SciELO - Scientific Electronic Library Online

 
vol.14 número1Taxonomic reassessment of the Little pocket mouse, Perognathus longimembris (Rodentia, Heteromyidae) of southern California and northern Baja CaliforniaAn 1896 specimen helps clarify the phylogenetic placement of the Mexican endemic Hooper’s deer mouse índice de autoresíndice de assuntospesquisa de artigos
Home Pagelista alfabética de periódicos  

Serviços Personalizados

Journal

Artigo

Indicadores

Links relacionados

  • Não possue artigos similaresSimilares em SciELO

Compartilhar


Therya

versão On-line ISSN 2007-3364

Therya vol.14 no.1 La Paz Jan. 2023  Epub 31-Jul-2023

https://doi.org/10.12933/therya-23-2236 

Special contributions

Revisiting species delimitation within Reithrodontomys sumichrasti (Rodentia: Cricetidae) using molecular and ecological evidence

Elizabeth Arellano1  * 

Ana L. Almendra1 

Daily Martínez-Borrego1 

Francisco X. González-Cózatl1 

Duke S. Rogers2 

1 Centro de Investigación en Biodiversidad y Conservación, Universidad Autónoma del Estado de Morelos. Avenida Universidad 1001, CP. 62209, Cuernavaca. Morelos, México. Email: elisabet@uaem.mx (EA), al.almendra@gmx.com (ALA), daily.marbo@gmail.com (DM-B), xavier@uaem.mx (FXG-C).

2 Life Science Museum and Department of Biology, Brigham Young University. CP. 84602, Provo. Utah, U.S.A. Email: duke_rogers@byu.edu (DSR).


Abstract

Reithrodontomys sumichrasti is distributed from central México to Panama. Previous studies using DNA sequences suggest the existence of distinct clades that may deserve species-level recognition. Here, we use multiple methods of species delimitation to evaluate if this taxon is a complex of cryptic species. DNA sequences from the genes Cyt-b, Fgb-I7, and Acp5 were obtained from GenBank to perform molecular analyses. Species boundaries were tested using the bGMYC, STACEY, and BPP species delimitation methods. Divergence times were estimated as well as the Cyt-b genetic distances. We developed Ecological Niche Models and tested hypotheses of niche conservatism. Finally, we estimated the spatiotemporal history of lineage dispersal. The bGMYC proposed two species while STACEY and BPP proposed 4 species (genetic distances ranged from 5.43 % to 7.52 %). The ancestral position of clade I was recovered, with a Pleistocene diversification time within R. sumichrasti at ~2.15 Ma. For clade pairwise niche comparisons, the niche identity hypothesis was rejected. The ancestral distribution of R. sumichrasti was centered in Central America and spread to the west crossing the Isthmus of Tehuantepec and extending to the mountain regions of Central México. Our taxonomic considerations included the recognition of four clades as distinct species within R. sumichrasti.

Keywords: Cryptic species; harvest mice; integrative taxonomy; Mesoamerican highlands; phylogeographic patterns

Resumen

Reithrodontomys sumichrasti se distribuye desde el centro de México hasta Panamá. Estudios previos con secuencias de ADN sugieren la existencia de clados distintos y su posible reconocimiento como especies. En este estudio, probamos diferentes métodos de delimitación de especies para evaluar si este taxón constituye un complejo de especies crípticas. Las secuencias de ADN de los genes Cyt-b, Fgb-I7 y Acp5 fueron descargadas de GenBank y utilizadas en análisis moleculares. Los límites de especies fueron probados utilizando los métodos de delimitación bGMYC, STACEY y BPP. Se estimaron tiempos de divergencia y distancias genéticas para el gen Cyt-b. Además, construimos Modelos de Nicho Ecológico y probamos hipótesis de conservadurismo de nicho. Finalmente, reconstruimos la historia espaciotemporal de la dispersión de los linajes. El bGMYC propuso dos especies, mientras que STACEY y BPP propusieron 4 especies (las distancias genéticas oscilaron entre 5.43 % y 7.52 %). Se recuperó la posición ancestral del clado I, ubicando en el Pleistoceno la diversificación dentro de R. sumichrasti, hace ~2.15 Ma. En las comparaciones de nicho por pares de clados fue rechazada la hipótesis de identidad de nicho. La distribución ancestral de R. sumichrasti se centró en América Central desde donde comenzó a extenderse hacia el oeste cruzando el Istmo de Tehuantepec y extendiéndose hacia las regiones montañosas del centro de México. Nuestras consideraciones taxonómicas incluyeron el reconocimiento de cuatro clados como especies distintas dentro de R. sumichrasti.

Introduction

A special issue of Therya dedicated to Dr. Alfred L. Gardner for his long research career on the diversity of neotropical mammals, especially for his work in México, honors this outstanding scientist by contributing important advances to the knowledge of mammalogy. Our contribution adds to the mission of modern systematic biology: the discovery, description, and classification of the biodiversity on the planet from an evolutionary perspective (Daly et al. 2012). This task involves subjects under debate over the past three decades, such as the species concept (what a species is) and species delimitation (how a species is recognized). Both subjects are closely related but conveniently divided for practical applications (see review by de Queiroz 2007), and over time, species delimitation has taken priority over species concepts (Sites and Marshall 2003, 2004). Given the current rate of species loss, it is urgent to accurately delimit species inasmuch they are the fundamental unit in studies of ecology, systematic, and conservation biology, among other research areas. From the evolutionary standpoint, species delimitation includes the understanding of population-level mechanisms that can be complex (Huang 2020). Populations differentiation through multiple stages at different rates, in part dependent on factors such as generation time, selection pressure, and gene flow. Tracing the process with an acceptable level of certainty depends on the use of appropriate markers (preferably multiple and independent) and the criteria of evaluation (de Queiroz 2007). One of the most reliable strategies is to use multiple sources of evidence (morphology, genetics, ecology, geography, among others) and to base conclusions on their consistency (Knowles and Carstens 2007; Rissler and Apodaca 2007; Carstens et al. 2013).

There are both regions as well as biological groups, which are amenable to test hypotheses about species delimitation. The Mesoamerican region has been repeatedly used as a study model because of its complex physiography and biogeographical history, which is reflected by high biological diversity, including many endemic species (Myers et al. 2000), particularly for highland groups. As for groups of organisms, rodents, reptiles, and insects, among others have served as models to test hypotheses about evolutionary patterns and processes (e. g.Doody et al. 2009; Gilbert and Manica 2015; Maestri et al. 2017). Some species of rodents have been assessed by evaluating their phylogenetic relationships and further used to illuminate the vicariant biogeography of Mesoamerica (e. g.Sullivan et al. 2000; Leon-Paniagua et al. 2007; Almendra et al. 2018; León-Tapia et al. 2021). Such is the case of Reithrodontomys sumichrasti (Family Cricetidae; Bradley 2017), with a particular interest in the high levels of intraspecific divergence reported (Sullivan et al. 2000; Urbina et al. 2006; Hardy et al. 2013).

Reithrodontomys sumichrasti is distributed along the highlands of Mesoamerica, from central México at 1,200 masl to Panama above 3,400 masl, inhabiting temperate pine-oak and cloud forests. Seven subspecies are recognized, which are distributed in three disjunctive spots (Hooper 1952; Hall 1981; Figure 1). The range of R. s. sumichrasti includes portions of the Sierra Madre Oriental, the Mexican Transvolcanic Belt, and the Oaxacan Highlands (type locality El Mirador, Veracruz, México). The distribution of R. s. nerterus is restricted to the west portion of the Mexican Transvolcanic Belt (type locality Nevado de Colima, Jalisco, México) whereas R. s. luteolus is found in the Sierra Madre del Sur (type locality Juquila, Oaxaca, México). R. s. dorsalis occurs in the mountains of the Mexican states of Chiapas and Guatemala (type locality Tonicapan, Guatemala) and R. s. modestus in the highlands of El Salvador, Honduras, and western Nicaragua (type locality Jinotega, Nicaragua). The southernmost distribution of the species includes the Cordillera Central and Cordillera de Talamanca in Costa Rica for R. s. australis (type locality Cartago, Costa Rica) and the extreme east of Costa Rica and high mountains of western Panama for R. s. vulcanius (type locality Chiriquí, Panama; Hooper 1952).

Previous phylogenetic studies using DNA sequences of the mitochondrial Cytochrome b (Cyt-b) gene (Sullivan et al. 2000), or also incorporating the seventh intron of nuclear gene beta-fibrinogen (Fgb-I7) and the second intron of the acid phosphatase type V (Acp5; Hardy et al. 2013) have revealed the existence of several distinct clades that may deserve species-level recognition. Lineages on either side of the Isthmus of Tehuantepec in México were proposed as distinct biological species, but this pattern has been supported by only mtDNA sequences (Sullivan et al. 2000; Hardy et al. 2013). Although it was difficult to elucidate the relationships among networks of populations from central México (Hardy et al. 2013; Figure 2), there was a clear pattern of phylogenetic structure.

Here, we evaluate species delimitation within R. sumichrasti using different methods of analysis than those previously employed to test the hypothesis that R. sumichrasti represents a complex of cryptic species. We also comment on the diversification processes in the region and make taxonomic suggestions.

