<<

Western North American Naturalist

Volume 70 Number 3 Article 3

10-11-2010

High levels of gene flow in the olev ( californicus) are consistent across spatial scales

Rachel I. Adams Stanford University, Stanford, California, [email protected]

Elizabeth A. Hadly Stanford University, Stanford, California, [email protected]

Follow this and additional works at: https://scholarsarchive.byu.edu/wnan

Recommended Citation Adams, Rachel I. and Hadly, Elizabeth A. (2010) "High levels of gene flow in the California (Microtus californicus) are consistent across spatial scales," Western North American Naturalist: Vol. 70 : No. 3 , Article 3. Available at: https://scholarsarchive.byu.edu/wnan/vol70/iss3/3

This Article is brought to you for free and open access by the Western North American Naturalist Publications at BYU ScholarsArchive. It has been accepted for inclusion in Western North American Naturalist by an authorized editor of BYU ScholarsArchive. For more information, please contact [email protected], [email protected]. Western North American Naturalist 70(3), © 2010, pp. 296–311

HIGH LEVELS OF GENE FLOW IN THE CALIFORNIA VOLE (MICROTUS CALIFORNICUS) ARE CONSISTENT ACROSS SPATIAL SCALES

Rachel I. Adams1,2 and Elizabeth A. Hadly1

ABSTRACT.—Gene flow links the genetic and demographic structures of species. Despite the fact that similar genetic and demographic patterns shape both local population structure and regional phylogeography, the 2 levels of population connectivity are rarely studied simultaneously. Here, we studied gene flow in the California vole (Microtus californicus), a small-bodied with limited vagility but high local abundance. Within a 4.86-km2 preserve in central California, genetic diversity in 6 microsatellites was high, and Bayesian methods indicated a single genetic cluster. However, indi- vidual-based genetic analysis detected a clear signal for isolation-by-distance (IBD) and fine-scale population structure. Mitochondrial cytochrome b sequencing revealed 11 unique haplotypes from the one local area where we sequenced 62 individuals. Phylogeographic analysis of these individuals combined with those sampled from the northern geographic range of the species (the range of the species spans western North America from southern to northern Mexico and is centered geographically within the state of California) again indicated a lack of structure but a signal for IBD. Pat- terns of gene flow thus are consistent across spatial scales: while dispersal of the California vole is limited across geo- graphic distance, there is nonetheless considerable movement across the landscape. We conclude that in this species, high local population abundances overcome the potential genetic and demographic effects of limited dispersal.

Key words: fine-scale genetic structure, population genetics, phylogeography, Jasper Ridge Biological Preserve, cytochrome b, Bayesian analysis, SIMCOAL.

The geographic distribution of genetic diver- individuals across space and time—the demo- sity in natural populations harbors a wealth of graphic characteristics at the different scales information on species’ evolutionary and eco - vary, and consequently the patterns may be logical histories, demographic parameters, and quite distinct. Nevertheless, the 2 levels of gene conservation statuses. Gene flow, mediated by flow are rarely studied concurrently. For a spe - the movement of reproductive individuals and cies with limited vagility and strong philopatry therefore genes across space, is a prominent but high abundance, expectations for signals determinant of the genetic composition and of gene flow may differ depending on spatial diversity within a species and ultimately of the scale considered. A strong signal of genetic species’ survival. The role of gene flow in shap- structure is expected on both a microgeographic ing the pattern of gene frequencies depends and macrogeographic scale for organisms with on ecological, historical, and demographic fac- limited vagility, high fidelity to natal site, and tors. Therefore, measurements of gene flow specific habitat requirements (Lowe et al. 2004). give insight into the behavior, migration, and However, species with high abundances have mating patterns of natural populations. Gene the potential to override factors limiting gene flow is traditionally considered at 2 different flow through sheer force of numbers, and temporal scales: for short distances and times, therefore, genetic differentiation on a microgeo- gene flow is a process, along with genetic drift, graphic scale, but not necessarily on a macro- mutation, and natural selection, that shapes geographic scale, would be lacking in these population genetic structure (Hartl and Clark species. We set out to test whether signals of 1997); at longer distances and times, gene flow, gene flow would be similar at different spatial and the lack thereof, determine a species’ phy- scales for a species characterized by low disper- logeographic pattern. sal, high philopatry, and large population sizes. While both local population structure and By using 2 genetic markers of variable muta- phylogeography ultimately study the same tion rates (microsatellites and cytochrome b) evolutionary process—that of movement of from the same Microtus californicus individuals

1Department of Biology, Stanford University, Stanford, CA 94305-5020. 2E-mail: [email protected]

296 2010] GENE FLOW ACROSS SPATIAL SCALES 297 in one locality and linking these data to a study Heske 1987, Salvioni and Lidicker 1995), and of the phylogeography of the species that used this variation is known to affect the genetic cytochrome b (Conroy and Neuwald 2008), we diversity and structure of populations. Using were able to examine genetic structure across allozyme data from M. californicus, Bowen a small scale and place the locality into the phy - (1982) found significant genetic differentiation logeographic structure of the species—effec- between populations separated by only 50–200 tively synthesizing patterns of gene flow at 2 m when population density was low. This differ- spatial scales. ential disappeared when density rebounded. A The California vole (Microtus californicus) is a similar recovery of connectivity during peri- small rodent (<100 g) whose demographic and ods of high density has been observed in another ecological features have been well characterized. vole, terrestris (Berthier et al. 2006). Individual dispersal distances are estimated to However, other vole species show increased be small, at 50 m or less (e.g., Lidicker 1973, gene flow during periods of low density (Andrea - Bowen 1982), and philopatry high, with male ssen and Ims 2001, Hadly et al. 2004). home ranges slightly larger than female home On a larger spatial scale, Microtus californicus ranges (e.g., 180 m2 and 120 m2, respectively; has 17 distinct subspecies (Hall 1981) that span Heske 1987). California , like many other the range of the species and are based on pel - microtine , undergo 2- to 4-year popula- age, skull shape, and body size (Kellogg 1918). tion cycles (Krebs 1966). In addition, the Cali- However, further work based on distinct mor- fornia vole is limited to areas of heavy ground phological differences, decreased fertility in cover, predominantly and oak wood- crosses, and allozyme variation has shown a lands, where it lives in underground burrows break between only 2 broad regions—northern but travels aboveground to forage for grasses, and southern California (Gill 1980). Recently, a herbs, and sedges. It breeds throughout the year, phylogeographic study of the California vole although reproduction peaks whenever food and corroborated this geographic barrier, reporting cover are most abundant. Litter size averages strong genetic differentiation between popula- 4 individuals but ranges from 1 to 9, and fe - tions on the northern and southern sides of the males produce 2–5 litters per year. Females Transverse Mountain range in southern Cali- reach sexual maturity at an average of 29 days. fornia (Conroy and Neuwald 2008). The California vole’s circadian activity, along Thus, previous work with California voles with its widespread and abundant distribution, indicates that barriers to gene flow occur on make it an important prey for carnivores. Al- temporal scales as well as local and regional though it has a broad range, from Baja Califor- spatial scales, but never before have the same nia throughout western and central California individuals been used to determine gene flow to southern Oregon, the vole’s populations are at the 2 spatial scales simultaneously. For this naturally fragmented into suitable habitat, which goal, our study sampled heavily in one locality, is patchily distributed across this range. the Jasper Ridge Biological Preserve (JRBP), The particular biology of the California vole San Mateo County, located in the and other voles leads to the presence of popu- Bay region of California. The preserve itself lation structure on a small spatial scale due to represents habitat within an increasingly de - habitat patchiness and landscape features (Neu- veloped area, and JRBP is varied in particular wald 2010, Berthier et al. 2005, Gauffre et al. landscape features. The preserve contains both 2008), as well as demography (Bowen 1982, native and nonnative grasslands. Nonnative Berthier et al. 2006, Gauffre et al. 2008). In grasses, predominantly of Mediterranean origin, southern California, a subspecies of the Cali- dominate the nonserpentine grasslands, which fornia vole which inhabits distinct marshes are located on greenstone, sandstone, and chert adjoining the Amargosa River in the Mojave substrates. And native grasses, which are Desert, M. c. scirpensis, showed reduced genetic adapted to the low nutrient content and the diversity and a greater degree of subdivision unique chemical content of serpentine grass- between subpopulations than a more widely lands, remain dominant on the serpentine-sub- distributed subspecies, M. c. sanctidiegi (Neu - strate outcroppings (Hobbs et al. 1997 and wald 2010). Moreover, dispersal distances can citations therein). It is not known whether vary depending on population density, season, small have responded to this shift and and microhabitat (Krebs 1966, Lidicker 1973, partitioning of types. Moreover, the 298 WESTERN NORTH AMERICAN NATURALIST [Volume 70

