Sie sind auf Seite 1von 114

Tierrztliche Hochschule Hannover

Institut fr Zoologie

Geographical variation in the granular
poison frog, Oophaga granulifera:
genetics, colouration, advertisement call
and morphology

DISSERTATION
zur Erlangung des Grades
eines Doktors der Naturwissenschaften
Doctor rerum naturalium
(Dr. rer. nat.)

vorgelegt von
Oscar Brusa
Cuggiono, Italien.

Hannover 2012





Wissenschaftliche Betreuung: Prof. Dr. Heike Prhl








1. Gutachterin: Prof. Dr. Heike Prhl
2. Gutachter: Dr. Sebastian Steinfartz

Tag der mndlichen Prfung: 21 September 2012







Eidesstattliche Erklrung

Hiermit versichere ich, dass ich die vorliegende Arbeit selbstndig und nur mit den
angegebenen Hilfsmitteln angefertigt habe. Alle Stellen, die dem Wortlaut oder Sinn
nach anderen Werken entnommen sind, wurden durch Angaben der Quellen als
Entlehnungen kenntlich gemacht.
Ich versichere, dass die vorliegende Dissertation nicht woanders eingereicht wurde.

Hannover, den 14/7/2012
Oscar Brusa














To:
All people who dream.


















Table of contents

Zusammenfassung 9
Summary 11

1. Introduction 13
1.1 Biological species concept and speciation mechanisms 13
1.2 Geographic variation in animals 14
1.3 Aposematism 18
1.4 The remarkable case of polymorphic toxic species 19
1.5 The biogeographical peculiarities of the lower Central American 21
1.6 Evolutionary studies on amphibians in lower Central American 23
1.7 Neotropical poison-dart frogs 25
1.8 The granular poison-dart frog, Oophaga granulifera 28
1.9 Aims of the study 29
1.10 Study Area 31

2. Material and Methods 33
2.1 Field activity 33
2.2 Tissue sampling and DNA extraction 33
2.3 Phylogeography 34
2.3.1 Mitochondrial DNA sequencing 34
2.3.2 Mitochondrial DNA analyses 35


2.4 Population genetics 37
2.4.1 Microsatellite loci genotyping and analyses 37
2.5 Colour description and measurements 38
2.6 Bioacoustics 38
2.7 Morphology 39

3. Results 40
3.1 Colouration 40
3.2 Mitochondrial DNA variation and divergence 46
3.3 Mitochondrial DNA phylogeography and population structure 48
3.4. Microsatellite loci variation and population structure 51
3.5 Bioacoustics 53
3.6 Morphology 57

4. Discussion 65
4.1 General discussion 65
4.2 Why should a cryptic morph evolve in an aposematic species? 70
4.2.1 General background 70
4.2.2 Role of genetic drift 71
4.2.3 Role of predation 71
4.2.4 Role of toxicity 73
4.2.5 Role of sexual selection 75

5. Conclusion 76
6. References 77


Additional Information 110
1. Sequences accession numbers 110
2. Population structure K results 113

Ackowledgments 114



































9

Geographische Variation in dem granulierten Giftfrosch,
Oophaga granulifera: Genetik, Frbung, Paarungsrufe und
Morphologie.

Oscar Brusa

Zusammenfassung
Die vorliegende Arbeit untersucht die intraspezifische genetische Divergenz und phnotypische
Diversifizierung des neotropischen Giftfrosches, Oophaga granulifera, innerhalb seines
Verbreitungsgebietes. Dabei werden phylogeographische Muster, Populationsstrukturen und
phnotypische Merkmale, die potentiell an reproduktiver Isolation beteiligt sind, analysiert. Die
Analysen zur Phylogeographie and der Populationsstruktur wurden mit Hilfe der
mitochondrialen Gene von Cytochrom b und 16S sowie sieben polymorphen Mikrosatelliten-
Loci von neun bis elf Populationen durchgefhrt. Fr die Analyse der Phnotypen erfasste ich
den Farbphnotyp in 189 Frschen und ma die Reflektionsspektren in fnf ausgewhlten
Populationen. Zudem wurden in zehn Populationen mnnliche Paarungsrufe aufgenommen und
zehn morphologische Parameter gemessen. Die phylogenetische Rekonstruktion untersttzte die
Aufspaltung in eine sdliche und eine nrdliche mitochondriale Linie von O. granulifera. Dieses
Ergebnis wurde durch die Mikrosatelliten-Analyse besttigt. Weiterhin zeigt die Art
hauptschlich zwei dorsale Farbvarianten, rot im sdlichen und grn im nrdlichen
Verbreitungsgebiet. Die Verteilung von Farbmorphen ist asymmetrisch zur mitochondrialen
Variation. Die sdliche Linie umfasste nur monomorphe rote Populationen, wohingegen die
nrdliche Linie rote, grne und polymorphe Populationen enthielt. Die polymorphen
10

Populationen kommen in einer bergangszone zwischen den beiden wichtigsten Farbmorphen
vor. Hier traten neue graduell intermedire Farbmorphen, von orange-selb bis grn-braun,
zusammen mit roten oder grnen Frschen auf. Trotzdem folgte der bergang von rot nach grn
keiner einfachen geographischen Kline. Das Vorkommen von graduell intermediren
Farbmorphen deutet auf signifikante Vermischung entlang der bergangszone hin. Beide
genetischen Linien unterschieden sich signifikant in ihren Paarungsrufen und ihrer Morphologie.
In der nrdlichen Linie unterschied sich die rote Farbmorphe von der grnen Farbmorphe in der
Pulsrate der Paarungsrufe und in der Krpergre. Die Unterschiede in den Lauten und der
Morphologie beider Farbmorphen liefern Potential fr assortative mating, wobei zueinander
passende Partner bevorzugt werden. Die mehrfachen Hinweise auf die tiefreichende
intraspezifische Divergenz deuten stark darauf hin, dass es sich bei den beiden identifizierten
Linien um verschiedene Arten handelt. Dem gegenber lsst sich aufgrund der fehlenden
genetischen Diversifizierung zwischen den Farbmorphen innerhalb der nrdlichen Linie und der
genetischen Zusammengehrigkeit der bergangszone auf eine erst krzlich aufgetretene
Diversifizierung der Krperfrbung schlieen. Ausgehend von einer Sd-Nord
Ausbreitungsbewegung sind unterschiedliche lokale Selektionsdrcke der Schlssel zur
Interpretation der beachtlichen Farbvariation in diesem aposematischen System.




11

Geographical variation in the granular poison frog, Oophaga
granulifera: genetics, colouration, advertisement call and
morphology.

Oscar Brusa

Summary

The present work investigates the intraspecific genetic divergence and phenotypic
diversification in the neo-tropical poison frog, Oophaga granulifera, across its distribution
range, analysing phylogeographic patterns, population structure and phenotypic traits
potentially involved in reproductive isolation. The phylogeographic and population structure
analyses were performed using the mitochondrial genes cytochrome b and 16S and seven
polymorphic microsatellite loci from nine to eleven populations. For phenotypic analysis, I
observed the colour phenotype in 189 frogs and measured the reflectance spectra in five
targeted populations. Moreover, male advertisement calls were recorded and ten
morphological variables measured in ten populations. The phylogenetic reconstruction
supported the divergence into southern and northern mitochondrial lineages in O. granulifera,
which was confirmed by the microsatellite analyses. The species presents two main dorsal
colour morphs, red in the south of the distribution and green in the north. The distribution of
colour phenotypes is asymmetric as compared to the mitochondrial variation. The southern
lineage included only monomorphic red populations while the northern lineage contained red,
green and polymorphic populations. The polymorphic populations occured in a transition area
between the main morphs in the northern lineage. Here novel gradual intermediate colour
12

variants, varying from orange-yellow to green-brown, occurred with the red or green
phenotypes. However the gradual change from red to green did not follow simple linear
geographic colour cline. The occurrence of these intermediate morphs suggests the presence
of gene flow between the red and green morphs as confirmed by the significant admixture
across the transition area. The two genetic lineages significantly differed in advertisement
calls and morphology. The red and green colour morphs in the northern lineage differed in
pulse rate of the advertisement call and body size revealing potential for assortative mating.
In conclusion, multiple lines of evidence show that O. granulifera presents deep intraspecific
divergence strongly suggesting that the two identified lineages may well represent different
species. On the contrary, the lack of genetic diversification among colour morphs within the
northern lineage and the genetic connectivity across the colour transition area suggest a recent
colour diversification. Originating from a south-north colonization movement, locally
diversified selection is likely to represent the key for interpreting the striking colour variation
in this aposematic system.







