Vous êtes sur la page 1sur 17

doi: 10.1111/jeb.

12255

REVIEW

Widespread phenotypic and genetic divergence along altitudinal gradients in animals


I. KELLER*, J. M. ALEXANDER*, R. HOLDEREGGER* & P. J. EDWARDS*
rich, Universit rich, Switzerland *Institute of Integrative Biology, ETH Zentrum CHN, ETH Zu atsstrasse 16, Zu Department of Fish Ecology and Evolution, EAWAG Swiss Federal Institute of Aquatic Science and Technology, Center of Ecology, Evolution and Biochemistry, Kastanienbaum, Switzerland Department of Aquatic Ecology and Macroevolution, Institute of Ecology and Evolution, University of Bern, Bern, Switzerland WSL Swiss Federal Research Institute, Birmensdorf, Switzerland

Keywords:
adaptation; common garden experiment; elevation; molecular adaptation; outlier scan; phenotypic divergence.

Abstract
Altitudinal gradients offer valuable study systems to investigate how adaptive genetic diversity is distributed within and between natural populations and which factors promote or prevent adaptive differentiation. The environmental clines along altitudinal gradients tend to be steep relative to the dispersal distance of many organisms, providing an opportunity to study the joint effects of divergent natural selection and gene ow. Temperature is one variable showing consistent altitudinal changes, and altitudinal gradients can therefore provide spatial surrogates for some of the changes anticipated under climate change. Here, we investigate the extent and patterns of adaptive divergence in animal populations along altitudinal gradients by surveying the literature for (i) studies on phenotypic variation assessed under common garden or reciprocal transplant designs and (ii) studies looking for signatures of divergent selection at the molecular level. Phenotypic data show that signicant between-population differences are common and taxonomically widespread, involving traits such as mass, wing size, tolerance to thermal extremes and melanization. Several lines of evidence suggest that some of the observed differences are adaptively relevant, but rigorous tests of local adaptation or the link between specic phenotypes and tness are sorely lacking. Evidence for a role of altitudinal adaptation also exists for a number of candidate genes, most prominently haemoglobin, and for anonymous molecular markers. Novel genomic approaches may provide valuable tools for studying adaptive diversity, also in species that are not amenable to experimentation.

Introduction
The geographical distribution of many species is so broad that various characteristics of their environment vary either abruptly or in a clinal manner within their range. A common pattern observed in response to such environmental heterogeneity is local adaptation, where, at a given location, the tness of local individuals is higher than that of immigrants from other
Correspondence: Irene Keller, Department of Clinical Research, Murtenstrasse 35, 3010 Bern, Switzerland. Tel.: +41 31 631 3018; fax: +41 31 631 4888; e-mail: irene.keller@dkf.unibe.ch

environments (Kawecki & Ebert, 2004). Local adaptation is possible if populations contain ecologically relevant genetic variation and if divergent selection between different environments is strong relative to the rate of gene ow (Morjan & Rieseberg, 2004). The distribution of adaptive genetic diversity and the factors promoting or preventing adaptive divergence are of fundamental interest to evolutionary ecologists but remain poorly characterized in most natural populations (see Hereford, 2009a for recent review). It is also largely unclear how consistently different species respond to similar selection pressures. A better understanding of these issues has direct implications for

2013 THE AUTHORS. J. EVOL. BIOL. JOURNAL OF EVOLUTIONARY BIOLOGY 2013 EUROPEAN SOCIETY FOR EVOLUTIONARY BIOLOGY

I. KELLER ET AL.

conservation and management, for example, by improving our ability to assess the future of populations in rapidly changing environments or to anticipate the effect of changes to population connectivity. In addition to the environmental changes observed in space, anthropogenic climate change is expected to lead to temporal changes, but its implications for local climatic conditions are likely to vary widely. Thermal environments at high latitudes, for example, may become more similar to the current thermal environments at lower latitudes. Other environmental variables, notably day length, are not expected to change, which may lead to completely novel environmental conditions. This suggests that range shifts (Parmesan & Yohe, 2003; Parmesan, 2006) may be insufcient for locally adapted populations to track their preferred (multidimensional) environment and additional responses are necessary. Phenotypic plasticity provides one mechanism to deal with environmental variability, but plastic responses may be possible only within certain limits, and evolutionary change may be necessary in the face of large and consistent environmental change (e.g. Gienapp et al., 2008). Indeed, a number of case studies report evidence of such microevolutionary changes in response to global warming in natural populations (reviewed in Bradshaw et al., 2006; Hoffmann & , 2011). Sgro Spatial gradients can serve as surrogates for at least some of the temporal changes anticipated under climate change (Reusch & Wood, 2007), providing an opportunity to investigate the historical and current responses of natural populations to climate-related selection pressures. Altitudinal gradients are particularly relevant in this context because they are also climate gradients. Some of the environmental changes along altitudinal gradients are specic to certain locations or biogeographical regions, whereas others, namely decreasing temperature, decreasing atmospheric pressure and rner, 2007), increasing intensity of solar radiation (Ko are physical properties shared by altitudinal gradients worldwide, allowing particular effects to be studied in numerous replicated systems. Altitudinal gradients offer a valuable contrast to latitudinal gradients, especially with respect to geographical scale. Altitudinal gradients are typically steep, with environmental transitions occurring at spatial scales that are small relative to the dispersal distances of many species. This has several important implications. First, it means that the effects of divergent selection may often be opposed by gene ow, which, if strong enough, acts to homogenize allele frequencies between environments. Altitudinal gradients thus provide the opportunity to investigate whether, and under which conditions, adaptive divergence is possible in the face of gene ow. Second, the small geographical scale of altitudinal gradients also implies that confounding effects, such as distinct regional evolutionary histories, are less

of an issue than in latitudinal surveys, which are often performed across thousands of kilometres (e.g. Balany a et al., 2006). Many examples exist of phenotypic transitions associated with the changing environment along altitudinal gradients. Plants often show conspicuous intraspecic differences in growth form or leaf morphology between rner, 2003), whereas phenohigh and low altitudes (Ko typic differences in animals can involve body size clines (Chown & Klok, 2003) or, perhaps more conspicuously, transitions from one generation per year at high altitudes to two or more at lower altitudes (Hodkinson, 2005). What is often less clear, however, is whether these phenotypic differences have a genetic basis or result entirely from plastic responses to the environment. As a step towards a better understanding of the distribution of adaptive genetic diversity along altitudinal gradients, we surveyed the literature for studies on genetic divergence in phenotypic traits and at functional loci in animal populations. Our particular goal was to identify general patterns emerging from these studies by asking (i) whether some taxa are more prone to differentiation than others, for example due to differences in specic species traits, (ii) whether adaptive divergence is apparent at all geographical scales or whether it tends to be rare between nearby populations where gene ow may be high and (iii) whether different species show similar responses to selection pressures associated with altitude.

Literature survey
We performed a literature survey to identify studies investigating genetic differentiation between animal populations (both aquatic and terrestrial) along altitudinal gradients. We used ISI Web of Knowledge to conduct a search for papers on (altitud* OR elevation*) AND (gradient OR transect OR cline) AND (genetic OR common garden OR transplant) NOT plant to obtain a list of ca. 500 publications. Based on the abstracts, we retained two types of studies. The rst consisted of papers reporting measurements of phenotypic traits for individuals from different altitudinal origins studied under common garden conditions or in a reciprocal transplant design. Data on phenotypic traits collected directly in the eld were included only in three cases where additional information supported a genetic basis for the trait (wing melanization, Ellers & Boggs, 2002, 2004a; macroptery, Fairbairn & King, 2009; and heatshock protein expression, Dahlhoff & Rank, 2000). The second group contained papers providing molecular evidence of adaptive genetic differences between populations at different altitudes. These studies typically investigated associations between genotypes and altitude or used outlier locus detection (e.g. Storz, 2005) to identify loci showing unusually high between-popula-

2013 THE AUTHORS. J. EVOL. BIOL. doi: 10.1111/jeb.12255 JOURNAL OF EVOLUTIONARY BIOLOGY 2013 EUROPEAN SOCIETY FOR EVOLUTIONARY BIOLOGY

Divergent altitudinal adaptation

tion differentiation. The focus was on intraspecic phenotypic or genetic diversity. Papers on incipient species pairs were included only if there was evidence of on-going gene ow between the two species. Additional studies were identied based on the bibliography in relevant papers as well as from thematically related review articles (Leinonen et al., 2008; Conover et al., 2009; Hereford, 2009; Nosil et al., 2009). Phenotypic data

affected by the native environment through maternal effects (Kawecki & Ebert, 2004).

Analysis and graphical overviews


We used the assembled data to investigate whether populations from different altitudes show genetically based differences in phenotypic traits, initially ignoring the adaptive signicance of these differences. Further, we investigated whether the altitudinal trends observed for a given trait were similar across species. This second analysis was conducted for 14 trait categories measured in at least ve different species, where each trait category included several related traits as indicated in Figs 1 and S1. Melanization was also included due to its potential relevance for thermoregulation, even though this trait was only studied in three species. To account for differences in the range of trait values, all observations were standardized within trait and study to mean 0 and variance 1. These standardized trait values were then regressed against the altitude of the source population assuming a linear relationship. Regressions were calculated separately for each trait, study, common garden environment and other subgroups (e.g. sex, age class or study year) if available. Figures 1 and S1 show the regression slopes for different trait categories, and Table 1 provides a summary of the main patterns. Studies including only one high- and one lowaltitude source population, where regression slopes were estimated based on only two data points, were distinguished from studies with multiple populations from each elevation. Additionally, we retained information about statistical signicances as reported in the original publications, distinguishing between signicant effects of altitude and of population. The latter category included (i) studies using only two source populations, in which case altitudinal and population effects could not be distinguished (e.g. Conover et al., 2009), (ii) reports that did not formally test for altitudinal effects or (iii) studies where signicant population differences were not associated with altitude. We did not distinguish between signicant main effects and signicant two-way interactions involving altitude or population. This inclusion of signicant interaction terms explains why slope estimates near zero are sometimes displayed as statistically signicant in the gures.