Figure 1 Map of México and Central America (adapted from Hall [1981] and Hardy et al. [2013]) showing geographic distribution of the seven recognized subspecies of Reithrodontomys sumichrasti. Dots represent the localities used in this study and follow the clade-color distinction described in Figure 2. 

Materials and methods

Data acquisition. DNA sequences from the mitochondrial gene Cyt-b, and the Fgb-I7 and Acp5 nuclear genes, representing Hardy et al. (2013) populations dataset of Reithrodontomys sumichrasti (n = 226) were obtained from GenBank. We sequenced an additional 11 specimens of R. sumichrasti, five of these from three new geographic localities (64 to 66; Appendix 1). Given the current availability of sequence data for outgroup taxa, we included samples of R. zacatecae, R. megalotis, R. chrysopsis, R. humulis, R. montanus, and R. raviventris from the R. megalotis species group (Musser and Carleton 2005). The updated DNA datasets were realigned with MAFFT v7 [L-INS-i refinement, gap penalty = 3, offset = 0.5] (Katoh et al. 2005) for nuclear markers, and manually for Cyt-b using Geneious Pro v6.1.6 (https://www.geneious.com). The optimal partition scheme (by gene) and models of nucleotide substitution (Cyt-b: GTR+I+G, Fgb-I7: HKY+I+G, Acp5: K80+I+G); were determined with Partition Finder (Lanfear et al. 2014).

Phylogenetic hypothesis. We considered the phylogenetic relationships proposed by Hardy et al. (2013) as our working hypothesis, where two geographic clades are supported as species-level lineages. One species (spA) split ~2.5 million years ago (Ma) and comprises populations from Chiapas south into Central America (clade I; Figure 2). Species (spB) includes 3 haplogroups restricted to México, west of the Isthmus of Tehuantepec (Figure 2), whose most recent common ancestor was placed ~1.36 Ma (see Hardy et al. 2013). To assess support for this phylogenetic hypothesis (Hardy et al. 2013), and for alternative topological arrangements, we applied three methods for assessing species boundaries and species tree estimation (see below) that do not require a guide topology or species assignments to be specified a priori.

Single locus species delimitation. A time-calibrated Bayesian Inference (BI) analysis of Cyt-b for R. sumichrasti samples was run in BEAST2 v.2.6.2 (Bouckaert et al. 2014). We employed a prior rate of evolution of 0.017 substitutions per site per million years (Arbogast et al. 2002) and fossil calibrations (R. moorei, R. wetmorei, R. galushai, R. pratincola, R. rexroadensis, and R. sp.) with an offset of exponential prior for the age (in Ma) of the root (mean = 2.25, offset = 1.3, HD = 95 % between 1.5 to 5.5 Ma; Dalquest 1978; Czaplewski 1987; Martin et al. 2002; Morgan and White 2005; Lindsay and Czaplewski 2011; Martin and Peláez-Campomanes 2014). BI analysis consisted of four Markov chain Monte Carlo (MCMC) chains of 10 million generations, sampling trees every 1,000 generations and with a burn-in of 20 % of the trees. The last 100 trees sampled from each run were analyzed with 1 million generations of the Bayesian General Mixed Yule-Coalescent (bGMYC) model (Reid and Carstens 2012) in the computing environment R (R Core Team 2018). As advised by Reid and Carstens (2012), outgroup taxa were not included in this analysis. For all Bayesian analyses reported herein, stabilization and appropriate Effective Sample Sizes (ESS ≥ 200) of the posterior distributions for model parameters were examined in Tracer 1.8 (Rambaut et al. 2018).

Figure 2 a) Map of México and Central America adapted from Hardy et al. (2013) showing collecting localities of Reithrodontomys sumichrasti superimposed on a map of the physiographic provinces they occupy. The four clades detected by the authors are demarcated with the colors purple (clade I), blue (clade II), red (clade III), and green (clade IV). Newly incorporated localities are shown as black dots (64-66; Appendix 1). b) Close-up of the area of sympatry of individuals from populations between clade II and clade III. c) Standing time-calibrated phylogenetic hypotheses of the evolutionary relationships among clades within the currently recognized extent of R. sumichrasti. Uncorrected Cytochrome-b genetic distances between sister clades are denoted in parentheses as a reference for the level of molecular divergence. 

Time-calibrated multiple loci species delimitation. The multiple loci multiple species dataset was analyzed simultaneously with the multi-tree multi-species coalescent method (Heled and Drummond 2010) and the assignment-free species delimitation technique implemented in STACEY (Jones 2017), using BEAST2. The search strategy implemented in STACEY uses a birth-death-collapse prior to approximate alternative delimitation models and node re-height MCMC move that aims to improve the convergence of the species tree estimation, therefore, its performance is subject to the accuracy of divergence times estimation. As recommended, the analysis was run twice, the second time sampling from the prior only; for 100 million generations, trees were sampled every 5,000 generations. A Fossilized Birth-Death model was set on the speciation rate (Heath et al. 2014), time-calibrated as specified above. Topologies and clock rates from individual loci were left unlinked, and substitution rates among branches were drawn from a log-normal distribution with a prior mean rate of 0.017 substitutions per site per million years for the Cyt-b (Arbogast et al. 2002).

Clock-like multiple loci species delimitation. We assessed the probability of alternative species delimitation models and species trees with the Bayesian Phylogenetics and Phylogeography method (BPPv3.2; Yang and Rannala 2014). This assumes a Jukes-Cantor evolutionary model (strict molecular clock) and applies a species tree search strategy that is grounded on the Nearest Neighbor Interchange (NNI) algorithm, followed by its characteristic rjMCMC move. Although it accounts for the uncertainty on estimated rates of evolution compared to *BEAST-STACEY, this method is applicable to inter- and intra-species datasets that meet the criteria of having clock-like evolutionary rates. For this analysis, uniform rooted species trees were assumed, with gamma priors for the population size (α, β) of Θ = (2, 2000) and root age (Tau = τ) τ0 = (4, 2, and 1). The rjMCMC was run with algorithm A11 with fine-tune parameter ε _= 2 (joint unguided species delimitation and species tree inference) for 500,000 generations with a sampling frequency of 200 after a burn-in period of 10,000.

Genetic distances. Cyt-b genetic distances using the Kimura 2-parameter (K2P; Kimura 1980) and the uncorrected P-distances were estimated between and within clades suggested as distinct species using MEGA X (Kumar et al. 2018). This allowed us to make genetic distance comparisons with other values reported for rodents and for R. sumichrasti by Bradley and Baker (2001) and Hardy et al. (2013), respectively.

Ecological niche equivalence. For each species-level clade (clades I-IV, see Results section), we developed present-time Ecological Niche Models (ENMs) with MAXENT 4. (Phillips and Dudik 2008). Correlation between the 19 environmental variables from the WORLDCLIM database (1 km2 resolution; Hijmans et al. 2005) was calculated with ENMtools v1.4.1 (Warren et al. 2010). Then, 9 environmental variables (correlation = r ≤ 0.80) and presence points confirmed with molecular data (Appendix 1) were employed to obtain the ENMs. For clades I-III, 10 bootstrap replicates of presence/background points assigning 15 % of the presence points for training were applied. For clade IV, 10-fold cross-validation replicates were applied because of the limited number of presence records.

To test the hypothesis of niche conservatism between the ENMs from sister clades, a null distribution of 99 estimates of the I Statistics (Warren et al. 2008) and the Schoener’s D (Schoener 1968) measures of niche overlap was generated for each pair of sister clades with the R package DISMO (Hijmans et al. 2017). In addition, a canonical discriminant function (CF) analysis was executed with the package candisc (Friendly and Fox 2015), to distinguish the potential affecting the extent to which their niches have been conserved. For this analysis, current time ENMs were reclassified so that each pixel predicted by each model would equal 1 and the rest of the grid 0. The resultant ENM masks were used to extract for each clade pixel-level data for the 9 environmental variables.

Lineage dispersal. To reconstruct the spatiotemporal history of lineage dispersal in R. sumichrasti we used the Relaxed Random Walk model (RRW; Lemey et al. 2010) as implemented in BEAST2. This model assumes an uncorrelated diffusion rate across the tree and infers the dispersal lineage history in space and time simultaneously, using both the phylogenetic tree and the geographic locations of the samples (Dellicour et al. 2021). To build the RRW we employed the geographic coordinates from each terminal collecting locality as a two-dimensional trait. We assumed a relaxed molecular clock (prior rate = 0.017, SD = 1.0), and the tree priors were calibrated as described above. To visualize the estimated phylogeographic reconstruction, space-time dispersal networks were created using SPREAD 1.0.6 (Bielejec et al. 2011).

Results

Phylogenetic hypothesis and species delimitation. The bGMYC species delimitation analysis of the Cyt-b recovered two species-level clades within R. sumichrasti (P ≥ 0.95), separated by the Isthmus of Tehuantepec (Figure 3; Hypothesis 1). In this phylogeny, samples from new populations 64 to 66 from Guerrero and Oaxaca formed part of clade II. For the BPP and STACEY multiple-loci methods, the highest probability values (BPP, pP = 0.56; STACEY, pP = 0.91) supported Hypothesis five which recovered four divergent clades at the species level (Figure 3). One of them (clade I) was confined to the east and south of the Isthmus of Tehuantepec in México and Central America and the other three (clades II, III, and IV) were restricted to México. The K2P genetic distance values ranged from 5.43 % to 7.52 %, with the lowest value between clades II and IV and the highest between clades I and IV (Table 1). Similar genetic distance values among clades were obtained with the uncorrected P-distances (Table 1).

