Accepted Manuscript

SPECIES BOUNDARIES IN THE CUSICKII COMPLEX DE- LIMITED USING MOLECULAR PHYLOGENETIC TECHNIQUES

J.C. Zimmers, M. Thomas, L. Yang, A. Bombarely, M.M. Mancuso, M.F. Wojciechowski, J.F. Smith

PII: S1055-7903(17)30039-8 DOI: http://dx.doi.org/10.1016/j.ympev.2017.06.004 Reference: YMPEV 5845

To appear in: Molecular Phylogenetics and Evolution

Received Date: 11 January 2017 Revised Date: 1 May 2017 Accepted Date: 7 June 2017

Please cite this article as: Zimmers, J.C., Thomas, M., Yang, L., Bombarely, A., Mancuso, M.M., Wojciechowski, M.F., Smith, J.F., SPECIES BOUNDARIES IN THE ASTRAGALUS CUSICKII COMPLEX DELIMITED USING MOLECULAR PHYLOGENETIC TECHNIQUES, Molecular Phylogenetics and Evolution (2017), doi: http:// dx.doi.org/10.1016/j.ympev.2017.06.004

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain. SPECIES BOUNDARIES IN THE ASTRAGALUS CUSICKII COMPLEX DELIMITED

USING MOLECULAR PHYLOGENETIC TECHNIQUES

J. C. Zimmers1, M. Thomas2, L. Yang2, A. Bombarely3, M. M. Mancuso1, M. F.

Wojciechowski4, and J. F. Smith1

1Department of Biological Sciences, Boise State University, 1910 University Drive, Boise,

Idaho 83725 U. S. A.

2Department of Biological Sciences, Idaho State University, 921 South 8th Avenue

Pocatello, Idaho, 83209-8007 U.S.A.

3Department of Horticulture, Virginia Tech, Latham Hall 216, 220 Ag Quad Lane, Blacksburg,

Virginia 24060 U.S.A.

4School of Life Sciences, Arizona State University, Tempe, Arizona 85287-4501, USA

Abstract— Understanding the source of phenotypic variability is a challenge in the biological sciences. Variation in phenotypes is the result of variation in the genetics and environment the organism experiences, but elucidating the relative contribution of these two parameters can pose problems, especially in the field of systematics. Systematists are challenged to classify biological diversity into groups that share common ancestry. Phenotypic variation can be useful to demonstrate common ancestry, but only when the primary contributor to the variation is under strong genetic control, and thus heritable. Cusick’s milkvetch (Astragalus cusickii) is a perennial forb endemic to the northwestern intermountain region of the United States. The species currently comprises four varieties based on subtle morphological dissimilarities, such as leaf size and density, and the size and shape of the seed pods. The taxonomic organization of the varieties of A. cusickii and related species of Astragalus were reexamined through phylogenetic analysis of low copy nuclear, nuclear-ribosomal, and chloroplast gene regions.

Maximum parsimony, maximum likelihood, Bayesian inference, the genealogical sorting index, and an approximately unbiased test were used to determine appropriate species boundaries under the phylogenetic species concept. The results support reclassification of A. cusickii var. packardiae and A. cusickii var. sterilis as separate species. Additionally, evidence suggests a chloroplast capture event may have occurred in one population of A. cusickii var.

2

packardiae.

Keywords — coalescence; reciprocal monophyly; species boundaries; varieties

1. Introduction

Distinct populations under different environmental conditions can have markedly different phenotypes. Such differences are the basis for recognizing distinct taxa, but can be a challenge if the phenotype is the result of the environmental conditions on a common genotype.

Such phenotypic plasticity, where environmental factors alter the phenotype despite identical genetic backgrounds, erodes the reliability of phenotypic variation as the sole method of diagnosing species boundaries (Mayr, 1969). The impact of phenotypic plasticity is especially critical in that lack the ability to directly move in the landscape and therefore must contend with the environment where they germinate. Consequently, plants may exhibit large- scale morphological and physiological responses to variations in environmental factors such as soil nutrient content, temperature, and water availability which can have profound effects on their phenotypes (Sultan, 2000; Gurevitch et al., 2002; Valladares et al., 2007). These challenges have been a source of confusion for hundreds of years and are being resolved with more analyses that examine the genotypic variation directly (Belton et al., 2014; Muggia et al.,

2014; DeBiasse and Heilberg, 2015). The advent of molecular systematics has revolutionized our classification of biodiversity from the most inclusive (Adl et al., 2005; Soltis et al., 2005;

Hibbett et al., 2007) to the least inclusive taxonomic levels (Stepansky et al., 1999;

3

Bronikowski and Arnold, 2001; Clark et al., 2003; Abbasi et al., 2005; Tomasello et al., 2014), and has been a major means of improving our overall concept of biological diversity.

One group of plants that has a complicated taxonomic history in terms of interpreting phenotypic diversity and its ability to resolve taxonomic boundaries is the genus Astragalus L.

(Polhill, 1981; Wojciechowski, 2005). Astragalus () is a diverse group of approximately 2,500 species in the Old and New Worlds, containing more recognized species

than any other genus of angiosperms (Frodin, 2004; Lock and Schrire, 2005; Mabberley, 2008).

Until recently many systematicists regarded Astragalus as a ‘wastebasket’ genus, likely to be paraphyletic (Polhill, 1981; Wojciechowski, 2005). However, while the monophyly of

Astragalus sensu stricto has been well-supported (Sanderson, 1991; Sanderson and Doyle,

1993; Wojciechowski et al., 1993, 1999), many species-level relationships within the genus remain poorly resolved. Species of Astragalus can be found on every continent except Australia and Antarctica (Lewis et al., 2005). The nearly cosmopolitan distribution and extreme morphological diversity of Astragalus make it a difficult genus for systematic studies

(Sanderson and Doyle, 1993; Scherson et al., 2008). Mating systems have been studied in fewer than 1% of the species within Astragalus (Watrous and Cane, 2011) and most species in the

Old World had not been revised since the late 19th century (Bunge, 1868, 1869; Taubert, 1894) until only recently (Podlech and Zarre, 2013).

Astragalus exhibits rich diversity in four geographic areas (southwest and south-central

Asia, the Sino-Himalayan region, the Mediterranean Basin, and western North America; in addition the Andes in South America have at least 100 species). The most diverse of these areas, home to approximately 2000 species, and assumed to be the place of origin of the genus,

4

are the steppes and mountains of southwest and south-central Asia and the Himalayan plateau

(Wojciechowski, 2005). Second to Eurasia in terms of species diversity is the New World, with approximately 400-450 species. The Intermountain Region of western North America

(Barneby, 1989) is especially diverse, and an estimated 70 species of Astragalus can be found in Idaho alone, including several endemic taxa (Mancuso, 1999).

Among the species of Astragalus in the North American Intermountain Region is

Astragalus cusickii. First described by Gray (1878), A. cusickii is a sparsely leafy, multi-

stemmed, perennial forb found in western Idaho, eastern Oregon, and the extreme southeast corner of Washington (Fig. 1). The species overall has small to moderately sized flowers and conspicuous papery inflated pods (Fig. 2). It is found primarily on relatively sparsely vegetated to canyon grassland sites, often steep hillsides, ash soils, and talus slopes (Barneby, 1989;

Mancuso, 1999). The species currently comprises four infraspecific taxa, varieties cusickii, flexilipes, sterilis, and packardiae. The inclusion of these four taxa into a single species was made on the basis of morphological similarity. However, it is possible that edaphic and microclimatic conditions where A. cusickii var. sterilis and A. cusickii var. packardiae occur may alter the morphologies of these varieties and as such, they may not merit taxonomic distinction.

Herein we resolve the taxonomic boundaries for the four varieties of A. cusickii using molecular phylogenetic techniques. Species status is evaluated using the criteria of the phylogenetic species concept, monophyly with diagnosable differences within a larger understanding of species as separately evolving meta-population lineages, as described in the unified species concept (de Queiroz, 2003). This concept has been selected because 1) monophyly can be assessed, 2) morphological differences are known and presumably the result

5

of inheritance from a common ancestor if monophyly is found, and 3) there are few studies to directly address the breeding system of these plants. Recent studies of species in Astragalus have also employed monophyly with diagnosable differences between populations as the criteria for recognizing species (Scherson et al., 2008; Riahi et al., 2011).

2. Materials and Methods

2.1 Study Species

Astragalus cusickii var. cusickii (Fig. 2) has the widest geographic distribution of the four varieties. It is found in western Idaho, eastern Oregon, and southeast Washington, although with a concentration in the Hells Canyon area (Fig. 1). Individuals of A. cusickii var. cusickii are generally the largest in physical size of the varieties. Notably, they possess an approximately even distribution of leaves throughout the length of the stems, in contrast to some of the other varieties (Barneby, 1989). The flowers are relatively larger and typically more abundant than those of the other varieties and the pods have a more inflated appearance and are often found in greater abundance after flowering than in the other varieties (although A. cusickii var. sterilis has pods that are similarly inflated). Astragalus cusickii var. flexilipes (Fig.

2) appears to be only weakly differentiated from A. cusickii var. cusickii by subtle morphological differences: small, purplish flowers, and oblique, half-ellipsoid pods (Barneby,

1989). It can be found in the vicinity of the lower Salmon River (Fig. 1). The distribution of

Astragalus cusickii var. sterilis is limited to a small geographic area in southeastern Oregon and adjacent southwestern Idaho (Fig. 1). This variety is distinguished by its smaller leaflets and bright red mottling on its pods (Fig. 2). It is considered rare and of conservation concern in both Oregon and Idaho.

6

Astragalus cusickii var. packardiae (Fig. 2) is distinguished by its relative paucity of leaves on the stems, particularly distally, its relatively small purplish flowers, and its small, narrow pods. It is considered one of the rarest taxa in Idaho, being restricted to an approximately 10 square mile area in Payette County, Idaho (Mancuso, 1999). It has become a high priority conservation concern due to its limited geographic distribution, small population size, habitat decline, and vulnerability of its habitat to multiple, ongoing disturbances and threats (Mancuso, 2016).

Astragalus cusickii var. packardiae occurs almost entirely on a visually distinct soil type that contrasts sharply to the surrounding areas. Small exposures of this relatively sparsely vegetated whitish substrate dot the landscape, especially on steep slopes. These exposures are sparsely covered in vegetation. Given the abrupt change in vegetation and visually distinct nature of these exposures, it is reasonable to assume that ecologically significant differences exist in the edaphic properties of the exposures versus the surrounding landscape. Unique edaphic conditions often commonly influence the narrow distribution of rare endemic plants

(Kruckeberg and Rabinowitz, 1985). It is important when considering the taxonomic organization of A. cusickii to consider the possibility that morphological variation observed between the different varieties may be phenotypic plasticity in response to different edaphic conditions. It is possible that the four varieties of A. cusickii represent a single meta-population, with gene flow across the entire distribution, and where individual populations exhibit observable phenotypic variation in response to the particular type of soil they are found on.

Given their geographic proximity and morphological similarity to the other varieties of A. cusickii, as well as the tendency of plants to exhibit strong physiological responses to different environments (Gurevitch et al., 2002), it is possible that the distinct morphological characters apparent in A. cusickii vars. packardiae and sterilis may be the result of their environments.

7

2.2 Data Collection and Sampling

Sequences included in this study originated from GenBank and DNA extracted from field-collected individuals (Appendix A). Field-collected individuals were gathered from southwestern Idaho and eastern Oregon (Fig. 1). Leaf material was collected from 20 individuals from five populations of Astragalus cusickii var. packardiae, nine individuals from three populations of A. cusickii var. sterilis, twelve individuals from four populations of A.

cusickii var. cusickii, and nine individuals from three populations of A. cusickii var. flexilipes.

In addition to the four varieties of A. cusickii (sect. Cusickiani), seven additional aneuploid species of Astragalus found in the western U.S. were collected and included in the study on the basis of similar leaf and/or fruit morphology to A. cusickii. These seven species are: A. ceramicus, A. filipes, and A. whitneyi var. confusus of sect. Cusickiani, A. lentiginosus of sect.