Data availability and methodological limitations


A total of 68 publications met our selection criteria for phenotypic data, and these contained data from 44 different species: 24 arthropods, 19 chordates and one mollusc. From 66 of these papers (Table S1), we were able to extract data for at least one phenotypic trait from tables or from gures using g3data (http://www. frantz./software/g3data.php). Among the arthropods, Diptera was the bestrepresented order with 14 different species, followed by Lepidoptera and Orthoptera with three species each. Among the chordates, two-thirds of the species were amphibians or reptiles. The maximum altitudinal distance between sampling locations ranged from 126 m n-Velarde et al., (Eales et al., 2010) to 4000 m (Leo 1996). Most studies were performed at a relatively small scale with a median Euclidian distance of 135 km between the two most distant source populations (range: 53900 km, distances were estimated using Google Maps if not provided in the original publication). The majority of papers reported data on phenotypic traits recorded from individuals reared under the same environmental conditions. Although such common garden experiments are valuable to investigate the genetic basis of traits (but see caveat below), the adaptive signicance of between-population differences cannot be inferred. Such an assessment would require results from reciprocal transplants or at least from multiple common garden experiments, which try to mimic the range of natural conditions associated with changes in altitude. Such reciprocal transplant experiments have been performed only for three species (the frog Rana sylvatica, Berven, 1982a,b; the buttery Colias philodice eriphyle, Ellers & Boggs, 2004b; the lizard Psammodromus algirus, Iraeta et al., 2006), in all cases in addition to common garden experiments. Further, to exclude effects of the native environment on phenotypes, experimental animals should be reared in a common environment for two or more generations prior to measurement (Kawecki & Ebert, 2004). This was the case in only about one-third of the studies, all of them on ies. Another 13% of the studies used wild-caught individuals (F0), whereas the majority used laboratory-reared offspring of wild-caught animals (F1; 38%). In these cases, the phenotype may still be

Molecular data
The second focus of our literature survey was on studies looking for signatures of divergent altitudinal selection at the molecular level. Many of these studies made use of outlier locus detection (e.g. Storz, 2005) to determine whether between-population divergence at a given gene or anonymous marker was signicantly higher than the genomic average. Additionally, we included studies showing altitudinal clines in the frequency of particular alleles at candidate genes or anonymous regions of the genome. Similar to the

2013 THE AUTHORS. J. EVOL. BIOL. doi: 10.1111/jeb.12255 JOURNAL OF EVOLUTIONARY BIOLOGY 2013 EUROPEAN SOCIETY FOR EVOLUTIONARY BIOLOGY

I. KELLER ET AL.

(a)

(b)

(c)

(d)

2013 THE AUTHORS. J. EVOL. BIOL. doi: 10.1111/jeb.12255 JOURNAL OF EVOLUTIONARY BIOLOGY 2013 EUROPEAN SOCIETY FOR EVOLUTIONARY BIOLOGY

Divergent altitudinal adaptation

Fig. 1 Observed changes in phenotypic traits in animals along altitudinal gradients. Shown are slopes from linear regression of trait value against altitude of the source population, with trait values standardized to mean 0 and variance 1 within trait and study. Separate estimates are shown for each data set, where data sets can be different common garden environments, latitudes, age classes, sexes, etc. Each row represents a different species and summarizes data from one or several studies, as indicated in the ref column. Several related traits were combined into each trait category as specied in the traits column. Shading indicates if the original publication reported statistically signicant effects of altitude (blue) or population (orange), a nonsignicant (n.s.) effect (red) or did not provide test results (empty circle). Note that we did not distinguish between main effects and interactions involving altitude or population. The inclusion of signicant interaction terms explains why some slope estimates near zero are displayed as statistically signicant. The size of the circles indicates the number of source populations available to estimate the slopes (small: 2 sources; large: > 2 sources). Asterisks indicate traits for which the sign of the slope was reversed to produce consistent patterns across traits. For the wing size traits, for instance, wing loading is the only trait where a smaller trait value would indicate a larger wing. Consequently, if unadjusted, an increase in wing size with altitude would result in a negative slope for wing loading but a positive slope for all remaining traits. (a) Traits: a, wing/thorax ratio; b, wing loading*; c, wing length; d, wing width; e, wing centroid size; f, wing area. Ref: 1 = Bears et al. (2008), 2 = Bubliy & Loeschke (2004), 3 = Dahlgaard et al. (2001), 4 = Norry et al. (2001), 5 = Sambucetti et al. (2006), 6 = Collinge et al. (2006) 7 = Pitchers et al. (2012), 8 = Stalker & Carson (1948), 9 = Tantowijoyo & Hoffmann (2011), 10 = Belen et al. (2004), 11 = Karan et al. (2000), 12 = Karl et al. (2008). *Sign of slope reversed. (b) Traits: a, chill coma recovery time*; b, cold shock survival; c, lower limiting T for embryonic development*. Ref: 1 = Beattie (1987), 2 = Bridle et al. (2009), 3 = Sarup et al. (2009), 4 = Sorensen et al. (2005), 5 = Collinge et al. (2006), 6 = Parkash et al. (2010), 7 = Karl et al. (2008). *Sign of slope reversed. (c) Traits: a, tadpole; b, at metamorphosis; c, pupa; d, at hatching; e, adult. Ref: 1 = Ficetola & De Bernardi (2005), 2 = Jasienski (2009), 3 = Sommer & Pearman (2003), 4 = Buckley et al. (2010), 5 = Bears et al. (2008), 6 = Stillwell & Fox (2009), 7 = Karan et al. (2000), 8 = Karl et al. (2008). (d) Traits: a, embryonic development time (dt); b, larval dt; c, postdiapause dt; d, pupal dt; e, egg-adult; f, hatching-adult. Ref: 1 = Beattie (1987), 2 = Jasienski (2009), 3 = Marquis & Miaud (2008), 4 = Tsuchiya et al. (2012), 5 = Bubliy & Loeschcke (2004), 6 = Folguera et al. (2008), 7 = Norry et al. (2001), 8 = Sambucetti et al. (2006), 9 = Collinge et al. (2006), 10 = Etges (1989), 11 = Belen & Alten (2006), 12 = Blanckenhorn (1997), 13 = Karl et al. (2008), 14 = Tanaka & Brookes (1983), 15 = Dingle et al. (1990), 16 = Berner et al. (2004). Table 1 Number of animal species showing signicant phenotypic differences between altitudinal populations reared in a common environment for different traits. Patterns are summarized based on the detailed gures. (A) An increase in trait value with altitude is supported by all statistically signicant tests based on > 2 populations (i.e. all large blue circles > 0 and no large orange circles). (B) A decrease in trait value with altitude is supported by all statistically signicant tests based on > 2 populations (i.e. all large blue circles < 0 and no large orange circles). (C) Statistically signicant differences between populations are not or not consistently associated with altitude. This category includes only species in which altitudinal effects were formally tested. In particular, estimates based only on two populations (small circles) are not considered. In parentheses, we indicate the number of studies relying on animals reared in a common environment for two or more generations before the experiments. Traits are listed in the order in which they appear in the text.
Number of species showing Expected altitudinal pattern Increase Decrease Increase Unclear Variable Increase (?) Increase (?) Decrease (?) Increase (?) Decrease (?) Unclear Unclear (C) Pop differences not associated with alt 2 2 2 3 2 1 0 3 1 3 2 3 (2) (2) (2) (2) (2) (1) (2) (0) (2) (2) (3) # different taxonomic groups* 3 3 3 4 1 6 4 5 3 2 3 4

Trait category Wing size Heat tolerance Cold tolerance HSP expression Desiccation tolerance Mass Body length Development time Growth rate Longevity Viability Fecundity

Putative agent(s) of selection Air density T T T, other stressors Water availability T, T, T, T, resource availability resource availability season length season length

(A) Consistent increase 3 (3) 0 3 (2) 0 0 3 (1) 0 0 0 0 0 1 (1)

(B) Consistent decrease 0 2 (1) 0 2 (0) 0 0 2 3 0 1 0 1

# species 9 5 6 5 5 7 7 13 6 6 11 11

Details Fig. Fig. Fig. Fig. Fig. Fig. Fig. Fig. Fig. Fig. Fig. Fig. 1a S1a 1b S1b S1c 1c S1d 1d S1e S1f S1g S1h

(0) (0) (0) (1)

HSP, heat-shock protein; T, temperature. *Number of different orders (for Arthropoda) or classes (for Chordata, Mollusca). HSP expression is a very complex and general response to cellular stress, and predictions are difcult. The observed response may depend heavily on common garden conditions (e.g. if these are closer to high- or low-altitude conditions). Water availability often changes with altitude, but the direction of the change may vary among regions. If mortality risks increase with altitude due to increased environmental stochasticity, this could select for higher fecundity early in life. However, the data set contains only two estimates of fecundity early in life and no trends can be inferred.
2013 THE AUTHORS. J. EVOL. BIOL. doi: 10.1111/jeb.12255 JOURNAL OF EVOLUTIONARY BIOLOGY 2013 EUROPEAN SOCIETY FOR EVOLUTIONARY BIOLOGY