Table 1 Matrix of mean genetic distances (%) for Cytochrome b gene sequence data among the 4 clades delimited in Reithrodontomys sumichrasti. Values above (uncorrected P-distances) and below (Kimura 2-parameter) the diagonal represent genetic distances between clades. Numbers on the diagonal represent Kimura 2-parameter genetic distances within a clade. 

R. sumichrasti Clade I Clade II Clado III Clado IV
Clade I 1.71 6.69 6.97 7.01
Clade II 7.16 1.66 5.74 5.17
Clade III 7.47 6.07 1.59 6.28
Clade IV 7.52 5.43 6.67 0.25

The species delimitation methods and the species tree (Figure 4) recovered the ancestral position of clade I (pP = 0.84), with a mean divergence time for the most recent common ancestor (MRCA) of ~2.15 Ma. The bGMYC supported the sister relationship between clades II and IV, whereas the multi-loci methods and the species tree supported the split of clade IV (pP = 0.79; mean divergence time 1.42 Ma), and a sister relationship between clades II and III (pP = 0.70; mean divergence time 0.90 Ma). In addition, the ancestral position of R. chrysopsis with respect to R. megalotis-R. zacatecae and R. sumichrasti was strongly supported (pP = 1.00), with an MRCA mean age estimated at 6.18 Ma. Also, a closer relationship was recovered between R. humulis and R. montanus-R. raviventris (pP = 1.00; mean divergence time 6.43 Ma), although the sister relationship of R. montanus-R. ravivientris received lower probabilities (pP = 0.86; mean divergence time 4.44 Ma).

Figure 3 a) Single locus [Hypothesis 1; discontinuous red-yellow heat-map represents the pP ≥ 0.95 of belonging to different species (red color)] and multiple-loci (Hypothesis 2- Hypothesis 5) species delimitation models for Reithrodontomys sumichrasti. Solid and dashed lines denote the species delimitation proposal supported by bGMYC (Hypothesis 1; spA and spB). b) Amount of support for each model in the posterior sample (MCMC) of trees estimated with STACEY and BPP. The abbreviations of the physiographic provinces and clade colors follow Figure 2

Ecological niche equivalence. Ecological Niche Models generated for the four species-level clades within R. sumichrasti had AUC values above 0.90 for training data. The inter-clade predictability of the ENM of clade I ranged from 95 % when predicting known localities from clade III to 100 % when predicting known localities of clade IV (Figure 5). Clade IV had the most restricted ENM, and its inter-clade predictability ranged from 0 % when predicting clade III (and vice versa), to 18 % when predicting clade II. The ENMs of clades II and III showed the lowest intra-clade predictability values with 90 % and 95 %, respectively. Quantification of niche overlap with the I and Schoener’s D statistics (from here forward I and D) revealed small amounts of overlap between each clade pair. For all clade pairwise comparisons, the niche identity (niche equivalency) hypothesis was rejected regardless of the similarity measure (I or D; Table 2).

Table 2 Niche comparisons between sister clades of Reithrodontomys sumichrasti. The I statistics and Schoener’s D represent the observed niche overlap values and the Identity tests represent the comparison of niche equivalency between each clade. 

R. sumichrasti Clade Schoener’s D I statistics Identity test
Clade I II 0.1322 0.3075 niche non-equivalency
III 0.4369 0.7547 niche non-equivalency
IV 0.2722 0.5371 niche non-equivalency
Clade II III 0.3803 0.6456 niche non-equivalency
IV 0.1872 0.3900 niche non-equivalency
Clade III IV 0.0260 0.0843 niche non-equivalency

Figure 4 Time-calibrated species tree estimated with *BEAST-STACEY for Reithrodontomys sumichrasti and the outgroup taxa. Values above branches indicate the mean divergence times (millions of years) and below are the Bayesian posterior probabilities for clades. White bars represent the 95% highest posterior density intervals. Colors follow the clade-color distinction described in Figure 2. Specimens assigned to the collapsed terminal taxa are listed in Appendix 1. 

The canonical variable analysis did not discriminate significantly among the ENMs of the clades (Figure 6). The first and second canonical functions accounted for 97.3 % of the variance and the meaningful structure coefficients (> 0.3) were exclusively related to temperature (BIO1, BIO2, BIO4, BIO5, BIO6, BIO7). Overall, there was more similarity among mean values of each climatic variable between the ENM of clades II and III, whereas the area that occupied clade IV displayed extreme values for the Max Temperature of Warmest Month (BIO5; 27.4 °C), Annual Precipitation (BIO12; 1086 mm), and Precipitation of Driest Quarter (BIO17; 14.86 mm; Table 3).

Table 3 Coefficients of the three first canonical discriminant functions derived from the bioclimatic variables used in the ecological analyses in Reithrodontomys sumichrasti. Mean values of the bioclimatic variables based on the environmental information from occurrence records are given for each clade. 

Climatic Variable Function 1 Eigen=0.261 Function 2 Eigen=0.035 Function 3 Eigen=0.008 Clade I Clade II Clade III Clade IV
BIO1 0.689 0.402 0.028 17.11 16.75 14.15 18.44
BIO2 -0.054 0.409 0.023 11.82 12.23 12.18 12.96
BIO4 0.632 0.086 0.021 104.09 124.54 185.44 164.25
BIO5 0.239 0.486 0.379 24.88 25.24 23.17 27.47
BIO6 -0.385 0.280 0.252 9.00 8.16 4.40 8.30
BIO7 -0.614 0.671 0.015 15.88 17.09 18.77 19.17
BIO11 0.421 0.149 0.619 15.70 15.11 11.64 16.16
BIO12 -0.257 0.116 0.056 1723.79 1237.19 1157.14 1086
BIO17 -0.196 0.232 0.302 79.52 34.47 98.99 14.86
EV (%) 85.575 11.724 2.700

BIO1 = Annual Mean Temperature; BIO2 = Mean Diurnal Range (Mean of monthly (max temp - min temp)); BIO4 = Temperature Seasonality (standard deviation *100); BIO5 = Max Temperature of Warmest Month; BIO6 = Min Temperature of Coldest Month, BIO7 = Temperature Annual Range (BIO5-BIO6); BIO11 = Mean Temperature of Coldest Quarter; BIO12 = Annual Precipitation; BIO17 = Precipitation of Driest Quarter; EV (%) = Percent of explained variance.

Lineage dispersal. The RRW model predicted the ancestral distribution of R. sumichrasti was centered in the SMdC physiographic region (abbreviations described in Figure 2), within the current extent of clade I (Figure 7). This clade started to spread at ~1.80 - 1.75 Ma to the west crossing the Tehuantepec Isthmus towards both the Oaxacan Highlands (OH) and Sierra Madre del Sur (SMdS) where the MRCA of clades II, III, and IV originated. Subsequently (between 1.53 - 1.25 Ma), the MRCA of clade III extended to the Sierra Madre Oriental (SMOr), while clade I colonized the Costa Rican Seasonal Moist Forest (TR*) and Talamancan Range (TR) regions. By ~1.25 to 0.65 Ma, the ancestor of clade IV expanded to the west of the Mexican Transvolcanic Belt (CT as named in Hardy et al. 2013), and by ~ 0.11 Ma most dispersal events occurred when clade II expanded through the central and east of the CT, but also seemed to expand towards the east by the OH (Figure 7).

Discussion

Species delimitation. The use of innovative tools and methodologies to assess species boundaries has helped to clarify taxonomic problems while facilitating the generation of robust hypotheses to reveal cryptic species and describe the speciation processes (Dayrat et al. 2005; Padial et al. 2010). Such is the case of mammals distributed in Mesoamerica, characterized by a peculiar evolutionary history that is linked to the environmental and biogeographical characteristics of this region (see Almendra and Rogers 2012). We used the cricetid rodent R. sumichrasti because it is a good model to evaluate the biogeographical and ecological niche conservatism hypotheses linked to vicariant speciation events in México to Central America. This approach was addressed by other authors (Sullivan et al. 2000; Martínez-Gordillo et al. 2010; Hardy et al. 2013), but this is the first time that the use of mathematical methods for species delimitation and phylogeographic reconstruction is put into practice for this species.

Our results show that the species delimitation methods support the phylogenetic hypotheses one and five with higher posterior probabilities, suggesting that R. sumichrasti is a complex of multiple species. In both hypotheses, clade I was identified as a distinct species, as this result was congruent among the three species delimitation methods. Recognition of clade I at the species level has been suggested previously due to its position in the molecular phylogenies (Sullivan et al. 2000; Hardy et al. 2013), and to the P-distances to the remaining clades (6.15 % to 9.10 %; Hardy et al. 2013). We agree with this species-level suggestion since this clade was placed as an independent sister lineage to the other clades of R. sumichrasti in our phylogenetic trees and also showed the highest genetic divergence (both K2P and P-distances) compared to clades II-IV. The populations belonging to this clade are distributed southeast of the Isthmus of Tehuantepec, from the Sierra Madre de Chiapas, México to western Panama (Hall 1981), and were the first to diverge from a common ancestor ~2.15 Ma. This mean age is close to that reported by Hardy et al (2013; ~2.56 Ma), placing the species diversification within R. sumichrasti at the Plio-Pleistocene boundary (see discussion below).