Diphysi, A. mulfordiae and A. yoder-williamsii of sect. Neonix, A. solitarius of sect. Solitarii.

Astragalus purshii of sect. Argophylli, which does not exhibit similar leaf and/or fruit morphology to A. cusickii, was also included in the study. Leaf material was preserved in silica gel. DNA was extracted from frozen and pulverized leaf tissue with Qiagen DNeasy plant mini kits (Valencia, CA) according to manufacturer’s instructions.

Four gene regions were chosen for investigation based on success in previous molecular systematics studies and their potential utility for species-level resolution (Table 1). Low-copy nuclear, nuclear ribosomal, and chloroplast gene regions were targeted to include a broad survey of the genome, a necessary approach due to a lack of species-level resolution in previous studies (Riahi et al., 2011; Scherson et al., 2008; Wojciechowski, 2005).

Four regions have been used in previous phylogenetic analyses of Astragalus. The

8 nuclear rDNA internal transcribed spacer (nrDNA ITS) region was investigated using the C26A and NC18s10 primers (Wen and Zimmer, 1996). The utility of nrDNA ITS for systematic studies of plants was recognized by Baldwin (1992) and has been important in phylogenetic studies at the species level in many groups, including: Rosaceae: Prunus (Lee and Wen, 2001),

Betulaceae: Corylus (Whitcher and Wen, 2001), Rutaceae: Haplophyllum (Navarro et al.,

2004), Apiaceae: Sium, Cryptotaenia (Spalik and Downie, 2006, 2007; Apioideae Zhou et al.,

2009; Lomatium Carlson et al., 2011; George et al., 2014), Lamiaceae: Dicerandra (Oliveira et

al., 2007), Piperaceae: Piper, Peperomia (Smith et al., 2008; Jaramillo et al., 2008),

Crassulaceae: Thompsonella (Carillo-Reyes et al., 2008), Saxifragaceae: Saniculaphyllum

(Xiang et al., 2012), Gesneriaceae: Columnea (Smith et al., 2013), Papaveraceae: Fumaroideae

(Pérez-Gutiérrez et al., 2015) and Astragalus (Wojciechowski et al., 1993, 1999). The external transcribed spacer (ETS) region was initially developed for phylogenetic analyses by Baldwin and Markos (1998) and investigated here using the primers developed by Riahi et al. (2011) for

Astragalus. The cyclic nucleotide-gated channel 4 (CNGC4) protein-coding gene was developed by Choi et al. (2004, 2006) and its utility in phylogenetic studies of Astragalus was demonstrated by Scherson et al. (2005, 2008). The trnS-G intergenic spacer (trnS-trnG) region of cpDNA was developed by Shaw et al. (2005). Riahi et al. (2011) successfully used this chloroplast spacer in a phylogenetic analysis of Old World Astragalus, and it has been used in studies of other taxa in Fabaceae (e.g., Zhang et al., 2009). DNA was amplified by polymerase chain-reaction (PCR) using the methods of Smith et al. (1997).

Sequence data from all individuals of Astragalus cusickii var. packardiae were included in the analysis, as this taxon is the focus of the study. Field-collected individuals from the other varieties of A. cusickii and other species of Astragalus were represented by a single individual

9 from each population from which they were gathered. GenBank sequence data for nrDNA ITS and CNGC4 from eleven more aneuploid Astragalus species from western North America were added to supplement the field-collected individuals: A. allochrous, A. arizonicus, A. asymmetricus, A. asclepiadoides, A. brandegei, A. calycosus, A. douglasii, A. falcatus, A. inyoensis, A. lonchocarpus, A. mollissimus, A. nothoxys, A. oxyphysus, A. pachypus, A. preussii, A. tetrapterus, and A. woodruffii (Appendix A). To place A. cusickii and the other western North American Astragalus species in a larger phylogenetic context, GenBank

sequences from species occurring in other geographic areas were added (Appendix A). These include four euploid Astragalus species native to North America, but are not exclusively found in the west: A. adsurgens, A. alpinus, A. americanus, and A. canadensis, 17 Astragalus aneuploid species native to South America: A. amatus, A. arnottianus, A. berteroanus, A. cruckshanksii, A. cryptobotrys, A. curvicaulis, A. cysticalyx, A. darumbium, A. edmonstonei, A. johnstonii, A. looseri, A. monticola, A. nivicola, A. patagonicus, A. pehuenches, A. uniflorus, and A. vagus, and seven Old World euploid Astragalus species: A. atropilosulus, A. cerasocrenus, A. complanatus, A. corrugatus, A. epiglottis, A. peristereus, and A. vogelii (syn.

Podlechiella vogelii Maassoumi & Kaz. Osaloo). Oxytropis sericea was included as the outgroup

(Scherson et al., 2008).

2.3 Whole Plastid Sequencing

A small set of samples was selected to generate whole or nearly whole chloroplast genomes for phylogenetic analyses. These individuals were chosen to represent the four varieties of A. cusickii and close relatives to determine if a monophyletic A. cusickii could be recovered with greater data sampling. These included two accessions of A. whitneyi var. confusus, and A. solitarius. The published chloroplast genome of Glycyrrhiza glabra was used

10 as the outgroup. For each species, approximately 50 ng of DNA were prepared with the

Nextera DNA sample prep kit (Illumina, San Diego, CA) according to the manufacturer protocol. Chloroplast DNA was sequenced by an Illumina MiSeq (2×250 read length) at the

Idaho State University Molecular Research Core Facility. Illumina reads were processed using

Fastq-mcf with a minimum qscore of 30 and a minimum length of 50 (-q 30 –l 50). Reads were mapped to the Glycyrrhiza glabra chloroplast genome (GenBank accession: NC_024038) using

Bowtie2 v2.2.4 (Langmead and Salzberg, 2012; Aronesty, 2013). Mapped reads were filtered

and sorted using Samtools v1.1 (Li et al., 2009). A reference based chloroplast assembly was constructed using Samtools mpileup. Gaps were filled using GapCloser from the

SOAPdenovo2 package (Luo et al., 2012).

Reference based assemblies were aligned using Muscle (Edgar, 2004). Two independent methodologies were used for the phylogenetic analysis:1) A maximum likelihood

(ML) analysis was performed using PhyML v20131022 (Guindon et al., 2010) with 100 bootstrap replicates; 2) A Bayesian inference (BI) analysis was performed using BEAST v1.7

(Drummond et al., 2012) with a MCMC of 10,000,000 generations. The substitution model was previously optimized using jModelTest 2.1.7 (Darriba et al., 2012). As a result, the model GTR

+  + I was used for both approaches. Phylogenetic trees were plotted using FigTree

(http://tree.bio.ed.ac.uk/software/figtree/).

Read processing, mapping, assembly, alignment, and phylogenetic analyses were performed in an Ubuntu 14.04 Linux server (128 Gb of RAM, 4 Tb hard drive, 64 cores) purchased at Red Barn computers.

11

2.4 Alignment and Phylogenetic Analyses

Sanger sequence data were manually aligned and edited for quality using Phy-DE

(Müller at al., 2010). There was a 30 base pair reverse complement of a section of trnS-G in some individuals. The sequence was manually reversed for the individuals that had the minority version, to match the majority. The presence of the majority or minority version of this sequence was coded as a single character state appended to the sequence data for each individual in the maximum parsimony (MP) analyses.

Sequence data were not available for all individuals and all gene regions. Three concatenated super matrices were assembled to determine the extent to which the missing data would affect phylogenetic analyses. A preliminary maximum parsimony analysis of individual gene regions was undertaken prior to concatenation. An incongruence was detected in the placement of three individuals of A. cusickii var. packardiae (JZ-052, JZ-053, and JZ-054)

(Figs. 3, 4). These individuals were removed from all further gene matrices. Matrix 1 included all available data for all individuals (Appendix A), and included Oxytropis sericea as the outgroup. Matrix 2 included individuals for which sequence data from at least two gene regions

(Appendix A) were available and also included Oxytropis sericea as the outgroup. Matrix 3 included only individuals for which sequence data was available for nrDNA ITS, ETS,

CNGC4, and trnS-G (Appendix A). Consequently, matrix 3 does not contain any sequences from GenBank, including the outgroup Oxytropis sericea, as sequence data from GenBank was only available for, at most, two of the gene regions examined. For this reason, matrix 3 used

Astragalus purshii as the outgroup in place of Oxytropis sericea, based on the phylogenetic position of A. purshii in preliminary analysis of all data.

Simple indel coding was conducted with the SeqState plug-in for Phy-DE. Maximum

12

parsimony analysis was conducted on all three matrices both with and without indel coding, using TNT version 1.1 (Goloboff et al., 2008). Default settings were used unless otherwise specified. A strict consensus tree was generated with TNT. Branch support values for the strict consensus tree were found with bootstrapping (Felsenstein, 1985) for 10,000 replicates using a tree bisection-reconnection swapping algorithm. Branch arrangements with bootstrap values ≥

75 were considered statistically significant. Matrix 1 was chosen as the dataset to proceed with all subsequent analyses (see results).

Maximum likelihood (ML) topology was reconstructed using RAxML-HPC2

(Stamatakis, 2014) on XSEDE on the Cipres Science Gateway (Miller et al., 2010). Oxytropis sericea was specified as the outgroup. One hundred bootstrap iterations were used. Branch arrangements with bootstrap values ≥ 75 were considered statistically significant. The

GTRCAT model, as the default in RAxML-HPC2 was used. The dataset was not partitioned.

FigTree v. 1.4.2 (Rambaut, 2006) was used to visualize the topology of the best tree with bootstrap values.

Prior to the Bayesian inference (BI) analysis, appropriate partitions were found using

PartitionFinder (Lanfear et al., 2012). Nucleotide substitution models for each partition were found using jModelTest 2.1 (Darriba et al., 2012). Bayesian inference was conducted using

MrBayes 3.2.3 (Altekar et al., 2004; Huelsenbeck and Ronquist, 2001; Ronquist and

Huelsenbeck, 2003) on XSEDE on the Cipres Science Gateway (Miller et al., 2010). Two independent analyses were conducted using four Metropolis-coupled Markov chains (MCMC), one hot and three cold chains, all with the same temperature (Geyer, 1991; Hastings, 1970;

Metropolis et al., 1953) each for ten million generations. Burn-in was set at 50,000 generations.

Metropolis-coupled Markov chain analysis completion was tested with AWTY (Nylander et al.,

13

2007) and Tracer v. 1.6 (Rambaut et al., 2014). FigTree v. 1.4.2 (Rambaut, 2006) was used to visualize the topology of the majority-rule consensus tree with branch posterior probabilities.

Branch arrangements with posterior probability values ≥ 0.95 were considered statistically significant.

2.5 Testing Alternative Hypotheses

To further investigate the potential that the four varieties of A. cusickii represent a monophyletic group, an approximately unbiased (AU) test (Shimodaira, 2002) was performed with CONSEL (Shimodaira and Hasegawa, 2001). Two additional AU tests were conducted to investigate the relative power of the AU test as applied to Astragalus at the species level. In these additional tests the four varieties of A. cusickii were specified as belonging to a monophyletic group with a morphologically similar species (A. whitneyi var. confusus) and, separately, with a much less morphologically similar species (A. solitarius). Site likelihoods generated in PAUP* (Swofford, 2002) were input into CONSEL. The significance level was designated at α = 0.05 (Lang et al., 2002; Dantrakool et al., 2004; Shannon et al., 2005; Gill and

Fast, 2006; Heiss and Keeling, 2006; Struck et al., 2007; Ernst et al., 2008; Helmkampf et al.,

2008; Kuo et al., 2008; Yu et al., 2008; Gao et al., 2009; Ishiwata et al., 2011; Zhang et al.,

2011).