13


1. Introduction

1.1 Biological species concept and speciation mechanisms
The scientific definition of species in biology is a complex matter not free from controversy.
The topic has been the object of extensive investigation and reviewing (Ereshefsky, 1992;
Mallet, 1995; Dover, 1995, Mayden, 1997; Burton, 2004; Coyne & Orr, 2004; Hausdorf,
2011). If we refer to sexual organisms, two species are considered distinguished when they
are reproductively isolated (Mayr, 1947, 1963, 1970). The reproductive isolation can be pre-
zygotic (Bailey et al., 2004, Dopman et al., 2010) or post-zygotic (Snchez-Guilln et al.,
2012). Pre-zygotic mechanisms avoid interbreeding among populations and the consequent
gene exchange through behavioural, physiological or morphological divergence (Schluter,
2001, 2009; Fitzpatrick et al., 2009). If the reproductive isolation persists over time, post-
zygotic isolation characterised by sterility or inviability of hybrids may arise following the
accumulation of genetic differences (Welch, 2004; Orr & Turelli, 2001). Genetic drift and
natural selection play an important role in speciation acting at pre- and post-zygotic levels
(Lande, 1976; Templeton, 1981; Schluter, 2001). Sexual selection is the other key factor able
to drive population differentiation, typically at the pre-zygotic stage (pre and post-mating)
(Noor, 1999; Panhuis et al., 2001; Kirkpatrick & Ravign, 2002), through a differential
reproductive output among individuals determined by mate choice, mostly by females, that
determines a selective pressure on the opposite sex (Lande, 1981; Panhuis et al., 2001). The
speciation process is also influenced by the geographic distribution of the diverging
populations. Three biogeographic mechanisms of species formation have been identified:
allopatry (Feder et al., 2003; Kirkpatrick & Barton, 2006; Feder & Nosil, 2009), parapatry
(Gavrilets et al., 2000; Turelli et al., 2001; Doebeli & Dieckmann, 2004) and sympatry
14

(Schliewen et al., 2006; Barluenga et al., 2006; Mallet et al., 2009). If different populations
are separated by a large distance as compared to the species vagility or an insurmountable
physical barrier such as a river or a mountain range, that is allopatry by definition. Allopatric
speciation is a classic model and there is convergence among scientists on its importance in
speciation. Parapatric speciation occurs to populations occupying adjacent areas following
behavioural or ecological diversification. Sympatric speciation concerns populations found in
the same area when reproductive isolation arises by natural or sexual selection (White, 1968,
Endler, 1977; Feder et al., 2005; Barluenga et al, 2006). Sympatric models have been
supported by some studies (Schliewen et al., 2006; Barluenga et al., 2006; Mallet et al., 2009)
but the possibility that populations may diverge in the presence of potential gene flow is not
universally accepted (e.g. Fitzpatrick et al., 2008).

1.2 Geographic variation in animals
Intraspecific geographical variation in animals is a widespread phenomenon closely linked to
the speciation process (Mayr, 1942; Gould, 1972; Endler, 1977; Coyne & Orr, 2004). For
instance, populations or individuals occurring in different areas diverge in morphology
(Grant, 1999), colour (Zink & Remsen, 1986) or calls (Wilczynski & Ryan, 1999).
Geographical variation in body size and body shape has been reported in all vertebrate taxa
including, amphibians [e.g. newts (Arntzen & Wallis, 1999); anurans (Castellano & Giacoma,
1998; Schuble, 2004)], birds [e.g. barnacle goose (Larsson & Forslund, 1991); Galapagos
finches (Grant & Grant, 1989)], mammals [e.g. bats, (Marchan-Rivadeneira et al., 2012);
rodents, (Colangelo et al., 2010)], reptiles [e.g. lizards, (Wiens et al., 2002; Malhotra &
Thorpe, 1997], fish (e.g. Barlow, 1961). Colour variation has been observed as well in several
species [e.g. fish (Kohda et al., 1996); amphibians (Hoffmann & Blouin, 2000); birds (Hill,
1993); reptiles (Rosenblum, 2004); mammals (Hoekstra et al., 2004)] and variation in
15

acoustic signals was found in amphibians (Gerhardt & Huber, 2002) and birds (Slabbekoorn
& Smith, 2002). In anurans, body size is relatively conserved within species (Cherry et al.,
1978; Green, 1993) but may vary along habitat, altitudinal and latitudinal gradients (Berven,
1982; Morrison & Hero, 2003; Rosso et al., 2004). Measuring variation in body size across
historically divergent lineages allows to infer whether population differences are explained by
geographical isolation alone, or whether localized selection contributes to fine-scale
population variation. Colouration and male advertisement calls are often heritable and subject
to selection (Endler, 1992; Ryan et al., 1996; Kingston et al., 2003; Hoekstra et al., 2004;
Hoffman et al., 2006). In polychromatic species individuals of the same sex and age display
colour variants that originate from the interplay between genetic and environmental
components (Roulin, 2004). Colour polymorphism can arise and be maintained by different
evolutionary processes. For instance, a reduction in gene flow among ecologically distinct
populations can lead to phenotypic divergence (Gray & McKinnon, 2006; Rueffler et al.,
2006). Similarly, male advertisement calls can diverge as a result of local adaptation (West-
Eberhard, 1983; Ryan & Wilczynski, 1991; Panhuis et al., 2001; Couldridge & van Staaden,
2004). All these traits have a partial (body size and colour) or exclusive (call) signalling
function which is important in intraspecific communication, being therefore potentially under
sexual selection (Arak, 1983; Forester & Czarnowsky, 1985; Ryan & Rand, 1990; Ryan &
Wilczynski, 1991; Ryan, 1999; Gabor et al., 2000). Body size can have an effect on mating
selection when larger males gain primary access to preferred females through greater
competitive ability (Berven, 1981; Elwood et al., 1987; Harari et al., 1999; Gabor et al., 2000;
Fisher & Cockburn, 2006) or when females selectively mate with larger males likely to have
superior fitness (Wedell & Sandberg, 1995; Harari et al., 1999; Gabor & Page, 2003). On the
other hand, males can also mate selectively with larger females because of fecundity
advantages (Verrell, 1985; Olsson, 1993; Harari et al., 1999; Dosen & Montgomerie, 2005;
Preston et al., 2005). Mate selection based on colour has received a great deal of attention for
16

the conspicuous traits that can arise in males through the selective process. For instance, the
blue structural colour of crown feathers (reflecting also in the ultraviolet range) has a direct
fitness effect for male blue tits because the females prefer to mate with those having a higher
reflectance (Andersson et al., 1998). Carotenoid-based colour traits of males are also selected
by females, for example in birds (Hill, 1990; MacDougall & Montgomerie, 2003) and fish
(Endler, 1983; Seehausen & van Alphen, 1998; Van Oppen et al., 1998; Barluenga et al.,
2006). Studies have shown that these colour traits are a cue for foraging ability, immune
system efficiency and overall body condition, characterising an honest signal. This depends
on the fact that carotenoid pigments cannot be synthesised endogenously and are a limited
resource in the environment. Therefore, a bright carotenoid-based colouration can only be
obtained diverting these pigments from their immune function according to a trade-off
mechanism (Blount et al., 2003; Baeta et al., 2008). Overall, the phenotypic variation
observed across the distribution of an animal species may depend on different factors that can
be summarised into three main components: genetic drift (Wright, 1948), natural (Hoekstra et
al., 2004) and sexual selection (Tilley et al., 1990; Ryan et al., 1996). These factors can act in
combination (Kirkpatrick & Ravigne, 2002). Genetic drift was introduced by Wright (1948)
and represents the stochastic genetic variation in a population. As it can be intuitively
understood, genetic drift is conceptually the opposite of selection, because the latter
modulates the expression of a certain trait according to its fitness outcome. Despite a
controversial evaluation of the actual importance of drift in the phenotypic evolution of
animal populations (Coyne, 1992; Gavrilets, 1997) it is well established that in some
circumstances, when significant numerical reductions of populations occur, such as in islands
colonization or after a bottleneck, random fluctuations in allele frequencies are able to
determine evolutionary effects (Mayr, 1963; Carson & Templeton, 1984). It seems likely that
a diversification process within a species that eventually leads to speciation may often depend
on an interaction between selection and genetic drift (Mayr, 1963; Carson, 1975; Templeton,
17

1980). For instance, local genetic drift may determine intraspecific divergence in temporarily
isolated populations (Wright, 1948; Avise, 2000) but prezygotic isolation mechanisms, such
as mate choice, are necessary to impede the possible hybridisation between the diverging
populations if gene flow is re-established (Lande, 1982; Panhuis et al., 2001). The
intraspecific divergence in selected traits and its importance for speciation is well documented
(Butlin, 1987; Endler & Houde, 1995; Boughmann, 2002). Nonetheless, the vast majority of
studies have analysed these traits independently while comprehensive integrative analyses of
phenotypic geographical variation and genetic differentiation are lacking. The importance of
geographical variation in advertisement calls and in visual signals and its role in reproductive
isolation has been shown in anuran amphibians (Blair, 1958; Littlejohn et al., 1968; Gerhardt
& Huber, 2002; Hoskin et al., 2005), but very few studies investigated the relationships
among clines in visual signals, acoustic signals and genetic population structure (Prhl et al.
2006; Amezquita et al., 2009). It is important to understand that the quantification of
geographical genetic and phenotypic diversity among populations represents a valuable
approach to get insights into the historical roles of drift and selection and evaluate the
potential for population diversification and speciation (Foster & Endler, 1999; Panhuis et al.,
2001; Gray & McKinnon, 2006; Hoffman et al., 2006). As mentioned, phenotypic
differentiation may lead to intra-specific divergence when populations are geographically
isolated or under diversifying selection (Dobzhansky, 1940; Thorpe et al., 1995; Hoekstra et
al., 2005; Prhl et al., 2006). Therefore, it is necessary to analyse the geographical variation
considering historical and recent factors to understand the relationship between phenotypic
differentiation and genetic divergence, (Lougheed et al., 2006). Despite a possible
discordance in terms of direction and rate of evolution (Marroig & Cheverud, 2001; Wiens &
Penkrot, 2002), when analysed conjunctly genetic and phenotypic traits can reveal patterns of
covariation providing information on the relative importance of gene flow and local selection
in shaping spatial and temporal diversification (Lu & Bernatchez, 1999; McKay & Latta,
18

