Botany

Phylogeography and climate-associated morphological variation in the endemic white Quercus deserticola () along the Trans-Mexican Volcanic Belt

Journal: Botany

Manuscript ID cjb-2017-0116.R1

Manuscript Type: Article

Date Submitted by the Author: 26-Oct-2017

Complete List of Authors: Rodríguez-Gómez, Flor; Universidad Nacional Autonoma de Mexico, Escuela Nacional de Estudios Superiores Morelia Oyama, Ken;Draft Escuela Nacional de Estudios Superiores (ENES) Unidad Morelia, UNAM, Laboratorio Nacional de Análisis y Síntesis Ecológica para la Conservación de los Recursos Genéticos Ochoa-Orozco, Magaly; Universidad Nacional Autonoma de Mexico, Instituto de Investigaciones en Ecosistemas y Sustentabilidad Mendoza-Cuenca, Luis; Universidad Michoacana de San Nicolas de , Facultad de Biología Gaytán-Legaria, Ricardo; Universidad Nacional Autonoma de Mexico, Instituto de Investigaciones en Ecosistemas y Sustentabilidad González-Rodríguez, Antonio; Universidad Nacional Autonoma de Mexico, Instituto de Investigaciones en Ecosistemas y Sustentabilidad

Is the invited manuscript for consideration in a Special N/A Issue? :

Keyword: gene flow, Mexican , Ecological Niche Modeling, Last Glacial Maximum

https://mc06.manuscriptcentral.com/botany-pubs Page 1 of 48 Botany

Phylogeography and climate-associated morphological variation in the endemic

white oak Quercus deserticola (Fagaceae) along the Trans-Mexican Volcanic Belt

Flor RodríguezGómez1, Ken Oyama1, Magaly OchoaOrozco2, Luis MendozaCuenca3

Ricardo GaytánLegaria2 and Antonio GonzálezRodríguez2*

1Escuela Nacional de Estudios Superiores Unidad Morelia, UNAM, Antigua Carretera a

Pátzcuaro 8701, Col. Ex Hacienda de San José de la Huerta, 58190, Morelia, Michoacán,

México. FRG: [email protected]; KO: [email protected]

2Instituto de Investigaciones en EcosistemasDraft y Sustentabilidad, Universidad Nacional

Autónoma de México, Antigua Carretera a Pátzcuaro No. 8701, Col. ExHacienda de

San José de la Huerta, 58190 Morelia, Michoacán, México. MOO: [email protected];

RGL: [email protected]

*Correspondence: Antonio GonzálezRodríguez, Fax: +52 (443) 322 2719; Email:

[email protected]

3 Laboratorio de Ecología de la Conducta, Facultad de Biología, Universidad

Michoacana de San Nicolás de Hidalgo, Avenida Francisco J. Múgica S/N, Morelia,

Michoacán, México. LMC: [email protected]

1 https://mc06.manuscriptcentral.com/botany-pubs Botany Page 2 of 48

ABSTRACT

Mexico is a center of diversification for the genus Quercus, with an important number of

taxa occurring along the TransMexican Volcanic Belt (TMVB). However, the impact of

the interaction between historical and current climatic variation and geological

heterogeneity in the TMVB on the genetic and phenotypic diversification within oak

species has been scarcely investigated. We used chloroplast DNA microsatellites and a

geometric morphometrics analysis of leaf shape to understand differentiation between populations of Quercus deserticola, which inhabits dry highlands along the TMVB.

Ecological niche modeling (ENM) for presentday conditions and projections into past

scenarios were performed to evaluateDraft the influence of environmental variables on the

evolutionary history of the species. Results showed high genetic diversity (hS =0.774) and

high genetic structure (RST =0.75) and the morphological subdivision of populations into

two clusters, corresponding to the west/south and east/north sectors of the Q. deserticola geographic distribution. ENM indicated that the potential distribution of the species has remained similar from the late Pleistocene to the present. Seemingly, the phylogeographic structure of the species has been shaped by low seedmediated gene flow and mostly local migration patterns. In turn, leaf shape is responding to climate differences either through phenotypic plasticity or local adaptation.

Keywords: gene flow, Mexican oaks, Ecological Niche Modeling, Last Glacial

Maximum.

2 https://mc06.manuscriptcentral.com/botany-pubs Page 3 of 48 Botany

INTRODUCTION

The TransMexican Volcanic Belt (TMVB) is a center of diversification, endemism and

biogeographic transition of the Mexican biota (Halffter and Morrone 2017). The TMVB

is a volcanic mountain chain with nearly 8000 volcanic structures, extending about 1200

km west to east through central Mexico, from the Pacific coast to the Gulf of Mexico

coast. Furthermore, the TMVB has a large environmental heterogeneity and has

experienced important climatic changes from the Pliocene and Pleistocene to the present

(GómezTuena et al. 2007; Ferrari et al. 2012). Based on age, orogeny and tectonic

features, the TMVB has been dividedDraft into four sectors (western, central, eastern and

easternmost), each with its own characteristics (GómezTuena et al. 2005; Ferrari et al.

2012). The TMVB has also been considered as a complex biogeographic unit (i. e. it

shows a high degree of species endemism and diversity), with two sectors, west and east

(Gámez et al. 2012; TorresMiranda et al. 2013). Four main episodes of volcanic activity

of the TMVB have occurred during different periods from the early Miocene to the

present, affecting this region asynchronously, first the western and later the eastern

sectors (GómezTuena et al. 2005; Gámez et al. 2012; Ferrari et al. 2012).

Several studies have found that the physiographic context of the TMVB has been

important in the genetic structuring and phenotypic divergence of different species and

how climatic and geologic events have modified their distributions in various time

periods (JaramilloCorrea et al. 2008; Gámez et al. 2012; RuizSánchez and Specht 2013;

TorresMiranda et al. 2013; MastrettaYanes et al. 2015; RodríguezGómez and Ornelas

3 https://mc06.manuscriptcentral.com/botany-pubs Botany Page 4 of 48

2015). The TMVB has been shown to be a geographic barrier that limits the dispersion of and animals that inhabit to the north and south of the barrier (AguirrePlanter et al.

2000; McCormack et al. 2011; ParraOlea et al. 2012; Gándara and Sosa 2014; Moreno

Letelier et al. 2014; Grummer et al. 2015; Jiménez and Ornelas 2016). Also, climatic

oscillations and geological activity during the Miocene and Pleistocene promoted

alternating periods of connection and disconnection of habitats along the TMVB,

impacting the genetic diversity of the populations. In particular, the colder periods of the

Pleistocene may have promoted increased connectivity among temperate habitats,

allowing the expansion of associated and animal populations across the TMVB,

which could have acted repeatedly as a corridor for the migration and possible gene flow

of organisms and even as a continuousDraft connection between the Sierra Madre Occidental

(SMOc) and the Sierra Madre Oriental (SMOr) regions (GonzálezRodríguez et al. 2004;

RuizSánchez and Specht 2013; MastrettaYanes et al. 2015; RodríguezGómez and

Ornelas 2015). Phylogeographic studies for species in this region have also shown that

historical isolation into multiple refugia played an important role in structuring genetic

diversity on the TMVB, sometimes followed by population expansion and increased

connectivity among habitats (ParraOlea et al. 2012; RuizSánchez et al. 2012; Velo

Antón et al. 2013; Ornelas and González 2014).

Mexico is considered a center of species diversification for the genus Quercus

(Fagaceae) (Manos et al. 1999; Hipp et al. 2017). This genus is represented in Mexico by

approximately 161 species, and 109 of these are endemic (Valencia 2004). Oaks can be

found throughout the country, except for the Yucatan Peninsula. However, the highlands,

such as the SMOc, SMOr, TMVB, Sierra Madre del Sur (SMS) and Sierras de Chiapas,

4 https://mc06.manuscriptcentral.com/botany-pubs Page 5 of 48 Botany

are the areas with the greatest oak species diversity (Rzedowski 1978; RodríguezCorrea

et al. 2015; RamírezToro et al. 2017). The TMVB, for example, contains 29% of the

total species of oaks in Mexico. A recent biogeographic study suggested an important

role of the TMVB on the distribution of oak species, identifying several areas of

endemism related to the presence of a wide variety of climatic zones, which may have

allowed the establishment of oak species with different climatic requirements

(RodríguezCorrea et al. 2015). However, evolutionary patterns and processes have been

studied for a few Mexican oak species, mainly from tropical lowlands (CavenderBares et

al. 2011, 2015), cloud and temperate forests (GonzálezRodríguez et al. 2004; Tovar

Sánchez et al. 2008; RamosOrtiz et al. 2016), and disturbed areas with a xerophytic

scrub type of vegetation at mid to highDraft altitude (ValenciaCuevas et al. 2014).