Figure 5 Map projection of the Ecological Niche Models for the 4 clades of Reithrodontomys sumichrasti indicating the within-clade and inter-clade localities predictability values. Color dots represent the presence records of each clade and follow the clade-colors in Figure 2. Dark and light colors on the maps represent the suitable and non-suitable areas of each clade, respectively. 

The proposal that clade I evolved independently was better supported by molecular data than by ecological data. The environmental niche space that this clade occupies predicted the potential distribution areas of the remaining clades with high percentages, although the inverse was not true. In general, R. sumichrasti sensu lato inhabits brush and grass in pine-oak and cloud forests throughout its geographical distribution. However, Hooper (1952) reported a greater diversity of habitats for the subspecies that encompass clade I, particularly for R. s. dorsalis and R. s. australis. This apparently broad environmental range could explain the high percentages of predictability we found, which was also evidenced in the canonical analysis. Nevertheless, non-equivalency of niche was found in the niche identity test. The remaining ecological analyses showed a relatively high similarity between this clade and clades II-IV, suggesting that their differentiation at the species level within R. sumichrasti sensu lato was more favored by geography than by ecology (Peterson et al. 1999).

Figure 6 Graphic of the first two discriminant functions among Ecological Niche Models of clades I to IV of Reithrodontomys sumichrasti. Colored crosses represent the centroid of each clade environmental niche. Colors follow the clade-color distinction described in Figure 2. Black arrows denote the power and direction of the discrimination for that bioclimatic variable (see text and Table 3 for descriptions of bioclimatic variables). 

The species delimitation methods were not consistent in the delimitation of clades II, III, and IV. The single-locus bGMYC (Cyt-b) proposed that the three clades form a single species, while the multiple-loci BPP and STACEY (Cyt-b + Fgb-I7 + Acp5) considered each clade as a distinct species. Molecular delimitation methods are considered a valuable complement to taxonomy based on morphological traits and are often used as part of an integrative approach to validate putative species (Luo et al. 2018). The three delimitation methods used in our study have been recognized for their high performance for this purpose (Jones 2017; Luo et al. 2018), but only two of them (BPP and STACEY) were consistent in this work. The performance and accuracy of each method can be affected by factors including both biological (variation in population size, uninterrupted gene flow) and methodological (input tree), among others, so they can over or underestimate the number of species (Rannala 2015; Luo et al. 2018). For this reason, the use of different molecular delimitation methods is highly recommended with species hypotheses based on the congruence among them (Carstens et al. 2013). In accordance with this suggestion, Hypothesis five (which is based on multiple loci) should be accepted and therefore each clade distributed west of the Isthmus of Tehuantepec constitutes a distinct species-level entity. Hypothesis five (Fig. 2) was also supported by the amount of Cyt-b genetic differentiation among clades. The K2P genetic distance values between pairwise clades II-III, II-IV, and III-IV were 6.07, 5.43, and 6.67, respectively, which are greater than the 5 % value associated with sister species recognition in mammals (Baker and Bradley 2006) including rodents (ranged from 2.70 % to 19.23 %; Bradley and Baker 2001).

Phylogenetic relationships among clades II, III, and IV were different between the Cyt-b tree topology and the species tree, but generally with weak nodal support. In the first case, II and IV were recovered as sister clades, while in the second, clades II and III were more closely related. These results partially coincide with the topologies obtained by Hardy et al. (2013), in which their concatenated DNA tree is consistent with our species tree. On the other hand, none of our phylogenies (gene tree or species tree) recovered sister relationships between clades III and IV, such as those obtained in the Cyt-b tree of Hardy et al. (2013). This is also supported by the ecological results where there is a greater ecological similarity (based on both directions of area predictability) between clades II and III than between clades II and IV or III and IV.

Figure 7 Spatiotemporal dynamics of the Reithrodontomys sumichrasti lineages diffusion for 1.80 Ma, 1.75 Ma, 1.53 Ma; 1.25 Ma, 0.65 Ma, and 0.11 Ma. Lines represent the branches of the Maximum Clade Credibility Tree and circles the location of occurrence records of the terminal labels (Appendix 1). An overlay of the sum of current, Last Glacial Maximum, and Last Interglacial ENMs was added to denote areas of relative environmental stability. Line and circle colors follow the clade-color distinction described in Figure 2. Maps were generated using Google Earth (http://earth.google.com). 

The ecological niche characteristics (from the bioclimatic variables used) of clade II showed high predictability percentages of the ecological suitability areas of clades III and IV, but these tended to have low or null values when the inverse analysis was performed. For example, clade IV predicted only 18 % of clade II and 0 % of clade III. The geographical distribution of each clade could explain the different percentages of predictability of the environmental niche. The wide geographical distribution of clade II includes localities of the CT, SMdS, extreme south of SMOr, and OH, while clade III is distributed in the SMOr, and clade IV is restricted to Coalcomán and Dos Aguas localities, in Michoacán (Hall 1981; Hardy et al. 2013; Figure 1, 2).

Niche pairwise comparisons showed low observed values for D and I similarity indices, mainly between clades III and IV. This is based on the fact that these indices can take values from 0 (no niche overlap) to 1 (total niche overlap; Warren et al. 2008). Closely related species are predicted to share characteristics of their environmental niche due to their common ancestry (Peterson et al. 1999), but niche differentiation can occur when allopatric populations exist, and gene flow is assumed to have been disrupted in the past (Avise 2000; Martínez-Gordillo et al. 2010). This could explain the non-equivalency of niche between these clades, as well as the low values of area predictability, which coincides with reports of Martínez-Gordillo et al. (2010) for different rodent species, including R. sumichrasti.

Bioclimatic data show that clade II shared similar characteristics to the other clades depending on the variable being analyzed. Moreover, clade III was characterized by low temperatures and the second-highest value of annual mean precipitation. These bioclimatic characteristics correspond to the habitat description of R. s. sumichrasti, mainly associated with pine and pine-oak forests, in “areas frequently bathed by clouds and rain (Hooper 1952:72)”. In contrast, clade IV was associated with higher temperatures and lower precipitation values, showing extreme values with respect to the other clades in at least five of the nine variables analyzed. Hardy et al. (2013) highlighted the presence of geographical barriers such as low-lying river drainages that have isolated clade IV populations from other R. sumichrasti sensu lato populations, which could justify our molecular and ecological results regarding the species recognition of this clade.

Phylogeographic history. Our results suggest that the common ancestor of the R. sumichrasti sensu lato originated in the montane regions of northern Central America ~2 Ma ago and expanded to where this species complex currently occurs. Various geographic and environmental factors may have favored and/or limited its dispersal in Central America and México (for more details see Hardy et al. 2013). The montane and intermontane Central America regions have a deep tectonic and volcanic history, which may have influenced the origin and diversification of montane species such as Peromyscus guatemalensis, P. bakeri, and P. carolpattonae (Álvarez-Castañeda et al. 2019). Also, the Pleistocene glacial cycles may have played a key role, due to favorable climatic conditions (Ceballos et al. 2010), which allowed the colonization of new areas and in some cases new habitats, followed by post-glacial isolation that limited the gene flow between populations (Martin 1961). This has been reported in several groups such as plants (e. g.Ramírez-Barahona and Eguiarte 2013), reptiles and amphibians (e. g.Church et al. 2003; Howes et al. 2006), birds (e. g.Johnson and Cicero 2004; Baker 2008), and mammals (e. g.Ceballos et al. 2010; Chiou et al. 2011) including other species of Reithrodontomys (Martínez-Borrego et al. 2022). In addition, geographic regions such as the Isthmus of Tehuantepec seem to have acted as an efficient barrier limiting gene flow between populations that are distributed on both sides of the Isthmus, an accepted explanation for R. sumichrasti and other rodent species (e. g.Sullivan et al. 2000; León-Paniagua et al. 2007; Ordoñez-Garza et al. 2010; Hardy et al. 2013).

The lineage dispersal in México was from populations in the west of the OH and SMdS that currently belong to the clade II, which spread into SMOr (clade III) and the west of CT (clade IV) as well as through the central and east of the CT (clade II). This model would explain the wide geographical distribution of clade II, and also its greater number of haplotypes compared to the other clades (Hardy et al. 2013). Although these dispersal events seem to have occurred relatively recently, the physiographic characteristics of the Mexican mountainous regions (Morrone 2005; Escalante et al. 2009) could have favored relatively faster speciation processes within R. sumichrasti complex, leading to differentiation, at least genetically and ecologically, among each clade analyzed here. This seems to be a common pattern in several species of small mammals, where the allopatric effect and the habitat characteristics each ancestral species occupied resulted in complete speciation of lineages, often associated with cryptic speciation processes (e. g.Arellano et al. 2005; Rogers et al. 2007; León-Tapia et al. 2021; Martínez-Borrego et al. 2022).