I. KELLER ET AL.

genome scan approaches, these patterns should ideally be compared to those at putatively neutral genetic markers to exclude the possibility that clines result from purely neutral processes (e.g. isolation by distance; Storz, 2002). Such data from neutral loci were, however, not always available (see comments in Table 2). We identied 30 studies that present evidence of divergent selection under the criteria outlined above (Table S1). These studies investigate 22 different species, and the taxonomic focus was less biased towards particular groups (i.e. Diptera) than in the phenotypic data set. Genetically based phenotypic variation along altitudinal gradients In many studies, phenotypic traits measured in common garden environments varied signicantly between populations from different altitudinal origins. Among the statistical tests performed in the original publications, 73% detected signicant differences between source populations or altitudes (including main and interaction effects; Fig. 2). A very similar proportion of signicant test results was observed when we considered only studies using experimental animals bred in a common environment for at least two generations (70%; based on 142 tests from 27 studies). We then asked whether traits measured in multiple species tended to show the same changes along altitudinal gradients. For the trait categories measured in ve or more species, clear predictions of the variation with altitude could be formulated for three (Table 1), whereas for several additional trait categories, the expected patterns were more difcult to predict (Table 1). In the following discussion, we particularly focus on species showing statistically signicant associations with altitude for particular traits (as reported in the original study), especially if these are consistent across data subsets or studies (columns A and B in Table 1). We additionally report all species for which signicant between-population differences in a given trait were detected, but for which trait values did not change linearly with altitude (column C in Table 1). This latter category includes only species for which altitudinal effects were formally tested. The results from additional studies in which the experimental design precluded testing altitudinal effects or where such effects were either not tested or not signicant are shown in Figs 1 and S1 but not considered in Table 1 and the following discussion. First, air density decreases with altitude, and in addition to its signicance for respiration, this also implies that more power is needed for ight. A possible adaptive response includes an increase in wing size relative to body size (e.g. Dillon et al., 2006). In our data set, traits related to wing size were investigated in seven y

species, a buttery and a bird (Fig. 1a), but only for two traits wing to thorax ratio (a in Fig. 1a) and wing loading (b in Fig. 1a) relative to body size. In all three studies reporting signicant altitudinal effects, wing size increased consistently with altitude (Fig. 1a; Table 1). Air temperature is a second environmental variable that shows consistent altitudinal clines, dropping an rner, 2007). Not suraverage 5.5C per 1000 m (e.g. Ko prisingly, traits potentially relevant to thermal adaptation were well represented in our data set, including diverse morphological, physiological, developmental and behavioural traits. The most obvious prediction is that the average cold tolerance of individuals should increase with altitude, whereas heat tolerance should decrease (Table 1). For heat tolerance, some signicant between-population differences were reported in all ve species (Fig. S1a), but only in two cases were these differences related to altitude. In both, heat tolerance decreased with altitude as predicted. Cold tolerance was investigated in four Drosophila species, one frog and one buttery (Fig. 1b). In three of these species, an increase in cold tolerance was observed in populations from higher altitudes, while in a further two signicant population differences were reported that were not associated with altitude (Table 1). Interestingly, the latter studies were all conducted with populations from equatorial regions (< 30 north/south), whereas all studies with higher latitude populations did nd a positive correlation of cold tolerance with altitude. Heat-shock proteins (Fig. S1b), which are involved in general responses to cellular stress (Morris et al., 2013), showed either decreasing expression levels with altitude (two species) or between-population differences that were not consistently associated with altitude (three species; Table 1). Clinal change with altitude is also predicted for body melanization, a morphological trait that probably has thermoregulatory relevance, as darker bodies absorb more energy (Clusella Trullas et al., 2007). Body melanization along altitudinal gradients was studied in only three species, with the expected positive correlation with altitude being found in each case (buttery Colias philodice eriphyle, Ellers & Boggs, 2002, 2004a; Drosophila melanogaster, Sub-Saharan Africa: Pool & Aquadro, 2007; India: Parkash et al., 2008, 2010; Drosophila americana, Wittkopp et al., 2011). In addition to the thermoregulatory advantages, darker individuals could be better protected against elevated UV radiation at higher altitudes, and in some Drosophila populations, body pigmentation also shows a strong positive correlation with desiccation tolerance (Parkash et al., 2008; but not Wittkopp et al., 2011; see Fig. S1c for results on desiccation tolerance). The altitudinal patterns expected for traits related to body size are less clear (Table 1). According to Bergmanns rule (e.g. Gardner et al., 2011), endotherms tend to be larger in cooler environments due to the

2013 THE AUTHORS. J. EVOL. BIOL. doi: 10.1111/jeb.12255 JOURNAL OF EVOLUTIONARY BIOLOGY 2013 EUROPEAN SOCIETY FOR EVOLUTIONARY BIOLOGY

Table 2 Candidate loci and anonymous markers (outlier loci) showing evidence of adaptive differentiation along altitudinal gradients in animals.
Evidence of adaptive signicance Potential Gene selective Elevated Fst cline Comments Reference Altitudinal agent Gene/Marker function

Common

Species

name

Candidate genes Hemoglobin Hemoglobin Hemoglobin Hemoglobin ND3 Possibly temperature subunit 3 (mitochondrial) (OXPHOS) pathway Control region Non-coding Possibly temperature domain (HVII) of the mitochondrial control region EGLN1 hypoxia-inducible factor (HIF) pathway egl nine homolog 1 Enzyme in the O2 partial pressure Second hypervariable Non-shivering thermogenesis is inuenced by interaction between haplotype and sex Putative high-altitude alleles reduce hemoglobin concentration. High altitude variants differ between Tibetans and Andeans EPAS1 Transcription factor involved in the induction of genes regulated by oxygen Enzyme in the hypoxia-inducible factor (HIF) pathway; involved in nitric oxide (NO) synthesis, which plays a role in vasodilation and increased blood ow PPARA activated receptor alpha Peroxisome proliferatorEnzyme in the hypoxia-inducible factor (HIF) pathway O2 partial pressure Putative high-altitude alleles reduce hemoglobin concentration. Outlier status of PPARA was not conrmed in Peng et al. (2011) PRKAA1 activated, alpha 1 catalytic subunit Albumin Protein kinase, AMPEnzyme in the hypoxiainducible factor (HIF) pathway; regulation of cellular ATP Maintenance of biochemical equilibria in body uids Hemoglobin a-globin (HBA-T1, HBA- T2) and b-globin (HBB-T1, HBB-T2) O2 transport O2 partial pressure Storz et al. (2009) Unclear, possibly O2 partial pressure Storz & Dubach (2004) O2 partial pressure Bigham et al. (2010) Bigham et al. (2010), Simonson et al. (2010), Peng et al. (2011) O2 partial pressure containing protein 1 (= Hypoxia-inducible factor2a [HIF-2a]) NOS2A Nitric oxide synthase 2A Endothelial PAS domainO2 partial pressure Within Tibetans, EPAS1 alleles are correlated with erythrocyte count and hemoglobin concentration Bigham et al. (2010) Yi et al. (2010), Bigham et al. (2010), Peng et al. (2011) Bigham et al. (2010), Simonson et al. (2010), Peng et al. (2011) Ehinger et al. (2002), Fontanillas et al. (2005) phosphorylation NADH dehydrogenase Enyzme in oxidative aD, aA, bA subunits O2 transport O2 partial pressure aA, bA subunit O2 transport O2 partial pressure Elevated Fst only for bA aA, bA subunit O2 transport O2 partial pressure aA, bA subunit O2 transport O2 partial pressure Elevated Fst only for bA Wilson et al. (2013) McCracken et al. (2009) McCracken et al. (2009) Bulgarella et al. (2012) Cheviron & Brumeld (2009)

Anas canoptera

Cinnamon teal

Anas avirostris

Speckled teal

Anas georgica

Yellow-billed pintail

Lophonetta specularioides

Crested duck

Zonotrichia capensis

Rufous-collared sparrow

Crocidura russula

White-toothed shrew

Homo sapiens

Human (Tibetans &

Andeans)

Homo sapiens

Human (Tibetans)

Homo sapiens

Human (Andeans)

2013 THE AUTHORS. J. EVOL. BIOL. doi: 10.1111/jeb.12255 JOURNAL OF EVOLUTIONARY BIOLOGY 2013 EUROPEAN SOCIETY FOR EVOLUTIONARY BIOLOGY

Homo sapiens

Human (Tibetans)

Homo sapiens

Human (Andeans)

Peromyscus maniculatus

Deer mouse

Divergent altitudinal adaptation

Peromyscus maniculatus

Deer mouse

Table 2 (Continued)
Evidence of adaptive signicance Potential Gene selective agent Parasite community Fst cline Comments Reference Keller et al. (2011) Elevated Altitudinal Gene/Marker UBA untranslated tail of major histocompatibility complex (MHC) region IA a-Gpdh Energy metabolism in ight muscle altitude MDH cycle 4 candidate genes on chromosome 2 ebony Possibly climate biogenic amine synthesis pathway PGI isomerase Phosphoglucose Glycolytic enzyme Possibly temperature Several functions in cricklet, CG14591 neurogenesis temperature invected, mastermind, Metabolism, Unclear, possibly temperature Variation at these genes underlies altitudinal cline in larval developmental time Substitutions in the enhancer of the ebony locus are associated with abdominal melanisation Individuals from low sites with high-altitude-like PGI genotype resemble highaltitude individuals with respect to development rates and chill-coma recovery time Karl et al. (2008, 2009) Rebeiz et al. (2009) Mensch et al. (2010) Malate dehydrogenase Enzyme in citric acid Unclear, possibly Kraushaar et al. (2002) muscle decreases with Unclear Size of wing and ight dehydrogenase a-Glycerophosphate Drotz et al. (2001, 2012) Microsatellite within 3 Immune defense function

Common

Species

name

I. KELLER ET AL.

Salmo trutta spp.