The genealogical sorting index (GSI) (Cummings et al., 2007) was used to further investigate the possibility of a monophyletic Astragalus cusickii. A file containing the last 100

trees (the maximum number allowed by the GSI software) generated in the BI analysis, as well 14

as text files used to specify monophyletic constraints among taxa, were used as the inputs. The significance level was designated at α = 0.05 (Koopman and Baum, 2010; Kubatko et al., 2011;

Keith and Hedin, 2012; Levsen et al., 2012).

The following monophyletic arrangements were defined and tested against the BI data using the GSI (Table 2): 1) each variety of Astragalus cusickii as a single independent monophyletic group, 2) all four varieties of A. cusickii as a single monophyletic group, 3) A.

cusickii as three separate monophyletic groups: A. cusickii var. packardiae, A. cusickii var. sterilis, and a combined group containing A. cusickii vars. cusickii and flexilipes, 4) all four varieties of A. cusickii combined with A. whitneyi var. confusus as a monophyletic group, and

5) all four varieties of A. cusickii combined with A. solitarius as a monophyletic group. The rationale for combining A. cusickii vars. cusickii and flexilipes was based on the results of the phylogenetic analyses and their status as weakly differentiated varieties (Barneby, 1989).

Astragalus whitneyi, a morphologically similar species to A. cusickii, and A. solitarius, a less morphologically similar species to A. cusickii, were included in separate analyses as a test of the relative power of the GSI as applied to Astragalus at the species level.

3. Results

3.1 Amplification and Sequencing

NrDNA ITS and trnS-G were successfully amplified and sequenced for 43 individuals newly sampled for this study (Appendix A). NrDNA ITS sequences were available for an additional 46 individuals accessed from GenBank, and CNGC4 were available for an additional

11 individuals. ETS was successfully amplified and sequenced for 53% of individuals and

none were available on GenBank. CNGC4 was successfully amplified and sequenced for 98% 15

of the individuals.

3.2 Phylogenetic Analyses

Maximum parsimony analysis of matrix 1 (Fig. 5), with sequence data from 89 individuals, all with nrDNA ITS, 40 with sequences from trnS-G, 23 with ETS sequences, and 55 with CNGC4 (Appendix A) resulted in 30 equally most-parsimonious trees (length

= 689, CI = 0.805, RI = 0.869). There is strong support for a monophyletic Astragalus

cusickii var. packardiae (BS = 95) and a monophyletic A. cusickii var. sterilis (BS = 84).

Astragalus cusickii var. cusickii and A. cusickii var. flexilipes were not resolved as reciprocally monophyletic, and instead formed a combined clade lacking bootstrap support.

Maximum parsimony analysis of matrix 2 (Supplemental Fig. 1), with sequence data from 55 individuals, all with nrDNA ITS, 40 of them with sequences from trnS-G, 23 with sequences from ETS, and 55 sequences from CNGC4 (Appendix A) resulted in 20 equally most-parsimonious trees (length = 419, CI = 0.886, RI = 0.929). In general, the results are similar to those found in the analysis of matrix 1: strong bootstrap support for a monophyletic

A. cusickii var. packardiae (BS = 97) and a monophyletic A. cusickii var. sterilis (BS = 96), as well as an unsupported clade including A. cusickii var. cusickii and A. cusickii var. flexilipes.

Maximum parsimony analysis of matrix 3 (Supplemental Fig. 2), with sequence data from 23 individuals, with sequences from all four regions (Appendix A) resulted in a single most-parsimonious tree (length = 154, CI = 0.883, RI = 0.941). Matrix 3 showed high bootstrap support for the reciprocal monophyly of three of the varieties of Astragalus cusickii: A. cusickii

var. packardiae (BS = 100), A. cusickii var. sterilis (BS = 99), and A. cusickii var. cusickii (BS 16

= 92). One variety, A. cusickii var. flexilipes, was not included in matrix 3 as sequence data were not available for all regions.

Comparison of the three matrices shows a trend toward less support as additional taxa lacking data from one or more regions are added. The proportion of supported nodes is highest in matrix 3 (73%) and drops with additional taxa in matrices 2 and 1 (30% and 29%, respectively). Matrix 1 was chosen for all subsequent analyses. Although this matrix has the

lowest proportion of supported nodes, it included the most taxa, thereby maximizing the phylogenetic space available to resolve relationships within A. cusickii.

The results of the ML analysis was generally in agreement with MP (Fig. 5). Astragalus cusickii var. packardiae and A. cusickii var. sterilis each received strong bootstrap support as reciprocally monophyletic groups (96% and 100%, respectively). Astragalus cusickii var. cusickii and A. cusickii var. flexilipes form an unsupported (52) combined clade.

PartitionFinder indicated that each of the four gene regions should be partitioned, as well as each codon position in CNGC4 (a protein-coding gene region), for a total of six partitions. Each partition was assigned a unique model by jModelTest, with the exception of

CNGC4 codon positions one and two, which were assigned the same model (Table 3). The

MCMC trace plot did not have an apparent vertical trend, suggesting MCMC completion. A joint-marginal plot of two independent BI analyses is consistent with MCMC convergence.

Metropolis-coupled Markov chain completion was supported by AWTY. The first of the independent analyses had a posterior mean of -5520.91, and a posterior effective sample size

(ESS) of 3373. The second independent analysis had a posterior mean of -5520.74, and a posterior ESS of 3015. The BI analysis is in agreement with the MP and ML analyses (Fig. 5).

17

Astragalus cusickii var. packardiae and A. cusickii var. sterilis each received strong posterior probabilities (1.00 and 1.00, respectively) as reciprocally monophyletic groups. Astragalus cusickii var. cusickii and A. cusickii var. flexilipes form a combined clade with PP = 0.922.

3.3 Testing Alternative Hypotheses

None of the phylogenetic analyses recovered the four varieties of Astragalus cusickii as a single monophyletic group. The AU test could not reject a monophyletic Astragalus cusickii

containing all varieties (p = 0.911). The AU test also failed to reject the possibility of a monophyletic combination of A. cusickii and A. whitneyi var. confusus (p = 0.746), but did reject a monophyletic combination of A. cusickii and A. solitarius (p = 0.005).

The GSI supported separately monophyletic Astragalus cusickii var. packardiae and A. cusickii var. sterilis (Table 2). Both taxa received the maximum score of 1.00 from the GSI when each was constrained to be monophyletic, indicating coalescence had occurred. Astragalus cusickii var. cusickii and A. cusickii var. flexilipes were supported as comprising a single clade in the results of the GSI (0.989). When evaluated separately, A. cusickii var. cusickii received a GSI of 0.796, and A. cusickii var. flexilipes 0.442. Evaluated as a single group, the four varieties of A. cusickii received a GSI score of 0.874. Including A. whitneyi var. confusus returned a GSI score of 0.874; including A. solitarius with the varieties of A. cusickii returned a GSI score of 0.839.

3.4 Chloroplast Genome Sequencing

Nearly complete chloroplast genomes were assembled for the eight samples. The ML analyses provided strong support for all nodes. The BI tree matched the topology of the ML

tree, but support was weaker for the position of A. cusickii var. cusickii and the placement of A. 18

cusickii var. sterilis (Fig. 6).

4. Discussion

All three phylogenetic methods produced nearly identical topologies for the concatenated data from Sanger sequencing (Fig. 5) which were also seen in the MP analyses of data matrices with reduced taxon sampling (Supplemental Figs. 1, 2). The results provide strong support for some clades, typically species where more than a single individual was

sampled, but poor support for relationships among species (Fig. 5). In particular, the monophyly of both A. cusickii var. packardiae and A. cusickii var. sterilis were strongly supported (Fig. 5). In contrast, neither A. cusickii var. cusickii nor A. cusickii var. flexilipes were recovered as reciprocally monophyletic (Fig. 5). Likewise support for a monophyletic A. cusickii that included all varieties was lacking, although this set of relationships could not be rejected with AU tests. Most relationships among the species from western North America showed low levels of resolution or support. These results clearly demonstrate that both A. cusickii var. packardiae and A. cusickii var. sterilis are reciprocally monophyletic and distinct from A. cusickii var. cusickii and A. cusickii var. flexilipes. In contrast, these latter two varieties are not reciprocally monophyletic, nor distinct from each other which matches the morphological variation that characterized A. cusickii var. flexilipes as only weakly differentiated from the type variety (Barneby, 1989).

Phenotypic variation found in Astragalus cusickii vars. packardiae and sterilis is unlikely to stem exclusively from edaphic conditions. Recently, 160 individuals of A. cusickii var. packardiae were successfully grown in a greenhouse at the Idaho Botanical Garden (A.

DeBolt, pers. comm.). These individuals were grown in a mixture of soil from their natural

19 habitat and commercial potting soil. Despite conditions that are clearly different from their native habitat, these plants appear to exhibit similar morphology to wild individuals, and do not show phenotypic characters similar to any of the other varieties. The recovery of both varieties as well-supported monophyletic groups indicates that these two varieties have evolved independently from different most recent common ancestors. The morphological traits that distinguish them are likely to have evolved either as adaptations to the extreme arid environments they occupy or the result of drift. The areas that these two varieties occupy are

isolated and relatively distant from each other, and the remainder of the populations of A. cusickii. It is likely that these populations were founded by one to few individuals and the character states that define the varieties may be the result of chance alone. Phenotypic plasticity is a possible explanation for the subtle phenotypic differences that exist between

Astragalus cusickii vars. cusickii and flexilipes. These two varieties exist in overlapping territory (Fig. 1), and are considered weakly differentiated (Barneby, 1989). However, see discussion of whole plastid sequencing results below.

4.1 Taxonomic Status of the Varieties of Astragalus cusickii

Reciprocal monophyly is often a first step toward identifying species boundaries, however in this case reciprocal monophyly of A. cusickii var. packardiae, A. cusickii var. sterilis, and A. cusickii vars. cusickii/flexilipes is not sufficient to resolve whether these represent a single species or three distinct species, mainly due to the poor resolution and support among the species from western North America (Fig. 5). The AU test failed to reject a monophyletic Astragalus cusickii (p = 0.911), but it also could not reject a monophyletic group containing both A. cusickii and A. whitneyi var. confusus (p = 0.746). This is not to say that A.

whitneyi var. confusus should also be merged as a variety of A. cusickii. Monophyly is also 20

expected as speciation occurs from a single common ancestor with diversification.

The genealogical sorting index (GSI) (Cummings et al., 2007) was also used to investigate the potential for a monophyletic Astragalus cusickii. Any GSI value above zero with a corresponding p-value < 0.05 implies some degree of genealogical clustering, with a value of one indicating complete coalescence for the sequences tested. No commonly used GSI value exists by which to accept a particular group of individuals as monophyletic, though

values as low as 0.218 have been considered significant genealogical clustering (Koopman and

Baum, 2010). Wang et al. (2014) considered GSI ≥ 0.85 to imply a “high degree of exclusive ancestry.” Our GSI results supported a monophyletic Astragalus cusickii with a value of 0.874

(Table 2). However, as with the AU tests, equal support was found for a monophyletic group containing both A. cusickii and A. whitneyi var. confusus (GSI 0.874) (Table 2). Here again, we recover monophyly, but likely as the result of speciation from a shared common ancestor. The

GSI provided the most robust support when the varieties of A. cusickii were split into three groups: A. cusickii vars. cusickii and flexilipes as a combined group, and A. cusickii vars. packardiae and sterilis each representing separate groups (GSI 0.989, 1.000, 1.000, respectively) (Table 2), consistent with the results of the traditional phylogenetic analyses.

The phylogenetic species concept defines species not just as monophyletic groups, but the smallest such groups that are distinguishable from other groups. The results of the AU tests and GSI demonstrate that the A. cusickii vars. packardiae and sterilis are at least as distinct from each other as they are from taxa that have long been considered separate species. These two varieties are morphologically distinct from each other and A. cusickii var. sterilis was originally described at the species level (Barneby, 1949). It was only with the publication of

21

The Intermountain Flora when Barneby (1989) moved A. sterilis to the rank of variety and described A. cusickii var. packardiae. In the discussion of the treatment for A. cusickii,

Barneby (1989) acknowledged that recent collections led him to change his view of A. sterilis as a distinct species defined by "..its rhizomatous, subterranean caudex (resembling that of A. ceramicus) which gave rise to colonies of solitary or paired, short, and densely branched stems" to a distinct ecotype differentiated from typical A. cusickii only by "uniformly very short leaflets, and by the bright mottling of the pod, a syndrome too weak to support specific status."