Fig. 1. Jasper Ridge Biological Preserve, San Mateo County, showing its location in the San Francisco Bay area of cen- tral California. Microtus californicus was trapped in 7 sites representing different habitat types in the preserve. Sam- pling numbers at each site are given in parentheses. Individuals from the highlighted California counties were included in the phylogeographic analysis. preserve is varied in habitat, where areas of diversity and abundance of small popu- grasslands are of variable sizes and are sepa- lations at JRBP, a 4.86-km2 preserve in the rated by other, less preferred habitat types, such foothills of the Santa Cruz Mountains managed as those dominated by dense woody and/or by Stanford University (Fig. 1). Briefly, live shrubby vegetation. Sherman and Tomahawk traps were laid out in This study examined the genetic signature of 50-m-radius circular arrays or 100-m line tran- neutral microsatellite loci and examined the con- sects in different sites representing the diver- nectivity of California voles within a protected sity of habitats at the preserve. Trapping arrays preserve characterized by distinct juxtaposed in different habitats varied in separation from landscape features, including compositional 0.28 to 2.11 km with a median of 0.75 km. shifts in grassland types separated by unsuit- Individual traps were baited with a mixture able forested habitat. We also used mitochon- of oats and peanut butter and were opened drial DNA to determine how these same JRBP for 5 consecutive nights during March and California vole individuals were phylogeneti- April 2006. All trapped were scored for cally related to other populations across the species identity, sex, reproductive condition, species’ range. and weight. Tissue samples (initially toe clips, later ear snips) were collected and stored in METHODS 100% ethanol in a refrigerator until DNA was extracted. The subspecies distinction for the Study Site and Collection site of JRBP is M. c. californicus (Hall 1981), Tissue sampling was performed as part of a and all individuals trapped were classified as larger trapping study aimed at detecting the belonging this subspecies. Trapping methods 2010] GENE FLOW ACROSS SPATIAL SCALES 299 followed the American Society of Mammalo- 94 °C for 1 minute followed by 40 cycles con- gists guidelines and were approved by the sisting of 15 seconds at 94 °C, 15 seconds at 45 Institutional Animal Care and Use Committee °C, 45 seconds at 72 °C, and a final 3 minutes of at Stanford University. annealing at 72 °C. Products were sequenced using the 3730 DNA Analyzer (Applied Bio - Genetic Data Collection systems) at Cogenics, Inc. (Houston, TX). All 72 M. californicus tissue samples collected Alignment was straightforward and done with during the trapping study were included in the Sequencher 4.7 (GeneCodes), and unique analysis. Animal material was dissociated using sequences have been deposited into GenBank, tweezers prior to extraction, and genomic DNA accession numbers GQ168680–GQ168690. extraction followed the Qiagen (Valencia, CA) Complete cytochrome b sequences from indi- DNeasy protocol. viduals belonging to the northern clade of Cali- fornia voles (Conroy and Neuwald 2008) were Microsatellites retrieved from GenBank and used in phylo- A total of 6 nuclear microsatellite loci were geographic analyses. used. Neuwald (2010) optimized PCR conditions for 5 primers (MSMM-2, 3, 5, 6, and 8) originally Genetic Analysis developed for Microtus montebelli (Ishibashi MICROSATELLITE DIVERSITY.—The sampling et al. 1999); an additional polymorphic locus protocol designed to survey small mammals in (Moe2) was adapted from primers originally general, rather than M. californicus in particular, developed for Microtus oeconomus (Van de led to highly variable sample sizes for each Zande et al. 2000). All microsatellite loci contain predefined trapping locality, with many of the a dinucleotide repeat sequence. Tagged fluo- localities containing only a few individuals rescent primers were synthesized by Applied (Fig. 1). Therefore, individual-based, rather than Biosystems (Foster City, CA) to allow for mul- population-based estimates of genetic structure tiplexing: MSMM-2 and 8 were tagged with 6- were used for all microsatellite analyses. Popu- Fam (blue), MSMM-3 and 6 with VIC (green), lation-based analyses require the assignment and MSMM-5 and Moe2 with NED (yellow). of individuals to predefined populations; using Each PCR was carried out in a total volume of individual organisms as the unit of analysis has 10 μL containing up to 5 μL of DNA template, the advantage of utilizing all available genetic 2.5 μM of each primer, 2.5 mM dNTPs, 1.5 mM information and requires fewer a priori assump- μ MgCl2, 5 U of Taq DNA polymerase, and 1 L tions. Summary statistics and basic measures of 10X PCR buffer from Invitrogen (Carlsbad, CA). genetic diversity were generated in GenAlEx PCR cycling conditions were as follows: an ini - (Peakall and Smouse 2006). GenePop (Rousset tial denaturation at 93 °C for 2 minutes followed 2008) was used to test for deviations from by 35 cycles consisting of 30 seconds at 90 °C, Hardy-Weinberg equilibrium using an exact test 20 seconds at 50–52 °C, 20 seconds at 72 °C, and also for evidence of linkage disequilibrium and a final 3 minutes of annealing at 72 °C. Sam- using the G test. ples were prepared with Hi-Di™ formamide DETECTING GENETIC STRUCTURE.—At JRBP, and 500 LIZ® size standard and run on the several characteristics independent of distance 3730xl DNA Analyzer (Applied Biosystems, have the potential to limit vole dispersal. First, Foster City, CA) at the University of Califor- several of the grasslands where the California nia, San Francisco, Genome Core Facilities. voles were collected are separated by unsuitable habitat for this species, such as forest and dense Mitochondrial Cytochrome b chaparral. Second, the grasslands are repre- The primers MVZ 05 (Smith and Patton sented by 2 types—serpentine and nonserpen- 1993) and vole-14 (Hadly et al. 2004) were used tine—and it was unknown whether the voles to amplify the entire cytochrome b mitochon - discriminated between these 2 types. Therefore, drial region (1143 bps). PCR was carried out in we sought to determine if there was any a total volume of 10 μL containing 2 μL of DNA population subdivision within JRBP. We used template, 5 μM of each primer, 2.5 mM dNTPs, 2 Bayesian-based programs, the spatially non- 1.5 mM MgCl2, 5 U of Taq DNA polymerase, explicit Structure 2.2 (Pritchard et al. 2000) and 1 μL 10X PCR buffer. PCR cycling condi- and the spatially explicit Geneland (Guillot et tions were as follows: an initial denaturation at al. 2005, 2008), to cluster individuals into 300 WESTERN NORTH AMERICAN NATURALIST [Volume 70 predefined K populations that minimize link- 1985) and tested for significance with 1000 boot- age disequilibrium (LD) and maximize Hardy- straps. Similarly, Ritland’s estimator of relat- Weinberg equilibrium. Within Structure, 10 edness (Ritland 1996) was calculated for each independent runs for a fixed K = 1 to 7 were individual pairwise comparison, and this esti- carried out, with a maximum of 7 defined by the mate was correlated with geographic distance number of trapping arrays. We assumed admix- using Pearson’s product moment correlation. ture and a model of correlated allele frequencies We further sought to estimate the dispersal between populations (Falush et al. 2003) with a distance of voles in this population, and to this burn-in period of 100,000 iterations followed by end we utilized the relationship between pair- 1,000,000 Markov Chain Monte Carlo (MCMC) wise genetic distance and neighborhood size replicates. The assumption of correlated allele developed by Rousset (2000). Dispersal distance frequencies better handles departures from can be estimated according to σ 2 = 1/(4Dπb), Hardy-Weinberg equilibrium or LD; however, where σ is the mean parent-offspring dispersal it can also lead to overestimating K (Pritchard distance, D is the effective density of the popu- et al. 2007), so particular attention was paid to lation, and b is the slope of the regression of this possibility during analysis. Time