European trout

Agabus bipustulatus

Water beetle

Sepsis cynipsea

Dungy

Drosophila melanogaster

Fruity

D. melanogaster

Fruity

Lycaena tityrus

Copper buttery

Anonymous loci

Inversions Chromosome 2 Chromosome 3 Chromosome 3 All chromosomes Chromosomes 2,3,X Unknown Unknown Unknown Unknown Unknown Possibly climate Possibly climate Possibly climate Possibly climate Possibly climate

Drosophila buzzatti

Fruity

Rodriguez et al. (2000) No neutral reference loci Dobzhansky (1948) Dobzhansky (1948), Schaeffer et al. (2003) No neutral reference loci Karyotypic differences underlie variation in several life history traits Burla et al. (1986) Stalker & Carson (1948), Levitan (1978), Etges (1989) Possibly climate Ayala et al. (2011)

D. persimilis

Fruity

D. pseudoobscura

Fruity

D. subobscura

Fruity

D. robusta

Fruity

Anopheles funestus

Mosquitoes

Chromosomes 2,3

Unknown

AFLP and microsatellite loci 8 AFLP loci (2.0%)* 12 AFLP loci (0.8%) 22 AFLP loci (8.8%) Unknown Unknown Unknown Unknown Unknown Unknown, possibly plague presence 5 AFLP loci (2.2%) 1 Microsatellite 1 Microsatellite Unknown Unknown Unknown Unknown Unknown Unknown Plague is present only at higher altitudes Keller et al. (2012) Barker et al. (2011) Collinge et al. (2006) Bonin et al. (2006) Fischer et al. (2011) Tollenaere et al. (2011)

Rana temporaria

Common frog

Microtus arvalis

Voles

Rattus rattus

Rat

Salmo trutta spp.

European trout

Drosophila buzzatii

Fruity

Drosophila melanogaster

Fruity

2013 THE AUTHORS. J. EVOL. BIOL. doi: 10.1111/jeb.12255 JOURNAL OF EVOLUTIONARY BIOLOGY 2013 EUROPEAN SOCIETY FOR EVOLUTIONARY BIOLOGY

*PGI, phosphoglucose isomerase. *For AFLP-based studies, the percentage of outliers among all polymorphic loci is given in parentheses.

Divergent altitudinal adaptation

Fig. 2 Number of signicant (dark grey) and nonsignicant (light grey) results as reported in the original publications for different geographical scales in animal species. The results from studies performed at a scale of less than 100 km are plotted again at a higher resolution in the small inset and show that signicant phenotypic differences between populations can be observed even at very local scales. The numbers above the bars indicate the number of independent studies contributing to each distance class. Signicant effects include main effects or interactions involving population or altitude.

thermoregulatory advantages arising from a smaller surface to volume ratio. Some ectotherms also comply with Bergmanns rule, although the underlying mechanisms are likely to be different (Gardner et al., 2011); furthermore, the opposite pattern is also common (Blanckenhorn & Demont, 2004). Our data set, which contained information for seven ectotherms from different taxonomic groups, showed that in three species, body mass tended to increase with altitude (Fig. 1c). Body length changed in the opposite direction, with a signicant decrease with altitude reported from two insects (Fig. S1d; Table 1). In temperate regions, the period available for growth rner, 2007). At the shortens with increasing altitude (Ko same time, the completion of different developmental stages may take longer, especially in ectotherms, because lower ambient temperatures slow down physiological processes. An expected adaptation to these conditions involves a compensatory response, with high-altitude individuals developing or growing faster than low-altitude individuals under a given thermal regime (i.e. countergradient variation; e.g. Hodkinson, 2005). Such a pattern was indeed observed in all three species for which development time was found to be

signicantly associated with altitude (Fig. 1d; Table 1), although the same number of between-population differences was found that were unassociated with altitude. Similarly, none of the observed between-population differences in growth rate were associated with altitude (Fig. S1e; Table 1). Viability, longevity and fecundity were additional life-history traits, which were repeatedly investigated, almost exclusively in ies. A possible expectation here could be that more variable and unpredictable highaltitude environments favour a faster pace of life, characterized by high investments in reproduction early in life (Tieleman, 2009) and, perhaps, reduced longevity (Table 1). Fecundity early in life has been estimated in only two species (indicated by asterisks in Fig. S1h). All three traits showed some signicant between-population differences, although demonstrations of altitudinal patterns were rare (Fig. S1fh; Table 1). Remarkably, all four studies using multiple rearing temperatures found that the decrease in longevity with altitude was strongest in the coldest environment, that is, the conditions most strongly resembling high-altitude conditions. Overall, our literature survey provided clear evidence for signicant, genetically based phenotypic differences

2013 THE AUTHORS. J. EVOL. BIOL. doi: 10.1111/jeb.12255 JOURNAL OF EVOLUTIONARY BIOLOGY 2013 EUROPEAN SOCIETY FOR EVOLUTIONARY BIOLOGY

10

I. KELLER ET AL.

between populations of different altitudinal origin. Comparisons across species identied several traits for which parallel clinal patterns were observed in two or three species, and for which the phenotypic changes consistently occurred in the predicted direction (melanization, wing size, cold and heat tolerance, mass, development time; A or B in Table 1). However, in all of these cases, several species also showed signicant between-population differences that were not, or at least not consistently, associated with altitude (C in Table 1). Overall, the available data are clearly limited. The number of species for which data on a given trait category were available was typically small ( 13) and skewed towards particular taxonomic groups; furthermore, for several species, only two source populations had been studied, making it impossible to distinguish population effects from altitudinal effects. Evidence of adaptive genetic divergence from molecular studies Almost all vertebrates rely on haemoglobin (Hb) for the transport of oxygen. Hb is therefore an obvious candidate gene for adaptation to the changing O2 partial pressure along altitudinal gradients, and the genes coding for different Hb subunits have been studied in a number of species, most prominently birds (Table 2). All of these studies found evidence of divergent selection for at least some of these genes. In deer mice, Storz et al. (2009) further demonstrated that the b-globin variant common in high-altitude populations has indeed a higher O2 afnity. Additional genes with a potential role for adaptation to low O2 partial pressure have been detected in humans through genome scans (Table 2). For example, one enzyme from the hypoxia-inducible factor pathway (EGLN1) has been identied as a potential target of selection in both Tibetans and Andeans, but the haplotypes common at high altitudes differ between the two regions. Other genes have been implicated in high-altitude adaptation in only one of the two populations (Table 2). Evidence of divergent selection along altitudinal gradients is also available for other candidate genes, including mitochondrial loci and several allozymes. For many of these loci, a role in adapting to thermal conditions is very plausible (Table 2), and in some cases, a direct link between genotype and phenotype of relevance for altitudinal adaptation has been demonstrated. For example, Fontanillas et al. (2005) found that nonshivering thermogenesis (i.e. mitochondrial heat production in brown fat cells) in white-toothed shrews was inuenced by an interaction between sex and mitochondrial haplotype. Copper butteries from low-altitude sites but with high-altitude-like genotypes at an allozyme locus (phosphoglucose isomerase) resembled high-altitude individuals with respect to

development rates and chill-coma recovery time (Karl et al., 2008). And nally, in D. melanogaster, four candidate genes on chromosome 2 underlie altitudinal clines in developmental time (Mensch et al., 2010), and mutations in the cis regulatory elements of the ebony locus underlie altitudinal pigmentation clines (Rebeiz et al., 2009). Chromosomal inversion polymorphisms have also been repeatedly found to show altitudinal clines in ies, where the presence of large polytene chromosomes has made chromosomal rearrangements more amenable to study (Table 2). Several lines of evidence suggest that climatic variables may play an important role in maintaining spatial patterns in the frequency of particular inversion genotypes. For instance, inversion polymorphisms in Drosophila subobscura show similar latitudinal clines on three continents (Balany a et al., 2006). The position of the clines has shifted in recent decades, probably due to rising global temperatures (Balany a et al., 2006), and similar temporal changes have been observed for a D. melanogaster inversion polymorphism in Australia (Umina et al., 2005). Some inversion polymorphisms also show consistent clines with altitude and latitude (Etges et al., 2006 and references therein) or recurrent seasonal uctuations (Dobzhansky, 1943). Finally, several studies have demonstrated a link between particular inversion polymorphisms and resistance to extreme temperatures (reviewed in Hoffmann et al., 2004). Six studies have also screened panels of anonymous markers (e.g. AFLPs, microsatellite loci; Table 2) and identied loci showing patterns consistent with divergent selection along altitudinal transects (i.e. elevated differentiation and/or genotype-altitude associations). In the four AFLP-based studies, between 0.8% and 8.8% of all polymorphic loci showed evidence of adaptive differentiation (Table 2). It should be noted, however, that these studies used varying approaches for identifying outliers. Furthermore, rather than being the actual targets of selection, the anonymous markers detected using these approaches are more likely to be in linkage disequilibrium with unknown divergently selected loci. Synthesis

Pervasive evidence for genetically based phenotypic differentiation


Our survey of the literature provides evidence for genetically based phenotypic divergence along altitudinal gradients for a wide range of species and traits, including wing size, cold tolerance, mass and development time (Fig. 1). This nding suggests that phenotypic divergence between populations is not rare, even if its prevalence may be overestimated in our data set due to a possible bias towards publishing signicant results. Our analyses also show that genetically based

2013 THE AUTHORS. J. EVOL. BIOL. doi: 10.1111/jeb.12255 JOURNAL OF EVOLUTIONARY BIOLOGY 2013 EUROPEAN SOCIETY FOR EVOLUTIONARY BIOLOGY

Divergent altitudinal adaptation

11

phenotypic differentiation is taxonomically widespread, with some signicant differences between populations being detected in all the groups studied. Furthermore, in some cases, signicant phenotypic divergence occurred at local geographical scales (Fig. 2). For example, eight studies detected signicant divergence between populations separated by ten kilometres or less (Fig. 2, inset), which in several cases was well within the dispersal range of the species. In Anolis lizards, the morphological divergence was larger than neutral genetic divergence (FST < QST; Eales et al., 2010) and, similarly, Sarup et al. (2009) detected significant phenotypic divergence between Drosophila buzzatii and D. simulans populations that were undifferentiated at neutral molecular markers.

Are the observed differences adaptively relevant?