He also recognized that A. cusickii var. flexilipes was only weakly differentiated from the type variety and that A. cusickii var. packardiae was similar to A. cusickii var. flexilipes but was highly localized and differentiated by the loss of leaflets in the upper leaves and the narrowly and symmetrically ellipsoid fruit. Despite his views of the ranking of these taxa, Barneby

(1989) clearly recognized morphological differences of both A. cusickii vars. sterilis and packardiae. Therefore, following the concept of a monophyletic group that is diagnosably distinct from close relatives, we would advocate ranking both A. cusickii var. packardiae and A. cusickii var. sterilis at the status of species.

4.2 Poor Resolution among Species in Astragalus

Beyond Astragalus cusickii var. packardiae, and A. cusickii var. sterilis, there was a general lack of support for species-level relationships across the species tree, with many of the species grouped in polytomies (Fig. 5). Sanderson and Wojciechowski (1996) determined that

Astragalus has a similar diversification rate to closely related taxa (Oxytropis and seventeen other genera in Galegeae), yet Astragalus contains many more species than these other genera.

Therefore, another mechanism must explain the great diversity of Astragalus. The low level of phylogenetic resolution among species may reflect an adaptive radiation event that occurred

22 within Astragalus after colonization of the Intermountain West area of North America. Species colonizing a new habitat may rapidly diversify into multiple species and take advantage of available niches (Schluter, 2000). The high level of diversity of Astragalus found in the region

(Barneby, 1989) and the challenging nature of systematic studies of the genus (Scherson et al.,

2008) are consistent with an adaptive radiation event that occurred relatively recently, or perhaps is still in progress. The many poorly-resolved relationships found in the phylogenetic analyses may indicate that an insufficient amount of time has passed since the radiation for

complete coalescence to have occurred across the species in the region. Gene flow between some taxa may have been occurring until recently, or could still be occurring (see below). In this context, the strong support for a monophyletic A. cusickii var. packardiae and a monophyletic A. cusickii var. sterilis could be understood to reflect a greater degree of isolation relative to other Astragalus species in the region, allowing for the accumulation of mutations and, eventually, synapomorphic phenotypic traits.

4.3 Chloroplast Capture

One unexpected result was the discrepant phylogenetic placement of three individuals of A. cusickii var. packardiae between nrDNA ITS and trnS-trnG. The nrDNA ITS data resulted in a monophyletic variety (Fig. 3), but the chloroplast data, while mostly matching the nrDNA ITS data, placed three individuals outside of the clade that consisted of all other individuals of A. cusickii and A. whitneyi, and instead placed them closer to A. filipes (Fig. 4).

These three individuals were all from a single sub-population of A. cusickii var. packardiae.

Astragalus lentiginosus, A. filipes, A. eremeticus, and A. purshii can all be found in close proximity to A. cusickii var. packardiae, including co-occurring on the same outcrop in some places. Although other species were not found intermixed with A. cusickii var. packardiae it is

23 possible that they may have been present in the past. Although our data are not sufficiently robust to state that A. filipes is the source of chloroplast DNA for these individuals of A. cusickii var. packardiae, this species is relatively widespread and common in southwestern

Idaho and adjacent Oregon. Thus potential for gene flow between these is a validly potential hypothesis.

4.4 Relationships Based on Whole Plastid Genome Comparisons

Despite a multi-fold increase in the total number of base pairs examined, not all relationships among the eight accessions of Astragalus used in the analyses of the plastid genome data are strongly supported, at least based on Bayesian posterior probabilities.

Sampling was much lower for this portion of the study and only a single individual was used for most species. In sharp contrast to the phylogenetic analyses based on Sanger sequence data, whole chloroplast genomes did not resolve the relationship among A. cusickii var. cusickii and

A. cusickii var. flexilipes. The two accessions of A. cusickii var. flexilipes do not form a monophyletic group, nor is the close relationship between these varieties recovered as it was based on the four locus data analyzed (Fig. 5). This in part may stem from the fact that a combination of nuclear ribosomal, low copy nuclear and chloroplast DNA was used to evaluate relationships with the Sanger sequencing, but only chloroplast DNA was used for the these analyses. The trnS-trnG data, when analyzed apart from the other loci place the two accessions of A. cusickii var. flexilipes (JZ024, JZ027) used in the analyses of the plastid genome in a large polytomy with A. whitneyi and A. cusickii var. packardiae (Fig. 4). The third accession A. cusickii var. flexilipes that was not included in the whole plastid sequence analyses is in a clade with all other accessions of A. cusickii var. cusickii with BS = 86 (Fig. 4) based

only on this one cpDNA region. There is evidence for chloroplast capture in this group, and it 24

is possible that the accessions of A. cusickii var. flexilipes in the whole plastid sequence data

(Fig. 6) represent additional chloroplast capture events. Further analyses will need to include additional individuals from all varieties of A. cusickii as well as samples from other species and sampling more of the non-chloroplast data to be able to evaluate the role of chloroplast capture and incomplete lineage sorting in this group of species.

4.5 Conclusion

Based on the results of the phylogenetic analyses and the genealogical sorting index, we advocate elevation of both Astragalus cusickii var. packardiae, and A. cusickii var. sterilis, to species status. Both varieties were strongly supported as monophyletic groups in the MP, ML, and BI analyses (Fig. 5). Including loci from both nuclear ribosomal and low copy loci and chloroplast DNA, which are independently inherited, results in phylogenetic analyses that have greater power to resolve species trees (Corl and Ellegren, 2013). Combined with the outcome of the gene trees from the separate phylogenetic analyses displaying shared topology where support existed, these results give us confidence that the phylogenetic analyses represent true monophyletic groups.

Criteria for determination of species status and boundaries were determined a priori to align with those of the phylogenetic species concept. Because the varieties of Astragalus cusickii already exhibit morphological characters that can be used to distinguish them from each other, demonstration of the monophyly of any particular variety satisfies the criteria of the phylogenetic species concept for consideration of that variety as a separate species. The results of the various analyses employed in this study have demonstrated the monophyly, separately, of both A. cusickii var. packardiae and A. cusickii var. sterilis, thus showing merit for the

25

recognition of these two taxa as distinct species.

The addition of two species to our concept of Astragalus within the Intermountain West expands our understanding of the extreme diversity of the genus found in the region (Barneby,

1989). A few, but an expanding number of studies have used modern molecular-based techniques to explore the species boundaries among closely related taxa within Astragalus which are phenotypically similar and exist within close proximity to one another (Travis et al.,

1996; Lavin and Marriott, 1997; Kazempour Osaloo et al., 2003, 2005; Alexander et al., 2004;

Knaus et al., 2005; Scherson et al., 2008; Riahi et al., 2011; Bartha et al., 2013; Dizkirici et al.,

2014). Our appreciation of the tremendous diversity of Astragalus in the Intermountain West is largely informed by the taxonomic work done based on morphological analysis (Rydberg,

1929; Barneby, 1964, 1989; Isely, 1998), prior to the development of the field of molecular systematics. This understanding may or may not reflect an accurate estimate of the actual diversity found in the region. The results of this study demonstrate the possibility that the diversity of Astragalus as measured by described species may be underestimated.

There is an ongoing effort to reexamine previous taxonomic work conducted primarily on the basis of morphology using modern molecular techniques. Much of the taxonomic work previously done within Astragalus was performed by carefully examining and comparing morphology among taxa (e.g., Rydberg, 1929; Barneby, 1964, 1989; Podlech, 1982;

Isely, 1998). It is unclear if subtle morphological variation between populations is due to each population following its own unique evolutionary path, or simply due to environmental factors.

Molecular phylogenetic studies continue to discover cryptic species within what had previously been understood to be single species (Adjie et al., 2007; Heinrichs et al., 2010; Liao et al.,

2011; Carter, 2012; Dong et al., 2012; Carstens and Satler, 2013; Dauphin et al., 2014).

26

5. Acknowledgments

The authors would like to thank the US Fish and Wildlife Office in Boise Idaho and Boise State

University for financial support and Karen Colson for pointing us in the direction of this problem in the first place. We also thank Dr. Don Mansfield for assistance in field collections,

Ann DeBolt and Dr. Lynn Kinter for sharing unpublished results, and two anonymous reviewers and the editorial staff of Molecular Phylogenetics and Evolution.

Table 1. Regions investigated for this analysis

Region Citation Type Averag Matri PIC's

e Pair- x

Wise Lengt

Distanc h

e internal transcribed Wen and Nuclear 0.034 559 bp 87 bp spacer Zimmer, 1996 riboso

mal

27

external transcribed Baldwin and Nuclear 0.015 441 bp 15 bp spacer Markos, 1998 riboso

mal cyclic nucleotide- Choi et al., 2004, Nuclear 0.014 396 bp 33 bp gated channel 4 2006 trnS-G Shaw et al., 2005 plastid 0.006 562 bp 71 bp

28

Table 2. Genealogical sorting index scores and corresponding p-values. P-value < 0.05 results in rejection of null hypothesis that the defined monophyletic group is incorrect. Genealogical sorting index possible scores range from 0 to 1, with 1 indicating complete lineage sorting, and

0 indicating no lineage sorting.

GSI

Monophyletic Group Score P-value

A. cusickii var. cusickii 0.796 0.0001

A. cusickii var. flexilipes 0.442 0.001

A. cusickii var. packardiae 1.000 0.0001

A. cusickii var. sterilis 1.000 0.0001

A. cusickii var. cusickii + A. cusickii var. flexilipes 0.989 0.0001

28

A. cusickii all varieties 0.874 0.0001

A. cusickii all varieties + A. whitneyi var. confusus 0.874 0.0001

A. cusickii all varieties + A. solitarius 0.839 0.0001

29

Table 3. Models and parameters suggested by jModelTest.

Gene freq. freq. freq. freq. R(a) R(b) R(c) R(d) R(e) R(f) gamma

Region model -lnL K A C G T [AC] [AG] [AT] [CG] [CT] [GT] p-inv shape Kappa ti/tv nrDNA SYM +

ITS I + G 2538.8408 183 n/a n/a n/a n/a 1.1813 5.5134 2.4867 0.8975 3.7307 1.0000 0.3120 0.8120 n/a n/a

ETS TIM3 761.6996 50 0.3675 0.2880 0.2221 0.1223 0.2299 1.0659 1.0000 0.2299 0.2258 1.0000 n/a n/a n/a n/a

CNGC4 codon pos. 1 TPM3uf 342.7643 181 0.3260 0.1067 0.1962 0.3711 2.3301 2.5549 1.0000 2.3301 2.5549 1.0000 n/a n/a n/a n/a

CNGC4 codon pos. 2 TPM3uf 307.6324 181 0.3547 0.1727 0.2171 0.2554 0.0000 2.6512 1.0000 0.0000 2.6512 1.0000 n/a n/a n/a n/a

CNGC4 codon pos. 3 HKY 283.1729 180 0.3734 0.2066 0.1496 0.2703 n/a n/a n/a n/a n/a n/a n/a n/a 8.100429 3.6279

TIM2 + trnS-G I 838.2825 91 0.3907 0.1318 0.1320 0.3455 0.2564 0.1895 0.2564 1.0000 1.8973 1.0000 0.9200 n/a n/a n/a

30

Figure Legends

Figure 1. Map of the approximate ranges of the varieties of Astragalus cusickii, focused on an area spanning the borders between the states of Idaho, Oregon, and Washington of the United

States.