estimates pairwise genetic distance (Rousset’s a) against were determined by ensuring that parameter the logarithm of geographic distance. Rousset’s estimates converged before the end of the a was generated with the program SPAGeDi burn-in period, and the method developed by (Hardy and Vekemans 2002), and 2 values for Evanno et al. (2005) was employed to calculate D were used due to the cyclical nature of vole the most likely value of K. Geneland incorpo- populations, one during high animal density rates the geographic location of sampling in (0.101 voles ⋅ m–2) and one during low animal estimates of population structure. K was allowed density (0.017 voles ⋅ m–2). These estimates of to vary in 10 independent runs, and the most density were informed by a field study previ- probable value was estimated using 1,000,000 ously conducted at JRBP (Giusti 1985). Although MCMC iterations, with the maximum rate of the approximate linear relationship between the Poisson process fixed at 500, the maximum genetic and geographic distances is not expected number of nuclei in the Poisson-Voronoi tes- to hold as well at distances less than the esti- sellation fixed at 200, and use of the Dirichlet mated demographic dispersal distance (Rousset model for allele frequencies. 1997), we did not exclude any pairwise com- SCALES OF GENETIC STRUCTURE.—We sought parisons because the geographic center of the to identify the geographic scale of genetic struc- trapping array was used as the trapping point ture at JRBP independent of delimited popula- rather than the specific location of the trap. tions, and for this we employed a number of Regardless, the slopes of the linear regression analyses in GenAlEx. Principle coordinates have been shown to differ little whether pair- analysis (PCA) was used to plot the relationships wise comparisons less than the demographic among all individuals. In order to test whether dispersal distance are excluded or not (Rousset individuals collected from within serpentine 2000). grasslands were more genetically similar than To look for evidence of a recent reduction in individuals from across different grassland types, effective population size, we used the program we performed a Mantel test with genetic dis- Bottleneck v.1.2.02 (Piry et al. 1999). This pro- tance and serpentine/nonserpentine distinction. gram tests for a severe population reduction Isolation-by-distance (IBD) was tested in sev- using the principle that populations that have eral ways. To test the statistical relationship undergone a bottleneck demonstrate a reduction between geographic distance and the genetic in the number of alleles as well as in heterozy- distance matrices, we performed a Mantel test, gosity (sensu Nei 1987), with decrease in allele where significance was determined compared to number occurring faster than loss of heterozy- 1000 permutations. Moreover, in order to test gosity. Under this expectation, the observed for fine-scale genetic structure, we employed heterozygosity is higher than the expected het- a heterogeneity test for autocorrelation of pairs erozygosity from the allele number at mutation- of individuals within defined distance classes drift equilibrium (Cornuet and Luikart 1996). (Smouse et al. 2008). A distance class of 400 m Within Bottleneck, we implemented a Wilcoxon was chosen based on estimates for home ranges test assuming the 2-phase model (TPM) of of M. californicus measured at JRBP (Giusti microsatellite evolution with 95% of mutations 2010] GENE FLOW ACROSS SPATIAL SCALES 301 following a single-step model and 5% a multi- Conroy and Neuwald [2008], GenBank acces- step model. sion #EF506039) served as the outgroup, from We compared the results of the microsatellite which all northern samples were forced to be analyses with those reported in another study monophyletic. Parsimony analysis was performed examining these same loci in other populations using stepwise random addition sequence and of M. californicus (Neuwald 2010). While we TBR-branch swapping, and 1000 bootstrap rep- cannot compare absolute sizes of microsatellites licates were executed. A haplotype network because we did not standardize across genotyp- was generated using TCS version 1.21 (Clement ing runs and because of limited gene flow be - et al. 2000) with a 95% confidence interval for tween northern and southern clades of this assigning branches, excluding the outgroup, species (Conroy and Neuwald 2008), we are which required a lower threshold value. Repre- nevertheless able to compare relative numbers sentation of the haplotype network generated and range of alleles across the 2 studies. by TCS was subsequently generated by hand. CYTOCHROME B AND PHYLOGEOGRAPHY.—We Because the high diversity in mitochondrial looked for evidence of nuclear mitochondrial haplotypes from a single locality indicated that DNA inserts (Numts) by examining patterns of California voles were extremely abundant, substitution. We checked for the presence of Bayesian serial coalescent modeling was used premature stop codons and a bias in substitu- to estimate the effective population size of the tions in the 3rd codon position relative to the JRBP individuals. We used this technique to 1st and 2nd positions, which tend to be con- simulate the number of segregating sites and served because changes in these sites often re - examine the probability of detecting the ob- sult in amino acid replacements. EstimateS served level of diversity (i.e., 11 segregating (Colwell 2006) was used to generate a haplotype sites) under different mutation rates and effec- accumulation curve, where the expected rich- tive population sizes. We explored the effect of ness function, Mao Tau, and the standard de - deme size from 30,000 to 600,000 individuals, at viation were calculated using the analytical 3 mutation rates, μ, corresponding to low (2% formulas of Colwell et al. (2004) and scaled per million years per base pair [bp]), moderate against the computed individuals in the sam- (4%), and high (10%) values (Hadly et al. 2004). pling effort (Gotelli and Colwell 2001). We assumed 5 generations per year (Greenwald In order to incorporate individuals from 1957), a constant growth rate, a uniform muta- JRBP into the biogeographic history of the tion rate, and a sequence length of 1143 bp. species, we employed standard phylogeographic Running the model 10,000 times at each para- techniques. Because Conroy and Neuwald meter combination provided the expected dis- (2008) found little evidence for gene flow be - tribution of gene diversity. Bayesian Serial tween the northern and southern clades of M. SimCoal (http://www.stanford.edu/group/hadly californicus, and because we did not detect lab/ssc/BayeSSC.htm) is a modification of SIM- any sequences from the southern clade in this COAL 1.0 (Excoffier et al. 2000). study, we used complete sequences (n = 19) from just the northern clade. Mantel tests using RESULTS genetic distance between cytochrome b haplo- types were conducted in GenAlEx at 2 spatial Seventy-two M. californicus individuals were scales: JRBP individuals exclusively (n = 11) collected from 7 trapping sites (median individu- and JRBP individuals plus ones from across als captured = 7) across JRBP in predomi- the northern California range (n = 30). nantly grassland habitats (Fig. 1). The only vole Maximum parsimony with 1000 bootstrap captured outside a grassland was a pregnant replicates was determined in PAUP* (Swof- female, which was trapped in a woody scrub- ford 2003) to examine the evolutionary history land; other woodland and forest trapping locali- of California voles. We limited our phylogenetic ties contained no voles. The scrubland in which analysis to parsimony because of the reduced the vole was captured is situated adjacent to likelihood of homoplasy in our recently diverged small grassland pastures. samples. Exploration using maximum likelihood Genetic Variation and Bayesian techniques produced similar re - sults (results not shown). A representative from Genotype analyses revealed high levels of the southern clade (haplotype #S13 from genetic diversity (Table 1). A total of 72 alleles 302 WESTERN NORTH AMERICAN NATURALIST [Volume 70