2002; Thorpe & Stenson, 2003; Gbitz et al., 2005; Nosil & Crespi, 2006). Given the
numerous pathways for signal divergence, it is not surprising that differentiation in calls is
often observed at the population level (Ryan et al., 1996; Foster & Endler, 1999) as well as
among species (Blair, 1958; Sanderson et al., 1992; Roberts & Wardell-Johnson, 1995;
Parsons & Shaw, 2001; Smith et al., 2003). At both scales, behavioural and morphological
characters can retain evolutionary and phylogenetic signals therefore providing a valuable
integrative information in phylogeographic studies (de Queiroz & Wimberger, 1993; Foster &
Endler, 1999; Wiens & Penkrot, 2002; Price, et al., 2007).

1.3 Aposematism
Aposematic animals have conspicuous warning colour patterns informing predators that the
individual is toxic, unpalatable or otherwise unprofitable. The occurrence of species using
toxic or distasteful substances as a defence against predators is a known phenomenon,
observed for example in insects (Frazer & Rothschild, 1960) molluscs (Faulkner, 1992), frogs
(Daly, 1987) and snakes (Smith, 1977). These chemical defences are often associated with
bright colour making the bearer of such signals easily recognizable by potential predators
(Wallace, 1867; Poulton, 1890; Ruxton et al. 2004). The association of these two traits
constitutes the base for the evolutionary theory known as aposematism. The related scenario
is that a potentially costly trait (bright colour) evolved to maximise the effect of another
potentially costly trait (toxicity). Distastefulness and toxicity other than with a bright
colouration may also be coupled with other distinctive colour patterns, odours, sounds and
behaviours originating an overall conspicuous strategy aimed at maximising the efficacy of
the unprofitability signal to predators (Poulton, 1890; Cott, 1940; Edmunds, 1974). Both
theory and empirical evidences suggest that an effective warning signal should be easy to
detect, learn, recall and associate with the defence (Benson, 1972; Gittleman & Harvey, 1980;
19

Roper & Redstone, 1987; Guilford, 1990; Alatalo & Mappes, 1996; Gamberale-Stille &
Tullberg, 1999; Lindstedt et al., 2011). A key point of the aposematic theory (Mller, 1879) is
that populations of aposematic species should converge into one conspicuous phenotype
because in this way the predator will easily learn to avoid that specific colour, making fewer
recognition errors (e.g., Guilford, 1986; Mallet & Barton, 1989; Joron & Mallet, 1998;
Gamberale-Stille, 2000; Kapan 2001; Beatty et al., 2004; Rowland et al., 2007; but see Rowe
et al., 2004; Ihalainen et al., 2007). Accordingly, selection is expected to favour the most
conspicuous and most common colour morph and decrease variation in warning coloration.
On the contrary, the presence of polymorphic populations is considered evolutionary not
optimal because it would imply a longer learning time by the predator with consequent
potential detrimental effect on the aposematic species individual fitness. Being the bright
colour associated with toxicity it represents a signal that the bearer cannot manipulate. This is
the reason why aposematic colouration has been explained within the theoretical framework
of honest signalling (Blount et al., 2009). The described scenario applies if we consider the
evolution of bright colour successive to the evolution of toxicity. On the other hand, one
cannot rule out the possibility that the bright colour evolved first, independently by toxicity.
In this case, the bright colour would have made the animal more detectable and therefore
more exposed to predation attempts, making the phenotype counter-selected via natural
selection. Subsequently, to compensate for the higher predation pressure caused by a
conspicuous appearance, the species evolved a system of chemical defence so that to originate
a deterrent effect to the predator and re-establish an evolutionary optimal strategy.

1.4 The remarkable case of polymorphic toxic species
One of the most striking examples of divergence in visual signals in animals is colour
polymorphism, defined as the co-occurrence of different colour morphs in a population
20

(Mller, 1879). Colour polymorphism belong to a broader phenomenon definable as colour
diversity or polychromatism, that is the occurrence of intraspecific colour variation, at the
interpopulation but not necessarily at the intrapopulation level. As previously mentioned
colour traits may depend on structural characteristics of integuments or pigments or a
combination of the two (Auber, 1957; Sieffermann & Hill, 2003; Endler, 1983). There are two
major classes of pigments responsible for colouration in animals: melanins and carotenoids
(Griffith et al., 2006). Melanins are synthesised endogenously and are responsible for both
black (across taxa) and rufous (mammals and birds) colour background and patterning in the
variants of eumelanin and pheomelanin respectively (Roulin, 2004). On the contrary, animals
are not able to synthesise carotenoids de-novo and therefore depend on the diet for displaying
a colouration based on these pigments. Animals can produce red, brown, orange and yellow
pigmentation rearranging enzymatically the many molecular variants of carotenoids (over 600
split into two classes, xantophylles and carotenes). Colour diversity depends mainly on
pigments variation and is a relatively common phenomenon in vertebrates, observed for both
melanins and, less frequently, carotenoids (Roulin, 2004; Hoffman & Blouin, 2000). It is also
worth to mention a less common class of endogenously synthesised pigments called pteridines
that can originate a colour range overlapping with those dependent on carotenoids. Pteridines
have been found for example in wings of butterflies (Rutowski et al., 2005; Bezzerides et al.,
2007), skin of fish (Grether et al., 2001), amphibians (Obika & Bagnara, 1964) and reptiles
(Steffen & McGraw, 2006) and irises of birds (Hill & McGraw, 2006). Pteridines can be
synthetized in the body therefore they are not considered components of honest signalling
systems. A biochemical analysis able to discriminate between carotenoids and pteridines is
therefore necessary for a correct inference concerning the evolutionary ecology of
pigmentation. Notably, colour polymorphism has been observed in association with toxicity
challenging the Mllerian prediction of phenotypic convergence in aposematic species. In
fact, aposematic animals are usually chromatically monomorphic (Edmunds, 1974;
21

Greenwood et al., 1981). Nonetheless, colour polymorphism exists in aposematic insects such
as lepidopterans [(Brakefield & Liebert, 1985; Stimson & Berman, 1990 including Heliconius
butterflies (Mallet, 1986)] ladybirds (Hodek & Honk, 1996; Povoln, 1999) and true bugs
(Heteroptera; tys, 2004). Among vertebrates dendrobatid frogs provide the most significant
example, showing a striking array of colour morphs, (Daly & Myers, 1967; Summers &
Clough, 2001; Summers et al., 2003; Darst & Cummings, 2006). This unexpected pattern is
known, for example, in the strawberry poison frog [Oophaga pumilio (Schmidt, 1857)] whose
colour diversity spans the visual spectrum including even cryptic variants (Summers et al.,
2003; Prhl & Ostrowski, 2011). The presence of cryptic morphs is surprising because
predators are thought to learn to avoid unpalatable prey faster if they are conspicuous rather
than cryptic (Gittleman & Harvey, 1980; Roper & Redston, 1987; Alatalo & Mappes, 1996;
Lindstrm et al., 1999a) and if they are clearly contrasted against the average background
(Roper & Redston, 1987; Lindstrm, et al., 1999a; Gamberale-Stille, 2001; Prudic et al.,
2007). Several factors have been taken into consideration to explain the occurrence of colour
polymorphism in aposematic species. These include: genetic drift (Mallet & Singer, 1987;
Turner & Mallet, 1996; Mallet & Joron, 1999, Rudh et al., 2007), sexual selection (Summers
et al., 1999, Maan & Seehausen, 2011), variation in toxin availability (Wang, 2011) and
variation in predation regime (Mallet & Joron, 1999; Endler & Mappes, 2004; Nosil et al.,
2005; Ratcliffe & Nydam, 2008).