The nding that phenotypic differentiation can be maintained at least sometimes in the face of gene ow strongly suggests that some between-population differences are maintained by strong divergent natural selection. And this conclusion is supported by additional lines of evidence from both the phenotypic and molecular data sets. First, several phenotypic traits showed consistent clines in multiple species in the direction predicted from known environmental gradients, and secondly, many of the molecular studies identify loci showing signatures of divergent selection (Table 2). We can also predict that the level of genetic differentiation at loci with a putative role in altitudinal adaptation should increase with the altitudinal distance between sites, assuming that the latter provides a rough proxy for the intensity of divergent selection. Neutral loci, on the other hand, should not show such an association unless gene ow between different altitudi1.0

nal environments is reduced across the entire genome, for example due to dispersal barriers or immigrant inviability becoming stronger as the altitudinal contrast increases. Consistent with these predictions, we found that FST increased with maximum altitudinal distance at candidate loci, but not at neutral loci (Fig. 3). However, these analyses were based on a subset of only 15 studies, and the data set did not lend itself to statistical analysis. First, altitudinal distance in these data was highly correlated with geographical distance (Pearson correlation coefcient 0.77), making it impossible to distinguish between the effects of the two variables. Secondly, the representation of different taxonomic groups was very uneven across altitudinal distance classes (e.g. all studies at > 4000 m are from birds). Despite these limitations, it will be interesting for future studies to investigate whether genetic differentiation increases with altitudinal distance, and whether this increase is genome-wide or limited to genomic regions with a direct role in altitudinal adaptation. Most of the phenotypic traits discussed above were specically selected by researchers because their adaptive relevance seemed likely and, for some of the traits, clear expectations as to how they should respond to the environmental change associated with altitude could be formulated. As discussed above, a number of traits showed patterns consistent with these predictions (Table 1). However, in all cases, some species did not conform to our expectations, for example showing no signicant between-population differences or differences that were unassociated with altitude. There are many possible explanations for these conicting results. First, a given phenotypic trait may simply not be relevant for tness in populations that are diverging due to random drift. Alternatively, if trait differences are adaptively relevant, the link between a phenotypic trait
Residuals Fst ~ Geo distance
0.4 0.4 1000 0.2 0.0 0.2

Fst

0.0 1000

0.2

0.4

0.6

0.8

2000

3000

4000

2000

3000

4000

Max. altitudinal distance (m)

Max. altitudinal distance (m)

Fig. 3 Left panel: Genetic differentiation between animal populations (FST) increases with the maximum altitudinal distance between sampling sites at candidate loci (black diamonds), but not at neutral loci (grey squares). Each point represents a study, and in the majority of cases (12 of 15 studies), the type of molecular marker was the same for candidate and neutral loci. Note that we did not assess the statistical signicance of the observed patterns because of limitations of the data set. First, different taxonomic groups were unevenly represented in the different altitudinal distance classes. Secondly, altitudinal distance was highly correlated with geographical distance making it impossible to disentangle the effects of the two variables: there was no longer an association between altitudinal distance and FST after removing the effect of geographical distance (right panel: residuals from a linear regression of FST against geographical distance, plotted against altitudinal distance).
2013 THE AUTHORS. J. EVOL. BIOL. doi: 10.1111/jeb.12255 JOURNAL OF EVOLUTIONARY BIOLOGY 2013 EUROPEAN SOCIETY FOR EVOLUTIONARY BIOLOGY

12

I. KELLER ET AL.

Box 1: The promise of ecological genomics for testing the genetic basis of altitudinal adaptation

At present, genome scans and outlier locus detection are a commonly used approach to detect signals consistent with the action of divergent selection between populations (Schoville et al., 2012). Perhaps the most serious limitation of this approach is that by denition it detects loci showing elevated between-population differences. Adaptive divergence, however, does not always involve large allele frequency changes, especially for quantitative traits which can be inuenced by many loci and where interactions between loci can be more important than additive effects (e.g. McKay & Latta, 2002; Le Corre & Kremer, 2012). A second hurdle in nonmodel organisms is that the actual target of selection will (mostly) not be the outlier locus itself but rather a locus linked to it. Without a well-annotated reference genome, it will be difcult to identify nearby candidate genes of known function. Still, the detection of outlier loci, even if they remain anonymous, may provide a relatively costeffective and tractable way to gauge the extent of putatively adaptive differentiation between populations that may be relevant for designing conservation and management strategies (but see Allendorf et al., 2010 for additional limitations and caveats). In recent years, genome-scale analyses have become increasingly possible also in nonmodel species. Using next-generation sequencing of reduced representation libraries (e.g. restriction site associated DNA; Baird et al., 2008), for example, tens of thousands of single nucleotide polymorphisms (SNPs) can be identied and genotyped at moderate cost and without the need for a reference genome (Stapley et al., 2010). These data offer promising new opportunities to investigate the genetic basis of particular phenotypic traits, for example, using association mapping in natural populations without known pedigrees (Slate et al., 2010; Stapley et al., 2010). Once adaptively relevant variation has been identied, the frequency of particular variants can be monitored in space and/or time and related to environmental changes. Ideally, such surveys should be replicated to distinguish between general and local effects and to follow the fate of genetic variants in different genomic backgrounds. Here, altitudinal gradients may be particularly valuable because similar gradients are replicated across the globe. Strong barriers to dispersal may exist also within a mountain range, subdividing species into units that follow largely independent evolutionary trajectories (e.g. aquatic organisms in different drainages; Keller et al., 2012). A second feature of altitudinal gradients, namely the small spatial scale at which environmental changes occur, makes them particularly suited to investigate whether specic genomic architectures are overrepresented in cases where adaptive divergence occurs in the face of gene ow. The chromosomal location of loci involved in divergent adaptation can most easily be studied if a reference genome from a closely related species is available, but genetic maps can also be constructed by following the segregation of variants in a pedigree (Slate et al., 2010; Stapley et al., 2010). Of particular interest might be a comparison of the genomic architectures underlying adaptation along altitudinal vs. latitudinal gradients. The two types of gradients share some similarities with respect to the observed environmental transitions (e.g. temperature), but these changes occur across much larger spatial scales with latitude and divergence may consequently be less constrained by gene ow.

and tness in a given altitudinal environment may have been misjudged; for example, the optimal phenotype may be different from what we expect. Also, selection pressures may not change consistently along an altitudinal gradient. For instance, small-scale topographical features inuence ambient temperatures and can produce local patterns that oppose large-scale gradirner, 2011). Finally, the trait mean ents (Scherrer & Ko in a population can deviate from the local optimum due to, for example, indirect selection resulting from genetic correlations with other traits, lack of additive genetic variation or immigration from other environments (Lenormand, 2002; Hoffmann & Willi, 2008). Open questions and directions for future research

Are phenotypic differences between populations adaptively relevant and how does mean population tness change along altitudinal gradients?
Although the available studies provide some evidence of adaptive differences between populations, explicit tests of local adaptation along altitudinal gradients, and the ecological relevance of the observed interpopulation differences, are clearly needed. Reciprocal transplant experiments along altitudinal gradients would be

particularly valuable in this context, although we are aware of only three animal species for which such studies have been performed. All of these found that some trait differences persisted also in common natural environments (body size of frogs: Berven, 1982a; age and size at rst reproduction in frogs: Berven, 1982b; ight activity of butteries: Ellers & Boggs, 2004b; size and growth rate in lizards: Iraeta et al., 2006), but none actually demonstrated that tness (or any tness proxy) was indeed higher for local than nonlocal individuals. More thorough studies of local adaptation will also provide insights into how the relative and absolute tness of populations change along altitudinal gradients, which is largely unknown. If most populations are indeed adapted to their local environment, we might expect little variation in tness along the gradient; however, most species have a restricted altitudinal distribution, suggesting that there must be limits to adaptation (e.g. Bridle & Vines, 2007). It is also possible that the environment imposes constraints on the maximum tness that cannot be overcome by adaptation. For instance, fundamental thermodynamic constraints may lead to lower population growth rates in cold-adapted than warm-adapted species, even when both are tested at their thermal optimum (Frazier et al., 2006).

2013 THE AUTHORS. J. EVOL. BIOL. doi: 10.1111/jeb.12255 JOURNAL OF EVOLUTIONARY BIOLOGY 2013 EUROPEAN SOCIETY FOR EVOLUTIONARY BIOLOGY

Divergent altitudinal adaptation

13

Unfortunately, in many cases, it will remain difcult to perform well-designed experiments that speak directly to the extent and patterns of local adaptation along altitudinal gradients, as well as to the link between particular phenotypes and tness in a given environment. These are not easy questions to address, even in species that are experimentally tractable, and their study becomes particularly problematic in animals that cannot be reliably followed through time. However, recent methodological advances have opened up exciting new opportunities to investigate potential adaptation in a more diverse array of species using molecular approaches (see Box 1).

low-altitude conditions under global warming, we might predict that contemporary gene ow is usually asymmetrical, occurring mainly from low into high-altitude populations. The symmetry of gene ow has been assessed along latitudinal gradients (Paul et al., 2011; Fedorka et al., 2012), but we are unaware of similar studies along altitude. A markrecapture study in a buttery, however, found that dispersal was indeed more common from low- to high-altitude populations, probably because, at higher altitudes, host plants became available later in the season (Peterson, 1997).

What is the evolutionary potential of populations along altitudinal gradients?