Nevertheless, the great ecological diversity of the Mexican oaks represents an interesting

opportunity to evaluate and compare how geological and climatic factors affected

congeneric species differing in habitat affinities.

Here, we focused on the Mexican endemic white oak Quercus deserticola Trel.

that inhabits high, cold and dry regions, around the foothills of the mountains in

xerophytic areas (at altitudes between 2000 to 2800 m), with a geographical distribution

mostly spanning the TMVB. Few studies in Mexico have assessed the genetic structure

and population history in plants within xerophytic vegetation in the highlands (Sosa et al.

2009; RuizSanchez et al. 2012; Gándara and Sosa 2014; ValenciaCuevas et al. 2014). In

this study, we used chloroplast DNA microsatellites (cpSSRs) to investigate

phylogeographic patterns in populations of Q. deserticola. We also used species

distribution models (SDMs) to evaluate if populations contracted (disconnected) or

5 https://mc06.manuscriptcentral.com/botany-pubs Botany Page 6 of 48

expanded (connected) during past climate fluctuations as observed in other plant species

inhabiting the highlands of the TMVB (e. g. RuizSanchez and Specht 2014).

Disconnections should have led to restrictions in gene flow promoting genetic

differentiation. On the contrary, if populations expanded their distribution, this should

have led to contact and gene exchange.

On the other hand, in oaks it has been shown that much of the morphological variation in leaves between individuals of the same species results from the influence of ecological factors (e. g. temperature, precipitation and light conditions) (González

Rodríguez and Oyama 2005; UribeSalas et al. 2008). Common garden experiments in some oak species have shown that variation in leaf size and shape has a genetic component (Nicoli et al. 2006) and sometimesDraft geographically matches the neutral genetic structure of the species (e. g. CavenderBares et al. 2011).

Therefore, the specific questions addressed in this study were: 1) What are the phylogeographic and historical demographic patterns in Q. deserticola populations spanning the TMVB? 2) Are there variations in leaf morphology among Q. deserticola populations along their geographical distribution? 3) How are morphological and genetic variations in Q. deserticola populations associated to geographic and climatic factors?

MATERIALS AND METHODS

Sampling, DNA extraction and PCR amplification

We collected 144 individuals from 13 populations of Q. deserticola throughout its

6 https://mc06.manuscriptcentral.com/botany-pubs Page 7 of 48 Botany

geographic distribution (Table 1, Fig. 1) for genetic and morphological analyses. Total

genomic DNA was extracted from 100 mg of leaf material using the Qiagen DNeasy

Plant Mini Kit (Qiagen, Valencia, CA, USA). Seven chloroplast DNA microsatellite loci

previously developed for the Fagaceae family were used in polymerase chain reactions

(PCR): the primer pairs UDT3 and UKK4 (Deguilloux et al. 2003), and CMCS4,

CMCS5, CMCS6, CMCS7 and CMCS12 (Sebastiani et al. 2004). PCRs were performed

using the Qiagen Multiplex PCR kit in a volume of 5 µl containing 1X Multiplex PCR

Master Mix, 2 µМ each primer, dH20, and 20 ng template DNA. The thermal cycling

conditions consisted of 35 cycles, each at 94°C for 30 s, annealing at 45°C for 30 s, and

extension at 72°C for 2 min. A final extension at 72 °C for 10 min was included.

Multiplex PCR products were combinedDraft with a GeneScan500 LIZ size standard and then

run in an ABIPRISM 3100Avant sequencer (Applied Biosystems). Fragments were

analyzed and sized with the Peak Scanner program 1.0 (Applied Biosystems).

Genetic diversity and phylogeographic structure

Genetic variation in each population was measured by calculating the number of

haplotypes (H), the number of private haplotypes (P), allelic richness (Ar ) and haplotype

diversity (hs) with the software HAPLOTYPE ANALYSIS Version 1.05 (Eliades and

Eliades 2009) and SPAGeDi v1.1 (Hardy and Vekemans 2002). To depict genealogical

relationships among haplotypes, we constructed a haplotype network using the Median

joining algorithm in NETWORK v4.51.6 (Bandelt et al. 1999).

Genetic differentiation among populations was assessed by performing two

7 https://mc06.manuscriptcentral.com/botany-pubs Botany Page 8 of 48

hierarchal analyses of molecular variance (AMOVA), the first with the FST statistic

(based on the infinite alleles mutation model, IAM), and the second with RST (based on

the stepwise mutation model, SMM). Pairwise RST comparisons among populations were

also estimated in Arlequin v3.5 (Excoffier and Lischer 2010). The pairwise comparisons

and AMOVA significance tests were run with 10,000 permutations. The pattern of

genetic structure was further evaluated with the clustering analysis with linked loci

implemented in the BAPS v0.6 software (Corander et al. 2008). We ran the model four

times for K = 18 to ensure convergence and chose the K with the highest log likelihood

(Corander et al. 2008).

Phylogeographic structure was assessed by calculating genetic differentiation with

unordered alleles (GST) and ordered Draftalleles (NST) in SPAGeDi v1.1 (Hardy and Vekemans

2002). SPAGeDi implements a permutation test to evaluate if the values of GST and NST are significantly different. A higher value of NST than GST indicates phylogeographic

structure among populations, resulting from the presence of closely related haplotypes

within the same populations (Pons and Petit 1996). Significance of NST and GST values

was determined by 10,000 random permutations of individuals among populations

(Hardy and Vekemans 2002).

Demographic history

The demographic history of Q. deserticola populations was investigated by means of

mismatch distributions and neutrality tests carried out in Arlequin v3.5 (Excoffier and

Lischer 2010). The mismatch distribution (Harpending 1994) can be used to determine

8 https://mc06.manuscriptcentral.com/botany-pubs Page 9 of 48 Botany

whether a population has undergone a sudden population expansion. We tested for

deviations of the observed mismatch distributions from those expected under the model

of Schneider and Excoffier (1999) with 1000 bootstrap replicates. The validity of the

sudden expansion assumption was determined using the sum of squares differences

(SSD). We also used the Tajima´s D statistic (Tajima 1989) and Fu´s FS test (Fu 1997) to

assess demographic expansions. Both tests were run with 1000 bootstrap replicates. For

all these tests, the cpSSR data were binary coded following Navascués et al. (2009) with

the number of repeats coded as ‘1’ and shorter alleles being coded filling the differences

in repeats as ‘0’.

Morphological variation Draft

A geometric morphometrics approach was used for the analysis of morphological

variation. Analyses were performed on photographs (abaxial side) of ten randomly

chosen, fully extended, undamaged leaves from each individual. Coordinates ‘x, y’ of 29

unambiguous and repeatable anatomical marks (i. e. 12 landmarks and 17 semi

landmarks) were registered along the border of each leaf image using the program

TpsDig (Rohlf 2005). We constructed a ‘‘fan’’ (radial guidelines with equal angular

spacing on images) with 20 radial guidelines covering the whole leaf contour, which was

used to digitalize the 17 semilandmarks. Two additional marks were placed on the ruler

for size reference. The MakeFan6 program from the ‘‘Integrated Morphometrics

Package’’ IMP series (http://www.canisius.edu/~sheets/morphsoft.html) was used for this

procedure.

9 https://mc06.manuscriptcentral.com/botany-pubs Botany Page 10 of 48

A Procrustes superimposition analysis for the configuration of landmarks and semi

landmarks was developed with the CoordGen6 program in the IMP series

(http://www.canisius.edu/~sheets/morphsoft.html). This analysis allows the calculation of

leaf shape variation without the effect of the size. The first step of superimposing

configurations of landmarks in twodimensional shapes (x1, y1, x2, y2 . . .) is a

generalized least squares Procrustes superimposition that minimizes differences between

landmark configurations by translation, scaling, and rotation to remove all information

unrelated to shape and to obtain shape variables (Procrustes distances; Rohlf 1990). After

the superimposition, resulting Procrustes coordinates were averaged across all ten leaves by individual. Shape variables (Procrustes distances average) for all individuals were

used for a discriminant function analysisDraft to determine the variation in leaf shape among populations using SPSS 22.0 (IBM Corp. Released 2013).