1.5 The biogeographical peculiarities of the lower Central American region
The numerous studies carried out on Central America anuran amphibians have provided
insights into the great level of variation typical of this biogeographic region (Ryan et al.,
1996; Crawford, et al. 2007; Robertson et al., 2009). Due to its complex geomorphologic and
climatic history, Central America presents high biological diversity at the inter- and
22

intraspecific levels with the latter resulting in the evolution of different genetic lineages in
several species (Coates & Obando, 1996; Campbell, 1999; Kohlmann et al., 2002; Savage,
2002). Particularly, the history of Costa Rica and Panama has been profoundly influenced by
largescale landscape changes (Campbell, 1999; Kohlmann et al., 2002; Savage, 2002). A
fundamental geomorphologic trait of the area is the Cordillera de Talamanca, a mountain
chain extending across lower Central American that originated approximately 3 million years
ago dividing the area in Pacific and Atlantic slopes. This structure markedly altered the
geomorphology and climate of the Caribbean, Central and Pacific areas (Holdridge, 1947;
Campbell & Frost, 1993; Coates & Obando, 1996), producing a sharp geoclimatic
differentiation. This diversity likely contributed to the exceptionally high species richness and
endemicity characteristic of this area (Holdridge, 1947; Savage, 2002). The Cordillera de
Talamanca plaid a role in the diversification patterns across many Central American taxa
[terrestrial snakes (Zamudio & Greene, 1997; Garca-Pars et al. 2000); beetles (Kohlmann
et al. 2007); frogs (Crawford, 2003; Robertson & Robertson, 2008). In Costa Rica, the
southwestern region presents high level of endemism following the uplift of the Cordillera
and the consequent isolation from northwestern CR and Caribbean lowland forests
(Holdridge, 1947; Savage, 2002). Another biologically highly diverse area is the Osa
Peninsula, that may have derived from an offshore island drifted into mainland Costa Rica
approximately 2 million years ago (Kohlmann et al., 2002). Genetic isolation of the Osa
Peninsula populations from mainland Costa Rica has been documented in frogs (Crawford,
2003; Crawford et al., 2007; Robertson, 2008) and snakes (Zamudio & Greene, 1997). These
studies suggest that the Osa Peninsula has retained a signature of isolation, despite
reconnection to the mainland. In Panama, the Caribbean region of Bocas del Toro is a
significant biogeographic area for animal and plant taxa, including frogs (Crawford, 2003;
Weigt et al., 2005; Crawford et al., 2007; Hagemann & Prhl, 2007).

23

1.6 Evolutionary studies on amphibians in the lower Central American
region
Studies conducted on amphibians of Costa Rica and Panama returned a complex picture of the
interplay between genetic and phenotypic diversification. For instance, populations of a
species may be phenotypically highly diversified despite a lack of concordant genetic
divergence as identifiable using nuclear and mitochondrial markers (Hagemann & Prhl,
2007; Hauswaldt et al., 2011). On the contrary, populations rather similar in terms of
phenotype may be divergent when analysed at the molecular level (Hauswaldt et al., 2011;
present study). Therefore, a compelling aim of modern evolutionary studies is to clarify the
patterns of association between the variation in traits such as body size, colour, call and
historical and recent population genetic structure. This together with information on
ecological variation would allow to get insights into the possible evolutionary origin of the
stunning diversity of the area. Despite the need to consider all the relevant traits to have a
global description of the possible interrelations between genetic diversity and phenotype few
studies have really adopted a comprehensive approach. In the case of Neotropical anuran
amphibians, recent investigations have focussed on reconstructing the phylogeographic
patterns (Wang & Shaffer, 2008; Brown et al., 2010; Vences et al., 2003), some incorporated
data on recent population structure (Wang & Summers, 2010; Hauswaldt et al., 2011) and few
related those genetic traits to biological important components such as advertisement calls and
morphology (e.g. Prhl et al., 2006; Amezquita et al., 2009). A clear example of the lack of
such a comprehensive approach is represented by dendrobatid frogs, that present a remarkable
colour variation occurring in some species, for instance in the Panamanian populations of the
strawberry poison frog Oophaga pumilio (Maan & Cummings, 2012). Several studies have
dealt with dendrobatids behaviour (Prhl, 1997; Summers et al., 1999; Prhl & Hdl, 1999;
Prhl, 2003, 2005; Prhl et al., 2007; Richards-Zawacki & Cummings, 2011), aposematic
defense evolution [using theoretical (Ruxton et al., 2004, 2007; Lee et al., 2011), empirical
24

(Maan & Cummings, 2008; Darst et al., 2006; Noonan & Comeault, 2009) and phylogenetic
approaches (Santos et al., 2003; Brown et al., 2010; Wang, 2011)] and the phylogenetic
relations within the Oophaga (former Dendrobates) genus (Hagemann & Prhl, 2007;
Hauswaldt et al., 2011). Nonetheless, only a few studies have considered a key trait, for its
importance in intraspecific divergence, that is the male advertisement call (e.g. Prhl et al.,
2007), that has been shown to play a role in mate choice at the population level in O. pumilio
(Prhl et al., 2003) and in other Neotropical amphibians (Prhl et al., 2006). It is also
remarkable that other traits potentially important in sexual selection such as morphology have
been quite neglected to date. Moreover, significant levels of genetic divergence were found
among populations of Costa Rica and Panama leading to the definition of different lineages in
some dendrobatid species (Symula et al., 2003; Noonan & Wray, 2006; Hagemann & Prhl,
2007; Hauswaldt et al., 2011) as well as other amphibians (Ryan et al., 1996; Crawford et al.,
2007; Robertson et al., 2009). The complex paleogeographic history of the Neotropics
accounts for this high level of diversity (Lougheed et al., 1999; Graham et al., 2004; Elmer et
al., 2007; Roberts et al., 2007; Santos et al., 2009). The bright colour of many dendrobatids is
commonly associated with toxicity and explained according to the aposematism theory
(Summers & Clough, 2001; Darst & Cummings, 2006). Nonetheless, colour diversity among
populations and dull colour morphs exist (Savage, 1968). An instructive case is the strawberry
poison frog, Oophaga pumilio (Schmidt, 1857), that is monomorphic in Costa Rica but shows
a high number of colour morphs in Panama, in the Bocas del Toro archipelago (Daly &
Myers, 1967; Summers et al., 2003). A phylogeographic study found a southern and a
northern lineage with the former including all the polymorphic populations together with the
southern Costa Rica monomorphic populations (Hagemann & Prhl, 2007). The presence of
two main lineages in O. pumilio has been confirmed by the bioacoustic divergence in the male
advertisement call (Prhl et al., 2007). The importance of advertisement calls in anuran
reproductive isolation has been shown in several studies (Littlejohn et al., 1968; Gerhardt &
25

Huber, 2002; Hoskin et al., 2005). Nonetheless, there is a general lack of studies combining
phylogeographic reconstruction with an analysis of advertisement call variation in poison
frogs. Investigating the correlation between genetic distance and divergence in bioacoustic
traits is useful to evaluate the patterns of reproductive isolation and speciation (e.g. Prhl et
al., 2006). In fact, a low level of genetic differentiation between colour phenotypes has been
observed (Hagemann & Prhl, 2007; Wang & Shaffer, 2008; Brown et al, 2010; Hauswaldt
et al., 2011; but see Wang & Summers, 2010). In this scenario sexual selection is likely to
have played a key role in the fixation of the colour peculiarities of these populations, with
females preferring males of their local insular colour morph (Summers et al., 1999; Reynolds
& Fitzpatrick, 2007; Maan & Cummings, 2008; Richard-Zawacki & Cummings, 2010).
Despite the obvious phenotypic divergence, there is no clear evidence of genetic
differentiation between colour morphs in this species (Prhl et al., 2007; Hauswaldt et al.,
2011 but see Wang & Summers, 2010). Moreover, in dendrobatid species, studies
investigating the correlation between the geographical variation in colour and other traits
potentially involved in reproductive isolation are scant. Colour variation in poison frogs has
been related to several factors such as local variation in the predator communities, toxin
availability and assortative mating based on colour and call traits (Summers et al., 1997,
1999; Speed & Ruxton., 2004; Darts & Cummings, 2006; Saporito et al., 2007a; Speed &
Ruxton, 2007; Maan & Cummings, 2008, 2012). Within this framework, it is particularly
important to investigate the genetic and phenotypic variation across the species distribution to
have a clear picture of population-level diverging patterns.