TABLE 1. Number of alleles (A), expected heterozygosity (HE), observed heterozygosity (HO) averaged over individuals, and P values for Hardy-Weinberg equilibrium (HWE; determined using chi-square exact tests) for Microtus californicus in this study (M. c. californicus) and for Neuwald (2010; M. c. scirpensis and M. c. sanctidiegi).

______M. c. californicus (n = 71) ______M. c. scirpensis (n = 67) ______M. c. sanctidiegi (n = 27) HWE HWE HWE Locus A HE HO P value A HO P value A HO P value MSMM2 16 0.873 0.914 0.359 6 0.716 0.000 15 0.800 0.100 MSMM3 12 0.825 0.725 0.225 3 0.075 0.314 12 0.667 0.011 MSMM5 12 0.877 0.549 0.000 1 0 — 13 0.704 0.071 MSMM6 5 0.305 0.322 1.000 1 0 — 7 0.593 0.076 MSMM8 13 0.888 0.424 0.000 2 0.284 0.000 12 0.654 0.000 Moe2 14 0.881 0.871 0.001 —— — — — —

TABLE 2. Outcome of 2 individual-based Bayesian clus- endangered subspecies M. c. scirpensis. How- tering programs. In Structure, 10 independent runs of K ever, the level of genetic diversity we found most fixed from 1 to 7 were completed, and the means and standard deviations of the log likelihood values are closely mirrored the widespread subspecies M. reported here. In Geneland, 10 independent runs were c. sanctidiegi, which was assessed at a broader completed, and the estimated K and log likelihood values geographic scale—spanning 3 counties in south- are reported. ern California—but for fewer individuals (Neu - ______Structure ______Geneland wald 2010). K Log likelihood K Log likelihood Genetic Structure 1 –1669.29 +– 0.34 1 –1219.6157 Clustering analysis performed in Structure 2 –2029.33 +– 118.79 1 –1222.2049 3 –1800.14 +– 46.79 1 –1196.9087 and Geneland showed that Jasper Ridge repre- 4 –1892.94 +– 67.72 1 –1212.0824 sents a single genetic population (Table 2). In 5 –1986.00 +– 135.67 1 –1215.2921 Structure, the estimated logarithm of likelihood 6 –1982.31 +– 38.84 1 –1227.3241 + was highest for K = 1, and the method of 7 –2017.74 – 45.13 1 –1213.1726 Evanno et al. (2005) detected K = 1 cluster. 1 –1232.2259 1 –1214.4837 Individuals assigned to particular populations 1 –1211.5334 when K > 1 were distributed randomly across trapping sites. Likewise, 10 independent runs of Geneland all showed the greatest likelihood was found for the 6 microsatellite loci, with each at K = 1. Furthermore, PCA analysis corrobo- locus supporting between 5 and 16 alleles, with rated a lack of structure by demonstrating no an average of 12 (SE = 3.7) alleles per locus. obvious clustering of individuals by trapping Expected heterozygosity for each locus ranged locality or habitat type. from 0.31 to 0.89, with an average of 0.78 (SD A Mantel test between pairwise geographic = 0.23) for all loci. Observed heterozygosity at and genetic distances showed a significant result each locus ranged from 0.32 to 0.91, with an using microsatellites (P = 0.042). Similarly, average of 0.63 (SD = 0.24) for all loci. Three there was a significant and positive relationship of the 6 loci (MSMM-5, MSMM-8, Moe2) dem- between Rousset’s a and geographic distance onstrated a heterozygosity deficit and significant (P = 0.003). Using the slope of this regression deviation from Hardy-Weinberg equilibrium (b = 0.00826) and the 2 estimates of density expectations (Table 1). Lower than expected lev- (D), the indirect dispersal distance of voles (σ) els of heterozygosity may be common in species based on genetic data yields 2 values: 9.8 m at experiencing large interannual cycles in popu- high vole density and 23.8 m at low vole density. lation size (Delport et al. 2005). Three locus Spatial genetic autocorrelation analysis, in - pairs, each including locus MSMM-5, showed cluding a heterogeneity test, showed evidence significant linkage disequilibrium (P < 0.05). for microgeographic spatial genetic structure Analyses were explored without that locus, and (Fig. 2). Correlations were near zero for the first because results were similar (data not shown), distance classes, until the 1600-m distance in - analyses include all loci. The level of sam- terval, at which point correlations became highly pling, both in terms of individuals caught and negative. Interestingly, correlations were not size of geographic area, matched that of the positive at the short distance classes, as is often 2010] GENE FLOW ACROSS SPATIAL SCALES 303