Ecological niche modeling

An Ecological Niche Model (ENM) was constructed in order to compare the extent of the potential distribution of Quercus deserticola through time, i.e. between the present, the

Last Glacial Maximum (LGM, 21 ka) and the Last Interglacial (LIG, 120 ka). Presence

records of Q. deserticola were obtained from the Global Biodiversity Information Facility

(GBIF, http://data.gbif.org/species/browse/taxon/), and filtered according to the

geographical distribution of Q. deserticola proposed by Valencia (2004). To avoid

overfitting due to the spatial correlation of the occurrences resulting from unequal

collection effort across the species’ range (Boria et al. 2014), records were thinned to be

10 https://mc06.manuscriptcentral.com/botany-pubs Page 11 of 48 Botany

spaced at least 20 km from each other with the spThin package (AielloLammens et al.

2015) in R (R Core Team 2014). After these procedures, the potential distribution of Q.

deserticola was modeled with a total of 58 unique records. The area of accessibility (M)

was defined as the biogeographic provinces proposed by Morrone (2005), where there are

records of the species, that is, the TMVB, the SMOr, the SMOc, the Balsas Depression

(BD) and the SMS. We used M as a mask to reduce the overprediction of area suitability

(habitat suitability) as well as to perform a better model validation (Barve et al. 2011).

To construct the ENM, 19 bioclimatic variables were obtained from the WorldClim

Global Climate Data V. 1.4 (http://www.worldclim.org/version1) with a resolution of 30

arcsec (1km2). To reduce redundancy among variables, we simultaneously considered

those variables with the highest partialDraft contribution to the first two principal components

from a Principal Components Analysis (PCA) and with pairwise Spearman’s rank

correlations lower than 0.9. These analyses were performed with the JMP v13 software

(SAS Institute). The final dataset included 11 variables: mean diurnal range of the

temperature (BIO2), isothermality (BIO3), temperature seasonality (BIO4), maximum

temperature of the warmest month (BIO5), temperature annual range (BIO7), mean

temperature of the coldest quartet (BIO11), annual precipitation (BIO12), precipitation

seasonality (BIO15), precipitation of the driest quartet (BIO17), precipitation of the

warmest quartet (BIO18) and precipitation of the coldest quartet (BIO19).

The algorithm of maximum entropy implemented in MAXENT v3.3.3 (Phillips and

Dudik 2008) was used to construct the ENM using the eleven bioclimatic variables. To

evaluate the performance of the model, a random subset of 20% of the total unique

records was set aside, and the area under the curve (AUC) of the receiver operating

11 https://mc06.manuscriptcentral.com/botany-pubs Botany Page 12 of 48

characteristic (ROC) was measured. We ran 100 replicates of the model with a

convergence threshold of 105 and 500 iterations. In the settings, we disabled the extrapolation and clamping options to avoid overprediction. To compare present and past potential distribution of Q. deserticola, the ENM was projected into two general global circulation models (GCM) used as past climate scenarios for the LGM: the Community

Climate System Model (CCSM; Collins 2006) and the Model for Interdisciplinary

Research on Climate (MIROC; Hasumi and Emori 2004). Both models simulate climatic conditions as they are calculated to have been for the LGM, with a stronger temperature decrease assumed in CCSM compared to MIROC (OttoBliesner et al. 2007). Also, we projected the ENM into the LIG climate scenario. The GCM data for the LGM and the

LIG were downloaded from WorlClimDraft (http://www.worldclim.org/paleoclimate1) (Otto

Bliesner et al. 2006; Braconnot et al. 2007).

Association among genetic structure, morphological variation and environmental

distances

We examined the relationship of genetic and morphological differentiation with

environmental conditions and with geographic distances (Euclidean distances) across all

13 sampling localities. First, morphological and environmental distances were calculated

as dissimilarity matrices with Euclidean distances scaling to values between 0 and 1. We

obtained the environmental distance matrix from the values of the 19 climate variables

from WorldClim (Hijmans et al. 2005) for each locality. For the morphological distance

matrix, we performed a discriminant function analysis (DFA) using the locality as the

12 https://mc06.manuscriptcentral.com/botany-pubs Page 13 of 48 Botany

grouping variable, and the distance between localities was estimated as the Euclidean

distance between group centroids for the first discriminant function. These dissimilarity

matrices and DFA were estimated in SPSS 22.0 (IBM Corp. Released 2013). For genetic

distances we used the pairwise RST matrix. We used partial Mantel test to assess if the

genetic composition of the populations and morphological dissimilarity are associated

with environmental variables while controlling for the potential effects of geographic

distance (Mantel 1967). Mantel tests were performed in IBD (Jensen et al. 2005) with

1,000 randomizations.

RESULTS

Draft

Genetic diversity and phylogeographic structure

Fiftyfour haplotypes were found in the 13 sampled populations (Table 2, Fig. 1), with

most of the localities exhibiting more than one haplotype. The haplotype network (Fig.

1b) showed several closed loops and no clear haplogroups could be distinguished. The

most frequent haplotypes were H12 (12 individuals, 8.3%) and H20 (11 individuals,

7.6%) and the most widespread haplotype was H22, found in four populations (30.7%).

The number of haplotypes per population ranged from two to nine. The Amealco

population was the one that shared more haplotypes (four) with other populations (Table

2, Fig. 1a). Sierra de Agustinos and Chiluca populations did not share haplotypes with

other populations. Allelic richness (Ar) and haplotype diversity (hS) were higher for the

Amealco (3.73 and 0.955) and the Huichapan (3.73 and 0.956) populations and Sierra

13 https://mc06.manuscriptcentral.com/botany-pubs Botany Page 14 of 48

Agustinos showed the lowest values (1.8 and 0.378, respectively) (Table 2). The BAPS

clustering analysis for linked loci supported two main groups (K = 2, log likelihood =

666.47) (Fig. 2). One of the clusters is more frequent (blue in Fig. 2) and is present in all populations except in the Pablo Ixtayoc population (13). The second group (in red in Fig.

2) was found in nine populations. The two groups did not show a clear geographical structuring, but the red group seems to be in a higher frequency towards the east and in the northernmost population than in western/southern populations.

The mean of withinpopulation genetic diversity (hS, 0.774) and the total genetic

diversity (hT, 0.984) were high. The analysis of molecular variance (AMOVA) revealed significant genetic differentiation among populations; for the analysis under the IAM,

42.33% of the variation was explainedDraft by differences among populations and the 57.67% by differences within populations, and for the SMM, 75.97% was explained by

differences among populations and 24.03% within populations (Table 3) and

corresponding ΦST values were highly significant (0.42, P < 0.0001 and 0.75, P <0.0001, respectively for IAM and SMM) (Table 3). Pairwise RST values were higher and

significant for Sierra Agustinos, Chiluca and Pablo Ixtayoc (Table S1), as expected from

the result that Sierra Agustinos and Chiluca populations did not share any haplotype with

other populations, while the Pablo Ixtayoc population only shared one haplotype (H33)

with Xihuingo.

Values of both NST (0.75) and GST (0.235) indicated overall genetic differentiation

across populations. The permutation test for the comparison of NST and GST was significant (P< 0.003), indicating phylogeographic structure or that more closely related haplotypes tend to occur together in the same populations.

14 https://mc06.manuscriptcentral.com/botany-pubs Page 15 of 48 Botany

Demographic history

In the mismatch distribution analysis, the null model of population expansion was not

rejected except for the Sierra Agustinos and Tecajete populations, suggesting a

demographic expansion for the rest of the populations (Table 4). The values of Tajima´s

D statistic did not depart significantly from neutrality in any population, while Fu’s FS

test detected the signal of demographic expansion in the Amealco, Cerro los Pitos,

Chiluca and Huichapan populations (Table 4), all of them located in the eastern part of

the TMVB.

Draft

Morphological variation

The first and second axes of the discriminant function derived from the morphometric

data analysis explained 62% of total variation in leaf shape and significantly

discriminated between two morphological clusters (for F1, Wilks’ lambda = 0.002, d. f. =

312, P < 0.0001; for F2, Wilks’ lambda = 0.011, d. f. = 275, P < 0.0001). One cluster was

formed by populations on the west and south of the TMVB (Santa Fe, Volcán de Colima,

Agustinos, Chiluca and Pablo Ixtayoc) and the other cluster was formed by populations

on the east and north of the TMVB (Matehuala, Amealco, Cerro Pitos, Huichapan, Mesa

Burro, Pachuca, Tecajetes and Xihuingo) (Figs. 1 and 3).