1.7 Neotropical poison-dart frogs
During the last three decades the scientific knowledge on dendrobatid frogs, commonly
known as poison-dart frogs, has increased significantly. The number of described species
26

exponentially grew over the years and 247 are presently recognized. Dendrobatids are found
in dense primary and secondary forests, open areas, mountains and lowlands, and occupy a
variety of aquatic, terrestrial and arboreal habitats, in Central and South America. They are
distributed from Nicaragua to Bolivia and the Atlantic coast of Brazil and from the Pacific
coast of South America to the French Antilles (Savage, 2002). Approximately one-third of the
known species of dendrobatids have toxins in special dermal glands and are able to release
these toxins through the skin (Savage, 1968; Silverstone, 1975). The name poison-dart frogs
derives from the tradition of the native tribe of the Ember, inhabiting the Choc region of
western Colombia, to poison their hunting blowgun darts using the skin secretion of three
dendrobatid species, Phyllobates terribilis (Myers et al., 1978), P. aurotaenia (Boulenger,
1913), and P. bicolor (Dumril & Bibron, 1841). These three species contain in their skin two
of the most powerful toxins known, batrachotoxin and homobatrachotoxin (Myers et al.,
1978). Over 450 different lipophilic alkaloids have been isolated from dendrobatids and 24
major structural classes were defined (Daly et al., 1999). Many of these compounds found
application in medical research (Daly et al., 1997, 2000; Daly, 1998). It is known that
dendrobatids are not able to synthesise their toxins de-novo but they take them from the diet.
However, the details on the sequestration mechanism are still unknown (Daly et al., 1994a).
The dietary sources of alkaloids identified so far are formicine ants, a siphonotid millipede,
melyrid beetles and scheloribatid mites (Daly et al., 1987; Saporito et al., 2003; Dumbacher et
al., 2004; Saporito et al., 2004; Takada et al., 2005; Saporito et al., 2007a), while the
remaining alkaloids are still unknown elsewhere in nature. The vast majority of dendrobatids
alkaloids are lypophilic molecules but one hydrophilic alkaloid, called tetrodotoxin, has been
detected in Colostethus inguinalis (Cope, 1868) (Daly et al., 1994b). In addition to toxicity,
Neotropical poison-arrow frogs are characterised by many other peculiar traits (Daly &
Myers, 1967; Weygoldt, 1987). Dendrobatids, differently from the vast majority of
amphibian species, are diurnal (except Aromobates nocturnus, critically endangered and
27

poorly known species found in north-west Venezuela), they present bright and dull colours,
conspicuous and cryptic behaviour, territoriality with aggressive defence observed in both
males and females and complex mating rituals (e.g., Silverstone, 1973; Wells, 1978, 1980a,
1980b, 1980c; Weygoldt, 1987; Zimmermann & Zimmermann, 1988; Summers, 1989;
Aichinger, 1991; Caldwell, 1997; Fandino et al., 1997; Junc, 1998; Caldwell & de Oliveira,
1999; Summers et al., 1997; Summers et al., 1999; Siddiqi et al., 2004; Prhl, 2005; Maan &
Cummings, 2008; Lddecke, 2000; Bourne et al., 2001; Prhl & Berke, 2001; Prhl, 2003;
Narins et al., 2003, 2005; Summers & McKeon, 2004). They lay terrestrial eggs, either on the
ground or in water-filled leaf axils of phytotelmata plants (from the ancient greek phyto-,
meaning 'plant', and telma, meaning 'pond'). Many species are characterized by elaborate
reproductive behaviours, biparental care, eggs moistening, egg guarding, tadpoles
transportation on the dorsum of parent frogs and tadpole feeding with unfertilized oocytes by
the mother (genus Oophaga) and one or both sexes are involved in highly evolved parental
care (Weygoldt, 1980; Zimmermann & Zimmermann 1988; Prhl, 2005). In many species,
the male advertises a reproductive territory by emitting a species-specific call and defending it
actively from conspecific male intruders (Prhl, 2005). Despite a broad knowledge on the
different aspects of their biology, the phylogenetic relationships among dendrobatids at the
inter- and intraspecific level are not yet fully understood. Phylogenetic relationships have
been used to investigate the evolutionary origins of dendrobatid features (e.g., Summers et al.,
1999; Santos et al., 2003; Vences et al., 2003; Graham et al., 2004; Darst et al., 2005). The
growing awareness of the importance of well resolved phylogeny for dendrobatids has
recently brought to significant systematic revisions (e.g. Noonan & Gaucher, 2006;
Wollenberg et al., 2006; Grant et al., 2006; Santos et al., 2009).


28

1.8 The granular poison-dart frog, Oophaga granulifera
The granular poison frog, O. granulifera, shows a very limited distribution in south-central
pacific Costa Rica (Fig. 1). It inhabits the lowland forests of the pacific slope of Cordillera de
Talamanca in Costa Rica and possibly north Panama. It is endangered because of habitat
fragmentation and included in the IUCN Red List as a vulnerable species (IUCN Red List). In
the southern part of its distribution the species is characterised by a red dorsal patch and a pale
blue to grey colour on the remaining part of the body. The proportion of body area covered
with red increases from the south to the centre of the distribution. Three colour morphs were
known before the present study: a red colour morph in the south and central part of the
distribution, a green colour morph in the north and a yellow colour morph located in between
(Savage, 2002; Wang, 2011). The species was classified in the histrionica group (Myers et al.,
1984) together with other 8 species found in Costa Rica, Panama and Colombia [O. arborea
(Myers et al., 1984), O. histrionica (Berthold, 1845), O. lehmanni (Myers & Daly, 1976), O.
occultator (Myers & Daly, 1976), O. pumilio (Schmidt, 1857), O. speciosa (Schmidt, 1857),
O. sylvatica (Funkhouser, 1956) and O. vicentei (Jungfer et al., 1996). Currently the same
species are grouped in the Oophaga genus, a name that derives from the unusual feeding of
the tadpoles on unfertilized eggs provided by the mother (Grant et al., 2006). O. granulifera
(Taylor, 1958) shares several behavioural traits with the other Oophaga species, such as male
territorial behaviour, courtship behaviour and predominantly female parental care including
tadpole transportation and tadpole feeding (Goodmann, 1971; Crump, 1972; van
Wijngaarden, 1985; Bolaos, 1990; van Wijngarden & Bolaos, 1992; van Wijngarden &
Gool, 1994) other than tadpole morphological traits (Grant et al., 2006). However, it is rather
divergent from congeneric species in terms of bioacoustic properties of the male
advertisement call, particularly the length of the acoustic unit (call) within a call sequence
(Meyer, 1992, 1993). In addition, phylogenetic assessments concerning Oophaga spp. and
29

including a few samples of O. granulifera, found indications of a high genetic divergence of
the latter within the Oophaga genus (Clough & Summers, 2000; Hagemann & Prhl, 2007).

Figure 1 Sampling localities of Oophaga granulifera. The populations from Wang (2011) are
marked with *. Population codes as in Table 1.


1.9 Aims of the study
In the work here presented, I investigated the phylogeography, the population structure and
the geographical variation in colour, advertisement call and morphology in the granular
poison frog, Oophaga granulifera (Taylor, 1958). I aimed at evaluating the relationships
between historical and recent genetic structure, colour diversity and geographical variation in
30

traits which are potentially involved in reproductive isolation such as male advertisement call
and morphology. In particular, I focused on populations located in a very narrow colour
transition area between the two main colour morphs red and green, where intermorph crossing
was expected and intermediate colour phenotypes were more likely to occur. This assumption
was strengthened by the description of an intermediate yellow phenotype (Savage, 2002;
Wang 2011). I aimed at clarifying which of two possible scenarios may describe O.
granulifera colour diversification: the presence of intermediate gradual phenotypes as a result
of gene flow among the main morphs or, alternatively, the occurrence of the three discrete
colour morphs in reproductive isolation. I analysed the colour variation across the transition
area using field observations and spectrometric measurements and estimate the
interpopulation gene flow using microsatellite loci. In addition, I studied the geographical
variation in male advertisement call and morphology as potential sexual signals involved in
mate choice, to investigate how they relate to genetic and colour variation. Recently, Wang
(2011) identified two monophyletic clades in the species, one of southern red populations and
one of yellow and green populations nested within a northern clade which also contains a red
population. The author considered nine populations: four southern red, one central red, two
northern intermediate yellow and two northern green (Fig. 1). I analysed the phylogeography
of O. granulifera including three newly found polymorphic populations occurring in the
colour transition area. I also considered seven other populations: one northern green, three
northern red located within 20 km from the transition area and three red from the Osa
peninsula in the south (Table 1; Fig. 1). This population sampling allowed a detailed
evaluation of the historical genetic divergence across the colour transition area. Moreover, I
analysed the male advertisement calls and the morphological characteristics of the different
populations as key factors in mating recognition and consequently in reproductive isolation, to
unravel the pattern of interactions between morphological differentiation and speciation.
Summarising, the present study applied an integrative approach based on phylogeographic
31

analysis, bioacoustic and morphological analyses, detailed colour phenotype observations and
spectral reflectance measurements. Here, I provide the most comprehensive reconstruction of
O. granulifera phylogeography to date, including populations containing previously
undescribed colour morphs. I clarified the existence of two deep genetic lineages within the
species supported also by the differentiation in the male advertisement calls and morphology.

1.10 Study area
The study was carried out in south-western Costa Rica sampling in the lowland forest along
the pacific coast. The study area extends from the Osa Peninsula in the south to the central
part of the Puntarenas province between the towns of Quepos and Parrita and is rather
diversified in terms of climate and vegetation cover. The Osa peninsula is the most pristine
area of Costa Rica, (hosting for example the Corcovado national park, a world known
biodiversity hotspot) characterized by an extensive forest coverage and a low urbanization.
The area between the towns of Palmar and Quepos is characterised by a high human impact,
deforestation and consequent habitat fragmentation. The surroundings of Quepos also host
extended oil palm plantations. On the study area, many rivers are present including the Rio
Trraba (Fig. 1), the largest of Costa Rica that divides the town of Palmar into two parts,
named Palmar Norte and Palmar Sur.