Fig. 2. Fine-scale correlogram showing the genetic autocorrelation (r) as a function of distance partitioned into dis- crete size classes. Error bars bound the 95% CI for r determined with 1000 bootstrap trials for comparisons within size class; dashed lines show the 95% CI for the null hypothesis determined by 1000 random permutations of individuals among geographic locations. Note the abrupt change in r at distance classes >1600 m. the case (Gauffre et al. 2008, Smouse et al. 2008). haplotypes were treated as true mitochondrial Although the overall correlogram was significant se quence. Complete cytochrome b sequences (w = 25.63, P = 0.035), only one individual (1143 bp) were obtained from 62 individuals, distance class was significant, and it was posi- and 11 distinct haplotypes were observed. Nu - tive: 1200–1599 m (autocorrelation r = 0.013, cleotide base composition was similar to that P = 0.036). However, the distance class 1600– reported for other Microtus species: 29.0% 1999 m was marginally significant (autocorre- cytosine, 26.8% thymine, 32.3% adenine, and lation r = –0.080, P = 0.065). The confidence 12.0% guanine (Hadly et al. 2004, Spaeth et al. interval for the values increases at the large 2009). There were 11 total variable sites. Gene distance classes due to a reduced number of diversity (Hd = 0.85 +– 0.025) and nucleotide di- comparisons in these ranges. Relatedness, as versity (π = 0.0018 +– 0.0001) were similar to measured by Ritland’s RI (Ritland 1996), was those for other voles (Hadly et al. 2004, Spaeth negatively correlated with geographic distance et al. 2009). (correlation r = –0.047, P = 0.026). There was Individual-based rarefaction analysis indi- no evidence for a population bottleneck (Wil - cated that the sampling effort detected the coxon, P = 0.578), and a Mantel test showed a preponderance of genetic diversity at JRBP (Fig. nonsignificant (P = 0.251) correlation between 3). The number of haplotypes detected with pairwise genetic distance and serpentine/non- increased sampling effort had begun to saturate, serpentine sampling locality. as evidenced by a logarithmic line fit through the data points. This best-fit line predicts that Phylogeography and Cytochrome b sampling an additional 20 individuals would None of the cytochrome b genetic sequences uncover 1 further haplotype. The locality of had accumulated premature stop codons. The JRBP demonstrated a large amount of within- majority (73%) of base pair (bp) changes were site diversity, but detection of this haplotype located in the 3rd position and did not result diversity required intensive sampling. The in an amino acid change (64%). Therefore, entire set of 11 haplotypes appears in each of 304 WESTERN NORTH AMERICAN NATURALIST [Volume 70

Fig. 3. Haplotype accumulation curve for mitochondrial cytochrome b haplotypes at Jasper Ridge Biological Preserve. 95% CI bars are shown. Included is a logarithmic line through the observed data points, showing the projected modest detection of haplotypes with increased sampling effort. the 3 largest sampling localities, and the smaller counties of Marin and Mendocino. There is also sampling localities each contain a subset of notable relatedness between the northern and these types. The most abundant haplotype (n = southern extremes from the counties examined: 18) appeared in all sampling localities. All of the haplotypes from Santa Barbara and Siskiyou haplotypes detected at JRBP were undetected counties. in the larger phylogeographic study of the Bayesian serial coalescent modeling indicated species. that the effective population size for M. califor- A Mantel test for IBD within the mitochon- nicus is large, at least 120,000 individuals (Fig. drial locus was not significant for those indi- 6). Deme size at a low mutation rate peaked in viduals within JRBP (P = 0.340) but was highly likelihood at 400,000 individuals, at a moderate significant for all individuals across northern mutation rate at 210,000 individuals, and at a California (P = 0.001). high mutation rate at 120,000 individuals. Phylogeographic analysis with maximum par - simony analysis produced 5 most-parsimonious DISCUSSION trees. The 50% majority-rule consensus tree with bootstrap support is shown in Fig. 4. Mi- This study documents a high amount of gene crotus californicus individuals from JRBP fit flow in M. californicus, both on fine and coarse into the northern clade, but there is little struc- spatial scales. Microsatellites indicated that ture across the entire northern clade. Rather, the JRBP, a 4.86-km2 preserve, represents some geographic areas show in situ diversifi- one breeding population with high diversity, cation. In particular, haplotypes in Ventura nonetheless exhibiting IBD on a microgeo- County, Santa Barbara County, Jasper Ridge, graphic scale of distances <2 km. At the same Jasper Ridge / Yolo County, and Marin/Mendo- time, phylogeographic analysis revealed a com - cino counties demonstrate diversification from plete lack of structure across the entire north- a separate common ancestor. The haplotype ern clade of the species, further indicating a network (Fig. 5) corroborates the limited pres- high amount of movement and dispersal, while ence of structure by geography. Southern line - showing the pattern of IBD. Although nuclear ages from within the northern clade (Ventura microsatellites and mitochondrial sequence and Santa Barbara counties) are closely related, reveal gene flow on different timescales, both as are haplotypes from the northern central here show the same overall signal. 2010] GENE FLOW ACROSS SPATIAL SCALES 305

Fig. 4. Maximum parsimony tree with bootstrap support above the nodes generated from the complete mitochondrial cytochrome b nucleotide sequence. Numbers in parentheses indicate the unique haplotype identity used in the haplo- type network (see Fig. 5). Note the distribution of southern and northern variants throughout the tree.

The hypervariable microsatellites detected of vole individuals appear great enough to pre- a signal of increasing difference with distance, vent genetic differentiation. but did not detect clear population structure. Perhaps more suprisingly, the same pattern In this case, a significant signal for IBD did not is observed across a much wider spatial scale lead to overestimation of genetic structure, as within the entire northern clade of the species: can be the case when Bayesian methods are phylogeographic analysis further showed that employed (Frantz et al. 2009). A pattern of there is no geographic structure in individuals IBD without clearly limited subpopulations can spanning southern to northern California. Cali- be expected from a species whose distribution fornia vole movement across landscapes is high, is continuous rather than discrete (Futuyma such that the signal of IBD detected using 1998). Therefore, the unfavorable habitat nuclear microsatellite markers did not lead to a types—those containing woody and/or shrubby signal of phylogeographic structure using mito - vegetation—within JRBP do not represent sig- chondrial DNA. Despite the smaller effective nificant barriers to gene flow. Rather, from a population size of mitochondrial DNA relative genetic perspective, the voles represent a con- to nuclear DNA, gene flow is sufficiently high tinuous population: despite limited dispersal and on both a short- and long-term basis to prevent high philopatry, the movement and abundance significant population structure. 306 WESTERN NORTH AMERICAN NATURALIST [Volume 70

Fig. 5. Haplotype network for the complete mitochondrial cytochrome b nucleotide sequence. Numbers identify unique, detected haplotypes, and small circles without numbers represent inferred haplotypes. Each line represents one mutational step, and additional mutational steps are represented by hatch marks. The size of the circle represents the frequency of haplotype detected, where the smallest circle = 1 individual and the largest circle = 18 individuals. Geo- graphic location of samples is indicated with shading, where lighter circles are more southern (Ventura and Santa Bar- bara counties, haplotype #19–22, 25–28) and darker circles are more northern (Siskiyou County, haplotype #23–24). See Fig. 4 for the pairing of haplotype identity with precise sampling location. Haplotypes represent a panmictic popu- lation with some examples of in situ diversification within localities and of limited grouping by geographic location.

