaceae). International Journal of Plant Sciences 170: 93–106. Google Scholar

41.

Kopelman, N. M., J. Mayzel, M. Jakobsson, N. A. Rosenberg, and I. Mayrose. 2015. CLUMPAK: A program for identifying clustering modes and packaging population structure inferences across K. Molecular Ecology Resources 15: 1179–1191. Google Scholar

42.

Koontz, A., W. D. Pearse, and P. Wolf. 2022. Data from: Pronounced genetic separation among varieties of the Primula cusickiana species complex, a Great Basin endemic. Dryad Digital Repository.  https://doi.org/10.5061/dryad.12jm63xzvGoogle Scholar

43.

Lawlor, T. E. 1998. Biogeography of Great Basin mammals: Paradigm lost? Journal of Mammalogy 79: 1111–1130. Google Scholar

44.

Leinonen, T., R. B. O'Hara, J. M. Cano, and J. Merilä. 2008. Comparative studies of quantitative trait and neutral marker divergence: a meta-analysis. Journal of Evolutionary Biology 21: 1–17. Google Scholar

45.

Li, J., J. M. Cocker, J. Wright, M. A. Webster, M. McMullan, S. Dyer, D. Swarbreck, M. Caccamo, C. van Oosterhout, and P. M. Gilmartin. 2016. Genetic architecture and evolution of the S locus supergene in Primula vulgaris . Nature Plants 2: 16188. Google Scholar

46.

Mansfield, D. 1993. The status of “Primula wilcoxiana” and the Primula cusickiana complex in Southwestern Idaho. Bureau of Land Management, U.S. Department of the Interior. Google Scholar

47.

Mansfield, D. 2010. Vascular flora of the Owyhee River watershed in Oregon. Journal of the Idaho Academy of Science 46: 1–127. Google Scholar

48.

Mast, A. R. and J. L. Reveal. 2007. Transfer of Dodecatheon to Primula (Primulaceae). Brittonia 59: 79–82. Google Scholar

49.

Novembre, J. 2016. Pritchard, Stephens, and Donnelly on population structure. Genetics 204: 391–393. Google Scholar

50.

Nowak, C. L., R. S. Nowak, R. J. Tausch, and P. E. Wigand. 1994. Tree and shrub dynamics in northwestern Great Basin woodland and shrub steppe during the Late-Pleistocene and Holocene. American Journal of Botany 81: 265–277. Google Scholar

51.

Pritchard, J. K., M. Stephens, and P. Donnelly. 2000. Inference of population structure using multilocus genotype data. Genetics 155: 945–959. Google Scholar

52.

R Core Team. 2020. R: A language and environment for statistical computing. Vienna, Austria: R Foundation for Statistical Computing. Google Scholar

53.

Rajakaruna, N. 2018. Lessons on evolution from the study of edaphic specialization. Botanical Review 84: 39–78. Google Scholar

54.

Ren, G., R. G. Mateo, E. Conti, and N. Salamin. 2020. Population genetic structure and demographic history of Primula fasciculata in Southwest China. Frontiers in Plant Science 11: 986. Google Scholar

55.

Richards, A. J. 2003. Primula . Portland, Oregon: Timber Press. Google Scholar

56.

Schmidt-Lebuhn, A. N., J. M. de Vos, B. Keller, and E. Conti. 2012. Phylogenetic analysis of Primula section Primula reveals rampant non-monophyly among morphologically distinct species. Molecular Phylogenetics and Evolution 65: 23–34. Google Scholar

57.

Schorr, G., P. B. Pearman, A. Guisan, and J. W. Kadereit. 2013. Combining palaeodistribution modelling and phylogeographical approaches for identifying glacial refugia in alpine Primula. Journal of Biogeography 40: 1947–1960. Google Scholar

58.

Schott, H. W. 1852. Oesterreichische Primeln. Österreichisches Botanisches Wochenblatt 2: 35–36. Google Scholar

59.

Sukumaran, J. and L. L. Knowles. 2017. Multispecies coalescent delimits structure, not species. Proceedings of the National Academy of Sciences USA 114: 1607–1612. Google Scholar

60.

Theodoridis, S., D. Nogués-Bravo, and E. Conti. 2019. The role of cryptic diversity and its environmental correlates in global conservation status assessments: Insights from the threatened bird's-eye primrose (Primula farinosa L.). Diversity & Distributions 25: 1457–1471. Google Scholar

61.

Thompson, R. S. and J. I. Mead. 1982. Late Quaternary environments and biogeography in the Great Basin. Quaternary Research 17: 39–55. Google Scholar

62.