The available evidence suggests that there is some adaptively relevant genetic divergence between populations and implies that adaptation has occurred in the past. Whether adaptive change will be possible in the future will depend upon the availability of relevant genetic diversity within populations and the rate at which environmental conditions change (Bridle et al., 2008). Evidence from the molecular studies surveyed here suggests that populations may in fact often be genetically variable at loci with a potential role in adaptation to different environments; thus, the loci identied as outliers (Table 2) whereas showing large allele frequency differences between populations are rarely xed for alternative alleles. Often the alleles thought to be advantageous at low altitudes are observed at lower frequencies also in high-altitude populations and vice versa (data not shown). Similarly, for the phenotypic data, an average coefcient of variation (CV) of 10.7 was estimated across 651 observations for which this calculation was possible, suggesting some variation between individuals of a population reared in the same environment. Novel genetic variation can be introduced into a population not only through mutation but, perhaps more relevant for rapid adaptation (e.g. Abbott et al., 2013), also through gene ow. Altitudinal gradients tend to be steep relative to the dispersal distance of organisms, which means that immigrants will often come from different, but nearby environments. In such situations, the selection coefcients of variants in the different environments and the rate and symmetry of gene ow will determine whether between-population differences are maintained or lost (Lenormand, 2002). Although gene ow can hinder local adaptation by eroding allele frequency differences, it can sometimes also facilitate it by introducing novel and potentially benecial variants (Garant et al., 2007). Gene ow from lower towards higher altitudes, for example, could introduce genetic variants that have been pretested under warmer conditions. Consequently, if conditions at high-altitude sites indeed tend to become more similar to current

Does the extent of adaptive differentiation vary between species and, if so, what factors underlie these differences?
The available data show that genetic differences of potential adaptive relevance exist in a wide range of species, including highly mobile groups such as birds where population divergence is potentially maintained in the face of extensive gene ow. Still, it is important to keep in mind that the species covered in this review are probably a nonrepresentative sample and that the extent of intraspecic adaptive diversity may be low in many other species. This may be particularly true for species with narrow altitudinal distributions where opportunities for adaptive divergence might be more limited. To understand why local adaptations evolve in some cases, but not in others, it will be critical to study a diverse array of species with both narrow and broad altitudinal distributions. Of particular interest will be whether some species are somehow predisposed to evolving and maintaining adaptive differences between populations. Such a predisposition could involve a genomic architecture where adaptive traits are shaped by few loci of large effect and/or clusters of multiple loci in tight physical linkage, which is expected to facilitate adaptation in the face of gene ow (e.g. along altitudinal gradients; Yeaman & Whitlock, 2011).

Conclusions
Our literature survey on local adaptation to altitude detected extensive phenotypic and genetic diversity among animal populations sampled along altitudinal gradients, with several lines of evidence suggesting that these differences were, in part, adaptively relevant. Although these conclusions are based upon rather limited data, we remain convinced that altitudinal gradients provide very suitable model systems for investigating local adaptation, albeit systems that have not yet been used to their full advantage. Furthermore, we anticipate that methodological advances will enable future studies to address these phenomena in species that are not easily tractable experimentally. In the meantime, however, it seems prudent to assume that most populations show some adaptive differentiation along altitudinal gradients, sometimes at very local scales, and that these

2013 THE AUTHORS. J. EVOL. BIOL. doi: 10.1111/jeb.12255 JOURNAL OF EVOLUTIONARY BIOLOGY 2013 EUROPEAN SOCIETY FOR EVOLUTIONARY BIOLOGY

14

I. KELLER ET AL.

adaptively relevant differences should be considered in conservation and management efforts.

Acknowledgments
This study was carried out in the framework of GeneReach, a project lead by J. Bolliger (WSL) and funded by the Competence Center Environment and Sustain rich, Switzerland. IK would like to thank ability, ETH Zu O. Seehausen, M. Haesler, K. Lucek, D. Marques and J. Meier for support and helpful discussion.

References
Abbott, R., Albach, D., Ansell, S., Arntzen, J.W., Baird, S.J.E., Bierne, N. et al. 2013. Hybridization and speciation. J. Evol. Biol. 26: 229246. Allendorf, F.W., Hohenlohe, P.A. & Luikart, G. 2010. Genomics and the future of conservation genetics. Nat. Rev. Genet. 11: 697709. Ayala, D., Fontaine, M.C., Cohuet, A., Fontenille, D., Vitalis, R. & Simard, F. 2011. Chromosomal inversions, natural selection and adaptation in the malaria vector Anopheles funestus. Mol. Biol. Evol. 28: 745758. Baird, N.A., Etter, P.D., Atwood, T.S., Currey, M.C., Shiver, A.L., Lewis, Z.A. et al. 2008. Rapid SNP discovery and genetic mapping using sequenced RAD markers. PLoS ONE 3: e3376. Balany a, J., Oller, J.M., Huey, R.B., Gilchrist, G.W. & Serra, L. 2006. Global genetic change tracks global warming in Drosophila subobscura. Science 313: 17731775. Barker, J.S.F., Frydenberg, J., Sarup, P. & Loeschcke, V. 2011. Altitudinal and seasonal variation in microsatellite allele frequencies of Drosophila buzzatii. J. Evol. Biol. 24: 430439. Bears, H., Drever, M.C. & Martin, K. 2008. Comparative morphology of dark-eyed juncos Junco hyemalis breeding at two elevations: a common aviary experiment. J. Avian Biol. 39: 152162. Beattie, R.C. 1987. The reproductive biology of common frog (Rana temporaria) populations from different altitudes in northern England. J. Zool. 211: 387398. Belen, A. & Alten, B. 2006. Variation in life table characteristics among populations of Phlebotomus papatasi at different altitudes. J. Vector Ecol. 31: 3544. Belen, A., Alten, B. & Aytekin, A.M. 2004. Altitudinal variation in morphometric and molecular characteristics of Phlebotomus papatasi populations. Med. Vet. Entomol. 18: 343 350. Berner, D., Ko rner, C. & Blanckenhorn, W.U. 2004. Grasshopper populations across 2000m of altitude: is there life history adaptation. Ecography 27: 733740. Berven, K.A. 1982a. The genetic basis of altitudinal variation in the wood frog Rana sylvatica. I. An experimental analysis of life history traits. Evolution 36: 360369. Berven, K.A. 1982b. The genetic basis of altitudinal variation in the wood frog Rana sylvatica. II. An experimental analysis of larval development. Oecologia 52: 360369. Bigham, A., Bauchet, M., Pinto, D., Mao, X., Akey, J.M., Mei, R. et al. 2010. Identifying signatures of natural selection in

Tibetan and Andean populations using dense genome scan data. PLoS Genet. 6: e1001116. Blanckenhorn, W.U. 1997. Altitudinal life history variation in the dung ies Scathophaga stercoraria and Sepsis cynipsea. Oecologia 109: 342352. Blanckenhorn, W.U. & Demont, M. 2004. Bergmann and converse Bergmann latitudinal clines in arthropods: two ends of a continuum? Integr. Comp. Biol. 44: 413424. Bonin, A.A., Taberlet, P., Miaud, C. & Pompanon, F.F. 2006. Explorative genome scan to detect candidate loci for adaptation along a gradient of altitude in the common frog (Rana temporaria). Mol. Biol. Evol. 23: 773783. Bradshaw, W.E., Holzapfel, C.M. & Crowder, R. 2006. Evolutionary response to rapid climate change. Science 312: 1477 1478. Bridle, J.R., Gavaz, S. & Kennington, W.J. 2009. Testing limits to adaptation along altitudinal gradients in rainforest Drosophila. Proc. R. Soc. B Biol. Sci. 276: 15071515. Bridle, J.R. & Vines, T.H. 2007. Limits to evolution at range margins: when and why does adaptation fail? Trends Ecol. Evol. 22: 140147. Bridle, J.R., Polechov a, J. & Vines, T.H. 2008. Patterns of biodiversity and limits to adaptation in time and space. In: Speciation and Patterns of Biodiversity (R.K. Butlin & D. Schluter, eds), pp. 77101. Cambridge University Press, Cambridge. Bubliy, O.A. & Loeschcke, V. 2004. Variation of life-history and morphometrical traits in Drosophila buzzatii and Drosophila simulans collected along an altitudinal gradient from a Canary island. Biol. J. Linn. Soc. 84: 119136. Buckley, C.R., Irschick, D.J. & Adolph, S.C. 2010. The contributions of evolutionary divergence and phenotypic plasticity to geographic variation in the western fence lizard, Sceloporus occidentalis. Biol. J. Linn. Soc. 99: 8498. Bulgarella, M., Peters, J.L., Kopuchian, C., Valqui, T., Wilson, R.E. & McCracken, K.G. 2012. Multilocus coalescent analysis of haemoglobin differentiation between low- and high-altitude populations of crested ducks (Lophonetta specularioides). Mol. Ecol. 21: 350368. Burla, H., Jungen, H. & Ba chli, G. 1986. Population structure of Drosophila subobscura: Non-random microdispersion of inversion polymorphism on a mountain slope. Genetica 70: 915. Cheviron, Z.A. & Brumeld, R.T. 2009. Migration-selection balance and local adaptation of mitochondrial haplotypes in rufous-collared sparrows (Zonotrichia capensis) along an elevational gradient. Evolution 63: 15931605. Chown, S.L. & Klok, C.J. 2003. Altitudinal body size clines: latitudinal effects associated with changing seasonality. Ecography 4: 445455. Clusella Trullas, S., Van Wyk, J.H. & Spotila, J.R. 2007. Thermal melanism in ectotherms. J. Therm. Biol 32: 235245. Collinge, J.E., Hoffmann, A.A. & McKechnie, S.W. 2006. Altitudinal patterns for latitudinally varying traits and polymorphic markers in Drosophila melanogaster from eastern Australia. J. Evol. Biol. 19: 473482. Conover, D.O., Duffy, T.A. & Hice, L.A. 2009. The covariance between genetic and environmental inuences across ecological gradients: reassessing the evolutionary signicance of countergradient and cogradient variation. Ann. N. Y. Acad. Sci. 1168: 100129. Dahlgaard, J., Hasson, E. & Loeschcke, V. 2001. Behavioral differentiation in oviposition activity in Drosophila buzzatii