Gene Flow, Demography, and a growing body of evidence that high levels of Landscape Features genetic diversity and connectivity in voles can Many species occur in subdivided popula- be maintained in patchy habitats (Stewart et tions due to spatial heterogeneity of suitable al. 1999, Aars et al. 2006). For example, the water habitat. This fragmentation may be caused by a vole Microtus richardsoni maintains high genetic naturally patchy habitat, as can be the case when diversity despite its restriction to isolated environmental variation leads to discontinuous streamlets (Stewart et al. 1999). On the other distributions of ecosystems or habitat loss from hand, a subspecies of California voles residing in anthropogenic disturbance. Our data supports fragmented marshes demonstrates significant 2010] GENE FLOW ACROSS SPATIAL SCALES 307

Fig. 6. Frequency of outcomes in Bayesian serial coalescent modeling that match the observed level of mitochondrial gene diversity. Results are shown for 3 mutation rates at a range of effective population sizes. For a high mutation rate, the effective population size most likely matching the observed diversity is 120,000, while for a low mutation rate, the effective population size is approximately 400,000. population structure (Neuwald 2010). Especially large effective population size could prevent for species which occur in spatially discrete the development of both local and regional subpopulations, genetic structure can be com- genetic structure as long as corridors of migra- plex (Mayer et al. 2009) and seems to vary tion remain accessible. And in fact, the large depending on the population in consideration. deme size for individuals from JRBP indicates Fragmentation can constrain gene flow, which a high amount of genetic exchange with out- serves to increase the rate of genetic drift and side populations. A large effective population inbreeding (Frankham et al. 2002) over time. size as a factor overriding high philopatry in These processes decrease genetic diversity, shaping genetic structure has been suggested which can have direct impacts on the fitness of in kangaroo rats (Dipodomys spp.; Waser and a species (Soule 1980, Coltman et al. 1998) be - Elliott 1991). While the overall effective popu- cause genetic diversity and evolutionary poten- lation size has likely remained high through tial of a species are correlated (Lavergne and time in M. californicus, local census popula- Molofsky 2007). It is therefore important, from tions are known to experience boom-and-bust a conservation perspective, to describe the spa- cycles. Our JRBP study uncovers a snapshot of tial population structure of natural populations, the local genetic structure, and without trap- which can inform the extant level of gene flow, ping over time, we are unable to determine the in order to predict how species may respond to population abundance at the time of collection further habitat fragmentation. relative to the long-term average. However, The maintenance of high genetic diversity the lack of population structure across geo- at JRBP despite the spatial and demographic graphic distances spanning 2 km supports the characteristics that would limit genetic diver- conclusion that our sampling spanned a period sity indicate that M. californicus has a large of high abundance (Bowen 1982). effective population size at this site. And in Our estimates of gene flow are largely fact, findings from Bayesian serial coalescent concordant with indirect dispersal distances modeling corroborated these findings. Such a based on mark-recapture demographic studies. 308 WESTERN NORTH AMERICAN NATURALIST [Volume 70