Ecological niche modeling

15 https://mc06.manuscriptcentral.com/botany-pubs Botany Page 16 of 48

The ENM yielded a good fit to the current distribution of Q. deserticola (AUC > 0.96)

(Fig. 4a, b). The potential distribution predicted for the LGM scenarios suggest the presence of the species in roughly the same regions where it currently occurs (Figs. 4c, d), with some differences between the CCSM and the MIROC scenarios, particularly regarding the eastern and southern parts of the distribution, which appear to have higher suitability values under the CCSM scenario. In contrast, the potential distribution in the

LIG model appears contracted and shifted to the south, with some areas in the Sierra

Madre del Sur showing the highest habitat suitability values, and a decreased prediction of habitat suitability in the northernmost areas of the TMVB (Fig. 4e). The analysis of the contribution of each variable to the modelDraft indicated that the variables that contributed the most were temperature seasonality (BIO4) with 36.7% (permutation importance 52.7%), mean temperature of coldest quarter (BIO11) with 25% (permutation importance 27.5%) and the maximum temperature of warmest month (BIO5) with 19.9% (permutation importance 5.7%).

Association among genetic, morphological and environmental distances

We observed a positive correlation between morphological and environmental distances while controlling for geographic distance (r = 0.32, P = 0.031). However, we found no correlation between genetic (pairwise RST) and morphological distances (r = 0.10, P =

0.22), nor a relationship between genetic and environmental distances (r = 0.17, P =

0.77) while controlling for geographic distance. These analyses suggest that

16 https://mc06.manuscriptcentral.com/botany-pubs Page 17 of 48 Botany

environmental conditions are contributing in maintaining the morphological differences

between geographical regions.

DISCUSSION

Phylogeographic structure and demographic history

The populations of Q. deserticola were characterized by high genetic diversity and

considerable differentiation, as well as significant phylogeographic structure. Most

previous studies of oak species using cpDNA markers have found similar high

differentiation levels and a clear geographic segregation of haplotypes lineages, which

are usually explained by the historicalDraft migration patterns of the populations and the very

low dispersal capacity of acorns (Magni et al. 2005; Grivet et al. 2006; Magri et al. 2007;

Marsico et al. 2009). However, in Q. deserticola the distribution of haplotypes was rather

patchy, displaying strong local genetic structure but without obvious phylogeographic

breaks, contrasting with some previous studies in the TMVB that have found a westeast

pattern of lineage differentiation (e. g. ParraOlea et al. 2012; RuizSánchez and Specht

2014; PérezCrespo et al. 2017) consistent with the idea that climatic oscillations and

geological activity during the Miocene and Pleistocene promoted alternating periods of

connection and disconnection between different habitats along this mountain range

(JaramilloCorrea et al. 2008; Ornelas and González 2014; MastrettaYanes et al. 2015).

In the case of Q. deserticola, the observed palaeodistribution obtained from

ecological niche modeling supports the scenario in which populations of this species on

the TMVB probably expanded northwards from the LIG to the LGM, but were largely

17 https://mc06.manuscriptcentral.com/botany-pubs Botany Page 18 of 48

stable from the LGM to the present. However, changes in the altitudinal distribution,

resulting in the expansion and contraction of temperate forests might have occurred, as it

is suggested for different taxa and by palynological records that indicate abundance of

temperate forests in valleys and lowlands during the LGM (Metcalfe 2006; Caballero et

al. 2010; RamírezBarahona and Eguiarte 2013; VillanuevaAmadoz et al 2014;

MastrettaYanes et al. 2015). The relative historical range stability of Q. deserticola from the LGM to the present, in combination with the complex topography, the scattered distribution of suitable sites for the colonization of Q. deserticola and the limited dispersal capacity of acorns are factors that probably enhanced genetic drift as a predominant force shaping the distribution of cpDNA haplotypes in this species. These

results are similar to those observedDraft for Q. lobata in California, which also showed a

mosaic distribution of haplotypes and signs of range stability and local migration patterns

from the late Pleistocene to the present (Gugger et al. 2013).

The historical demography tests gave contrasting results. The mismatch

distribution analyses suggested demographic expansions for most populations, while Fu’s

Fs tests detected expansions only in four populations from the eastern part of the TMVB

and the values of Tajima’s D were not significant for any population. This might be

explained by the different sensitivity of the tests to demographic expansions and to

sample size (RamosOnsins and Rozas 2002). As previously mentioned, the species

distribution modelling suggested a northwards range expansion from the LIG to the

LGM, and the MIROC model suggested a modest eastwards expansion from the LGM to

the present along the TMVB. Therefore, it is possible that the populations to the north

and the east of the TMVB were more recently colonized than the populations to the south

18 https://mc06.manuscriptcentral.com/botany-pubs Page 19 of 48 Botany

and the west. Interestingly, a recent phylogeographic study of the mistletoe Psittacanthus

calyculatus (Loranthaceae) which frequently uses Q. deserticola as a host (PérezCrespo

et al. 2017), found evidence of a south to north expansion and colonization of the TMVB

from the Sierra Madre del Sur during the late Pleistocene; but afterwards the direction of

the expansion within the TMVB was from east to west. However, this parasitic plant has

a broad host range and seems to follow its own ecological niche rather than closely

tracking its hosts (RamírezBarahona et al. 2017).

Morphological variation

Two morphologically differentiated Draftpopulation groups were identified within Q.

deserticola. The first grouped populations from the west/south of the TMVB and the

second was formed by populations from the east/north. The two morphological groups

are partially concordant with the previous regionalization of the TMVB into western and

eastern biogeographic sectors (Gámez et al. 2012; TorresMiranda et al. 2013), and with

genetic subdivisions found in other organisms (ParraOlea et al. 2012; RuizSánchez and

Specht 2014; PérezCrespo et al. 2017). However, according to the partial Mantel tests, in

Q. deserticola the association between morphological and genetic distances among

populations was not significant while controlling for the geographic distance. While a

correlation between phenotypic variation and neutral genetic markers may occur if the

same neutral processes (e. g. gene flow and genetic drift) have historically affected the

spatial structure of both types of variation (e. g. CavenderBares et al. 2011), in this case

it seems like the variation patterns in cpDNA markers and leaf morphology of Q.

19 https://mc06.manuscriptcentral.com/botany-pubs Botany Page 20 of 48

deserticola are the result of different processes. As previously mentioned, the limited

dispersal capacity of acorns, coupled with the heterogeneous landscape, the sparse

distribution of suitable sites for the colonization of Q. deserticola and the relative historical range stability from the LGM to the present, have promoted a patchy distribution of haplotypes with a strong local genetic structure. In contrast, the geographical subdivision of leaf morphological variation was significantly associated to climate variation, while controlling for geographical or genetic distances among populations. For example, the sites that constitute the western group are characterized by a higher mean annual precipitation (mean ± standard error = 829 ± 55.3 mm) and a higher precipitation seasonality (93.7 ± 3.89), than the sites that constitute the eastern group

(537.2 ± 74.2 mm and 76.2 ± 5.22, respectively).Draft Similarly, previous studies on

California oaks (Riordan et al. 2016) and Q. rugosa in Mexico (UribeSalas et al. 2008) have also indicated strong and significant effects of climate on leaf morphology when controlling for geographical distances among populations. However, whether the association of leaf morphological variation with climate in Q. deserticola reflects phenotypic plasticity, common ancestry or adaptive differentiation among populations is

unclear. The incongruence between leaf morphology and genetic differentiation can be

explained because phenotypic plasticity as a response to environmental variation can

modify morphological characters without a corresponding genetic differentiation pattern.

Also, phenotypic differences could be due to local adaptation as a result of selection pressures at loci not linked to neutral genetic markers. This is particularly likely in our

case since such selected loci are expected to be in the nuclear DNA and largely

independent of cpDNA haplotype variation. In any case, the two morphological groups

20 https://mc06.manuscriptcentral.com/botany-pubs Page 21 of 48 Botany

identified in Q. deserticola are clearly reflecting the climatic differences between the

east/south and west/north portions of its distribution along the TMVB, but an

understanding of the causes of this relationship would require common garden

experiments or a landscape genomics approach (e. g. Sork et al. 2016).

CONCLUSIONS

In this study, we suggest that different evolutionary processes are likely to be involved in

shaping the genetic structure and phenotypic variation of Q. deserticola populations

distributed along of the TMVB. First, a pattern supported by local sharing of haplotypes

and a high genetic differentiation is Draftprobably due to low seed dispersal and restricted

range displacements during recent historical periods. Second, morphological variation is

also found between populations to the west/south and east/north sectors of the Q.

deserticola geographic distribution, as a response to environmental variables. The

genetic, morphological and environmental patterns found in Q. deserticola are significant

to continue understanding the evolutionary history of the species that inhabit the TMVB

region.