32

Table 1 Information for each study population: name, code, geographical location and
coordinates, number of individuals sequenced for 16S and cyt b and genotyped for
microsatellites, number of individuals captured for colour description, number of male
advertisement calls recorded and frogs measured for morphological analysis.
Population Code CR region GPS Cyt b 16S Msats Colour Call Morphology
Ventana VE Osa Peninsula
N842.457
W8329,589
2 7 16 9
Charcos CH Osa Peninsula
N840.319
W8330.320
3 3 8 19 8 19
Neotropica NE Osa Peninsula
N842.082
W8330.783
8 24 15 15
Palmar PA South-West
N858.112
W8327.356
3 3 7 14 4 14
Firestone FS Central-West
N916.668
W8352.167
3 2 8 15 5 15
Dominical DO Central-West
N915.065
W8351.292
3 2 8 15
Baru BA Central-West
N916.416
W8352.464
3 1 8 14 5 13
Matapalo MA Central-West
N920.913
W8352.167
4 1 10 20 3 20
Portalon PL Central-West
N920.951
W8356.277
3 2 15 19 4 19
Savegre SA Central-West
N922.039
W8358.277
1 1 8 20 6 19
Damitas DA North-West
N932.124
W8412.222
3 1 11 13 13




33

2. Material and Methods

2.1 Field activity
The data used in the present study were collected during three field seasons from 2008 to
2010 for a total of seven months. The work was carried out between April and July in the first
half of the rainy season when the frogs calling was intense. According to the frogs activity
field work was performed in most cases from 5:00 am in the morning to 1:00 pm when
temperature increased and the calling activity became weak. Data collection was also
performed during minor activity time of Oophaga granulifera in late afternoon until sunset. In
case of precedent very abundant rainfall or slight rain during the day, frogs remain active
throughout the day enabling continuous data collection.

2.2 Tissue sampling and DNA extraction
Tissues samples used for genetic analyses were obtained in the field by toe-clipping and
stored in 96% ethanol in 1.5 ml plastic tubes with screw cap. Toe-clipping is the removal of
one or more toes and is a common practice in amphibian studies (Hero, 1989; Waichman,
1992; Halliday, 1996). Other than for obtaining tissue for genetic analyses it is used to allow
individual recognition in behavioural studies removing several toes in specific combinations
and to obtain bone for estimating age using skeletochronology (Friedl & Klump, 1997;
Driscoll, 1998; McGuigan et al., 1998). Some studies revealed possible deleterious effects of
removing several toes in amphibians (Clarke, 1972; Humphries, 1979; Golay & Durrer, 1994;
Reaser & Dexter, 1996) while others did not (Lemckert, 1996; Williamson & Bull, 1996;
Parris & McCarthy, 2001). Nonetheless, it is recognized that the removal of one toe is likely
not to provoke any negative effect on survival (McCarthy & Parris, 2004). In the present
34

study only a portion of one toe (2-3 mm) was removed and no adverse effects were observed
after release. The sample was stored for some weeks in the 1.5 ml screw cap plastic tubes
filled with 96% ethanol. The preservation of tissue for DNA extraction in ethanol for a short
time is appropriate because ethanol inhibits the proteolytic enzymes responsible for tissue
degradation. Once in the laboratory, the samples were stored in the fridge for few days until
DNA extraction. The DNA was extracted in the genetic laboratory of the Institute of Zoology
at the TiHO (co-responsible Prof. Dr. Heike Prhl) in Hannover and at the Department of
Zoology of Cambridge University (evolutionary genetics group, responsible Dr. Nicholas
Mundy) using a Qiagen DNeasy extraction kits and a Qiagen QIAquick purification kit
respectively (Qiagen, Hilden, Germany) following the manufacturer's instructions. The
general procedure involved digestion of the tissue in 1.5 ml tubes containing a proteinase K
solution and incubation at 37C overnight. The proteinase K releases the DNA from the tissue
and inhibits the nucleases enzymes responsible for nucleic acid degradation. Then a binding
buffer was added and the samples were passed through filtering mini-columns embedded in
1.5 ml centrifuge tubes. After centrifugation, a washing buffer was added to the columns to
eliminate the impurities from the filter that binds the DNA and the flow-through discarded. As
a final step the DNA was eluted by adding an elution buffer in two successive aliquots each
followed by centrifugation. The eluate was stored at -20 C. Small known aliquots of purified
DNA mixed with a loading dye were run on 1% agarose gel stained with ethidium bromide to
quantify the DNA yield and prepare proper dilution for subsequent amplification steps.

2.3 Phylogeography
2.3.1 Mitochondrial DNA sequencing
Two fragments of mitochondrial DNA were considered for the phylogeographic analysis. I
amplified a 600 bp fragment of the cytochrome b gene (cyt b) from 28 individuals (10
35

populations) using primers MTAL (Hauswaldt et al., 2011) and cyt b-c (Bossuyt &
Milinkovitch, 2000). A portion (527 bp) of the 16S rRNA gene (16S) was amplified in a
subset of samples (N = 16; 9 population) using universal primers 16SA and 16SB (Palumbi et
al., 1991) (Table 1). PCRs were performed in 25 l volume using Taq DNA-polymerase
(Eppendorf, Hamburg, Germany). The thermocycling conditions for cyt b comprised an initial
denaturation at 95C for 2 min followed by 35 cycles of denaturation (95C for 45s),
annealing (52C for 45s), elongation (72C for 1min), and a final elongation at 72C for 10
min. The profile for 16S was identical except for an annealing temperature of 55C. PCR
products were purified using Qiagen purification kits and filtering columns system described
above. The purified PCR product was then precipitated and prepared for sequencing
according to standard procedure. The fragments were sequenced in both directions in 10 l
volume using BigDye 3.1. The sequences were run on an ABI sequencer 3500 (Applied
Biosystems, Darmstadt, Germany).

2.3.2 Mitochondrial DNA analyses
Sequences were aligned and edited with MEGA 4.1 (Tamura et al., 2007). Haplotypes were
extracted using the online tool DNA to haplotype collapser and converter available in
FABOX 1.40 (http://www.birc.au.dk/software/fabox/). I calculated the number of variable
sites as well as uncorrected pairwise distances between populations (p-distance) using MEGA
4.1. I performed Maximum Likelihood and Bayesian phylogenetic analyses with PAUP* 4.0
(Swofford, 2002) and MrBayes 3.1 (Ronquist & Huelsenbeck, 2003), respectively. All the cyt
b sequences of comparable size published for O. granulifera were included in the analyses
(Wang, 2011; see Additional Information for all accession numbers). The populations
considered in Wang (2011) were Corcovado (CO), Drake Bay (DB), Piedras Blancas (PB),
Rio Terribe (TE), Rio Baru (RB), Cerro Nara (CN), Rio Savegre (SV), Rio Damitas (DM) and
Fila Chonta (FC). The geographical areas of Costa Rica where the populations were found
36

and the correspondent colour phenotypes described in Wang (2011) were: south-west, red
(CO, DB, PB), central-south-west, red (TE), central-west, red (BR), central-north, yellow
(CN, SV), central-north, green (DM, FC) (Fig. 1). Hereafter, these populations will be cited in
the text using the abbreviations. I am not aware of any sampled population in common with
Wang (2011) (Fig. 1). As outgroups, I included eight 16S sequences from all but one of the
other species of the genus Oophaga (nomenclatural authorities) [O. arborea (Myers et al.,
1984), O. histrionica (Berthold, 1845), O. lehmanni (Myers & Daly, 1976), O. pumilio, O.
speciosa (Schmidt, 1857), O. sylvatica (Funkhouser, 1956) and O. vicentei (Jungfer et al.,
1996); two sequences for O. pumilio northern and southern lineages, one sequence for each
other species; O. occultator (Myers & Daly, 1976) was not included because there were no
sequences available], two 16S sequences of Dendrobates leucomelas and D. tinctorius and
two cyt b sequences of O. histrionica and O. sylvatica, (see Additional Information for all
accession numbers). For ML analyses, I selected the model of evolution that best fitted the
data using jMODELTEST 0.1 (Posada, 2008). Analyses were carried out using a heuristic
search with 10 random additions per replicate and TBR branch swapping. Node confidence
was assessed using 1000 bootstrap replicates. Bayesian posterior probabilities were calculated
using MrBayes 3.1 (MCMC sampling) applying the substitution model selected in
MrModeltest 3.7. Two markov chains were run from random trees for 4 10
6
generations and
monitored to ensure that the average standard deviation of split frequencies was <0.01. The
20% of the initial trees were discarded as burn-in. Finally, I reconstructed the cyt b haplotype
network using TCS 1.21 (Clement et al., 2000) and determined the minimum number of
mutational steps required to connect different clusters of haplotypes using the fix connection
limit option. I included the same published cyt b sequences (Wang, 2011) considered in the
phylogenetic analyses.