2013 THE AUTHORS. J. EVOL. BIOL. doi: 10.1111/jeb.12255 JOURNAL OF EVOLUTIONARY BIOLOGY 2013 EUROPEAN SOCIETY FOR EVOLUTIONARY BIOLOGY

Divergent altitudinal adaptation

15

from highland and lowland populations in Argentina: plasticity or thermal adaptation? Evolution 55: 738747. Dahlhoff, E.P. & Rank, N.E. 2000. Functional and physiological consequences of genetic variation at phosphoglucose isomerase: heat shock protein expression is related to enzyme genotype in a montane beetle. Proc. Natl. Acad. Sci. USA 97: 1005610061. Dillon, M.E., Frazier, M.R. & Dudley, R. 2006. Into thin air: physiology and evolution of alpine insects. Integr. Comp. Biol. 46: 4961. Dingle, H., Mousseau, T.A. & Scott, S.M. 1990. Altitudinal variation in life cycle syndromes of California populations of the grasshopper, Melanoplus sanguinipes (F.). Oecologia 84: 199206. Dobzhansky, T. 1943. Genetics of natural populations. IX. Temporal changes in the composition of populations of Drosophila pseudoobscura. Genetics 28: 162186. Dobzhansky, T. 1948. Genetics of natural populations. XVI. Altitudinal and seasonal changes produced by natural selection in certain populations of Drosophila pseudoobscura and Drosophila persimilis. Genetics 33: 158176. Drotz, M.K., Brodin, T., Saura, A. & Giles, B.E. 2012. Ecotype differentiation in the face of gene ow within the diving beetle Agabus bipustulatus (Linnaeus, 1767) in northern Scandinavia. PLoS ONE 7: e31381. Drotz, M.K., Saura, A. & Nilsson, A.N. 2001. The species delimitation problem applied to the Agabus bipustulatus complex (Coleoptera, Dytiscidae) in north Scandinavia. Biol. J. Linn. Soc. 73: 1122. Eales, J., Thorpe, R.S. & Malhotra, A. 2010. Colonization history and genetic diversity: adaptive potential in early stage invasions. Mol. Ecol. 19: 28582869. Ehinger, M., Fontanillas, P., Petit, E. & Perrin, N. 2002. Mitochondrial DNA variation along an altitudinal gradient in the greater white-toothed shrew, Crocidura russula. Mol. Ecol. 11: 939945. Ellers, J. & Boggs, C.L. 2002. The evolution of wing color in Colias butteries: heritability, sex linkage, and population divergence. Evolution 56: 836840. Ellers, J. & Boggs, C.L. 2004a. Evolutionary genetics of dorsal wing colour in Colias butteries. J. Evol. Biol. 17: 752758. Ellers, J. & Boggs, C.L. 2004b. Functional ecological implications of intraspecic differences in wing melanization in Colias butteries. Biol. J. Linn. Soc. 82: 7987. Etges, W.J. 1989. Chromosomal inuences on life-history variation along an altitudinal transect in Drosophila robusta. Am. Nat. 133: 83110. Etges, W.J., Arbuckle, K.L. & Levitan, M. 2006. Long-term frequency shifts in the chromosomal polymorphisms of Drosophila robusta in the Great Smoky Mountains. Biol. J. Linn. Soc. 88: 131141. Fairbairn, D.J. & King, E. 2009. Why do Californian striders y? J. Evol. Biol. 22: 3649. Fedorka, K.M., Winterhalter, W.E., Shaw, K.L., Brogan, W.R. & Mousseau, T.A. 2012. The role of gene ow asymmetry along an environmental gradient in constraining local adaptation and range expansion. J. Evol. Biol. 25: 16761685. Ficetola, G.F. & Bernardi, F. 2005. Supplementation or in situ conservation? Evidence of local adaptation in the Italian agile frog Rana latastei and consequences for the management of populations. Anim. Conserv. 8: 3340.

Fischer, M.C., Foll, M., Excofer, L. & Heckel, G. 2011. Enhanced AFLP genome scans detect local adaptation in high-altitude populations of a small rodent (Microtus arvalis). Mol. Ecol. 20: 14501462. Folguera, G., Ceballos, S., Spezzi, L., Fanara, J.J. & Hasson, E. 2008. Clinal variation in developmental time and viability, and the response to thermal treatments in two species of Drosophila. Biol. J. Linn. Soc. 95: 233245. Fontanillas, P., D epraz, A., Giorgi, M.S. & Perrin, N. 2005. Nonshivering thermogenesis capacity associated to mitochondrial DNA haplotypes and gender in the greater whitetoothed shrew, Crocidura russula. Mol. Ecol. 14: 661670. Frazier, M.R., Huey, R.B. & Berrigan, D. 2006. Thermodynamics constrains the evolution of insect population growth rates: warmer is better. Am. Nat. 168: 512520. Garant, D., Forde, S.E. & Hendry, A.P. 2007. The multifarious effects of dispersal and gene ow on contemporary adaptation. Funct. Ecol. 21: 434443. Gardner, J.L., Peters, A., Kearney, M.R., Joseph, L. & Heinsohn, R. 2011. Declining body size: a third universal response to warming? Trends Ecol. Evol. 26: 285291. Gienapp, P., Teplitsky, C., Alho, J.S., Mills, J.A. & Meril a, J. 2008. Climate change and evolution: disentangling environmental and genetic responses. Mol. Ecol. 17: 167178. Hereford, J. 2009. A quantitative survey of local adaptation and tness trade-offs. Am. Nat. 173: 579588. Hodkinson, I.D. 2005. Terrestrial insects along elevation gradients: species and community responses to altitude. Biol. Rev. 80: 489513.  , C.M. 2011. Climate change and evoluHoffmann, A.A. & Sgro tionary adaptation. Nature 470: 479485. Hoffmann, A.A. & Willi, Y. 2008. Detecting genetic responses to environmental change. Nat. Rev. Genet. 9: 421432. , C.M. & Weeks, A.R. 2004. ChromoHoffmann, A.A., Sgro somal inversion polymorphisms and adaptation. Trends Ecol. Evol. 19: 482488. Iraeta, P., Monasterio, C., Salvador, A. & D az, J.A. 2006. Mediterranean hatchling lizards grow faster at higher altitude: a reciprocal transplant experiment. Funct. Ecol. 20: 865872. Jasienski, M. 2009. Cogradient plasticity of growth in montane and lowland larvae of Rana temporaria (L.) at two levels of temperature. Pol. J. Ecol. 57: 353361. Karan, D., Dubey, S., Moreteau, B., Parkash, R. & David, J.R. 2000. Geographical clines for quantitative traits in natural populations of a tropical drosophilid: Zaprionus indianus. Genetica 108: 91100. Karl, I., Schmitt, T. & Fischer, K. 2008. Phosphoglucose isomerase genotype affects life-history traits and cold stress resistance in a Copper buttery. Funct. Ecol. 22: 887894. Karl, I., Schmitt, T. & Fischer, K. 2009. Genetic differentiation between alpine and lowland populations of a buttery is related to PGI enzyme genotype. Ecography 32: 488496. Kawecki, T.J. & Ebert, D. 2004. Conceptual issues in local adaptation. Ecol. Lett. 7: 12251241. Keller, I., Schuler, J., Bezault, E. & Seehausen, O. 2012. Parallel divergent adaptation along replicated altitudinal gradients in Alpine trout. BMC Evol. Biol. 12: 210. Keller, I., Taverna, A. & Seehausen, O. 2011. Evidence of neutral and adaptive genetic divergence between European trout populations sampled along altitudinal gradients. Mol. Ecol. 20: 18881904.

2013 THE AUTHORS. J. EVOL. BIOL. doi: 10.1111/jeb.12255 JOURNAL OF EVOLUTIONARY BIOLOGY 2013 EUROPEAN SOCIETY FOR EVOLUTIONARY BIOLOGY

16

I. KELLER ET AL.

rner, C. 2003. Alpine Plant Life: Functional Plant Ecology of Ko High Mountain Ecosystems. Springer, Berlin. rner, C. 2007. The use of altitude in ecological research. Ko Trends Ecol. Evol. 22: 569574. Kraushaar, U., Goudet, J. & Blanckenhorn, W.U. 2002. Geographical and altitudinal population genetic structure of two dung y species with contrasting mobility and temperature preference. Heredity 89: 99106. Le Corre, V. & Kremer, A. 2012. The genetic differentiation at quantitative trait loci under local adaptation. Mol. Ecol. 21: 15481566. Leinonen, T., OHara, R.B., Cano, J.M. & Meril a, J. 2008. Comparative studies of quantitative trait and neutral marker divergence: a meta-analysis. J. Evol. Biol. 21: 117. Lenormand, T. 2002. Gene ow and the limits to natural selection. Trends Ecol. Evol. 17: 183189. n-Velarde, F., De Muizon, C., Palacios, J.A., Clark, D. & Leo Monge, C. 1996. Hemoglobin afnity and structure in high-altitude and sea-level carnivores from Peru. Comp. Biochem. Physiol. A Physiol. 113: 407411. Levitan, M. 1978. Studies of linkage in populations. IX. The effect of altitude on X-chromosomal arrangement combinations in Drosophila robusta. Genetics 89: 751763. Marquis, O. & Miaud, C. 2008. Variation in UV sensitivity among common frog Rana temporaria populations along an altitudinal gradient. Zoology 111: 309317. McCracken, K.G., Barger, C.P., Bulgarella, M., Johnson, K.P., Kuhner, M.K., Moore, A.V. et al. 2009. Signatures of highaltitude adaptation in the major hemoglobin of ve species of Andean dabbling ducks. Am. Nat. 174: 631650. McKay, J.K. & Latta, R.G. 2002. Adaptive population divergence: markers, QTL and traits. Trends Ecol. Evol. 17: 285291. Mensch, J., Carreira, V., Lavagnino, N., Goenaga, J., Folguera, G., Hasson, E. et al. 2010. Stage-specic effects of candidate heterochronic genes on variation in developmental time along an altitudinal cline of Drosophila melanogaster. PLoS ONE 5: e11229. Morjan, C.L. & Rieseberg, L.H. 2004. How species evolve collectively: implications of gene ow and selection for the spread of advantageous alleles. Mol. Ecol. 13: 13411356. Morris, J.P., Thatje, S. & Hauton, C. 2013. The use of stress-70 proteins in physiology: a re-appraisal. Mol. Ecol. 22: 14941502. Norry, F.M., Bubliy, O.A. & Loeschcke, V. 2001. Developmental time, body size and wing loading in Drosophila buzzatii from lowland and highland populations in Argentina. Hereditas 135: 3540. Nosil, P., Funk, D.J. & Ortiz-Barrientos, D. 2009. Divergent selection and heterogeneous genomic divergence. Mol. Ecol. 18: 375402. Parkash, R., Rajpurohit, S. & Ramniwas, S. 2008. Changes in body melanisation and desiccation resistance in highland vs. lowland populations of D. melanogaster. J. Insect Physiol. 54: 10501056. Parkash, R., Sharma, V. & Kalra, B. 2010. Correlated changes in thermotolerance traits and body color phenotypes in montane populations of Drosophila melanogaster: analysis of within- and between-population variations. J. Zool. 280: 4959. Parmesan, C. 2006. Ecological and evolutionary responses to recent climate change. Annu. Rev. Ecol. Evol. Syst. 37: 637669. Parmesan, C. & Yohe, G. 2003. A globally coherent ngerprint of climate change impacts across natural systems. Nature 421: 3742.