Taxonomic considerations. Species delimitation methods and values of genetic divergence support the recognition of populations of R. sumichrasti at the east and south of the Isthmus of Tehuantepec, from Chiapas, México to Central America (Clade I), as a valid species which is different from everything occurring to the west of this geographical barrier. According to this hypothesis, then R. australis (Allen 1895) is the taxonomic name that has priority (Article 23; ICZN 1999). Subspecies distributed across this region of Mesoamerica, beyond the nominotypical would include R. a. dorsalis (Merriam 1901), R. a. modestus (Thomas 1907), and R. a. vulcanius (Bangs 1902).

In addition, the existence of an undescribed species represented by the populations included in clade IV, from Coalcomán and Dos Aguas in Michoacán, México (northwestern SMdS) is supported by species delimitation methods and values of genetic divergence. The disjoint distribution of this genetically distinct clade suggests that it does not belong to R. s. nerterus nor R. s. luteolus. The mountainous region inhabited by this new species is isolated from other mountain ranges in the area by lowlands of up to approximately 400 masl. This pattern of genetic differentiation coincides with the recent description of a new species of the genus Peromyscus (P. greenbaumi; Bradley et al. 2022; but see also León-Tapia et al. 2021). In order to make the formal description based on diagnostic characters that will derive in an appropriate species name, a morphological comparison would be necessary.

Molecular species delimitation and genetic distance values associated to populations from clades II and III indicate that these two lineages should be recognized as valid species. Nomenclatural suggestions are difficult to make due to the sympatry of individuals of some populations from both clades. This was already addressed by Hardy et al. (2013) through nested clade analysis. In our study a phylogeographic pattern of diffusion of the lineages (RRW model) suggests colonization after the separation of clades II and III. Nevertheless, in this work we propose populations comprising clade II should be recognized as R. nerterus (Merriam, 1901). Although we did not include specimens from the type locality of R. nerterus (El Nevado de Colima, Jalisco, México), we analyzed several individuals from sites reported by Hooper (1952) for this taxon. Because clade II includes populations of the known distribution of R. s. luteolus, this taxon should be considered as subspecies of R. nerterus. Clade III should be named as R. sumichrasti; here we also did not include individuals from the type locality (El Mirador, Veracruz, México), but we used specimens from localities that belong to this species. Populations from south Puebla and Northern Oaxaca (28, 1, and 10 in Figure 2), regarded originally as R. s. sumichrasti should be now R. n. luteolus. It remains necessary to evaluate sympatric populations from both clades in order to identify plausible evolutionary processes in this region.

Acknowledgements

We acknowledge financial support from the Consejo Nacional de Ciencia y Tecnologia (postdoctoral fellowship to ALA) and the Department of Biology, Brigham Young University (to DSR). We wish to thank D. D. Cruz for his assistance in compiling data for the Ecological Niche Modeling, E. C. Molina for gathering GeneBank data for the molecular analysis, and R. Núñez for his support in editing figures. We also thank two anonymous reviewers who kindly read drafts of this work and supplied valuable comments.

Literature cited

Allen, J. A. 1895. On the species of the genus Reithrodontomys. Bulletin of the American Museum of Natural History 7:7-143. [ Links ]

Almendra, A. L., and D. S. Rogers. 2012. Biogeography of Central American mammals. Pp. 203-229, in Bones, clones, and biomes: the history and geography of recent neotropical mammals (Patterson, B. D., and L. P. Costa, eds.). University of Chicago Press. Illinois, U.S.A. [ Links ]

Almendra, A. L. , et al. 2018. Evolutionary relationships and climatic niche evolution in the genus Handleyomys (Sigmodontinae: Oryzomyini). Molecular Phylogenetics and Evolution 128:12-25. [ Links ]

Álvarez-Castañeda, S. T., et al. 2019. Two new species of Peromyscus from Chiapas, Mexico, and Guatemala. Pp. 543-558, in From field to laboratory: a memorial volume in honor of Robert J. Baker (Bradley, R. D., H. H. Genoways, D. J. Schmidly, and L. C. Bradley, eds.). Special Publications, Museum of Texas Tech University. Lubbock, U.S.A. [ Links ]

Arbogast, B. S., et al. 2002. Estimating divergence times from molecular data on phylogenetic and population genetic timescales. Annual Review of Ecology and Systematics 33:707-740. [ Links ]

Arellano, E., F. González-Cozátl, and D. S. Rogers. 2005. Molecular systematics of Middle American harvest mice Reithrodontomys (Muridae), estimated from mitochondrial cytochrome b gene sequences. Molecular Phylogenetics and Evolution 37:529-540. [ Links ]

Avise, J. C. 2000. Phylogeography: the history and formation of species. Harvard University Press. Massachusetts, U.S.A. [ Links ]

Baker, A. J. 2008. Islands in the sky: the impact of Pleistocene climate cycles on biodiversity. Journal of Biology 7:1-4. [ Links ]

Baker, R. J., and R. D. Bradley. 2006. Speciation in mammals and the Genetic Species Concept. Journal of Mammalogy 87:643-662. [ Links ]

Bangs, O. 1902. Chiriqui Mammalia. Bulletin of the Museum of Comparative Zoology at Harvard College 39:17-51. [ Links ]

Bielejec, F., et al. 2011. SPREAD: spatial phylogenetic reconstruction of evolutionary dynamics. Bioinformatics 27:2910-2912. [ Links ]

Bouckaert, R., et al. 2014. BEAST 2: a software platform for Bayesian evolutionary analysis. PLoS Computational Biology 10:p.e1003537. [ Links ]

Bradley, R. D. 2017. Genus Reithrodontomys. Pp. 367-383, in Handbook of the Mammals of the World: Rodents II (Wilson, D. E., T. E. Lacher, and R. A. Mittermeier, eds.). Lynx Edicions. Barcelona, Spain. [ Links ]

Bradley, R. D. , and R. J. Baker. 2001. A test of the Genetic Species Concept: Cytochrome b sequences and mammals. Journal of Mammalogy 82:960-973. [ Links ]

Bradley, R. D. , et al. 2022. Two new species of Peromyscus (Cricetidae: Neotominae) from the Transverse volcanic belt of Mexico. Journal of Mammalogy 103:255-274. [ Links ]

Carstens, B. C., et al. 2013. How to fail at species delimitation. Molecular Ecology 22:4369-4383. [ Links ]

Ceballos, G., J. Arroyo-Cabrales, and E. Ponce. 2010. Effects of Pleistocene environmental changes on the distribution and community structure of the mammalian fauna of Mexico. Quaternary Research 73:464-473. [ Links ]

Chiou, K. L., et al. 2011. Pleistocene diversification of living squirrel monkeys (Saimiri spp.) inferred from complete mitochondrial genome sequences. Molecular Phylogenetics and Evolution 59:736-745. [ Links ]

Church, S. A., et al. 2003. Evidence for multiple Pleistocene refugia in the postglacial expansion of the eastern tiger salamander, Ambystoma tigrinum tigrinum. Evolution 57:372-383. [ Links ]

Czaplewski, N. J. 1987. Sigmodont rodents (Mammalia; Muroidea; Sigmodontinae) from the Pliocene (early Blancan) Verde Formation, Arizona. Journal of Vertebrate Paleontology 7:183-199. [ Links ]

Dalquest, W. W. 1978. Early Blancan mammals of the Beck Ranch local fauna of Texas. Journal of Mammalogy 59:269-298. [ Links ]

Daly, M., et al. 2012. Systematics Agenda 2020: the mission evolves. Systematic Biology 61:549-552. [ Links ]

Dayrat, B. 2005. Towards integrative taxonomy. Biological Journal of Linnean Society 85:407-415. [ Links ]

de Queiroz, K. 2007. Species concepts and species delimitation. Systematic Biology 56:879-886. [ Links ]

Dellicour, S., et al. 2021. Relax, keep walking-a practical guide to continuous phylogeographic inference with BEAST. Molecular Biology and Evolution 38:3486-3493. [ Links ]

Doody, J. S., S. Freedberg, and J. S. Keogh. 2009. Communal egg-laying in reptiles and amphibians: evolutionary patterns and hypotheses. The Quarterly Review of Biology 84:229-252. [ Links ]

Escalante, T., C. Szumik, and J. J. Morrone. 2009. Areas of endemism of Mexican mammals: reanalysis applying the optimality criterion. Biological Journal of the Linnean Society 98:468-478. [ Links ]

Friendly, M., and J. Fox. 2015. candisc: Visualizing generalized canonical discriminant and canonical correlation analysis. R package version 0.6-5. Available from: https://cran.r-project.org/web/packages/candisc/index.htmlLinks ]

Gilbert, J. D., and A. Manica. 2015. The evolution of parental care in insects: a test of current hypotheses. Evolution 69:1255-1270. [ Links ]

Hall, E. R. 1981. The Mammals of North America 2nd ed. John Wiley and Sons, Inc., New York, U.S.A. [ Links ]

Hardy, D. K., et al. 2013. Molecular phylogenetics and phylogeography structure of Sumichrast´s harvest mouse (Reithrodontomys sumichrasti: Cricetidae) based on mitochondrial and nuclear DNA sequences. Molecular Phylogenetics and Evolution 68:282-292. [ Links ]

Heath, T. A., J. P. Huelsenbeck, and T. Stadler. 2014. The fossilized birth-death process for coherent calibration of divergence-time estimates. Proceedings of the National Academy of Sciences 111:2957-2966. [ Links ]