ACKNOWLEDGMENTS

We thank V. Rocha, H. RodríguezCorrea and J. LlanderalMendoza for technical

support. FRG thanks the postdoctoral scholarship provided by Dirección General de

Asuntos del Personal Académico (DGAPA) of the Universidad Nacional Autónoma de

21 https://mc06.manuscriptcentral.com/botany-pubs Botany Page 22 of 48

México (UNAM). This project was partially supported by DGAPAPAPIIT IV201015 and CONACYT 240136 grants.

REFERENCES

AguirrePlanter, E., JaramilloCorrea, J.P., GómezAcevedo, S., Khasa, D.P., Bousquet,

J., and Eguiarte, L.E. 2012. Phylogeny, diversification rates and species boundaries of

Mesoamerican firs (Abies, Pinaceae) in a genuswide context. Mol. Phylogenet. Evol.

62(1): 263–274. Available from http://dx.doi.org/10.1016/j.ympev.2011.09.021

Aiello-Lammens, M. E., Boria, R.A.,Draft Radosavljevic, A., Vilela, B., and Anderson, R. P.

2015. spThin: an R package for spatial thinning of species occurrence records for use in ecological niche models. Ecography, 38(5): 541545.

Bandelt, H.J, Forster, P., and Röhl, A. 1999. Medianjoining networks for inferring intraspecic phylogenies. Mol. Biol. Evol. 16(1): 3748. Available from https://doi.org/10.1093/oxfordjournals.molbev.a026036

Barve, N., Barve, V., JiménezValverde, A., LiraNoriega, A., Maher, S. P., Peterson, A.

T., Soberón, J., and Villalobos, F. 2011. The crucial role of the accessible area in ecological niche modeling and species distribution modeling. Ecol. Model. 222(11):

18101819. doi.org/10.1016/j.ecolmodel.2011.02.011

22 https://mc06.manuscriptcentral.com/botany-pubs Page 23 of 48 Botany

Boria, R.A., Olson, L.E., Goodman, S.M. and Anderson, R.P. 2014. Spatial filtering to

reduce sampling bias can improve the performance of ecological niche model.

Ecol. Model. 275:7377. doi.org/10.1016/j.ecolmodel.2013.12.012

Braconnot, P., OttoBliesner, B., Harrison, S., et al. 2007. Results of PMIP2 coupled

simulations of the MidHolocene and Last Glacial Maximum – Part 2: feedbacks with

emphasis on the location of the ITCZ and mid and high latitudes heat budget. Clim. Past.

3: 279–296. doi:10.5194/cp32792007

Caballero, M., LozanoGarcía, S., VázquezSelem, L., and Ortega, B. 2010. Evidencias

de cambio climático y ambiental en Draftregistros glaciales y en cuencas lacustres del centro

de México durante el último máximo glacial. Bol. Soc. Geol. Mex. 62(3): 359–377.

Available from http://www.scielo.org.mx/scielo.php?pid=S1405

33222010000300005&script=sci_arttext&tlng=en

Cavender-Bares, J., González-Rodríguez, A., Eaton, D. A., Hipp, A. A., Beulke, A., and

Manos, P. S. 2015. Phylogeny and biogeography of the American live oaks (Quercus

subsection Virentes): A genomic and population genetics approach. Mol. Ecol. 24(14):

36683687. doi:10.1111/mec.13269

CavenderBares, J., GonzálezRodríguez, A., Pahlich, A., Koehler, K., and Deacon, N.

2011. Phylogeography and climatic niche evolution in live oaks (Quercus series Virentes)

23 https://mc06.manuscriptcentral.com/botany-pubs Botany Page 24 of 48

from the tropics to the temperate zone. J. Biogeogr. 38(5): 962981. doi:10.1111/j.1365

2699.2010.02451.x

Collins, W. D., Bitz, C., Blackmon, M., Bonan, G., Bretherton, C. S., et al. 2006. The

community climate system model Version 3 (CCSM3). J. Climate, 19, 2122–2143.

Available from http://dx.doi.org/10.1175/JCLI3761.1

Corander, J., Marttinen, P., Siren, J., and Tang, J. 2008. Enhanced Bayesian modelling in

BAPS software for learning genetic structures of populations. BMC Bioinformatics, 9:

539. doi.org/10.1186/147121059539

Draft

Deguilloux, M.F., DumolinLapègue, S., Gielly, L., Grivet, D., and Petit, R. 2003. A set

of primers for the amplification of chloroplast microsatellites in Quercus. Mol. Ecol.

Notes, 3(1): 24–27. doi:10.1046/j.14718286.2003.00339.x

Eliades, NGH., and Eliades, D.G. 2009. HAPLOTYPE ANALYSIS: Software for

analysis of haplotype data. Distributed by the authors. Forest Genetics and Forest Tree

Breeding, GeorgAugust University Goettingen, Germany. Available from

https://www.unigoettingen.de/en/134935.html (accessed 25 October 2016)

Excoffier, L., and Lischer, H.E.L. 2010. Arlequin suite ver 3.5: a new series of programs

to perform population genetics analyses under Linux and Windows. Mol. Ecol. Resour.

10(3): 564–567. doi:10.1111/j.17550998.2010.02847.x

24 https://mc06.manuscriptcentral.com/botany-pubs Page 25 of 48 Botany

Ferrari, L., OrozcoEsquivel, T., Manea, V., and Manea, M. 2012. The dynamic history

of the TransMexican Volcanic Belt and the Mexico subduction zone. Tectonophysics,

522: 122–149. doi:https://doi.org/10.1016/j.tecto.2011.09.018

Fu, Y.X. 1997. Statistical tests of neutrality against population growth, hitchhiking and

background selection. Genetics, 147: 915925. Available from

http://www.genetics.org/content/genetics/147/2/915.full.pdf

Gámez, N., Escalante, T., Rodríguez, G., Linaje, M., and Morrone, J.J. 2012.

Caracterización biogeográfica de la DraftFaja Volcánica Transmexicana y análisis de los

patrones de distribución de su mastofauna. Rev. Mex. Biodiv. 83(1): 258272. Available

from http://www.scielo.org.mx/scielo.php?script=sci_arttext&pid=S1870

34532012000100028

Gándara, E., and Sosa, V. 2014. Spatiotemporal evolution of Leucophyllum pringlei and

allies (Scrophulariaceae): A group endemic to North American xeric regions. Mol.

Phylogenet. Evol. 76: 93101. Available from

http://dx.doi.org/10.1016/j.ympev.2014.02.027

GómezTuena, A., OrozcoEsquivel, M.T., and Ferrari, L. 2005. Petrogénesis ígnea de la

Faja Volcánica Transmexicana. Bol. Soc. Geol. Mex. 3(57): 227283. Available from

http://s3.amazonaws.com/academia.edu.documents/30924814/94320266003.pdf?AWSAc

25 https://mc06.manuscriptcentral.com/botany-pubs Botany Page 26 of 48

cessKeyId=AKIAIWOWYYGZ2Y53UL3A&Expires=1488566256&Signature=XJIOqvp

7ljbQaFZzCeVc%2F7dVZKg%3D&responsecontent

disposition=inline%3B%20filename%3DPetrogenesis_ignea_de_la_faja_Volcanica.pdf

GómezTuena, A., OrozcoEsquivel, M.T., and Ferrari, L. 2007. Igneous petrogenesis of

the TransMexican Volcanic Belt. GSA Special Papers, 422 : 129–181.

doi1:0.1130/2007.2422(05)

GonzálezRodríguez, A., Arias, D.M., S. Valencia, S., and Oyama, K. 2004.

Morphological and RAPD analysis of hibridization between Quercus affinis and Q. laurina (Fagaceae), two Mexican redDraft oaks. Am. J. Bot. 91(3): 401409.

doi:10.3732/ajb.91.3.401

GonzálezRodríguez, A., and Oyama, K. 2005. Leaf morphometric variation in Quercus

affinis and Q. laurina (Fagaceae), two hybridizing Mexican red oaks. Bot. J. Linnean

Soc. 147(4): 427435. doi: 10.1111/j.10958339.2004.00394.x

Grivet, D., Deguilloux, M. F., Petit, R. J., and Sork, V. L. 2006. Contrasting patterns of

historical colonization in white oaks (Quercus spp.) in California and Europe. Mol. Ecol.