37

2.4 Population genetics
2.4.1 Microsatellite loci genotyping and analyses
Seven microsatellite loci [B9, C3, D4, E3, F1, G5, O1] previously isolated for O. pumilio
(Hauswaldt et al., 2009) were amplified via PCR from seven to 15 individuals belonging to
eleven populations. PCRs were performed in 25 l volume using Taq DNA-polymerase
(Eppendorf, Hamburg, Germany). The thermocycling conditions comprised an initial
denaturation step at 95C for 30 sec followed by 35 cycles of denaturation (95C for 30 sec),
annealing [48C (C3); 57C (B9, D4, F1); 59C (E3); 60C (G5); 64C (O1) for 45 sec] and
elongation (72C for 45 sec) followed by a final elongation at 72C for 10 min. Then
genotyping (electrophoresis) was performed on a ABI 3500 sequencer (GE Healthcare, Little
Chalfont, UK) using ladder GeneRuler 100bp Plus (Fermentas, Vilnius, Lithuania) and
fragments were scored with GeneMapper 4.1 software (Applied Biosystem, Foster City,
California, USA). I used GenAlEx 6.4 (Peakall & Smouse, 2006) to calculate observed and
expected heterozigosities (H
o
and H
e
) and the number of private alleles (A
P
). Average allelic
richness (A
R
), for at least seven individuals per population was calculated with FSTAT
(Goudet, 1995). I tested each locus and population for linkage disequilibrium and deviation
from Hardy-Weinberg equilibrium (HWE) with Genepop on the Web (Fishers exact tests,
Raymond & Rousset, 1995a,b; Rousset, 2008) using the Markov chain Monte Carlo (MCMC)
and default parameters (10,000 dememorizations, 1000 batches and 10,000 iterations per
batch). F
ST
pair-wise values (Weir & Cockerham, 1984) and the molecular variance
(AMOVA; Excoffier et al., 1992) were computed with Arlequin 3.11 (Excoffier et al., 2005).
Significance was tested with 10,000 permutations. Finally, I performed a population
assignment test based on Bayesian posterior probabilities as implemented in Structure 2.3.3
(Pritchard et al., 2000). Runs were performed with 110
6
burn-in period and MCMC of
310
6
, setting an admixture model (
initial
= 1.0;
max
= 10.0). I tested one to 13 possible
clusters (K), and 10 runs were conducted for each K. The uppermost hierarchical level of
38

population structure was determined according to Evanno et al. (2005), checking the greatest
rate of change in estimated likelihood between successive K-values.

2.5 Colour description and measurements
For colour description, 189 frogs were photographed with a digital camera Nikon Coolpix
4500 and colour phenotype characteristics recorded from 11 populations distributed across the
known distribution range of O. granulifera (Table 1; Fig. 1). Light reflectance measurements
were taken from 15 individuals each from populations Firestone, Dominical, Matapalo,
Portalon and Savegre using an Ocean Optics 2000+ (Ocean Optics Inc., Dunedin, FL, USA)
fiber optic spectrometer and a deuterium-tungsten DT-Mini-2-GS light source. Reflectance
spectra were obtained averaging five measurements each from the head, upper dorsum and
lower dorsum. Geographical coordinates and localities altitudes were documented with a
Garmin GPS 12 (Garmin Ltd., Schaffhausen, Switzerland).

2.6 Bioacoustics
Male advertisement calls were recorded using a Sennheiser directional microphone (MHK
416 P48; Sennheiser, Wedemark-Wennebostel, Germany) and a digital walkman Marantz
(PMD 671, Marantz Corporation, Kanagawa, Japan). The calls of 59 frogs from nine
populations were recorded (Table 1). Due to very low calling activity in the Dominical and
Damitas populations, the calls of frogs from these populations could not be recorded. In most
populations the frogs were found in close proximity to creeks, therefore many of the recorded
calls contained background noise, leading to low sample size in some populations. During
recording the distance to the frogs was no more than two metres to minimize environmental
noise interference. The calls were recorded mostly in the morning hours (5.00 to 13.00) when
the calling activity was more intense. After each recording, air temperature was measured to
39

allow analyses to be controlled for temperature effect on the calls. Pulse rate, call duration and
call rate were calculated using oscillograms. Dominant frequency was measured by using
spectrograms. For details on call parameters definition see Prhl (2003). Bioacoustic analyses
were performed using Avisoft (Avisoft Bioacoustics, Berlin, Germany). I used one-way
ANOVA to test for differences in call parameters among populations and colour morphs.

2.7 Morphology
Morphological measurements were collected from 147 frogs belonging to nine populations
(Table 1). Body weight was measured using a digital scale to the nearest 0.01g. Body (snout-
vent), femor, tibia, humerus, radio-ulna and head lengths, head width, tympanum diameter
and between-nostril distance were measured with a digital calliper to the nearest 0.01mm. I
collected all measurements personally to avoid inter-operator variation. I used one-way
ANOVA to test for differences in morphological parameters among populations and colour
morphs.







40

3. Results

3.1 Colouration
I present the results on coloration first because they are important for interpreting the genetic
results. I found that the red colour phenotype gradually changed from the south (Charcos,
Neotropica, Ventana) to the north of the species distribution. I observed an increase in the size
of red patch over the body (Palmar, Firestone, Dominical, Baru) and then a change to an olive
green phenotype. In the transition area, populations were polymorphic and the two main
colour morphs (red and green) co-occurred with an array of intermediate phenotypes (Fig. 2,
3). The intermediate phenotypes were found together with the red morph in Matapalo and
with the green morph in Portalon and Savegre. The polymorphic populations occupy a
remarkably small area of approximately five Km while the whole sampled area was
approximately 110 km long. The southern (red) Charcos, Neotropica, and Ventana were all
located in the Osa peninsula and showed a similar colour pattern with a bright red dorsal
patch beginning at the head and covering 2/3 of the dorsum. The remaining body parts were
covered by a pale blue colour (Fig. 3a). In these three populations 57%, 33%, 55% of frogs
respectively had small red patches on the arms. The colour of the ventrum and limbs varied
from dark to pale blue with black reticulation often present on the dorsal surface and flanks,
but rarely on the ventrum. In Palmar, the frogs were almost completely red on the dorsum and
also usually had extended red patches on the arms and red spots on the proximal portion of the
legs (Fig. 3b). The red phenotype in Firestone was mainly dark red (75%) with less orange
individuals (25%), in Dominical the frogs were mainly orange (70%) while a minority was
red (30%) (Fig. 3c,d,e). In Baru I observed a uniform dark red colour. In Firestone and Baru,
54% and 33% of frogs had red spots on the ventral surface. In Dominical, in all frogs the
entire trunk was red or orange, including in some cases hands and feet and most of the
41

ventrum (33% of frogs). The rest of the ventrum and legs was azure (Firestone, Baru) to azure
grey (Dominical). In the populations Matapalo, Portalon and Savegre we found previously
undescribed colour phenotypes. In Matapalo, brightly red frogs (three out of 20) were found
to coexist with uniformly orange-green individuals (17 out of 20). The intermediate frogs in
this population showed different degrees of saturation, being either darker or lighter (Fig. 3
f,g). In Portalon, I observed highly variable intermediate colour phenotypes (50% of the
frogs) ranging from green with brown patches to red-brown and orange similar to those
identified in Matapalo. Also frogs with three concentric areas of different colour (brown,
yellow and green) were observed. The other half of the observed individuals displayed the
green phenotype (Fig. 3l). In Savegre, more uniform orange to gold-yellow intermediate
phenotypes (25%) were found. In this population, the green morph was most common (75%
of frogs) (Fig. 3i). In Damitas I found mostly green frogs, although four out of 13 individuals
showed a brown patch on the head smaller than the patches observed in Matapalo, Portalon
and Savegre. The ventrum and leg regions ranged from azure to greenazure (Fig. 2l). A
northern green population, not included in the present study, was located near the city of
Parrita outside the known distribution of O. granulifera. Moreover, the altitudal range
occupied by the species is wider than reported (20-100m asl; Savage, 2002; IUCN Red List of
Threatened Species), with populations (Baru, Matapalo, Portalon, Savegre) extending up to
400m asl. The reflectance spectra confirmed the colour intergradation among Firestone and
Dominical. The spectra from Matapalo, Portalon and Savegre showed the presence of a
gradual variation from green to orange (Fig. 4). The colour variation in O. granulifera can be
broadly categorised in the following chromatic areas: red (red), orange (intermediate), orange-
green (intermediate), green-yellow-brown (intermediate), green-brown (intermediate) and
green (green) (Table 2).

42

Table 2 Grouping of O. granulifera colour phenotypes into six categories. Symbols describe
the colour variation occurring in the populations as follow: Red (), Orange (), Orange-
green (), Green-yellow-brown (), Green-brown (), Green ().
Population Colour
Charcos
Neotropica
Ventana
Palmar
Firestone
Baru
Dominical
Matapalo
Portalon
Savegre
Damitas




43

Figure 2 Sampling localities of Oophaga granulifera. The pie-graphs indicate the proportion
of red, intermediate and green phenotypes in each of the populations sampled for the present
study. The populations from Wang (2011) are marked with *. Population codes as in Table 1.