Figure 2. A. Astragalus cusickii var. cusickii photographed on 27 June 2013 on a steep, gravelly slope in Hells Canyon, Adams County, Idaho. Numerous inflated papery pods are evident. B. Conspicuous oblique, half-ellipsoid, papery pods on Astragalus cusickii var. flexilipes, photographed on 27 June 2013 on a steep, sandy slope near the top of a hill in Hells

Canyon, Adams County, Idaho. C. Conspicuous inflated, brightly-mottled, papery pods on

Astragalus cusickii var. sterilis, photographed on 11 June 2013 near Birch Creek, Malheur

County, Oregon. D. Astragalus cusickii var. packardiae photographed on 30 May 2014 on a hillside in Payette County, Idaho. Numerous slender pods are evident.

30

Figure 3. Strict consensus tree from maximum parsimony analysis of nrDNA ITS, with bootstrap values above branches. Varieties of Astragalus cusickii are highlighted in color: blue

= A. cusickii var. packardiae, purple = A. cusickii var. sterilis, red = A. cusickii var. cusickii, green = A. cusickii var. flexilipes. A bracket to the left of the tree marks the three individuals that are discrepant between nrDNA ITS and trnS-trnG. Seventy equally most-parsimonious trees were found (length = 381, CI = 0.750, RI = 0.806).

31

Figure 4. Strict consensus tree from maximum parsimony analysis of trnS-G, with bootstrap values above branches. Varieties of Astragalus cusickii are highlighted in color: blue = A. cusickii var. packardiae, purple = A. cusickii var. sterilis, red = A. cusickii var. cusickii, green =

A. cusickii var. flexilipes. A bracket to the left of the tree marks the three individuals that are discrepant between nrDNA ITS and trnS-trnG. Three equally most-parsimonious trees were found (length = 88, CI = 0.977, RI = 0.9910).

Figure 5A. Majority-rule tree generated from Bayesian inference analysis that is congruent with maximum parsimony and maximum likelihood analyses with sequence data from 89 individuals with nrDNA ITS, 40 of them also with sequences from trnS-G, 23 with ETS sequences, and 55 with CNGC4. Varieties of Astragalus cusickii are highlighted in color: blue

= A. cusickii var. packardiae, purple = A. cusickii var. sterilis, red = A. cusickii var. cusickii, green = A. cusickii var. flexilipes. Values above branches correspond to maximum parsimony bootstrap support, maximum likelihood bootstrap support, and Bayesian inference posterior probability, respectively. The triangle represents 17 individual Astragalus cusickii var. packardiae collapsed to save space.

Figure 5B. See caption for Fig. 5A. 31

Figure 6. Maximum likelihood and Bayesian Inference analysis of whole chloroplast genome sequences for a subset of specimens sampled with Sanger sequencing. Varieties of Astragalus cusickii are highlighted in color: blue = A. cusickii var. packardiae, purple = A. cusickii var. sterilis, red = A. cusickii var. cusickii, green = A. cusickii var. flexilipes.

32

Figure S1. Maximum parsimony analysis of Matrix 2 with sequence data from 55 individuals, all with nrDNA ITS, 40 of them with sequences from trnS-G, 23 with sequences from ETS, and

55 sequences from CNGC4 (Appendix 1).

Figure S2. Maximum parsimony analysis of Matrix 3 with sequence data from 23 individuals, with sequences from all four regions (Appendix 1).

32

33

LITERATURE CITED

Abbasi, M., Goodwin, S.B., Scholler, M., 2005 , phylogeny, and distribution of

Puccinia graminis, the black stem rust: New insights based on rDNA sequence data.

Mycosci. 46, 241-247

Adjie, B., Masuyama, S., Ishikawa, H., Watano, Y., 2007. Independent origins of tetraploid

cryptic species in the fern Ceratopteris thalictroides. J. Plant Res. 120, 129-138.

Adl, S. M., Simpson, A. G. B., Farmer, M. A., Andersen, R. A., Anderson, O. R., Barta, J. R.,

Bowser, S. S., Brugerolle, G., Fensome, R. A., Fredericq, S., James, T. Y., Karpov, S.,

Kugrens, P., Krug, J., Lane, C. E., Lewis, L. A., Lodge, J., Lynn, D. H., Mann, D. G.,

Mccourt, R. M., Mendoza, L., Moestrup, Ø., Mozley-Standridge, S. E., Nerad, T. A.,

Shearer, C. A., Smirnov, A. V., Speigel, F. W. Taylor, M. F. J. R., 2005. The new

higher jevel classification of Eukaryotes with emphasis on the taxonomy of Protists. J.

Euk. Microbiol. 52, 399–451.

Alexander, J.A., Liston, A., Popovich, S.J., 2004. Genetic diversity of the narrow endemic

Astragalus oniciformis (Fabaceae). Amer. J. Bot. 91, 2004-2012. 33

Altekar, G., Dwarkadas, S., Huelsenbeck, J.P., Ronquist, F., 2004. Parallel Metropolis-coupled

Markov chain Monte Carlo for Bayesian phylogenetic inference. Bioinform. 20, 407-

415.

Aronesty, E. 2013. TOBioiJ : "Comparison of Sequencing Utility Programs"

DOI:10.2174/1875036201307010001

Baldwin, B.G., 1992. Phylogenetic utility of the internal transcribed spacers of nuclear

34

ribosomal DNA in plants: an example from the Compositae. Mol. Phylogenet. Evol. 1,

3-16.

Baldwin, B.G., Markos, S., 1998. Phylogenetic utility of the external transcribed spacer (ETS)

of 18s-26s rDNA: congruence of ETS and ITS trees of Calycadenia (Compositae). Mol.

Phylogenet. Evol. 10, 449-463.

Barneby, R.C., 1949. Pugillus astragalorum XI: Two new species. Leafl. W. Bot. 5, 193-197.

Barneby, R.C., 1964. Atlas of North American Astragalus. Mem. New York Bot. Gard. 13, 1-

1188.

Barneby, R. C., 1989. Astragalus. Pp. 39-176 in Intermountain Flora, , vol. 3, part B,

eds. A. Cronquist, A. Holmgren, N. Holmgren, J. Reveal, and P. Holmgren. Bronx:

New York Botanical Garden.

Bartha, L., Dragos, N., Molnár V., A., Sramkó, G., 2013. Molecular evidence for reticulate

speciation in Astragalus (Fabaceae) as revealed by a case study from sect. Dissitiflori.

Bot. 91, 702-714.

Belton, G. S., van Reine, W. F. P., Huisman, J. M., Draisma, S. G. A., D. Gurgel, C. F., 2014.

Resolving phenotypic plasticity and species designation in the morphologically

challenging Caulerpa racemosa–peltata complex (Chlorophyta, Caulerpaceae). J. Phyc. 34

50, 32–54. doi: 10.1111/jpy.12132

Bronikowski, A.M., Arnold, S.J., 2001. Cytochrome b phylogeny does not match subspecific

classification in the western terrestrial garter snake, Thamnophis elegans. Copeia 2001,

508-513.

Bunge, A., 1868. Generis Astragali species Gerontogeae. Mem. Acad. Imp. Sci. St. Petersburg

11, 1-140.

35

Bunge, A., 1869. Generis Astragali species Gerontogeae. Mem. Acad. Imp. Sci. St. Petersburg

15, 1-254.

Carlson, K.M., Mansfield, D.H., Smith, J.F., 2011. A new species in the Lomatium

foeniculaceum (Apiaceae) clade revealed through combined morphometric and

phylogenetic analyses. Syst. Bot. 36, 495-507.

Carrillo-Reyes, P., Sosa, V., Mort, M.E., 2008. Thompsonella and the “Echeveria group”

(Crassulaceae): phylogenetic relationships based on molecular and morphological

characters. Taxon 57, 863-874.

Carstens, B.C., Satler, J.D., 2013. The carnivorous plant described as Sarracenia alata contains

two cryptic species. Biol. J. Linn. Soc. 109, 737-746.

Carter, B.E., 2012. Species delimitation and cryptic diversity in the moss genus Scleropodium

(Brachytheciaceae). Mol. Phylogenet. Evol. 63, 891-903.

Choi, H.K., Kim, D., Uhm, T., Limpens, E., Lim, H., Mun, J.H., Kalo, P., Penmetsa, R.V.,

Seres, A., Kulikova, O., Roe, B.A., Bisseling, T., Kiss, G.B., Cook. D.R., 2004. A

sequence-based genetic map of Medicago truncatula and comparison of marker

colinearity with M. sativa. Genet. 166, 1463-1502.

Choi, H.K., Luckow, M.A., Doyle, J., Cook. D.R., 2006. Development of nuclear gene-derived 35

molecular markers linked to legume genetic maps. Mol. Genet. Genomics 276, 56-70.

Clark, A.M., Moler, P.E., Possardt, E.E., Savitzky, A.H., Brown, W.S., Bowen, B.W., 2003.

Phylogeography of the timber rattlesnake (Crotalus horridus) based on mtDNA

sequences. J. Herp. 37, 145-154

Corl, A., Ellegren, H., 2013. Sampling strategies for species trees: The effects on phylogenetic

inference of the number of genes, number of individuals, and whether loci are

36

mitochondrial, sex-linked, or autosomal. Mol. Phylogenet. Evol. 67, 358-366.

Cummings, M.P., Neel, M.C., Shaw, K.L., 2007. A genealogical approach to quantifying

lineage divergence. Evol. 62, 2411-2422.

Dantrakool, A., Somboon, P., Hashimoto, T., Saito-Ito, A., 2004. Identification of a new type

of Babesia species in wild rats (Bandicota indica) in Chiang Mai province, Thailand. J.

Clin. Microbiol. 42, 850-854.

Darriba, D., Taboada, G.L., Doallo, R., Posada, D., 2012. jModelTest 2: More models, new

heuristics and parallel computing. Nat. Methods 9, 772.

Dauphin, B., Vieu, J., Grant, J.R., 2014. Molecular phylogenetics support widespread cryptic

species in moonworts (Botrychium s.s., Ophioglossaceae). Amer. J. Bot. 101, 128-140.

DeBiasse, M.B., Hellberg, M.E., 2015. Discordance between morphological and molecular

species boundaries among Caribbean species of the reef sponge Callyspongia. Ecol.

Evol. 5, 663–675.

Dizkirici, A., Ekici, M., Kaya, Z., 2013. Molecular phylogenetics of Astragalus L. sections

from Turkey with New World Astragalus species using nrDNA. Pl. Syst. Evol. 300,

163-175.

Dong, S., Schäfer-Verwimp, A., Meinecke, P., Feldberg, K., Bombosch, A., Pócs, T., Schmidt, 36

A.R., Reitner, J., Schneider, H., Heinrichs, J., 2012. Tramps, narrow endemics and

morphologically cryptic species in the epiphyllous liverwort Diplasiolejeunea. Mol.

Phylogenet. Evol. 65, 582-594.

Drummond, A.J., Suchard, M.A., Xie, D., Rambaut, A., 2012. Bayesian phylogenetics with

BEAUti and the BEAST 1.7 Mol. Biol. .Evol. 29, 1969-1973.

Edgar, R.C., 2004. MUSCLE: Multiple sequence alignment with high accuracy and high

throughput. Nucl. Acids Res. 32, 1792-1797 doi:10.1093/nar/gkh340.

37

Ernst, R., Agyei, A.C., Rödel, M., 2008. A new giant species of Arthroleptis (Amphibia:

Anura: Arthroleptidae) from the Krokosua Hills Forest Reserve, south-western Ghana.

Zootaxa 1697, 58-68.

Felsenstein, J., 1985. Confidence limits on phylogenies: An approach using the bootstrap. Evol.

39, 783–791.

Frodin, D.G., 2004. History and concepts of big plant genera. Taxon 53, 753-776.

Gao, S., Gong, J., Lynn, D., Lin, X., Song, W., 2009. An updated phylogeny of oligotrich and

choreotrich ciliates (Protozoa, Ciliophora, Spirotrichea) with representative taxa

collected from Chinese coastal waters. Syst. Biodivers. 7, 235-242.

George, E.E., Mansfield, D.H., Smith, J.F., Hartman, R.L., Downie, S.L., Hinchliff, C.E., 2014.

Phylogenetic analysis reveals multiple cases of morphological parallelism and

taxonomic polyphyly in Lomatium (Apiaceae). Syst. Bot. 39, 662-675.