15(13): 40854093. doi:10.1111/j.1365294X.2006.03083.x

Grummer, J. A., CalderónEspinosa, M. L., Nieto-Montes de Oca, A., Smith, E. N.,

Méndez de la Cruz, F. R., and Leaché, A. D. 2015. Estimating the temporal and spatial

26 https://mc06.manuscriptcentral.com/botany-pubs Page 27 of 48 Botany

extent of gene flow among sympatric lizard populations (genus Sceloporus) in the

southern Mexican highlands. Mol. Ecol. 24(7): 15231542. doi:10.1111/mec.13122

Gugger, P. F., Ikegami, M., and Sork, V. L. 2013. Influence of late Quaternary climate

change on present patterns of genetic variation in valley oak, Quercus lobata Née. Mol.

Ecol. 22(13): 35983612. doi: 0.1111/mec.12317

Halffter, G., and Morrone, J. J. 2017. An analytical review of Halffter's Mexican

transition zone, and its relevance for evolutionary biogeography, ecology and

biogeographical regionalization. Zootaxa, 4226(1): 146. Available from

http://mapress.com/j/zt/article/view/zootaxa.4226.1.1Draft

Hardy, O. J., and Vekemans, X. 2002. SPAGeDi: a versatile computer program to

analyse spatial genetic structure at the individual or population levels. Mol. Ecol. Notes,

2(4): 618–620. doi:10.1046/j.14718286.2002.00305.x

Harpending, H. C. 1994. Signature of ancient population growth in a lowresolution

mitochondrial DNA mismatch distribution. Hum. Biol. 66: 591600. Available from

http://www.jstor.org/stable/41465371

Hasumi, H., and Emori, D. 2004. K1 coupled GCM (MIROC) description. Center for

Climate System Research, University of Tokyo, Tokyo, Japan.

27 https://mc06.manuscriptcentral.com/botany-pubs Botany Page 28 of 48

Hijmans, R. J., Cameron, S. E., Parra, J. L., Jones, P. G., and Jarvis, A. 2005. Very high

resolution interpolated climate surfaces for global land areas. Int. J. Climatol. 25(15):

1965–1978. doi:10.1002/joc.1276

Hipp, A. L., Manos, P. S., González-Rodríguez, A., Hahn, M., Kaproth, M., McVay, J.

D., Valencia Avalos, S., and CavenderBares, J. 2017. Sympatric parallel diversification

of major oak clades in the Americas and the origins of Mexican species diversity. New

Phytol. doi: 10.1111/nph.14773

JaramilloCorrea, J. P., AguirrePlanter, E., Khasa, D. P., Eguiarte, L. E., Piñero, D.,

Furnier, G. R., and Bousquet, J. 2008.Draft Ancestry and divergence of subtropical montane

forest isolates: molecular biogeography of the genus Abies (Pinaceae) in southern Mexico

and Guatemala. Mol. Ecol. 17(10): 2476–2490. doi:10.1111/j.1365294X.2008.03762.x

Jensen, J. L., Bohonak, A. J., and Kelley, S. T. 2005. Isolation by distance, web service.

BMC Genetics, 6:13. v.3.23. Available from http://ibdws.sdsu.edu/ (accessed 22

September 2016)

Jiménez, R. A. and Ornelas, J. F. 2016. Historical and current introgression in a

Mesoamerican hummingbird species complex: a biogeographic perspective. PeerJ, 4: e1556. Available from https://doi.org/10.7717/peerj.1556

28 https://mc06.manuscriptcentral.com/botany-pubs Page 29 of 48 Botany

Magri, D., Fineschi, S., Bellarosa, R., Buonamici, A., Sebastiani, F., Schirone, B.,

Simeone, M. C., and Vendramin, G. G. 2007. The distribution of Quercus suber

chloroplast haplotypes matches the palaeogeographical history of the western

Mediterranean. Mol. Ecol. 16: 52595266. doi:10.1111/j.1365294X.2007.03587.x

Magni, C. R., Ducousso, A., Caron, H., Petit, R. J., and Kremer, A. 2005. Chloroplast

DNA variation of Quercus rubra L. in North America and comparison with other

Fagaceae. Mol. Ecol. 14(2): 513524. doi:10.1111/j.1365294X.2005.02400.x

Manos, P. S., Doyle, J. J., and Nixon, K. C. 1999. Phylogeny, biogeography, and

processes of molecular differentiationDraft in Quercus subgenus Quercus (Fagaceae). Mol.

Phylogenet. Evol. 12(3): 333349. Available from

http://dx.doi.org/10.1006/mpev.1999.0614

Mantel, N. 1967. The detection of disease clustering and a generalized regression

approach. Cancer Res. 27: 209–220. Available from

http://cancerres.aacrjournals.org/content/27/2_Part_1/209.fulltext.pdf

Marsico, T. D., Hellman, J. J., RomeroSeverson, J. 2009. Patterns of seed dispersal and

pollen flow in Quercus garryana (Fagaceae) following postglacial climatic changes. J.

Biogeogr. 36:929941. doi:10.1111/j.13652699.2008.02049.x

29 https://mc06.manuscriptcentral.com/botany-pubs Botany Page 30 of 48

Mastretta-Yanes, A., Moreno-Letelier, A., Pinero, D., Jorgensen, T. H., and Emerson,

B.C. 2015. Biodiversity in the Mexican highlands and the interaction of geology, geography and climate within the TransMexican Volcanic Belt. J. Biogeogr. 42(9):

15861600. doi:10.1111/jbi.12546

McCormack, J. E., Heled, J., Delaney, K. S., Peterson, A. T., and Knowles, L. L. 2011.

Calibrating divergence times on species trees versus gene trees: implications of

speciation history of Aphelocoma jays. Evolution, 65(1): 184–202. doi:10.1111/j.1558

5646.2010.01097.x

Metcalfe, S. E. 2006. Late QuaternaryDraft environments of the northern deserts and central

Transvolcanic Belt of Mexico. Ann. Mo. Bot. Gard. 93(2): 258273. Available from

http://dx.doi.org/10.3417/00266493(2006)93[258:LQEOTN]2.0.CO;2

MorenoLetelier, A., MastrettaYanes, A., and Barraclough, T.G. 2014. Late Miocene

lineage divergence and ecological differentiation of rare endemic Juniperus blancoi: clues for the diversification of North American conifers. New Phytol. 203(1): 335347.

doi:10.1111/nph.12761

Morrone, J. J. 2005. Hacia una síntesis biogeográfica de México. Rev. Mex.

Biodiv. 76:(2) 207252. Available from http://scielo.unam.mx/scielo.php?pid=s1870

34532005000200006&script=sci_arttext

30 https://mc06.manuscriptcentral.com/botany-pubs Page 31 of 48 Botany

Navascués, M., Hardy, O.J., and Burgarella, C. 2009. Characterization of demographic

expansions from pairwise comparisons of linked microsatellite haplotypes. Genetics,

181:1013–1019. doi: https:/doi.org/10.1534/genetics.108.098194

Nicoli, N. P., Krstic, B. D., Pajevic, S. P., and Orlovic, S. S. 2006. Variability of leaf

characteristics in different pedunculate oak genotypes (Quercus robur L.). Proc. Nat. Sci.

Matica Srpska Novi Sad. 111: 95–105. doi:10.2298/ZMWSPN0611095N

Ornelas, J. F., and González, C. 2014. Interglacial genetic diversification of Moussonia

deppeana (Gesneriaceae), a hummingbirdpollinated, cloud forest shrub in northern

Mesoamerica. Mol. Ecol. 23(16): 4119–4136.Draft doi:10.1111/mec.12841

OttoBliesner, B. L., Hewitt, C.D., Marchitto, T. M., Brady, E., AbeOuchi, A., Crucifix,

M., Murakami, S., and Weber, S.L. 2007. Last Glacial Maximum ocean thermohaline

circulation: PMIP2 model intercomparisons and data constraints. Geophys. Res. Lett.

34(12): L12706. doi:10.1029/2007GL029475

OttoBliesner, B. L., Marshall, S. J., Overpeck, J. T., Miller, G. H. and Hu, A. X. and

CAPE Last Interglacial Project members. 2006. Simulating Arctic climate warmth and

icefield retreat in the Last Interglaciation. Science, 311(5768): 1751–1753.

doi:10.1126/science.1120808

31 https://mc06.manuscriptcentral.com/botany-pubs Botany Page 32 of 48

ParraOlea, G., Windfield, J. C., Velo Antón, G., and Zamudio, K. R. 2012. Isolation in

habitat refugia promotes rapid diversification in a montane tropical salamander. J.