Trift, I., M. Källersjö, and A. A. Anderberg. 2002. The monophyly of Primula (Primulaceae) evaluated by analysis of sequences from the chloroplast gene rbcL. Systematic Botany 27: 396–407. Google Scholar

63.

Weir, B. S. and C. C. Cockerham. 1984. Estimating F-Statistics for the analysis of population structure. Evolution 38: 1358–1370. Google Scholar

64.

Williams, L. O. 1936. Revision of the Western Primulas. American Midland Naturalist 17: 741–748. Google Scholar

65.

Wolf, P. G. and R. B. Sinclair. 1997. Highly differentiated populations of the narrow endemic plant Maguire Primrose (Primula maguirei). Conservation Biology 11: 375–381. Google Scholar

66.

Wright Smith, W. and H. R. Fletcher. 1949. The genus Primula: Sections Cuneifolia, Floribundae, Parryi, and Auricula. Transactions of the Royal Society of Edinburgh 61: 631–686. Google Scholar

67.

Yang, L., H. Kong, J. P. Huang, and M. Kang. 2019. Different species or genetically divergent populations? Integrative species delimitation of the Primulina hochiensis complex from isolated karst habitats. Molecular Phylogenetics and Evolution 132: 219–231. Google Scholar

68.

Zhang, L., H. P. Comes, and J. W. Kadereit. 2004. The temporal course of Quaternary diversification in the European high mountain endemic Primula sect. Auricula (Primulaceae). International Journal of Plant Sciences 165: 191–207. Google Scholar

69.

Zhang, L. B. and J. W. Kadereit. 2004. Classification of Primula sect. Auricula (Primulaceae) based on two molecular data sets (ITS, AFLPs), morphology and geographical distribution. Botanical Journal of the Linnean Society 146: 1–26. Google Scholar

Appendices

Appendix 1. List of voucher specimens included in this study. Order of data is as follows: Species, Voucher, Herbarium. Institutional barcodes or accession numbers are included as parenthetical values following the voucher, when available.

Ingroup: Primula cusickiana var. cusickiana, 25330978, Intermountain Herbarium; Primula cusickiana var. cusickiana, 25330990, Intermountain Herbarium; Primula cusickiana var. cusickiana, 25331045, Intermountain Herbarium; Primula cusickiana var. cusickiana, 25331062, Intermountain Herbarium; Primula cusickiana var. cusickiana, 25331021, Intermountain Herbarium; Primula cusickiana var. cusickiana, 25331015, Intermountain Herbarium; Primula cusickiana var. cusickiana, 25331018, Intermountain Herbarium; Primula cusickiana var. cusickiana, 25331034, Intermountain Herbarium; Primula cusickiana var. cusickiana, 25331004, Intermountain Herbarium; Primula cusickiana var. cusickiana, 25330994, Intermountain Herbarium; Primula cusickiana var. cusickiana, 25330991, Intermountain Herbarium; Primula cusickiana var. maguirei, 25331026, Intermountain Herbarium; Primula cusickiana var. maguirei, 25331039, Intermountain Herbarium; Primula cusickiana var. maguirei, 25331041, Intermountain Herbarium; Primula cusickiana var. nevadensis, 25331101, Intermountain Herbarium; Primula cusickiana var. nevadensis, 25331106, Intermountain Herbarium; Primula cusickiana var. nevadensis, 25331092, Intermountain Herbarium; Primula cusickiana var. domensis, 25331066, Intermountain Herbarium; Primula cusickiana var. domensis, 25331070, Intermountain Herbarium; Primula cusickiana var. domensis, 25331077, Intermountain Herbarium; Primula cusickiana var. domensis, 25331083, Intermountain Herbarium.

Outgroups: Primula capillaris, 770850 (ASU0020421), Arizona State University Vascular Plant Herbarium; Primula capillaris, 3025822 (UTC00138833), Intermountain Herbarium; Primula parryi, 25331110, Intermountain Herbarium; Primula parryi, 25331112, Intermountain Herbarium.

© Copyright 2022 by the American Society of Plant Taxonomists
Austin Koontz, William D. Pearse, and Paul G. Wolf "Pronounced Genetic Separation among Varieties of the Primula cusickiana Species Complex, a Great Basin Endemic," Systematic Botany 47(3), 887-897, (16 September 2022). https://doi.org/10.1600/036364422X16573019348175
Published: 16 September 2022
KEYWORDS
allopatry
biogeography
cryptic speciation
edaphic endemism
heterostyly
Primula
RADseq
Back to Top