Geyer, C.J., 1991. Markov chain Monte Carlo maximum likelihood, in: Keramidas, E.M. (Ed.),

Computing Science and Statistics: Proceedings of the 23rd Symposium on the Interface.

Fairfax Station: Interface Foundation, pp. 156-163.

Gill, E.E., Fast, N.M., 2006. Assessing the microsporidia-fungi relationship: combined

phylogenetic analysis of eight genes. Gene 375, 103-109. 37

Goloboff, P.A., Carpenter, J.M., Arias, J.S., Esquivel, D.R.M., 2008. Weighting against

homoplasy improves phylogenetic analysis of morphological data sets. Cladistics 24, 1-

16.

Gray, A., 1878. Contributions to the botany of North America. Proc. Amer. Acad. Arts. 13,

361-374.

38

Guindon S., Dufayard J.F., Lefort V., Anisimova M., Hordijk W., Gascuel O., 2010. New

algorithms and methods to estimate maximum-likelihood phylogenies: Assessing the

performance of PhyML 3.0. Syst. Biol. 59, 307-21.

Gurevitch, J., Scheiner, S., Fox, G., 2002. The ecology of plants, second ed. Sinauer

Associates, Sunderland.

Hastings, W.K., 1970. Monte Carlo sampling methods using Markov chains and their

applications. Biom. 57, 97-109.

Heinrichs, J., Hentschel, J., Bombosch, A., Fiebig, A., Reise, J., Edelmann, M., Kreier, H.P.,

Schäfer-Verwimp, A., Caspari, S., Schmidt, A.R., Zhu, R.L., von Konrat, M., Shaw, B.,

Shaw, A.J., 2010. Mol. Phylogenet. Evol. 56, 1105-1114.

Heiss, A.A., Keeling, P.J., 2006. The phylogenetic position of the Oxymonad Saccinobaculus

based on SSU rRNA. Protist 157, 335-344.

Helmkampf, M., Bruchhaus, I., Hausdorf, B., 2008. Multigene analysis of lophophorate and

chaetognath phylogenetic relationships. Mol. Phlyogenet. Evol. 46, 206-214.

Hibbett , D.S., Binder, M., Bischoff, J.F., Blackwell, M., Cannon, P.F., Eriksson, O.E.,

Huhndorf, S., James, T., Kirk, P.M., Lücking, R., Lumbsch, H.T., Lutzoni, F., Matheny,

P.B., McLaughlin, D.J., Powell, M.J., Redhead, S., Schoch, C.L., Spatafora, J.W., 38

Stalpers, J.A., Vilgalys, R., Aime, M.C., Aptroot, A., Bauer, R., Begerow, D., Benny,

G.L., Castlebury, L.A., Crous. P.W., 2007 A higher-level phylogenetic classification of

the Fungi. Mycol. Res. 111, 509–547

Hillis, D., 1987. Molecular versus morphological approaches to systematics. Annu. Rev. Ecol.

Syst. 18, 23-42.

Huelsenbeck, J.P., Ronquist, F., 2001. MRBAYES: Bayesian inference of phylogeny.

39

Bioinform. 17, 754-755. lsely, D. 1998. Native and naturalized Leguminosae (Fabaceae) of the United States. M. L.

Bean Life Science Museum, Brigham Young University, Provo, Utah.

Ishiwata, K., Sasaki, G., Ogawa, J., Miyata, T., Su, Z.H., 2011. Phylogenetic relationships

among insect orders based on three nuclear protein-coding gene sequences. Mol.

Phylogenet. Evol. 58, 169-180.

Jaramillo, M.A., Callejas, R., Davidson, C., Smith, J.F., Stevens, A.C., and Tepe, E., 2008. A

phylogeny of the tropical genus Piper (Piperaceae) using ITS and the chloroplast intron

psbJ-petA. Syst. Bot. 33, 647-660.

Kazempour Osaloo, S., Maasoumi, A.A., Murakami, N., 2003. Molecular systematics of the

genus Astragalus L. (Fabaceae):Phylogenetic analyses of nuclear ribosomal DNA

internal transcribed spacers and chloroplast gene ndhF sequences. Pl. Syst. Evol. 242,

1-32.

Kazempour Osaloo, S., Maasoumi, A.A., Murakami, N., 2005. Molecular systematics of the

Old World Astragalus (Fabaceae) as inferred from nrDNA ITS sequence data. Brittonia

57, 367-381.

Keith, R., Hedin, M., 2012. Extreme mitochondrial population subdivision in southern 39

Appalachian paleoendemic spiders (Araneae: Hypochilidae: Hypochilus), with

implications for species delimitation. J. Arachnol. 40, 167-181.

Knaus, B. J., Cronn, R.C., Liston, A., 2005. Genetic characterization of three varieties of

Astragalus lentiginosus (Fabaceae). Brittonia 57, 334-344.

Koopman, M.M., Baum, D.A., 2010. Isolating nuclear genes and identifying lineages without

monophyly: An example of closely related species from southern Madagascar. Int. J.

40

Plant Sci. 171, 761-771.

Kruckeberg, A.R., Rabinowitz, D., 1985. Biological aspects of endemism in higher plants.

Annu. Rev. Ecol. Syst. 16, 447-479.

Kubatko, L.S., Gibbs, H.L., Bloomquist, E.W., 2011. Inferring species-level phylogenies and

taxonomic distinctiveness using multilocus data in Sistrurus rattlesnakes. Syst. Biol. 60,

393-409.

Kuo, C.H., Wares, J.P., Kissinger, J.C., 2008. The Apicomplexan whole-genome phylogeny:

An analysis of incongruence among gene trees. Mol. Biol. Evol. 25, 2689-2698.

Lanfear, R., Calcott, B., Ho, S.Y.W., Guindon, S., 2012. PartitionFinder: Combined selection

of partitioning schemes and substitution models for phylogenetic analyses. Mol. Biol.

Evol. 29, 1695-1701.

Lang, B.F., O’Kelly, C., Nerad, T., Gray, M.W., Burger, G., 2002. The closest unicellular

relatives of animals. Curr. Biol. 12, 1773-1778.

Langmead, B., Salzberg, S. 2012. Fast gapped-read alignment with Bowtie 2. Nature Methods.

9, 357-359.

Lavin, M., Marriott, H., 1997. Astragalus molybdenus s.l. (Leguminosae): Higher taxonomic

relationships and identity of constituent species. Syst. Bot. 22, 199-217. 40

Lee, S., Wen, J., 2001. A phylogenetic analysis of Prunus and the Amygdaloideae (Rosaceae)

using ITS sequences of nuclear ribosomal DNA. Amer. J. Bot. 88, 150-160.

Levsen, N.D., Tiffin, P., Olson, M.S., 2012. Pleistocene speciation in the genus Populus

(Saliaceae). Syst. Biol. 61, 401-412.

Lewis, G.P., Schrire, B.D., Mackinder, B.A., Lock, M., 2005. Legumes of the world. Royal

Botanic Gardens, Kew.

41

Li, H., Handsaker, B., Wysoker, A., Fennell, T., Ruan, J., Homer, N., Marth, G., Abecasis, G.,

Durbin, R., 1000 Genome Project Data Processing Subgroup, 2009. The sequence

alignment/map (SAM) format and SAMtools. Bioinf. 25, 2078-9. [PMID: 19505943]

Liao, Y.Y., Yang, X.Y., Motley, T.J., Chen, J.M., Wang, Q.F., 2011. Phylogeographic analysis

reveals two cryptic species of the endangered fern Ceratopteris thalictroides (L.)

Brongn. (Parkeriaceae) in China. Conserv. Genet. 12, 1357-1365.

Lock, J.M., Schrire, B.D., 2005. Galegeae, in: Lewis, G., Schrire, B., Mackinder, B., Lock, M.

(Eds.), Legumes of the World. Richmond, Surrey, pp. 475-487.

Luo, R., Liu, B., Xie, Y., Li, Z., Huang, W., Yuan, J., He, g., Chen, Y., Pan, Q., Liu, Y., Tang,

J., Wu, G., Zhang, H., shi, Y., Liu, Y., Yu, C., Wang, B., Lu, Y., Han, C., Cheung,

D.W., Yiu, S.-M., Peng, S., Xiaoqian, Z., Liu, G., Liao, X., Li, Y., Yang, Y., Yang, H.,

Wang, J., Lam, T.-W., Wang, J., 2012. SOAPdenovo2: An empirically improved

memory-efficient short-read de novo assembler. GigaScience 1, 18.

Mabberly, D.J., 2008. Mabberley's plant-book: A portable dictionary of plants, their

classifications, and uses, third ed. Camb. Univ. Press, Cambridge.

Mancuso, M., 1999. The status of Astragalus cusickii var. packardiae (Packard’s milkvetch).

Idaho Department of Fish and Game Natural Resource Policy Bureau. 41

Mancuso, M. 2016. Monitoring Packard’s milkvetch (Astragalus cusickii var. packardiae) in

southwestern Idaho, 2015 results. Report prepared for the Idaho Bureau of Land

Management, Boise ID. 47 pp plus appendices.

Mayr, E., 1969. Principles of systematic zoology, McGraw-Hill, New York.

Metropolis, N., Rosenbluth, A.W., Rosenbluth, M.N., Teller, A.H., Teller, E., 1953. Equations

of state calculations by fast computing machines. J. Chem. Phys. 21, 1087-1091.

42

Miller, M.A., Pfeiffer, W., Schwartz, T., 2010. Creating the CIPRES science gateway for

inference of large phylogenetic trees. Proceedings of the Gateway Computing

Environments Workshop, New Orleans, LA, 14 Nov 2010, pp 1–8.

Muggia, L., Pérez-Ortega, S., Fryday, A., Spribille, T., Grube, M., 2014. Global assessment of

genetic variation and phenotypic plasticity in the lichen-forming species Tephromela

atra. Fungal Div. 64, 233-251

Müller, J., Müller, K., Neinhuis, C., Quandt, D., 2010. PhyDE v. 0.9971, http://www.phyde.de/

Navarro, F.B., Suárez-Santiago, V.N., Blanca, G., 2004. A new species of Haplophyllum A.

Juss. (Rutaceae) from the Iberian peninsula: evidence from morphological, karyological

and molecular analyses. Ann. Bot. 94, 571-582.

Nylander, J.A.A., Wilgenbusch J.C., Warren D.L., Swofford D.L., 2007. AWTY: A system for

graphical exploration of MCMC convergence in Bayesian phylogenetics. Bioinform.

24, 581-583.

Oliveira, L.O., Huck, R.B., Gitzendanner, M.A., Judd, W.S., Soltis, D.E., Soltis, P.S., 2007.

Molecular phylogeny, biogeography, and systematics of Dicerandra (Lamiaceae), a

genus endemic to the southeastern United States. Amer. J. Bot. 94, 1017-1027.

Pérez-Gutiérrez, M.A., Romero-García, A.T., Fernández, M.C., Blanca, G., Salinas-Bonillo, 42

M.J., Suárez-Santiago, V.N., 2015. Evolutionary history of fumitories (subfamily

Fumarioideae, Papaveraceae): an old story shaped by the main geological and climatic

events in the northern hemisphere. Mol. Phylogenet. Evol. 88, 75-92.

Podlech, D. 1982. Neue aspekte zur evolution und glie- derung der gattung Astragalus L.

Mitteilungen der Botanischen Staatssammlung Munchen 18: 359-378.

Podlech, D., and Zarre, S. 2013. A taxonomic revision of the genus Astragalus L.

43

(Leguminosae) in the Old World. Naturhistorisches Museum, Wien, Austria, pp. 1-

2440.

Polhill, R.M., 1981. Galegeae, in: Polhill, R.M., Raven, P.H. (Eds.), Advances in legume

systematics, Part 1. Royal Botanic Gardens, Kew, pp. 357-363. de Queiroz, K., 2003. A unified concept of species and its consequences for the future of

taxonomy. Pp. 196-215 in Biodiversity, past, present and future and the future of

taxonomy. San Francisco, California Academy of Sciences.