Biogeogr. 39(2): 353–370. doi:10.1111/j.13652699.2011.02593.x

Pérez-Crespo, M. J., Ornelas, J. F., GonzálezRodríguez, A., Ruiz-Sánchez, E.,

Vásquez-Aguilar, A. A., and Ramírez-Barahona, S. 2017. Phylogeography and population differentiation in the Psittacanthus calyculatus (Loranthaceae) mistletoe: a

complex scenario of climate–volcanism interaction along the TransMexican Volcanic

Belt. J. Biogeogr. doi 10.1111/jbi.13070

Phillips, S. J., and Dudik, M. 2008. DraftModeling of species distributions with Maxent: new

extensions and a comprehensive evaluation. Ecography, 31(2): 161–175.

doi:10.1111/j.09067590.2008.5203.x

Pons, O., and Petit, R. J. 1996. Measuring and testing genetic differentiation with ordered

versus unordered alleles. Genetics, 144(3): 1237–1245. Available from

http://www.genetics.org/content/144/3/1237.articleinfo

RamírezBarahona, S., and Eguiarte, L. E. 2013. The role of glacial cycles in promoting

genetic diversity in the Neotropics: the case of cloud forests during the Last Glacial

Maximum. Ecol. Evol. 3(31): 725–738. doi:10.1111/14421984.12109

32 https://mc06.manuscriptcentral.com/botany-pubs Page 33 of 48 Botany

Ramírez-Barahona, S., González, C., González-Rodríguez, A., and Ornelas, J. F. 2017.

The influence of climatic niche preferences on the population genetic structure of a

mistletoe species complex. New Phytol. 214(4): 17511761. doi 10.1111/nph.14471

RamírezToro, W., TorresMiranda, A., GonzálezRodríguez, A., RuizSánchez, E.,

LunaVega, I., and Oyama, K. 2017. A Multicriteria Analysis for Prioritizing Areas for

Conservation of Oaks (Fagaceae: Quercus) in , Southern Mexico. Trop. Conserv.

Sci. 10, 1940082917714227. doi: 10.1177/1940082917714227

RamosOnsins, S. E., and Rozas, J. 2002. Statistical properties of new neutrality tests

against population growth. Mol. Biol.Draft Evol. 19(12):2092–2100.

https://doi.org/10.1093/oxfordjournals.molbev.a004034

Ramos-Ortiz, S., Oyama, K., Rodríguez-Correa, H., and González-Rodríguez, A. 2016.

Geographic structure of genetic and phenotypic variation in the hybrid zone between

Quercus affinis and Q. laurina in Mexico. Plant Spec. Biol. 31(3): 219232.

doi:10.1111/14421984.12109

Riordan, E. C., Gugger, P. F., Ortego, J., Smith, C., Gaddis, K., Thompson, P., and Sork,

V. L. 2016. Association of genetic and phenotypic variability with geography and climate

in three southern California oaks. Am. J. Bot. 103(1): 7385. doi:10.3732/ajb.1500135

33 https://mc06.manuscriptcentral.com/botany-pubs Botany Page 34 of 48

RodríguezCorrea, H., Oyama, K., MacGregorFors, I., and GonzálezRodríguez, A.

2015. How are oaks distributed in the neotropics? A perspective from species turnover,

areas of endemism, and climatic niches. Int. J. Plant. Sci. 176(3): 222231.

doi:10.1086/679904

Rodríguez-Gómez, F., and Ornelas, J. F. 2015. At the passing gate: past introgression in

the process of species formation between Amazilia violiceps and A. viridifrons

hummingbirds along the Mexican Transition Zone. J. Biogeogr. 42(7): 13051318.

doi:10.1111/jbi.12506 Rohlf, F.J. 1990. Rotational fit (Procrustes)Draft methods. In Proceedings of the Michigan Morphometrics Workshop. Edited by F. J. Rohlf, F. Bookstein. University of Michigan

Museums of Zoology, Ann Arbor. pp. 227–236.

Rohlf, F. J. 2005. tpsDig, digitize landmarks and outlines, ver 2.04. Department of

Ecology and Evolution, State University of New York at Stony Brook.

RuizSánchez, E., RodríguezGómez, F., and Sosa, V. 2012. Refugia and geographic barriers of populations of the desert poppy, Hunnemannia fumariifolia (Papaveraceae).

Org. Divers. Evol. 12(2): 133143. doi:10.1007/s131270120089z

RuizSánchez, E., and Specht, C.D. 2013. Influence of the geological history of the

TransMexican Volcanic Belt on the diversification of Nolina parviflora (Asparagaceae:

Nolinoideae). J. Biogeogr. 40(7): 13361347. doi:10.1111/jbi.12073

34 https://mc06.manuscriptcentral.com/botany-pubs Page 35 of 48 Botany

RuizSánchez, E., and Specht, C.D. 2014. Ecological speciation in Nolina parviflora

(Asparagaceae): lacking spatial connectivity along of the TransMexican Volcanic

Belt. PloS one, 9(6): e98754. Available from

http://dx.doi.org/10.1371/journal.pone.00987544

Rzedowski, J. 1978. La vegetación de México, Limusa, México

Schneider, S., and Excoffier, L. 1999. Estimation of demographic parameters from the

distribution of pairwise differences when the mutation rates vary among sites:

Application to human mitochondrial DNA. Genetics, 152:1079–1089. Available from

http://www.genetics.org/content/152/3/1079.short Draft Sebastiani, F., Carnevale, S., and Vendramin G. G. 2004. A new set of mono and

dinucleotide chloroplast microsatellites in Fagaceae. Mol. Ecol. Notes. 4(2): 259–261.

doi:10.1111/j.14718286.2004.00635.

Sork, V. L., Squire, K., Gugger, P. F., Steele, S. E., Levy, E. D., Eckert, A. J. 2016.

Landscape genomic analysis of candidate genes for climate adaptation in a California

endemic oak, Quercus lobate. Am. J. Bot. 103(1): 3346. doi: 10.3732/ajb.1500162

Sosa, V., Ruiz-Sánchez, E., and Rodríguez-Gómez, F. C. 2009. Hidden phylogeographic

complexity in the Sierra Madre Oriental: the case of the Mexican tulip poppy

Hunnemannia fumariifolia (Papaveraceae). J. Biogeogr. 36(1): 1827.

doi:10.1111/j.13652699.2008.01957.x

35 https://mc06.manuscriptcentral.com/botany-pubs Botany Page 36 of 48

Tajima, F. 1989. Statistical method for testing the neutral mutation hypothesis by DNA polymorphism. Genetics, 123: 585–595. Available from

http://www.genetics.org/content/genetics/123/3/585.full.pdf

TorresMiranda, A., LunaVega, I., and Oyama, K. 2013. New approaches to the biogeography and areas of endemism of red oaks (Quercus L., Section Lobatae). Syst.

Biol. 62(4): 555573. Available from https://doi.org/10.1093/sysbio/syt021

TovarSánchez, E., MussaliGalante, P., EstebanJiménez, R., Piñero, D., Arias, D.M.,

Dorado, O., and Oyama, K. 2008. Chloroplast DNA polymorphism reveals geographic

structure and introgression in QuercusDraft crassifolia × Q. crassipes complex in México.

Botany, 86(3): 228–239. Available from http://www.nrccnrc.gc.ca/main_e.html

UribeSalas, D., SáenzRomero, C., GonzálezRodríguez, A., TéllezValdéz, O., and

Oyama, K. 2008. Foliar morphological variation in the white oak Quercus rugosa Née

(Fagaceae) along a latitudinal gradient in Mexico: Potential implications for management

and conservation. Forest Ecol. Manag. 256(12): 21212126. doi:10.1016/j.foreco.2008.08.002

ValenciaCuevas, L., Piñero, D., MussaliGalante, P., ValenciaÁvalos, S., and Tovar

Sánchez, E. 2014. Effect of a red oak species gradient on genetic structure and diversity

of Quercus castanea (Fagaceae) in Mexico. Tree Genet. Genomes, 10(3) 641652. doi:

10.1007/s1129501407108

36 https://mc06.manuscriptcentral.com/botany-pubs Page 37 of 48 Botany

Valencia, S. 2004. La diversidad del género Quercus (Fagaceae) en México. Bol. Soc.