Heled, J., and A. J. Drummond. 2010. Bayesian inference of species trees from multilocus data. Molecular Biology and Evolution 27:570-580. [ Links ]

Hijmans, R. J., et al. 2005. Very high-resolution interpolated climate surfaces for global land areas. International Journal of Climatology 25:1965-1978. [ Links ]

Hijmans, R. J., et al. 2017. Package ‘dismo’. Circles 9:1-68. [ Links ]

Hooper, E. T. 1952. A systematic review of harvest mice (genus Reithrodontomys) of Latin America. Miscellaneous Publications Museum of Zoology, University of Michigan 77:1-255. [ Links ]

Howes, B. J., B. Lindsay, and S. C. Lougheed. 2006. Range-wide phylogeography of a temperate lizard, the five-lined skink (Eumeces fasciatus). Molecular Phylogenetics and Evolution 40:183-194. [ Links ]

Huang, J. P. 2020. Is population subdivision different from speciation? From phylogeography to species delimitation. Ecology and Evolution 10:6890-6896. [ Links ]

International Commission on Zoological Nomenclature. 1999. International code of zoological nomenclature. 4th ed. International Trust for Zoological Nomenclature, London, U.K. [ Links ]

Johnson, N. K., and C. Cicero. 2004. New mitochondrial DNA data affirm the importance of Pleistocene speciation in North American birds. Evolution 58:1122-1130. [ Links ]

Jones, G. 2017. Algorithmic improvements to species delimitation and phylogeny estimation under the multispecies coalescent. Journal of Mathematical Biology 74:447-467. [ Links ]

Katoh, K., et al. 2005. MAFFT version 5: improvement in accuracy of multiple sequence alignment. Nucleic Acids Research 33:511-518. [ Links ]

Kimura, M. 1980. A simple method for estimating evolutionary rates of base substitutions through comparative studies of nucleotide sequences. Journal of Molecular Evolution 16:111-120. [ Links ]

Knowles, L. L., and B. C. Carstens. 2007. Delimiting species without monophyletic gene trees. Systematic Biology 56:887-895. [ Links ]

Kumar, S., et al. 2018. MEGA X: molecular evolutionary genetics analysis across computing platforms. Molecular Biology and Evolution 35:1547-1549. [ Links ]

Lanfear, R., et al. 2014. Selecting optimal partitioning schemes for phylogenomic datasets. BMC Evolutionary Biology 14:1-14. [ Links ]

Lemey, P., et al. 2010. Phylogeography takes a relaxed random walk in continuous space and time. Molecular Biology and Evolution 27:1877-1885. [ Links ]

León-Paniagua, L., et al. 2007. Diversification of the arboreal mice of the genus Habromys (Rodentia: Cricetidae: Neotominae) in the Mesoamerican highlands. Molecular Phylogenetics and Evolution 42:653-664. [ Links ]

León‐Tapia, M. Á., et al. 2021. Role of Pleistocene climatic oscillations on genetic differentiation and evolutionary history of the Transvolcanic deer mouse Peromyscus hylocetes (Rodentia: Cricetidae) throughout the Mexican central highlands. Journal of Zoological Systematics and Evolutionary Research 59:2481-2499. [ Links ]

Lindsay, E. H., and N. J. Czaplewski. 2011. New rodents (Mammalia, Rodentia, Cricetidae) from the Verde Fauna of Arizona and the Maxum Fauna of California, USA, early Blancan Land Mammal Age. Palaeontología Electrónica 14:1-16. [ Links ]

Luo, A., et al. 2018. Comparison of methods for molecular species delimitation across a range of speciation scenarios. Systematic Biology 67:830-846. [ Links ]

Maestri, R., et al. 2017. The ecology of continental evolutionary radiation: Is the radiation of sigmodontine rodents adaptive? Evolution 71:610-632. [ Links ]

Martin, P. S. 1961. Southwestern animal communities in the late Pleistocene. Biology of the arid and semiarid lands of the Southwest. New Mexico Highlands University Bulletin 212:56-66. [ Links ]

Martin, R. A., et al. 2002. Blancan Lagomorphs and rodents of the Deer Park assemblages, Meade County, Kansas. Journal of Paleontology 76:1072-1090. [ Links ]

Martin, R. A., and P. Peláez-Campomanes. 2014. Diversity dynamics of the Late Cenozoic rodent community from south-western Kansas: the influence of historical processes on community structure. Journal of Quaternary Science 29:221-231. [ Links ]

Martínez-Borrego, D., et al. 2022. Molecular systematics of the Reithrodontomys tenuirostris group (Rodentia: Cricetidae) highlighting the Reithrodontomys microdon species complex. Journal of Mammalogy 103:29-44. [ Links ]

Martínez-Gordillo, D., O. Rojas-Soto, and A. Espinosa-De Los Monteros. 2010. Ecological niche modelling as an exploratory tool for identifying species limits: an example based on Mexican muroid rodents. Journal of Evolutionary Biology 23:259-270. [ Links ]

Merriam, C. H. 1901. Descriptions of 23 new harvest mice (genus Reithrodontomys). Proceedings of the Washington Academy of Sciences 3:547-558. [ Links ]

Morgan, G. S., and R. S. White Jr. 2005. Miocene and Pliocene vertebrates from Arizona. Pp. 114-135, in Vertebrate Paleontology in Arizona (Heckert A. B. and S. G. Lucas, eds.). New Mexico Museum of Natural History and Science Bulletin. Albuquerque, New Mexico, U.S.A. [ Links ]

Morrone, J. J. 2005. Toward a synthesis of Mexican biogeography. Revista Mexicana de Biodiversidad 76:207-252. [ Links ]

Musser, G. G., and M. D. Carleton. 2005. Superfamily Muroidea. Pp. 894-1531, in Mammal species of the world: a taxonomic and geographic reference. 3rd ed. (Wilson, D. E., and D. M. Reeder, eds.). Johns Hopkins University Press. Baltimore, U.S.A. [ Links ]

Myers, N., et al. 2000. Biodiversity hotspots for conservation priorities. Nature 403:853-858. [ Links ]

Ordóñez-Garza, N., et al. 2010. Patterns of phenotypic and genetic variation in three species of endemic Mesoamerican Peromyscus (Rodentia: Cricetidae). Journal of Mammalogy 91:848-859. [ Links ]

Padial, J., et al. 2010. The integrative future of taxonomy. Frontiers in Zoology 7:16. [ Links ]

Peterson, A. T., J. Soberón, and V. Sánchez-Cordero. 1999. Conservatism of ecological niches in evolutionary time. Science 285:1265-1267. [ Links ]

Phillips, S. J., and M. Dudík. 2008. Modeling of species distributions with Maxent: new extensions and a comprehensive evaluation. Ecography 31:161-175. [ Links ]

R Development Core Team. 2018. R: A language and environment for statistical computing. R Foundation for Statistical Computing. Vienna, Austria. [ Links ]

Rambaut, A., et al. 2018. Posterior summarization in Bayesian phylogenetics using Tracer 1.7. Systematic Biology 67:901-904. [ Links ]

Ramírez-Barahona, S. and L. E. Eguiarte. 2013. The role of glacial cycles in promoting genetic diversity in the Neotropics: the case of cloud forests during the Last Glacial Maximum. Ecology and Evolution 3:725-738. [ Links ]

Rannala, B. 2015. The art and science of species delimitation. Current Zoology 61:846-853. [ Links ]

Reid, N. M., and B. C. Carstens. 2012. Phylogenetic estimation error can decrease the accuracy species delimitation: a Bayesian implementation of the General Mixed Yule Coalescent model. BMC Evolutionary Biology 12:196. [ Links ]

Rissler, L. J., and J. J. Apodaca. 2007. Adding more ecology into species delimitation: ecological niche models and phylogeography help define cryptic species in the black salamander (Aneides flavipunctatus). Systematic Biology 56:924-942. [ Links ]

Rogers, D. S., et al. 2007. Molecular phylogenetic relationships among crested-tailed mice (genus Habromys). Journal of Mammalian Evolution 14:37-55. [ Links ]

Sites, J. W., and J. C. Marshall. 2003. Delimiting species: a renaissance issue in systematic biology. Trends in Ecology and Evolution 18:462-470. [ Links ]

Sites, J. W. , and J. C. Marshall. 2004. Operational criteria for delimiting species. Annual Review of Ecology, Evolution, and Systematics 199-227. [ Links ]

Schoener, T. W. 1968. The Anolis lizards of Bimini: resource partitioning in a complex fauna. Ecology 49:704-726. [ Links ]

Sullivan, J., E. Arellano, and D. S. Rogers. 2000. Comparative phylogeography of Mesoamerican highland rodents: concerted versus independent response to past climatic fluctuations. The American Naturalist 155:755-768. [ Links ]

Thomas, O. 1907. On Neotropical mammals of the genera Callicebus, Reithrodontomys, Ctenomys, Dasypus, and Marmosa. The Annals and Magazine of Natural History, 7:161-168. [ Links ]

Urbina, S. I., et al. 2006. Karyotypes of three species of harvest mice (genus Reithrodontomys). The Southwestern Naturalist 51:564-568. [ Links ]