Rambaut, A., 2006. FigTree v1.4.2, http://tree.bio.ed.ac.uk/

Rambaut, A., Suchard, M.A., Xie, D., Drummond, A.J., 2014. Tracer v1.6,

http://beast.bio.ed.ac.uk/Tracer

Riahi, M., Zarre, S., Maassoumi, A.A., Osaloo, S.K., Wojciechowski, M.F., 2011. Toward a

phylogeny for Astragalus section Caprini (Fabaceae) and its allies based on nuclear and

plastid DNA sequences. Plant Syst. Evol. 293, 119-133.

Ronquist, F., Huelsenbeck, J.P., 2003. MRBAYES 3: Bayesian phylogenetic inference under

mixed models. Bioinform. 19, 1572-1574.

Rydberg, P. A. 1929. Astragalanae. North American Flora 24: 251-462.

Sanderson, M.J., 1991. Phylogenetic relationships within North American Astragalus L. 43

(Fabaceae). Syst. Bot. 16, 414-430.

Sanderson, M.J., Doyle, J.J., 1993. Phylogenetic-relationships in North American Astragalus

(Fabaceae) based on chloroplast DNA restriction site variation. Syst. Bot. 18, 395-408.

Sanderson, M.J., Wojcichowski, M.F., 1996. Diversification rates in a temperate legume clade:

Are there "so many species" of Astragalus (Fabaceae). Amer. J. Bot. 83, 1488-1502.

Scherson, R.A., Choi, H., Cook, D.R., Sanderson, M.J., 2005. Phylogenetics of New World

44

Astragalus: Screening of novel nuclear loci for the reconstruction of phylogenies at low

taxonomic levels. Brittonia 57, 354–366.

Scherson, R.A., Vidal, R., Sanderson, M.J., 2008. Phylogeny, biogeography, and rates of

diversification of new world Astragalus (Leguminosae) with an emphasis on South

American radiations. Amer. J. Bot. 95, 1030-1039.

Schluter, D., 2000. The ecology of adaptive radiation. Oxf. Univ. Press, New York.

Shannon, A.J., Browne, J.A., Boyd, J., Fitzpatrick, D.A., Burnell, A.M., 2005. The

anhydrobiotic potential and molecular phylogenetics of species and strains of

Panagrolaimus (Nematoda, Panagrolaimidae). J. Exp. Biol. 208, 2433-2445.

Shaw, J., Lickey, E.B., Beck, J.T., Farmer, S.B., Liu, W., Miller, J., Siripun, K.C., Winder,

C.T., Schilling, E.E., Small, R.L., 2005. The tortoise ad the hare II: relative utility of 21

noncoding chloroplast DNA sequences for phylogenetic analysis. Amer. J. Bot. 92, 142-

166.

Shimodaira, H., Hasegawa, M., 2001. CONSEL: For assessing the confidence of phylogenetic

tree selection. Bioinform.17, 1246-1247.

Shimodaira, H., 2002. An approximately unbiased test of phylogenetic tree selection. Syst.

Biol. 51, 492-508. 44

Smith, J.F., Wolfram, J.C., Brown, K.D., Carroll, C.L., Denton, D.S., 1997. Tribal relationships

in the Gesneriaceae: evidence from DNA sequences of the chloroplast gene ndhF. Ann.

Mo. Bot. Gard. 84, 50-66.

Smith, J.F., Stevens, A.C., Tepe, E.J., Davidson, C., 2008. Placing the origin of two species-

rich genera in the late Cretaceous with later species divergence in the Tertiary: A

phylogenetic, biogeographic and molecular dating analysis. Pl. Syst. Evol. 275, 9-30.

45

Smith, J.F., Ooi, M.T., Schulte, L., Amaya-Marquez, M., Clark, J.L., 2013. Searching for

monophyly in the subgeneric classification systems of Columnea (Gesneriaceae).

Selbyana 31, 126-142.

Soltis, D.E., Soltis, P.S., Endress, P.K., Chase, M.W., 2005. Phylogeny and evolution of

angiosperms. Sinauer Assoc. Inc., Sunderland, MA.

Spalik, K., Downie, S.R., 2006. The evolutionary history of Sium sensu lato (Apiaceae):

Dispersal, vicariance, and domestication as inferred from ITS rDNA phylogeny. Amer.

J. Bot 93, 747-761.

Spalik, K., Downie, S.R., 2007. Intercontinental disjunctions in Cryptotaenia (Apiaceae,

Oenantheae): An appraisal using molecular data. J. Biogeogr. 34, 2039-2054.

Stepansky, A., Kovalski, I., Perl-Treves, R., 1999. Intraspecific classification of melons

(Cucumis melo L.) in view of their phenotypic and molecular variation. Pl. Syst. Evol.

217, 313-332.

Struck, T.H., Schult, N., Kusen, T., Hickman, E., Bleidorn, C., McHugh, D., Halanych, K.M.,

2007. Annelid phylogeny and the status of Sipuncula and Echiura. BMC Evol. Biol. 7,

61-73.

Swofford, D.E., 2002. PAUP*. Phylogenetic analysis using parsimony (* and other methods). 45

Sinauer Associates, Sunderland MA.

Sultan, S.E., 2000. Phenotypic plasticity for plant development, function and life history.

Trends Pl. Sci. 5, 537–542.

Taubert, P., 1894. Leguminosae, in: Engler, A., Prantl, K. (Eds.), Die Naturlichen

Pflanzenfamilien. Vol. III. Verlag von W. Engelmann, Leipzig, pp. 70-385.

Tomasello, S., Álvarez, I., Vargas, P., Oberprieler, C., 2014. Is the extremely rare Iberian

46

endemic plant species Castrilanthemum debeauxii (Compositae, Anthemideae) a ‘living

fossil’? Evidence from a multi-locus species tree reconstruction. Mol. Phylogen. Evol.

82, 118–13.

Travis, S. E., Maschinski, J., Keim, P., 1996. An analysis of genetic variation in Astragalus

cremnophylax var. cremnophylax a critically endangered plant, using AFLP markers.

Mol. Ecol. 5, 735-745.

Valladares, F., Gianoli, E., Gómez, J. M., 2007. Ecological limits to plant phenotypic plasticity.

New Phytol. 176, 749–763. doi:10.1111/j.1469-8137.2007.02275.x

Wang, X., Zang, R., Yin, Z., Kang, Z., Huang, L., 2014. Delimiting cryptic pathogen species

causing apple Valsa canker with multilocus data. Ecol. Evol. 4, 1369-1380.

Watrous, K.M., Cane, J.H., 2011. Breeding biology of the Threadstalk milkvetch, Astragalus

filipes (Fabaceae), with a review of the genus. Am. Midl. Nat. 165, 225-240.

Wen, J., Zimmer, E., 1996. Phylogeny and biogeography of Panax L. (the ginseng genus

Araliaceae): inferences from ITS sequences of nuclear ribosomal DNA. Mol.

Phylogenet. Evol. 6, 167-177.

Whitcher, I.N., Wen, J., 2001. Phylogeny and biogeography of Corylus (Betulaceae):

Inferences from ITS sequences. Syst. Bot. 26, 283-298. 46

Wojciechowski, M.F., 2005. Astragalus (Fabaceae): A molecular phylogenetic perspective.

Brittonia 57, 382-396.

Wojciechowski, M.F., Sanderson, M.J., Baldwin, B.G., Donoghue, M.J., 1993. Monophyly of

aneuploid Astragalus (Fabaceae) – evidence from nuclear ribosomal DNA internal

transcribed spacer sequences. Amer. J. Bot. 80, 711-722.

Wojciechowski, M.F., Sanderson, M.J., Hu, J.M., 1999. Evidence on the monophyly of

47

Astragalus (Fabaceae) and its major subgroups based on nuclear ribosomal DNA ITS

and chloroplast DNA trnL intron data. Syst. Bot. 24, 409-437.

Xiang, C.L., Gitzendanner, M.A., Soltis, D.E., Peng, H., Lei, L.G., 2012. Phylogenetic

placement of the enigmatic and critically endangered genus Saniculiphyllum

(Saxifragaceae) inferred from combined analysis of plastid and nuclear DNA

sequences. Mol. Phylogenet. Evol. 64, 357-367.

Yu, G., Rao, D., Yang, J., Zhang, M., 2008. Phylogenetic relationships among Rhacophorinae

(Rhacophoridae, Anura, Amphibia), with an emphasis on the Chinese species. Zool. J.

Linn. Soc. 153, 733-749.

Zhang, M., Fritsch, P.W., Cruz, B.C., 2009. Phylogeny of Caragana (Fabaceae) based on DNA

sequence data from rbcL, trnS-trnG, and ITS. Mol. Phylogenet. Evol. 50, 547-559.

Zhang, Q., Miao, M., Strüder-Kypke, M.C., Al-Rasheid, K.A.S., Al-Farraj, S.A., Song, W.,

2011. Molecular evolution of Cinetochilum and Sathrophilus (Protozoa, Ciliophora,

Oligohymenophorea), two genera of ciliates with morphological affinities to

scuticociliates. Zoologica Scripta 40, 317-325.

Zhou, J., Gong, X., Downie, S.R., Peng, H., 2009. Towards a more robust molecular phylogeny

of Chinese Apiaceae subfamily Apioideae: additional evidence from nrDNA ITS and 47

cpDNA intron (rpl16 and rps16) sequences. Mol. Phylogenet. Evol. 53, 56-68.

APPENDIX A. Voucher and GenBank information. All individuals were used in Matrix 1.

48

Individuals marked with a superscript 2 were used in Matrix 2 and individuals marked with a 3 were used in Matrix 3. NA indicates the sequence was not available, either from GenBank or did not amplify.

48

49

Individu Voucher Collecte Internal al ID ID d From Transcr

ibed

Spacer

GenBan

k

Accessio

n (two

labels

indicate

GenBan

k

sequence

split into

nrDNA

Gated Channel 4 GenBank Accession 4 Channel Gated GenBank

-

ITS1 and 49

GenBank Accession

nrDNA trnG

ITS2) -

Voucher Voucher Location Spacer GenBank External Transcribed Accession trnS Cyclic Nucleotide A. Wojciec AF1216 NA NA NA adsurge howski 74 ns Pall. and

Sanders

50

on 267

A. Sanders AF1217 NA NA NA allochro on 953 07 us A.

Gray

A. not HQ6133 NA NA NA alpinus vouchere 80

L. d

A. Scherson EU2829 NA NA NA amatus 106 76

Clos.

A. not U50492, NA NA NA america vouchere U50493 nus M.E. d

Jones

A. Sanders AF1216 NA NA NA

arizonic on 968 90 50

us A.

Gray

A. Scherson EU2829 NA NA DQ1072 arnottia 100 75 27 nus

Gillies2

51

A. Sanders AF1217 NA NA DQ1072 asclepia on 996 25 35 doides

M.E.

Jones2

A. Sanders AF1217 NA NA NA asymmet on 981 10 ricus E.

Sheld.

A. Yamashi AB0519 NA NA NA atropilos ta et al. 39 ulus 1068

Hochst.

A. Scherson EU2829 NA NA NA berteroa 113 83 nus

Moris 51

A. Wojciec L10768, NA NA DQ1072 brandeg howski L10769 37 eei and

Porter2 Sanders

on 157

A. Sanders AF1216 NA NA NA

52 calycosu on 975 91 s Torr. ex S.

Watson

A. Wojciec L10770, NA NA DQ1072 canaden howski L10771 40 sis L.2 and

Sanders

on 302

A. B. SRP Iron Co., KT2024 NA KT2024 KT2023 ceramic Franklin Utah 71 73 59 us E. 7679

Sheld.2

A. Blaine SRP Sheridan KT2024 NA KT2024 KT2023 ceramic H.M. Co., 72 74 60 us E. Mooers Montana

Sheld.2 1129 52

A. not U50514 NA NA NA cerasocr vouchere enus d

Bunge

A. not EU5919 NA NA NA complan vouchere 95

53 atus R. d

Br. ex

Bunge

A. not HQ6133 NA NA NA corrugat vouchere 78 us d

Bertol.