Bot. Méx. 75: 33–53. Available from http://www.redalyc.org/articulo.oa?id=57707503

VeloAntón, G., Parra, J. L., Parra-Olea, G., and Zamudio, K. R. 2013. Tracking climate

change in a dispersal-limited species: reduced spatial and genetic connectivity in a

montane salamander. Mol. Ecol. 22(12): 32613278. doi:10.1111/mec.12310

VillanuevaAmadoz, U., CalvilloCanadell, L., and CevallosFerriz, S. R. S. 2014.

Síntesis de los trabajos paleobotánicos del Cretácico en México. Bol. Soc. Geol.

Mex. 66 (1): 97121. Available fromDraft http://www.scielo.org.mx/scielo.php?pid=S1405

33222014000100009&script=sci_arttext

37 https://mc06.manuscriptcentral.com/botany-pubs Botany Page 38 of 48

Table 1. Geographical localization of thirteen studied Q. deserticola populations.

Locality State Elevation (masl) Latitude Longitude

1. Nevado de Colima 2870 19.47 103.56

2. Santa Fe Jalisco 2546 19.27 102.50

3. Matehuala San Luis Potosí 2640 22.58 101.50

4. Sierra Agustinos 2600 20.23 100.67

5. Amealco Querétaro 2595 20.43 100.27

6. Xihuingo Hidalgo 2702 20.72 99.86

7. Huichapan Hidalgo 2337 20.23 99.51

8. Chiluca Estado de México 2565 19.54 99.31

9. Mesa del Burro Hidalgo Draft2400 20.10 99.03

10. Tecajete Hidalgo 2582 19.93 98.91

11. Cerro de los Pitos Hidalgo 2700 19.48 98.79

12. Pachuca Hidalgo 2649 19.95 98.60

13. Pablo Ixtayoc Estado de México 2670 19.28 98.55

38 https://mc06.manuscriptcentral.com/botany-pubs

hs 39

, H31 = Allelic richness, richness, = Ar Allelic H32, H41, H22 H41, H32,

H14, H19, H18 H18 H19, H14, , H49, H50, H10 H10 H50, , H49, H28, H45, H53, H25, H27 H25, H53, H45, H28, , H37, H46, H47, H48, H51 H48, H47, H46, , H37,

H40, H31, H41 H31, H40, H29, H13, H16, H34 H34 H16, H13, H29, , H30, H17, H54, H52, H52, H54, H17, , H30, H43 , H9, H7 number of private haplotypes, haplotypes, private of number

H20 P = P , H31, H40, H31, H40 H44, H22, H11 H33, H22, H12, H12, H44, H22, H41, H20, H21, H32, H21 H44 H11, Haplotypes Haplotypes H7, H12 H12 H7, Botany

= sample size, sample = Draft https://mc06.manuscriptcentral.com/botany-pubs Q. deserticola.Q. n 10 10 4 2.98 0.778 6 12 12 6 2.88 0.758 6 H26, H2 H42, H5, H38, H23, 5 5 0 2 0.600 2 n n P Ar hs H number of haplotypes in each population. Shared haplotypes among populations are in bold. in populations are among Shared population. haplotypes in haplotypes each of number H H = 11. Cerro de los Pitos de Cerro 11. 16 4 3.38 0.883 9 4. Sierra Agustinos Agustinos Sierra 4. Amealco 5. 10 3 1.8 12 0.378 5 3 3.73 H15 H3, H1, 0.955 9 7. Huichapan Huichapan 7. 10 5 3.73 0.956 8 6. Xihuingo Xihuingo 6. 17 3 2.98 0.794 7 = genetic diversity, diversity, genetic = Locality 12. Pachuca Pachuca 12. 10 2 2.61 0.711 4 3. Matehuala Matehuala 3. 1. Nevado de Colima de Nevado 1. Fe Santa 2. 4 2 3 0.833 3 H36, H6, Table 2. Population genetic variability of variability genetic Population 2. Table 10. Tecajete Tecajete 10. 11 2 2.83 0.764 5 H39, H24, 8. Chiluca Chiluca 8. del Burro Mesa 9. 17 3 2.82 0.750 7 Page 39 of 48 Page 40 of 48 40

H33 Botany Draft https://mc06.manuscriptcentral.com/botany-pubs 11 11 3.2 2.86 0.754 5.6 Mean Mean 13. Pablo Ixtayoc Pablo 13. Ixtayoc 10 3 2.46 0.644 4 H35, H4, H8, Page 41 of 48 Botany

Table 3. Analysis of molecular variance (AMOVA) for the seven chloroplast microsatellites of Q. deserticola.

Source of variation d. f. Sum of Variance Percentage Fixation index squares components of variation ΦST FST (IAM)

Among populations 12 95.44 0.63 42.33 0.42*

Within populations 133 115.46 0.86 57.67

Total 145 210.11 1.50

RST (SMM) Among populations 12 30719.5Draft 223.82 75.97 0.75* Within populations 32 9414.2 70.78 24.03

Total 145 40133.7 294.61

*P < 0.00001

41 https://mc06.manuscriptcentral.com/botany-pubs Botany Page 42 of 48

Table 4. Summary statistics of historical demography analyses of Quercus deserticola. D

= Tajima’s D and FS = Fu’s Fs. Significant negative values (**P < 0.01 and *P < 0.05 in bold) indicate historical demographic expansion events. SSD = differences in the sum of squares or mismatch distribution. Significant (P ≤ 0.05 in bold) SSD values indicate deviations from the sudden expansion model.

Locality D Fs SSD 1. Nevado de Colima 0.2124 0.5563 0.1248 2. Santa Fe 1.2247 0.6261 0.0542 3. Matehuala 0.4331 0.9254 0.0204 4. Sierra Agustinos 0.6909 0.5938 0.2143 5. Amealco 0.4224 -3.6791* 0.0019 6. Xihuingo 0.5035 Draft 1.5221 0.0224 7. Huichapan 0.0248 -4.2247** 0.0223 8. Chiluca 0.4219 -2.9799** 0.0129 9. Mesa del Burro 1.0558 1.7538 0.0286 10. Tecajete 1.2177 0.8131 0.2257 11. Cerro los Pitos 0.0020 -3.9531** 0.0018 12. Pachuca 0.1542 0.8067 0.0584 13. Pablo Ixtayoc 1.1494 0.9275 0.0082

42 https://mc06.manuscriptcentral.com/botany-pubs Page 43 of 48 Botany

Figure legends.

Figure 1. (a) Frequency and geographic distribution of cpDNA haplotypes in 13 localities

of Q. deserticola in the TMVB, the number correspond to those in Table 1. The pie charts

represent the frequency of occurrence of haplotypes in each population, and colors

correspond to those shown in the haplotype network below. The black line encircles the

populations belonging to cluster one (west/south) of the morphological analysis (triangles

in Fig 2). (b) Haplotype statistical parsimony network. Each circle indicates an individual

haplotype and its size is proportional to the frequency of the haplotype. Lines represent a

single mutational change and black Draftcircles correspond to unsampled haplotypes (loops

are showed in the network).

Figure 2. Distribution map of genetic clusters inferred from BAPS based on chloroplast

microsatellites (cpSSRs) from Q. deserticola. The population numbers correspond to

those in the Table 1. Two clusters (K = 2) were retrieved, cluster 1 in blue and cluster 2 in

red.

Figure 3. Morphological differentiation between Q. deserticola populations. The first and

second discriminant functions are shown in x and yaxes. Colors represent the different

populations. Triangles correspond to western populations and circles to eastern of the

TMVB. Locality numbers correspond those in Table 1.

Figure 4. MAXENT analyses showing distribution points (a), and species distribution

models for Q. deserticola, at present (b), Last Glacial Maximum (LGM, CCSM, 21 ka)

43 https://mc06.manuscriptcentral.com/botany-pubs Botany Page 44 of 48

(c), Last Glacial Maximum (LGM, MIROC, 21 ka) (d) and Last Interglacial (120 ka) (e).

Light gray indicates areas with low suitability values for the occurrence of the species and dark gray indicates high suitability values.

Draft

44 https://mc06.manuscriptcentral.com/botany-pubs Page 45 of 48 Botany

Draft

264x342mm (300 x 300 DPI)

https://mc06.manuscriptcentral.com/botany-pubs Botany Page 46 of 48

Draft

144x101mm (300 x 300 DPI)

https://mc06.manuscriptcentral.com/botany-pubs Page 47 of 48 Botany

Draft

279x361mm (300 x 300 DPI)

https://mc06.manuscriptcentral.com/botany-pubs Botany Page 48 of 48

Draft

378x320mm (300 x 300 DPI)

https://mc06.manuscriptcentral.com/botany-pubs