Warren, D. L., R. E. Glor, and M. Turelli. 2008. Environmental niche equivalency versus conservatism: quantitative approaches to niche evolution. Evolution: International Journal of Organic Evolution 62:2868-2883. [ Links ]

Warren, D. L. , R. E. Glor, and M. Turelli. 2010. ENMTools: a toolbox for comparative studies of environmental niche models. Ecography 33:607-611. [ Links ]

Yang, Z., and B. Rannala. 2014. Unguided species delimitation using DNA sequence data from multiple loci. Molecular Biology and Evolution 31:3125-3135. [ Links ]

0Associated editor: Jake Esselstyn and Giovani Hernández Canchola

Appendix 1

Population numbers (corresponding in Figure 2), specimen identification numbers (museum voucher or collector numbers), Collecting locality information; GenBank accession numbers and related clade for each sample of Reithrodontomys sumichrasti individuals included in this study. Museum or collector abbreviations are as follows: ASNHC = Angelo State Natural History Collection; BYU = Brigham Young University; CMC = Colección de Mamíferos del CIByC, Universidad Autónoma del Estado de Morelos; MSB = Museum of Southwestern Biology; ROM = Royal Ontario Museum; TTU = Texas Tech University; CWK = C. William Kilpatrick (University of Vermont); JAG = José A. Guerrero (Universidad Autónoma del Estado de Morelos). Country abbreviations are as follows: CR = Costa Rica; GM = Guatemala; HD = Honduras; MX = México; NI = Nicaragua; PN = Panamá. New sequences are denoted by an asterisk. 