Demographic studies have found dispersal dis- distinct morphological features in different areas tances to be quite small, predominantly less around the bay. In the case of the California than 10 m (Lidicker 1973, Salvioni and Lidicker vole, subspeciation status was defined based on 1995), but a few individuals were caught at phenotypic differences of pelage, size of brain distances of 50 m from their home range (Bowen case/skull, size of rostrums and nasals, and over- 1982). Other studies have shown genetic esti- all size (length). Four subspecies were delimited mates to be in general concordance (Gauffre et in the San Francisco Bay Area. As is typical for al. 2008) to nearly half that of empirical esti- other phylogeographic analyses based on neutral mates (Broquet et al. 2006). How dispersal dis- genetic differences, subspeciation distinctions tances differ depending on population sizes is based on phenotype were not detected in this an interesting component to this analysis. That study, indicating that the species is either phe- voles may disperse different distances depend- notypically plastic or, more likely, prone to lo- ing on the size of the population has been cal adaptations or drifts of the phenotype. As suggested for the California vole based on recently as 18,000 years ago, however, the San demographic sampling (e.g., Heske 1987), with Francisco Bay was a large continuous grassy the expectation that dispersal is a stochastic valley (Atwater 1979), and the California vole process with higher overall success during would have been linked to neighboring regions. periods of high abundance that “wash out” Therefore, the submerge that created the San potential subdivisions based on limited vagility Francisco Bay has led to phenotypic differ- and high philopatry (Bowen 1982). Here, how- ences but has not resulted in phylogeographic ever, genetic analysis from Rousset’s a indicates isolation, as the individuals from JRBP on the that a longer dispersal distance might be more San Francisco peninsula are part of a panmic- common and successful during periods of low tic northern California population. Neverthe- density, since dispersal distance was nearly less, the lack of any overlap in haplotypes 2.5 times lower during periods of high density between the JRBP population and the larger (9.8 m at high density versus 23.8 m at low geographic range, without any unique structure density). Other studies have found that a lower for the JRBP population within the northern population size leads to a higher probability of range, indicates that California voles demon- successful distal gene flow (Andreassen and Ims strate genetic differences, differences that are at 2001, Hadly et al. 2004). Dispersal may then least partially based on geographic distance, but mimic and be linked to the probability of repro- that the isolation is either too recent or incom- ductive success, which in voles can also be plete, or both, to result in marked differences higher when density is lower (Krebs 1966). If in phylogeographic structure. In comparison California voles simultaneously demonstrate a to the genetic pattern of a subspecies of Cali- pattern of high reproduction (Krebs 1966) and fornia vole that has had its habitat limited by high dispersal (this study) during low popula- geologic change (Neuwald 2010), the genetic tion density, then this could result in sustained pattern seen in JRBP individuals indicates that or even elevated gen etic admixture during un - the population has had an uninterrupted pres- favorable conditions. Such a trait may prove ence in the area. It will be interesting to highly beneficial to a species experiencing en- determine whether the partitioning of vole vironmental change. In contrast to voles, for populations around the Bay has lead to ge- example, do not disperse readily dur- netic subdivision detectable by nuclear micro - ing years of low abundance, and therefore, satellites, since this pattern was not detected environmental changes which reduce popu- in the slower-evolving mitochondrial genome. lation size have a much higher chance of en - The current genetic sampling technique for dangering local populations (Hadly et al. 2004). phylogeographic analysis heavily undersamples existing haplotype diversity, which depends on Phylogeography sample size and overall genetic diversity. From Over longer time scales, the San Francisco this one locality, we detected 11 unique haplo- Bay has acted a barrier to dispersal, and the types. While some of these haplotypes indicate result of this barrier on the speciation and in situ diversification, other variants are as subspeciation of rodents has been studied pre- closely related to other JRBP individuals as viously (Hooper 1944). Microtus californicus, they are to individuals from more distant locali- like many rodents of the area, demonstrated ties. Because of the lack of overall structure in 2010] GENE FLOW ACROSS SPATIAL SCALES 309 this species across its northern distribution, in - to RIA from the Andrew W. Mellon Foundation creased sampling did not alter phylogeographic through Jasper Ridge Biological Preserve and conclusions; however, this may not be the case by NSF grant #EAR-0719429 to EAH. when structure is present. In conclusion, linking population genetics LITERATURE CITED and phylogeography has the potential to eluci- date important components of a species’ ecology AARS, J., J.F. DALLAS, S.B. PIERTNEY, F. MARSHALL, J.L. GOW, S. TELFER, AND X. LAMBIN. 2006. Widespread and evolutionary history (e.g., genetic response gene flow and high genetic variability in populations of populations to perturbations such as climate of water voles Arvicola terrestris in patchy habitats. change). By combining fine-scale genetic analy- Molecular Ecology 15:1455–1466. ses with longer-term phylogeographic signals in ANDREASSEN, H.P., AND R.A. IMS. 2001. Dispersal in patchy vole populations: role of patch configuration, the California vole, we are able to show that density dependence, and demography. Ecology 82: gene flow is consistently high across multiple 2911–2926. spatial scales. Despite potential impediments to ATWATER, B.F. 1979. Ancient processes at the site of gene flow, such as dispersal distance and the Southern San Francisco Bay: movement of the crust fragmentation of suitable habitat, the California and changes in sea level. Pages 31–45 in T.J. Cono- mos, editor, San Francisco Bay: the urbanized estuary. vole represents a continuous population within Pacific Division/American Association for the Advance- its northern range, marked by isolation-by-dis- ment of Science, San Francisco, CA. tance but no vicariances. Overall, we are en - BERTHIER, K., N. CHARBONNEL, M. GALAN, Y. CHAVAL, AND couraged by what we see at JRBP: despite J.F. COSSON. 2006. Migration and recovery of the genetic diversity during the increasing density phase habitat encroachment from development, the in cyclic vole populations. Molecular Ecology 15: population shows no signs of bottlenecking or 2665–2676. reduced genetic diversity, and the population BERTHIER, K., M. GALAN, J.C. FOLTETE, N. CHARBONNEL, remains continuous within the JRBP borders AND J.F. COSSON. 2005. Genetic structure of the and, at least in the past, remained connected to cyclic fossorial water vole (Arvicola terrestris): land- scape and demographic influences. Molecular Ecology other populations throughout its northern range. 14:2861–2871. Disturbances to the favored grassland habitat BOWEN, B.S. 1982. Temporal dynamics of microgeographic at JRBP are both natural—in the form of habi- structure of genetic variation in Microtus californicus. tat diversity creating a potential barrier to mi- Journal of Mammalogy 63:625–638. BROQUET, T., C.A. JOHNSON, E. PETIT, I. THOMPSON, F. gration—and anthropogenic—in the form of BUREL, AND J.M. FRYXELL. 2006. Dispersal and genetic conversion of native grassland into one domi- structure in the American marten, Martes americana. nated by invasive grasses. Yet the rodents ap - Molecular Ecology 15:1689–1697. parently move across both grassland types and CLEMENT, M., D. POSADA, AND K.A. CRANDALL. 2000. TCS: a computer program to estimate gene genealogies. through unfavorable habitats. The demographic Molecular Ecology 9:1657–1659. and behavioral characteristics of this species— COLTMAN, D.W., W.D. BOWEN, AND J.M. WRIGHT. 1998. high philopatry and short-distance dispersal— Birth weight and neonatal survival of harbour seal create a moderate but distinct pattern of pups are positively correlated with genetic variation increasing genetic difference with increasing measured by microsatellites. Proceedings of the Royal Society of London Series B, Biological Sciences 265: geographic distance, but without sharp bound- 803–809. aries leading to delimited populations, both on COLWELL, R.K. 2006. EstimateS: statistical estimation of a local and regional scale. species richness and shared species from samples. Version 8.0 user’s guide and application. Available from: http://viceroy.eeb.uconn.edu/EstimateS ACKNOWLEDGMENTS COLWELL, R.K., C.X. MAO, AND J. CHANG. 2004. Interpolat- ing, extrapolating, and comparing incidence-based We are grateful to Jennifer Neuwald and species accumulation curves. Ecology 85:2717–2727. Chris Conroy for valuable discussion and assis- CONROY, C.J., AND J.L. NEUWALD. 2008. Phylogeographic tance, to Nona Chiariello for providing research study of the California vole, Microtus californicus. support, to Hillary Young for spearheading the Journal of Mammalogy 89:755–767. CORNUET, J.M., AND G. LUIKART. 1996. Description and small-mammal survey, and to Trevor Hebert power analysis of two tests for detecting recent popu - for assistance with maps and GIS software. lation bottlenecks from allele frequency data. Genet- We thank Yvonne Chan, Jessica Blois, Sarah ics 144:2001–2014. McMenamin, and an anonymous reviewer for DELPORT, W., J.W. FERGUSON, AND P. BLOOMER. 2005. Characterization of six microsatellite loci in the African helpful comments on previous versions of the wild silk moth (Gonometa postica, Lasiocampidae). manuscript. Funding was provided by a grant Molecular Ecology Notes 5:860–862. 310 WESTERN NORTH AMERICAN NATURALIST [Volume 70