A. Scherson EU2829 NA NA NA crucksha 101 89 nksii

Hook &

Arn.

A. Scherson EU2829 NA NA NA cryptobo 108 80 trys I.M.

Johnst.

A. Scherson EU2829 NA NA NA 53

curvicau 112 84 lis Clos.

A. Liston AF1216 NA NA NA cysticaly 961 82 x Ledeb.

A. Jay SRP Malheur KT2024 NA KT2024 KT2023

54 cusickii Zimmers Co., 27 76 62

A. Gray 033 Oregon var. cusickii2

A. Jay SRP Washing KT2024 NA KT2024 KT2023 cusickii Zimmers ton Co., 28 77 63

A. Gray 039 Idaho var. cusickii2

A. Jay SRP Washing KT2024 KT2024 KT2024 KT2023 cusickii Zimmers ton Co., 29 05 78 64

A. Gray 045 Idaho var. cusickii2,

3

A. Jay SRP Washing KT2024 KT2024 KT2024 KT2023

cusickii Zimmers ton Co., 30 04 79 65 54

A. Gray 048 Idaho var. cusickii2,

3

A. Don CIC KT2024 KT2024 KT2024 KT2023 cusickii Mansfiel 64 22 75 61

55

A. Gray d 13-100 var. cusickii2,

3

A. Jay SRP Adams KT2024 NA KT2024 KT2023 cusickii Zimmers Co., 31 80 66

A. Gray 024 Idaho var. flexilipes

Barneby

2

A. Jay SRP Adams KT2024 NA KT2024 KT2023 cusickii Zimmers Co., 32 81 67

A. Gray 027 Idaho var. flexilipes

Barneby 55

2

A. Jay SRP Washing KT2024 NA KT2024 KT2023 cusickii Zimmers ton Co., 33 82 68

A. Gray 042 Idaho var. flexilipes

56

Barneby

2

A. Jay SRP Payette KT2024 KT2024 KT2024 KT2023 cusickii Zimmers Co., 34 06 83 69

A. Gray 001 Idaho var. packardi ae

Barneby

2, 3

A. Jay SRP Payette KT2024 KT2024 KT2024 KT2023 cusickii Zimmers Co., 35 07 84 70

A. Gray 002 Idaho var. packardi ae

Barneby 56

2, 3

A. Jay SRP Payette KT2024 KT2024 KT2024 KT2023 cusickii Zimmers Co., 36 08 85 71

A. Gray 003 Idaho var. packardi

57 ae

Barneby

2,3

A. Jay SRP Payette KT2024 KT2024 KT2024 KT2023 cusickii Zimmers Co., 37 09 86 72

A. Gray 004 Idaho var. packardi ae

Barneby

2,3

A. Jay SRP Payette KT2024 NA KT2024 KT2023 cusickii Zimmers Co., 38 87 73

A. Gray 005 Idaho var. packardi

ae 57

Barneby

2

A. Jay SRP Payette KT2024 NA KT2024 KT2023 cusickii Zimmers Co., 39 88 74

A. Gray 006 Idaho var.

58 packardi ae

Barneby

2

A. Jay SRP Payette KT2024 NA KT2024 KT2023 cusickii Zimmers Co., 40 89 75

A. Gray 007 Idaho var. packardi ae

Barneby

2

A. Jay SRP Payette KT2024 NA KT2024 KT2023 cusickii Zimmers Co., 41 90 76

A. Gray 008 Idaho var.

packardi 58

ae

Barneby

2

A. Jay SRP Payette KT2024 NA KT2024 KT2023 cusickii Zimmers Co., 42 91 77

A. Gray 009 Idaho

59 var. packardi ae

Barneby

2

A. Jay SRP Payette KT2024 NA NA NA cusickii Zimmers Co., 43

A. Gray 052 Idaho var. packardi ae

Barneby

A. Jay SRP Payette KT2024 NA NA KT2023 cusickii Zimmers Co., 44 78

A. Gray 053 Idaho var.

packardi 59

ae

Barneby

A. Jay SRP Payette KT2024 NA NA KT2023 cusickii Zimmers Co., 45 79

A. Gray 054 Idaho var.

60 packardi ae

Barneby

A. Jay SRP Payette KT2024 KT2024 KT2024 KT2023 cusickii Zimmers Co., 46 10 92 80

A. Gray 055 Idaho var. packardi ae

Barneby

2,3

A. Jay SRP Payette KT2024 KT2024 KT2024 KT2023 cusickii Zimmers Co., 47 11 93 81

A. Gray 056 Idaho var. packardi

ae 60

Barneby

2,3

A. Jay SRP Payette KT2024 KT2024 KT2024 KT2023 cusickii Zimmers Co., 48 12 94 82

A. Gray 057 Idaho var.

61 packardi ae

Barneby

2,3

A. Jay SRP Payette KT2024 NA KT2024 KT2023 cusickii Zimmers Co., 49 95 83

A. Gray 058 Idaho var. packardi ae

Barneby

2

A. Jay SRP Payette KT2024 NA KT2024 KT2023 cusickii Zimmers Co., 50 96 84

A. Gray 059 Idaho var.

packardi 61

ae

Barneby

2

A. Jay SRP Payette KT2024 NA KT2024 KT2023 cusickii Zimmers Co., 51 97 85

A. Gray 060 Idaho

62 var. packardi ae

Barneby

2

A. M. N/A Payette KT2024 NA KT2024 KT2023 cusickii Mancuso Co., 61 98 86

A. Gray 71 Idaho var. packardi ae

Barneby

2

A. M. N/A Payette KT2024 KT2024 KT2024 KT2023 cusickii Mancuso Co., 62 20 99 87

A. Gray 72 Idaho

var. 62

packardi ae

Barneby

2,3

A. Jay SRP Malheur KT2024 KT2024 KT2025 KT2023 cusickii Zimmers Co., 52 13 00 88

63

A. Gray 010 Oregon var. sterilis

(Barneb y)

Barneby

2,3

A. Jay SRP Malheur KT2024 KT2024 KT2025 KT2023 cusickii Zimmers Co., 53 14 01 89

A. Gray 015 Oregon var. sterilis

(Barneb y)

Barneby

2,3

A. Jay SRP Malheur KT2024 KT2024 KT2025 KT2023 63

cusickii Zimmers Co., 54 15 02 90

A. Gray 018 Oregon var. sterilis

(Barneb y)

64

Barneby

2,3

A. Scherson EU2829 NA NA NA darumbi 105 73 um

Bertero ex Colla

A. Sanders AF1217 NA NA NA douglasi on 980 09 i Torr. &

A. Gray

A. Scherson EU2829 NA NA NA edmonst 110 78 onei

(Hook. f.) B.L.

Rob. 64

A. Podlech AB0519 NA NA NA epiglotti 45851 10 s L.

A. Weber U50488, NA NA DQ1072 falcatus 15359 U50489 41

Lam.2

65

A. filipes Don CIC KT2024 NA KT2025 KT2023

Torr. ex Mansfiel 65 03 91

A. Gray2 d 13-005

A. filipes James F SRP KT2024 KT2024 KT2025 KT2023

Torr. ex Smith 66 23 04 92

A. 10762

Gray2,3

A. Wojciec AF1217 NA NA DQ1072 inyoensi howski 37 32 s E. 527

Sheld.2

A. Scherson EU2829 NA NA NA johnston 102 88 ii

Gomez-

Sosa

A. Jay SRP Malheur KT2024 KT2024 KT2025 KT2023 65

lentigino Zimmers Co., 55 17 05 93 sus 030 Oregon

Dougl. ex

Hook.2,3

A. Jay SRP Malheur KT2024 KT2024 KT2025 KT2023

66 lentigino Zimmers Co., 56 18 06 94 sus 036 Oregon

Dougl. ex

Hook.2,3

A. Jay SRP Washing KT2024 KT2024 KT2025 KT2023 lentigino Zimmers ton Co., 57 19 07 95 sus 051 Idaho

Dougl. ex

Hook.2,3

A. Wojciec AF1216 NA NA DQ1072 lonchoca howski 89 30 rpus and

Torr.2 Sanders

on 143

A. Scherson EU2829 NA NA NA 66

looseri 104 74

I.M.

Johnst.

A. Sanders AF1217 NA NA NA mollissi on 950 19 mus

67

Torr.

A. Scherson EU2829 NA NA NA monticol 103 82 a Phil.

A. James F SRP KT2024 KT2024 KT2025 KT2023 mulfordi Smith 67 24 08 96 ae M.E. 10725

Jones2,3

A. Scherson EU2829 NA NA NA nivicola 111 85

Gomez-

Sosa

A. Wojciec AF1216 NA NA DQ1072 nothoxys howski 88 31

A. Gray2 and

Sanders

on 177 67

A. Sanders AF1217 NA NA NA oxyphys on 979 08 us A.

Gray

A. Sanders AF1217 NA NA DQ1072 pachypu on 984 22 36

68 s

Greene2

A. Sanders AF1217 NA NA NA patagoni on 2515 46 cus Phil.

A. Scherson EU2829 NA NA NA pehuenc 107 81 hes

Niederl.

A. not U50494, NA NA NA peristere vouchere U50495 us Boiss. d

&

Hausskn.

A. Sanders AF1217 NA NA DQ1072 preussii on 999 26 34

2 A. Gray 68

A. James F SRP KT2024 KT2024 KT2025 KT2023 purshii Smith 68 25 09 97

Dougl. 10726 ex Hook.

2,3

A. James F SRP KT2024 KT2024 KT2025 KT2023

69 purshii Smith 69 26 10 98

Dougl. 10746 ex

Hook.2,3

A. Jay SRP Malheur KT2024 NA KT2025 KT2023 solitariu Zimmers Co., 58 11 99 s M.E. 013 Oregon

Peck2

A. Jay SRP Malheur KT2024 KT2024 KT2025 KT2024 solitariu Zimmers Co., 59 16 12 00 s M.E. 021 Oregon

Peck2,3

A. Sanders AF1217 NA NA DQ1072 tetrapter on 1006 28 28 us A.

Gray2

A. not EU2829 NA NA NA 69

uniflorus vouchere 86

DC. d

A. vagus Scherson EU2829 NA NA NA

Reiche 109 79

A. Mozaffar AB0519 NA NA NA vogelii ian et.al. 11

70

(Webb) 39103

Bornm.

A. Jay SRP Harney KT2024 NA KT2025 KT2024 whitneyi Zimmers Co., 60 14 02

A. Gray 061 Oregon var. confusus

Barneby

2

A. James F SRP KT2024 KT2024 KT2025 KT2024 whitneyi Smith 63 21 13 01

A. Gray 10946 var. confusus

Barneby

2,3

A. Sanders AF1217 NA NA NA 70

woodruff on 995 24 ii M.E.

Jones

A. Beth SRP KT2024 NA KT2025 KT2024 yoder- Corbin 70 15 03 williamsi 1550

71 i

Barneby

2

Oxytropi Wojciec AF1217 NA NA DQ1072 s sericea howski 57 39

Nutt.2 and

Sanders

on 255

71

72

72

73

73

74

74

75

75

76

76

77

77

78

78

79

Highlights

 Phylogenetic analyses indicate that 4 varieties of Astragalus cusickii do not form a monophyletic group  Genealogical sorting index results indicate recognizing 3 species is best fit to data  Chloroplast capture is likely to have occurred in at least one variety

79

80

Graphical abstract

80

本文献由“学霸图书馆-文献云下载”收集自网络,仅供学习交流使用。

学霸图书馆(www.xuebalib.com)是一个“整合众多图书馆数据库资源,

提供一站式文献检索和下载服务”的24 小时在线不限IP 图书馆。 图书馆致力于便利、促进学习与科研,提供最强文献下载服务。

图书馆导航:

图书馆首页 文献云下载 图书馆入口 外文数据库大全 疑难文献辅助工具