Pop. Num. Voucher number Country: State Locality GenBank accession numbers Clade
Cyt-b Fgb-I7 Acp5
1 BYU15437 MX: Oaxaca 1.5 km S Puerto de la Soledad, 2200 m (18.1623667; -96.9975333) AF211911     II
BYU15438 AF211905     II
BYU16249 HQ269530 HQ269737 HQ269468 II
BYU15433 HQ269531 II
BYU15434 AF211915     II
2 BYU20806 MX: Oaxaca El Polvorín, 5.3 km turn off Lachao Viejo, 1735 m (16.1999333; -97.1339667) HQ269532 HQ269738 HQ269469 II
BYU20808 HQ269534     II
BYU20807 HQ269533 HQ269739 HQ269470 II
3 CMC912 MX: Oaxaca Finca Copalita, Copalita, 1025 m (15.9655833; 96.4574667) HQ269535 HQ269740 HQ269471 II
CMC913 HQ269536     II
CMC914 HQ269537     II
CMC915 HQ269538 HQ269741 HQ269472 II
4 CMC991 MX: Oaxaca Río Molino, 2353 m (16.0796667; -96.4708833) HQ269539 HQ269742 HQ269473 II
CMC992 HQ269540 HQ269743 HQ269474 II
CMC993 HQ269541     II
CMC994 HQ269542     II
CMC995 HQ269543     II
CMC996 HQ269544     II
CMC997 HQ269545 HQ269744 HQ269475 II
CMC998 HQ269546     II
CMC999 HQ269547     II
CMC1000 HQ269548     II
CMC1001 HQ269549 HQ269745 HQ269476 II
CMC1002 HQ269550     II
CMC1003 HQ269551     II
CMC1004 HQ269552 HQ269746 HQ269477 II
CMC1005 HQ269553     II
CMC1006 HQ269554     II
CMC1007 HQ269555     II
CMC1008 HQ269556     II
CMC1009 HQ269557     II
CMC1010 HQ269558     II
5 CMC172 MX: Oaxaca Santa María Yacochi, Cerro Zempoaltepec, 2300 m (17.1583333; -96.0166667) HQ269559     II
6 CMC1650 MX: Oaxaca La Cumbre, 1.2 km SE 0.6 km S Agua Fría Juxtlahuaca, 1950 m (17.209; -97.9786667) HQ269560     II
7 TTU54952 MX: Oaxaca 3.0 mi S. Suchixtepec (16.0166667; -96.4666667) AF211920     II
8 CMC989 MX: Oaxaca 0.7 km E La Soledad (15.9823; -96.5198167) HQ269561     II
CMC990 HQ269562     II
9 CMC734 MX: Oaxaca La Cumbre, 18.5 km S Sola de Vega, 2175 m (16.4529; -97.00235) HQ269563     II
10 CWK1009 MX: Oaxaca Orizaba (17.8333333; -97.2333333) AF211895     II
11 FAC1112* MX: Guerrero 6.1 km SW Omiltemi, 2490 m (17.5491667; -99.721) AF211907     II
FAC1117* AF211913     II
FAC1118 AF211906     II
FAC1119 AF211908     II
BYU20801 HQ269564 HQ269747 HQ269478 II
BYU20802 HQ269565     II
CWK1019* AF211921     II
CWK1025* AF211901     II
12 BYU20799 MX: Guerrero 3.4 km W Carrizal, 2480 m (17.6004167; -99.8248333) HQ269566 HQ269748 HQ269479 II
CMC710 HQ269567     II
13 CMC1628 MX: Guerrero 3 km E El Tejocote, 2620m (17.3048667; -98.6511167) HQ269568     II
CMC1629 HQ269569 HQ269749 HQ269480 II
CMC1630 HQ269570 HQ269750 HQ269481 II
CMC1631 HQ269571     II
CMC1632 HQ269572     II
CMC1633 HQ269573     II
CMC1634 HQ269574     II
CMC1635 HQ269575     II
CMC1636 HQ269576     II
CMC1637 HQ269579     II
CMC1638 HQ269580     II
CMC1639 HQ269581     II
CMC1640 HQ269582     II
CMC1641 HQ269583     II
CMC1642 HQ269584     II
CMC1643 HQ269585     II
CMC1644 HQ269586     II
CMC1645 HQ269587     II
CMC1646 HQ269577     II
CMC1647 HQ269578     II
CMC1648 HQ269588     II
CMC1649 HQ269589     II
14 TK93354 MX: Guerrero 4 mi SSW Filo de Caballo (17.8166667; -99.6166667) AY293810     II
TK93363 AY293811     II
15 BYU20800 MX: Guerrero 1.1 km E Cruz Nueva, 2650 m (17.513483; -100.0295167) HQ269590 HQ269751 HQ269482 II
CMC712 HQ269591     II
CMC713 HQ269592 HQ269752 HQ269483 II
16 BYU15967 MX: Veracruz La Colonia, 6.5 km W Zacualpan, 6200 ft (20.4666667; -98.3666667) HQ269594     III
BYU 15968 AF211916     III
BYU15969 HQ269595 HQ269754 HQ269485 III
BYU15970 HQ269596     III
BYU 15971 AF211902     III
BYU15972 HQ269593 HQ269753 HQ269484 III
17 CMC873 MX: Veracruz Las Cañadas, 1340 m (19.1878333; -96.9834) HQ269597 HQ269755 HQ269486 III
CMC875 HQ269598 HQ269756 HQ269487 III
CMC876 HQ269599     III
18 CMC878 MX: Veracruz 3.5 km E Puerto del Aire, 2524 m (18.6715667; -97.3318667) HQ269600 HQ269757 HQ269488 II
CMC879 HQ269601 HQ269758 HQ269489 II
CMC880 HQ269602 HQ269759 HQ269490 II
19 CMC840 MX: Veracruz 2.9 km E Puerto del Aire, 2524 m HQ269603     III
CMC843 HQ269604     III
CMC847 HQ269605     III
CMC1403 HQ269606     III
CMC1405 HQ269607     II
20 CWK1007 MX: Veracruz Xometla, 2615 m (18.97775; -97.1910833) AF211914     III
CMC849 HQ269608 HQ269760 HQ269491 III
CMC850 HQ269609     III
CMC851 HQ269610     III
CMC853 HQ269611     III
CMC854 HQ269612     III
CMC855 HQ269613     III
CMC856 HQ269614     III
CMC857 HQ269615 HQ269761 HQ269492 III
CMC858 HQ269616     III
CMC859 HQ269617     III
CMC860 HQ269618 HQ269762 HQ269493 III
CMC861 HQ269619     III
CMC862 HQ269620     III
CMC863 HQ269621 HQ269763 HQ269494 III
CMC864 HQ269622     III
CMC866 HQ269623 HQ269764 HQ269495 III
CMC867 HQ269624     III
CMC869 HQ269625 HQ269765 HQ269496 III
CMC870 HQ269626     III
CMC871 HQ269627     III
21 CMC1378 MX: Veracruz Mesa de la Yerba, 3.4 km SW desviación a Mazatepec, 2040 m (19.5593; -97.0185) HQ269628     III
CMC1379 HQ269629     III
CMC1380 HQ269630     III
CMC1381 HQ269631     III
CMC1395 HQ269632     III
CMC1396 HQ269633     III
CMC1397 HQ269634     III
CMC1398 HQ269635     III
CMC1399 HQ269636     III
CMC1400 HQ269637     III
CMC1401 HQ269638     III
CMC1402 HQ269639     III
22 CMC1446 MX: Veracruz Cruz Blanca, 2180 m (19.4712; -97.0842) HQ269640     III
CMC1447 HQ269641     III
CMC1448 HQ269642     III
CMC1449 HQ269643     III
23 CMC1476 MX: Veracruz Xico Viejo, 1756 m (19.4517667; -97.0583) HQ269644     III
CMC1477 HQ269645     III
CMC1478 HQ269646     III
CMC1480 HQ269648     III
CMC1481 HQ269649     III
24 CMC1073 MX: Puebla 4.7 km NE Teziutlán, 1750 m (19.8353167; -97.34135) HQ269650 HQ269766 HQ269497 III
CMC1074 HQ269651     III
CMC1075 HQ269652 HQ269767 HQ269498 III
25 CMC1070 MX: Puebla El Durazno, 0.5 km Libramiento Parada, 1830m (19.8220833; -97.3399833) HQ269653     III
26 CMC1093 MX: Puebla 3 km W Cerro Chignaulta, 2176 m HQ269654 HQ269768 HQ269499 III
27 CMC1992 MX: Puebla Rancho 22 de Marzo, marker 75.8 km Carretera Ahuazotepec-Zacatlán, 2270 m (19.6677; -97.9890333) HQ269656 HQ269769 HQ269500 III
CMC1997 HQ269658     III
CMC2006 HQ269655     III
CMC2007 HQ269657     III
CMC2008 HQ269659     III
CMC2009 HQ269660     III
CMC2010 HQ269661     III
28 CMC2005 MX: Puebla Alhuaca, 8 km NE Vicente Guerrero, 2680 m (18.5705167; -97.1660833) HQ269662     II
29 CMC1711 MX: Puebla 2 km NW Cuautlamingo, 2171 m (19.7678667; -97.5403333) HQ269663     III
30 CMC1710 MX: Puebla Los Parajes, 2555 m (19.7664667; -97.4384667) HQ269664     III
31 CMC1860 MX: Michoacán 11 km NW Coalcomán, 1600 m (18.803; -103.2261667) HQ269665     IV
CMC1862 HQ269666     IV
CMC1863 HQ269667     IV
32 CMC1859 MX: Michoacán 10.9 km NW Coalcomán, 1680 m (18.7966667; -103.2303333) HQ269668     IV
33 CMC1855 MX: Michoacán 0.8 km NNE Dos Aguas, 2220 m (18.8075; -102.9263333) HQ269669 HQ269770 HQ269501 IV
34 CMC1856 MX: Michoacán 4.2 km NNE Dos Aguas, 2370 m (18.8358333; -102.9256667) HQ269670 HQ269771 HQ269502 IV
35 CMC1857 MX: Michoacán 9.2 km NNE Dos Aguas, 2245 m (18.8046667; -102.9775) HQ269671     IV
36 BYU16242 MX: Michoacán 10 km S Pátzcuaro, 2350 m (19.4535; -101.6027333) HQ269672     II
BYU16243 HQ269673     II
BYU16244 HQ269674     II
BYU16245 HQ269675     II
BYU16246 HQ269676     II
BYU16247 HQ269677 HQ269772 HQ269503 II
37 CMC1870 MX: Michoacán 9.6 km S Pátzcuaro, 2350 m (19.45695; -101.6075833) HQ269678     II
38 CMC1871 MX: Michoacán 4.9 km S Santa Clara, 2415 m (19.3611667; -101.6116667) HQ269679     II
CMC1872 HQ269680     II
39 CWK1014 MX: Michoacán 2.9 mi E Opopeo (19.4; -101.6) AF211896     II
CWK1015 AF211923     II
40 CWK1011 MX: Michoacán 9.9 km NW Mil Cumbres, 2820 m (19.6476667; -100.793) AF211900     II
CMC1864 HQ269681 HQ269773 HQ269504 II
CMC1865 HQ269682     II
CMC1866 HQ269683 HQ269774 HQ269505 II
CMC1867 HQ269684     II
CMC1868 HQ269685     II
41 CWK1056 MX: Michoacán Villa Escalante (19.4; -101.65) AF211898     II
42 CMC2001 MX: Hidalgo Río Chíflón, 9.7 km ENE Crucero los Tules, 1750 m (20.4013333; -98.3840833) HQ269688     III
CMC2000 HQ269687     III
CMC2002 HQ269689     III
CMC1982 HQ269686 HQ269775 HQ269506 III
43 CMC2003 MX: Hidalgo 5 km ENE Crucero los Tules, 2070 m (20.3834; -98.3647333) HQ269690     III
CMC2004 HQ269691 HQ269776 HQ269507 III
44 CMC1071 MX: Hidalgo 22 km NE Metepec, 2210 m (20.3158667; -98.23535) HQ269693     III
CMC1092 HQ269692 HQ269777 HQ269508 III
45 BYU15417 MX: Hidalgo La Mojonera, 6 km S Zacualtipan, 2010 m (20.65; -98.6) HQ269694     III
BYU15418 HQ269695     III
BYU15419 HQ269696     III
BYU15420 HQ269697     III
BYU 15421 AF211904     III
BYU 15422 AF211918     III
46 BYU 15415 MX: Hidalgo El Potrero, 10 km SW Tenango de Doria, 2200 m (20.65; -98.0666667) AF211899     III
BYU15416 HQ269699     III
BYU15414 HQ269698     III
47 CWK1027 MX: Hidalgo 5.0 Km N Zacualtipán (20.65; -98.6) AF211922     III
48 CWK1036 MX: Hidalgo 0.5 Km N Molango (20.7833333; -98.7166667) AF211903     III
49 CMC1786 MX: Estado de México 9 km SW Zacualpán, 2400 m (18.6882667; -99.80595) HQ269703     II
CMC1787 HQ269700 HQ269778 HQ269509 II
CMC1788 HQ269701 HQ269779 HQ269510 II
50 BYU17083 MX: Chiapas Cerro Mozotal, 2930 m (15.4311; -92.3379) HQ269704 HQ269780 HQ269511 I
BYU20784 HQ269707 HQ269781 HQ269512 I
CMC682 HQ269706     I
BYU17084 HQ269705     I
51 BYU20795 MX: Chiapas Rancho la Providencia, 1775 m (15.0913333; -92.0831) HQ269710 HQ269784 HQ269515 I
BYU20794 HQ269709 HQ269783 HQ269514 I
CMC694 HQ269708 HQ269782 HQ269513 I
52 CNMA 35505 MX: Chiapas San Cristobal (16.75; -92.6333333) AF211909     I
CNMA 35508 AF211910     I
CNMA 35514* AF211917     I
NMA 35506* AF211919     I
53 ASNHC2150 MX: Chiapas 9 km S Rayón (17.2; -93) AF211894     I
ASNHC2151 AF211897 HQ269785 HQ269516 I
54 TTU82780 MX: Chiapas Yalentay (16.7333333; -92.775) HQ269711     I
TTU82781 HQ269712     I
55 ECOSCM1220 MX: Chiapas El Vivero, Parque Nacional Lagos de Montebello, 3.55 km NNW El Vivero, 1452 m (16.25; -92.1333333) HQ269713 HQ269786 HQ269517 I
56 ROM98287 GM: Huehuetenango 10 km NW Santa Eulalia (15.75; -91.4833333) HQ269714     I
ROM98383 HQ269715 HQ269787 HQ269518 I
57 ROM98384 GM: Chimaltenango 15 km NW Santa Apolonia (14.7913833; -90.9708333) HQ269716 HQ269788 HQ269519 I
58 TTU83709 HD: Copán Picacho (13.9833333; -88.1833333) HQ269717     I
59 TTU84602 HD: Intibuca Santa Rosa (14.77; -88.78) HQ287797     I
60 JAG417 NI: Esteli Reserva de Miraflor, 3 km SE Miraflor (13.3683667; -86.4023) HQ269718     I
61 BYU 15246 CR: San José El Cascajal de Coronado, 1650 m (9.9166667; -84.0666667) AF211912     I
62 ROM113151 CR: Cartago Volcán Irazú, Route 8 Hwy Sign 28 km, La Pastora (9.8666667; -83.9166667) HQ269720 HQ269790 HQ269521 I
ROM113178 HQ269724     I
MSB61880 HQ269719 HQ269789 HQ269520 I
ROM113180 HQ269726     I
ROM113153 HQ269722 HQ269792 HQ269523 I
ROM113181 HQ269727     I
ROM113179 HQ269725     I
ROM113152 HQ269721 HQ269791 HQ269522 I
ROM113154 HQ269723     I
63 MSB130128 PN: Chiriqui Bugaba, Parque Nacional Volcán Baru-Intermedia (8.85; -82.5666667) HQ269728 HQ269793 HQ269524 I
64* unavailable MX: Guerrero Las Truchas, 3 km SE Carrizal de Bravo, 2400 m (17.359739; -99.489833) AB618727     II
unavailable AB618732     II
unavailable AB618730     II
65* unavailable MX: Guerrero Carrizal de Bravo, 2.5 km SE, 2400 m (17.609715; -99.820829) AB618729     II
66* CNMA42283 MX: Oaxaca Municipio Tlahuitoltepec, vicinity Santa María Yacochi, 2,300 m (17.158419; -96.030241) AY859471     II

Received: September 07, 2022; Revised: October 26, 2022; Accepted: December 07, 2022

*Corresponding author: https://orcid.org/0000-0001-8709-9514

Creative Commons License This is an open-access article distributed under the terms of the Creative Commons Attribution License