Paul, J.R., Sheth, S.N., Angert, A.L. & Geber, S.E.M. 2011. Quantifying the impact of gene ow on phenotype-environment mismatch: a demonstration with the scarlet monkeyower Mimulus cardinalis. Am. Nat. 178: S62S79. Peng, Y., Yang, Z., Zhang, H., Cui, C., Qi, X., Luo, X. et al. 2011. Genetic variations in Tibetan populations and highaltitude adaptation at the Himalayas. Mol. Biol. Evol. 28: 10751081. Peterson, M.A. 1997. Host plant phenology and buttery dispersal: causes and consequences of uphill movement. Ecology 78: 167180. Pitchers, W., Pool, J.E. & Dworkin, I. 2012. Altitudinal clinal variation in wing size and shape in African Drosophila melanogaster: One cline or many? Evolution 67: 438452. Pool, J.E. & Aquadro, C.F. 2007. The genetic basis of adaptive pigmentation variation in Drosophila melanogaster. Mol. Ecol. 16: 28442851. Rebeiz, M., Pool, J.E., Kassner, V.A., Aquadro, C.F. & Carroll, S.B. 2009. Stepwise modication of a modular enhancer underlies adaptation in a Drosophila population. Science 326: 16631667. Reusch, T.B.H. & Wood, T.E. 2007. Molecular ecology of global change. Mol. Ecol. 16: 39733992. Rodriguez, C., Piccinali, R., Levy, E. & Hasson, E. 2000. Contrasting population genetic structures using allozymes and the inversion polymorphism in Drosophila buzzatii. J. Evol. Biol. 13: 976984. Sambucetti, P., Loeschcke, V. & Norry, F.M. 2006. Developmental time and size-related traits in Drosophila buzzatii along an altitudinal gradient from Argentina. Hereditas 143: 7783. Sarup, P., Frydenberg, J. & Loeschcke, V. 2009. Local adaptation of stress related traits in Drosophila buzzatii and Drosophila simulans in spite of high gene ow. J. Evol. Biol. 22: 11111122. Schaeffer, S.W., Goetting-Minesky, M.P., Kovacevic, M., Peoples, J.R., Graybill, J.L., Miller, J.M. et al. 2003. Evolutionary genomics of inversions in Drosophila pseudoobscura: Evidence for epistasis. Proc. Natl. Acad. Sci. USA 100: 83198324. rner, C. 2011. Topographically controlled Scherrer, D. & Ko thermal-habitat differentiation buffers alpine plant diversity against climate warming. J. Biogeogr. 38: 406416. Schoville, S.D., Bonin, A., Franc ois, O., Lobreaux, S., Melodelima, C. & Manel, S. 2012. Adaptive genetic variation on the landscape: methods and cases. Annu. Rev. Ecol. Evol. Syst. 43: 2343. Simonson, T.S., Yang, Y., Huff, C.D., Yun, H., Qin, G., Witherspoon, D.J. et al. 2010. Genetic evidence for high-altitude adaptation in Tibet. Science 329: 7275. Slate, J., Santure, A.W., Feulner, P.G.D., Brown, E.A., Ball, A.D., Johnston, S.E. et al. 2010. Genome mapping in intensively studied wild vertebrate populations. Trends Genet. 26: 275284. Sommer, S. & Pearman, P.B. 2003. Quantitative genetic analysis of larval life history traits in two alpine populations of Rana temporaria. Genetica 118: 110. Srensen, J.G., Norry, F.M., Scannapieco, A.C. & Loeschcke, V. 2005. Altitudinal variation for stress resistance traits and thermal adaptation in adult Drosophila buzzatii from the New World. J. Evol. Biol. 18: 829837. Stalker, H.D. & Carson, H.L. 1948. An altitudinal transect of Drosophila robusta Sturtevant. Evolution 2: 295305. Stapley, J., Reger, J., Feulner, P.G.D., Smadja, C., Galindo, J., Ekblom, R. et al. 2010. Adaptation genomics: the next generation. Trends Ecol. Evol. 25: 705712.

2013 THE AUTHORS. J. EVOL. BIOL. doi: 10.1111/jeb.12255 JOURNAL OF EVOLUTIONARY BIOLOGY 2013 EUROPEAN SOCIETY FOR EVOLUTIONARY BIOLOGY

Divergent altitudinal adaptation

17

Stillwell, R.C. & Fox, C.W. 2009. Geographic variation in body size, sexual size dimorphism and tness components of a seed beetle: local adaptation versus phenotypic plasticity. Oikos 118: 703712. Storz, J.F. 2002. Contrasting patterns of divergence in quantitative traits and neutral DNA markers: analysis of clinal variation. Mol. Ecol. 11: 25372551. Storz, J.F. 2005. Using genome scans of DNA polymorphism to infer adaptive population divergence. Mol. Ecol. 14: 671688. Storz, J.F. & Dubach, J.M. 2004. Natural selection drives altitudinal divergence at the albumin locus in deer mice, Peromyscus maniculatus. Evolution 58: 13421352. Storz, J.F., Runck, A.M., Sabatino, S.J., Kelly, J.K., Ferrand, N., Moriyama, H. et al. 2009. Evolutionary and functional insights into the mechanism underlying high-altitude adaptation of deer mouse hemoglobin. Proc. Natl. Acad. Sci. USA 106: 1445014455. Tanaka, S. & Brookes, V.J. 1983. Altitudinal adaptation of the life cycle in Allonemobius fasciatus DeGeer (Orthoptera: Gryllidae). Can. J. Zool. 61: 19861990. Tantowijoyo, W. & Hoffmann, A.A. 2011. Variation in morphological characters of two invasive leafminers, Liriomyza huidobrensis and L. sativae, across a tropical elevation gradient. J. Insect Sci. 11: 116. Tieleman, B.I. 2009. High and low, fast or slow: the complementary contributions of altitude and latitude to understand life-history variation. J. Anim. Ecol. 78: 293295. Tollenaere, C., Duplantier, J.-M., Rahalison, L., Ranjalahy, M. & Brouat, C. 2011. AFLP genome scan in the black rat (Rattus rattus) from Madagascar: detecting genetic markers undergoing plague-mediated selection. Mol. Ecol. 20: 1026 1038. Tsuchiya, Y., Takami, Y., Okuzaki, Y. & Sota, T. 2012. Genetic differences and phenotypic plasticity in body size between

high- and low-altitude populations of the ground beetle Carabus tosanus. J. Evol. Biol. 25: 18351842. Umina, P.A., Weeks, A.R., Kearney, M.R., McKechnie, S.W. & Hoffmann, A.A. 2005. A rapid shift in a classic clinal pattern in Drosophila reecting climate change. Science 308: 691693. Wilson, R.E., Peters, J.L. & McCracken, K.G. 2013. Genetic and phenotypic divergence between low- and high-altitude populations of two recently diverged cinnamon teal subspecies. Evolution 67: 170184. Wittkopp, P.J., Smith-Winberry, G., Arnold, L.L., Thompson, E.M., Cooley, A.M., Yuan, D.C. et al. 2011. Local adaptation for body color in Drosophila americana. Heredity 106: 592602. Yeaman, S. & Whitlock, M.C. 2011. The genetic architecture of adaptation under migration-selection balance. Evolution 65: 18971911. Yi, X., Liang, Y., Huerta-Sanchez, E., Jin, X., Cuo, Z.X.P., Pool, J.E. et al. 2010. Sequencing of 50 human exomes reveals adaptation to high altitude. Science 329: 7578.

Supporting information
Additional Supporting Information may be found in the online version of this article: Figure S1 Genetically based phenotypic changes along altitudinal gradients for (a) heat tolerance, (b) heatshock protein expression, (c) desiccation tolerance, (d) body length, (e) growth rate, (f) longevity, (g) viability, and (h) fecundity. Table S1 List of publications included in the review.
Received 3 May 2013; revised 26 August 2013; accepted 27 August 2013

2013 THE AUTHORS. J. EVOL. BIOL. doi: 10.1111/jeb.12255 JOURNAL OF EVOLUTIONARY BIOLOGY 2013 EUROPEAN SOCIETY FOR EVOLUTIONARY BIOLOGY

Vous aimerez peut-être aussi