EVANNO, G., S. REGNAUT, AND J. GOUDET. 2005. Detecting HARTL, D.L., AND A.G. CLARK. 1997. Principles of popula- the number of clusters of individuals using the soft- tion genetics. 3rd edition. Sinauer Associates, Sun- ware STRUCTURE: a simulation study. Molecular derland, MA. Ecology 14:2611–2620. HESKE, E.J. 1987. Spatial structuring and dispersal in a EXCOFFIER, L., J. NOVEMBRE, AND S. SCHNEIDER. 2000. high density population of the California vole Micro- SIMCOAL: a general coalescent program for the tus californicus. Holarctic Ecology 10:137–148. simulation of molecular data in interconnected popu- HOBBS, R.J., S. YATES, AND H.A. MOONEY. 1997. Long-term lations with arbitrary demography. Journal of Heredity data reveal complex dynamics in grassland in relation 91:506–509. to climate and disturbance. Ecological Monographs FALUSH, D., M. STEPHENS, AND J.K. PRITCHARD. 2003. 77:545–568. Inference of population structure using multilocus HOOPER, E.T. 1944. San Francisco Bay as a factor influenc- genotype data: linked loci and correlated allele fre- ing speciation in rodents. University of Michigan quencies. Genetics 164:1567–1587. Press, Ann Arbor, MI. FRANKHAM, R., J.D. BALLOU, AND D.A. BRISCOE. 2002. ISHIBASHI, Y., Y. YOSHINAGA, T. SAITOH, S. ABE, H. IIDA, AND Introduction to conservation genetics. Cambridge M.C. YOSHIDA. 1999. Polymorphic microsatellite DNA University Press, Cambridge. markers in the field vole Microtus montebelli. Mole- FRANTZ, A.C., S. CELLINA, A. KRIER, L. SCHLEY, AND T. cular Ecology 8:163–164. BURKE. 2009. Using spatial Bayesian methods to KELLOGG, R. 1918. A revision of the Microtus californicus determine the genetic structure of a continuously group of meadow mice. University of California Pub- distributed population: clusters or isolation by dis- lications in Zoology 21:1–42. tance? Journal of Applied Ecology 46:493–505. KREBS, C.J. 1966. Demographic changes in fluctuating FUTUYMA, D.J. 1998. Evolutionary biology. 3rd edition. populations of Microtus californicus. Ecological Mono- Sinauer Associates, Sunderland, MA. graphs 36:239–273. GAUFFRE, B., A. ESTOUP, V. BRETAGNOLLE, AND J.F. COSSON. LAVERGNE, S., AND J. MOLOFSKY. 2007. Increased genetic 2008. Spatial genetic structure of a small rodent in a variation and evolutionary potential drive the suc- heterogeneous landscape. Molecular Ecology 17:4619– cess of an invasive grass. Proceedings of the National 4629. Academy of Sciences of the United States of America GILL, A.E. 1980. Partial reproductive isolation of sub- 104:3883–3888. species of the California vole Microtus californicus. LIDICKER, W.Z. 1973. Regulation of numbers in an island Pages 105–118 in N.N. Vorontsov and J.M. Van population of the California vole: a problem in com- Brink, editors, Animal genetics and evolution: munity dynamics. Ecological Monographs 43:271– selected papers of the 14th International Congress 302. of Genetics. BV Publishers: The Hague, Nether- LOWE, A., S. HARRIS, AND P. ASHTON. 2004. Ecological lands; Moscow, USSR. 21–30 August 1978. genetics: design, analysis, and application. Blackwell GIUSTI, G.A. 1985. Home ranges, activity patterns and Science, Oxford. relative densities of meadow mice (Microtus califor- MAYER, C., K. SCHIEGG, AND G. PASINELLI. 2009. Patchy nicus) and harvest mouse (Reithrodontomys megalo- population structure in a short-distance migrant: tis) on Jasper Ridge Biological Preserve. Master’s evidence from genetic and demographic data. Mole- thesis, San Francisco State University, San Francisco, cular Ecology 18:2353–2364. CA. NEI, M. 1987. Molecular evolutionary genetics. Columbia GOTELLI, N.J., AND R.K. COLWELL. 2001. Quantifying biodi- University Press, New York, NY. versity: procedures and pitfalls in the measurement NEUWALD, J.L. 2010. Population isolation exacerbates and comparison of species richness. Ecology Letters conservation genetic concerns in the endangered 4:379–391. Armagosa vole, Microtus californicus scirpensis. Bio- GREENWALD, G.S. 1957. Reproduction in a coastal Califor- logical Conservation 143:2028–2038. nia population of the field mouse, Microtus californi- PEAKALL, R., AND P.E. SMOUSE. 2006. GENALEX 6: genetic cus. University of California Publications in Zoology analysis in Excel. Population genetic software for 54:421–446. teaching and research. Molecular Ecology Notes GUILLOT, G., F. MORTIER, AND A. ESTOUP. 2005. GENE - 6:288–295. LAND: a computer package for landscape genetics. PIRY, S., G. LUIKART, AND J.M. CORNUET. 1999. BOTTLE- Molecular Ecology Notes 5:712–715. NECK: a computer program for detecting recent GUILLOT, G., F. SANTOS, AND A. ESTOUP. 2008. Analysing reductions in the effective population size using allele georeferenced population genetics data with Gene- frequency data. Journal of Heredity 90:502–503. land: a new algorithm to deal with null alleles and a PRITCHARD, J.K., M. STEPHENS, AND P. DONNELLY. 2000. friendly graphical user interface. Bioinformatics 24: Inference of population structure using multilocus 1406–1407. genotype data. Genetics 155:945–959. HADLY, E.A., U. RAMAKRISHNAN, Y.L. CHAN, M. VAN TUINEN, PRITCHARD, J.K., X. WEN, AND D. FALUSH. 2007. Documen- K. O’KEEFE, P.A. SPAETH, AND C.J. CONROY. 2004. tation for structure software. Version 2.2. Available Genetic response to climatic change: insights from from: http://pritch.bsd.uchicago.edu/structure.html ancient DNA and phylochronology. PLoS Biology 2: RITLAND, K. 1996. Estimators for pairwise relatedness and 1600–1609. individual inbreeding coefficients. Genetical Research HALL, E.R. 1981. The mammals of North America. 2nd edi- 67:175–185. tion. Wiley, New York, NY. ROUSSET, F. 1997. Genetic differentiation and estimates of HARDY, O.J., AND X. VEKEMANS. 2002. SPAGeDi: a versatile gene flow from F-statistics under isolation by dis- computer program to analyse spatial genetic structure tance. Genetics 145:1219–1228. at the individual or population levels. Molecular ______. 2000. Genetic differentiation between individuals. Ecology Notes 2:618–620. Journal of Evolutionary Biology 13:58–62. 2010] GENE FLOW ACROSS SPATIAL SCALES 311

______. 2008. GENEPOP007: a complete re-implementation and glacially dynamic landscape of the Central Rocky of the GENEPOP software for Windows and Linux. Mountains. Journal of Mammalogy 90:571–584. Molecular Ecology Resources 8:103–106. STEWART, W.A., S.B. PIERTNEY, F. MARSHALL, X. LAMBIN, SALVIONI, M., AND W.Z. LIDICKER. 1995. Social organization AND S. TELFER. 1999. Metapopulation genetic struc- and space use in California voles: seasonal, sexual, and ture in the water vole, Arvicola terrestris, in NE Scot- age-specific strategies. Oecologia 101:426–438. land. Biological Journal of the Linnean Society 68: SMITH, M.F., AND J.L. PATTON. 1993. The diversification of 159–171. South American murid rodents: evidence from mito- SWOFFORD, D.L. 2003. PAUP*: phylogenetic analysis using chondrial DNA sequence data for the akodontine tribe. parsimony (*and other methods). Version 4. Sinauer Biological Journal of the Linnean Society 50:149– Associates, Sunderland, MA. 177. VAN DE ZANDE, L., R.C. VAN APELDOORN, A.F. BLIJDEN- SMOUSE, P.E., R. PEAKALL, AND E. GONZALES. 2008. A STEIN, D. DE JONG, W. VAN DELDEN, AND R. BIJLSMA. heterogeneity test for fine-scale genetic structure. 2000. Microsatellite analysis of population structure Molecular Ecology 17:3389–3400. and genetic differentiation within and between popu- SOULE, M.E. 1980. Thresholds for survival: maintaining lations of the root vole, Microtus oeconomus, in the fitness and evolutionary potential. Pages 151–170 in Netherlands. Molecular Ecology 9:1651–1656. M.E. Soule and B.A. Wilcox, editors, Conservation WASER, P.M., AND L.F. ELLIOTT. 1991. Dispersal and ge - biology: an evolutionary-ecological perspective. Sin- netic structure in kangaroo rats. Evolution 45:935– auer Associates, Sunderland, MA. 943. SPAETH, P.A., M. VAN TUINEN, Y.L. CHAN, D. TERCA, AND E.A. HADLY. 2009. The phylogeography of the long- Received 27 August 2009 tailed vole, Microtus longicaudus, in the tectonically Accepted 19 February 2010