44

Figure 3 Dorsal colour variation in Oophaga granulifera. (a) Osa peninsula [], (b) Palmar
[], (c) Firestone [], (d) Firestone [] , (e) Dominical [], (f,g) Matapalo [] , (h)
Portalon [] , (i) Savegre [], (l) Damitas []. In [ ] colour symbols as in Table 2.
Photographs by Oscar Brusa




45

Figure 4 Average reflectance spectra of O. granulifera from populations (a) Firestone, (b)
Dominical, (c) Matapalo, (d) Portalon, (e) Savegre.



46

3.2 Mitochondrial DNA variation and divergence
There were 17 variable sites and five haplotypes identified in 16S (N = 16) and 44 variable
sites and twelve haplotypes identified in cyt b (N = 30). Both markers identified a southern
(Osa peninsula) and a northern (all remaining populations) genetic lineages. For 16S, the
average genetic distance between the population Charcos (Osa peninsula) and all the other
populations was 1.9% (Table 3). The distances for 16S among the other seven Oophaga spp.
included in the analyses varied between 0.8% and 3.2% with a mean of 1.9%. The average
distance between O. granulifera and the other Oophaga spp. was 5.5%, while the average
distance between Oophaga spp. and the species D. leucomelas and D. tinctorius was 9.2%
and 11.5% respectively (Table 3). Interpopulation genetic distances (p-distance) for cyt b
varied from 0.0 to 4.3% (Table 4). The average genetic difference between southern and
northern populations for the same marker was 4%, while the divergence among the northern
populations varied from 0.1% to 0.8% (Table 4). In the northern lineage, the average genetic
distance between Palmar and the other populations (0.8%) was higher than among the latter
(0.2%). From the AMOVA, most of the variation was among lineages (91.65%, ct = 0.92, P
< 0.001), whereas variation within populations (3.75%, st = 0.97, P < 0.001) and within
southern and northern lineages (4.60%, sc = 0.55, P < 0.001) were much lower.





47

Table 3 16S sequence divergence (uncorrected p-distances) between the two lineages of Oophaga granulifera,
seven other Oophaga spp. and two Dendrobates spp. O.gran. south indicates the O. granulifera southern
lineage; O.gran. n indicates the O. granulifera northern lineage.

Table 4 Cyt b sequence divergence (uncorrected p-distances) among the investigated populations of Oophaga
granulifera. Populations abbreviated as in Table 1.
Species D.leu D.tin O.arb O.his O.leh O.pum O.spe O.syl O.vic O.g. S
D.leucomelas
D.tinctorius 0.062
O.arborea 0.085 0.106
O.histrionica 0.094 0.115 0.008
O.lehmanni 0.092 0.110 0.014 0.014
O.pumilio 0.096 0.119 0.020 0.028 0.030
O.speciosa 0.092 0.108 0.014 0.018 0.020 0.026
O.sylvatica 0.092 0.110 0.012 0.016 0.014 0.032 0.022
O.vicentei 0.096 0.110 0.016 0.018 0.020 0.024 0.018 0.024
O.gran. S 0.109 0.133 0.048 0.054 0.049 0.065 0.059 0.050 0.055
O.gran. N 0.106 0.124 0.049 0.053 0.053 0.070 0.064 0.051 0.059 0.019
Population CH VE PA FS DO BA MA PL SA
Charcos
Ventana 0.004
Palmar 0.036 0.035
Firestone 0.040 0.041 0.007
Dominical 0.041 0.042 0.008 0.001
Baru 0.042 0.043 0.009 0.004 0.005
Matapalo 0.040 0.041 0.007 0.001 0.002 0.004
Portalon 0.042 0.042 0.009 0.002 0.001 0.006 0.003
Savegre 0.041 0.042 0.008 0.001 0.000 0.005 0.002 0.001
Damitas 0.041 0.042 0.008 0.001 0.000 0.005 0.002 0.001 0.000
48

3.3 Mitochondrial DNA phylogeography and population structure
Maximum-likelihood (ML) and Bayesian analyses revealed a deep genetic separation of the
southern populations, Charcos and Ventana, from all the other populations for both
mitochondrial markers (Fig. 5). The phylogeny reconstructed with 16S sequences indicated a
clear differentiation between the two lineages of southern and northern O. granulifera with
100% node support for both ML and Bayesian trees. I also found a higher level of divergence
of the species compared with the other Oophaga spp. (Fig. 5a). Similarly, in the cyt b tree the
node separating the Osa peninsula from the other populations was highly supported by
bootstrap and posterior probability values (ML = 98, BI = 1.00, Fig. 5b). The cyt b tree also
showed moderate support for a derived monophyletic clade including Firestone, Dominical,
Baru, Matapalo, Portalon, Savegre and Damitas populations. Within this clade, the divergence
of two haplotypes from Baru and two from BR was supported by significant values of
bootstrap and posterior probability (ML = 86, BI = 0.99, Fig. 5b). Despite a moderately
supported divergence from the other northern populations, Palmar was included in the
northern lineage (Fig. 5b). The haplotype network (cyt b) confirmed the genetic separation of
the populations located south and north of the river Ro Trraba (Fig. 6), with the two clusters
being separated by a minimum of 20 mutational steps. Eleven out of 16 haplotypes (seven in
the northern and four in the southern lineage) were private. Within the northern lineage,
haplotypes are shared among intermediate and monomorphic red or green populations except
for haplotypes H11 and H12 belonging to Portalon and Matapalo respectively (Fig. 6).






49

Figure 5 Phylogenies obtained analysing (a) 16S and (b) cyt b sequences. The ML bootstrap
values and Bayesian posterior probabilities are reported at nodes. For each O. granulifera
sequence we indicated: the species name, the population code (as in Table 1) and sample
number and in (b) the haplotype code (H1 to H16); Wang indicates the sequences from
Wang (2011) included in the analyses; font colours correspond to O. granulifera colour
morphs as follows: red = red morph, yellow = intermediate morphs, green = green morph. S =
southern lineage, N = northern lineage. * indicates sequences from Palmar, ** indicate
sequences from TE.
(a) (b)


50

Figure 6 Network of cyt b haplotypes of Oophaga granulifera. Haplotypes (H) are indicated
by circles whose areas correspond to the haplotype frequency; S and N indicate southern and
northern lineages; the dashed double arrow shows the minimum separation between lineages.
Correspondences between colours and populations are indicated in figure; * distinguishes the
populations sampled for this study (coded as in Table 1) from the populations sampled in
Wang (2011) included in the analyses (Fig. 1).

51

3.4. Microsatellite loci population structure
All microsatellite loci except one were polymorphic for the eleven O. granulifera populations.
Locus D4 was monomorphic in the three populations sampled south of the river Ro Trraba
(Charcos, Ventana and Neotropica). No deviation in linkage disequilibrium was found (P >
0.05). In eight populations one to two loci deviated from HWE after Bonferroni correction
(Table 5). Within-population diversity was overall higher in southern populations than in
northern ones (Table 5). Allelic richness ranged from 5.11 ( 2.00) in Matapalo to 8.00 (
3.27) in Ventana and was higher in the southern populations and Palmar (6.97 on average)
than in the northern populations (5.12). Private alleles were mainly found in the three
southern populations, Charcos, Ventana and Neotropica, as well as in Palmar sampled just
north of the river Terraba (9.5 on average), while they were scarce in other northern
populations (2.3 on average). Significant F
ST
values (P < 0.05) were detected in 52 out of 55
pairwise comparisons among populations (Table 6). Mean F
ST
value was 0.11 ( 0.05),
ranging from 0.01 (between Baru and Firestone) to 0.18 (between southern population
Charcos and the northernmost green population Damitas). Population of the northern mtDNA
lineage had higher average F
ST
value (0.11 0.05) than those belonging to the southern
lineage (0.05 0.01). AMOVA results markedly differed from those with the mitochondrial
cyt b marker. Most of the variation was found within populations (84.53%, F
ST
= 0.15, P <
0.01), whereas variation among lineages (9.63%, F
CT
= 0.10, P < 0.001) and among
population within lineages (5.84% F
SC
= 0.06, P < 0.001) were much lower. Bayesian
assignment analysis identified two well defined clusters (Fig. 7; see Additional Information
for K results). The first one included the Osa Peninsula populations, while the second
grouped all the populations belonging to the northern mitochondrial lineage except Palmar.
Palmar presented an intermediate population structure having genotypes similar to both the
southern and the northern clusters (Fig. 7). The differentiation of Palmar from Charcos,
Neotropica, Ventana grouped in the southern cluster is confirmed by high pairwise Fst values
52

(Table 6). One individual from the northern population Dominical showed a genetic profile
close to the southern populations. Admixture was high within the clusters and low among
clusters (Fig. 7).

Table 5 Analyses of seven microsatellite loci of Oophaga granulifera; n = number of
individuals analysed; A
R[14]
, mean allelic richness ( SD) estimated for a minimum of seven
individuals; H
o
, observed heterozygosity ( SD); H
e
, expected heterozygosity ( SD); HWE,
loci out of HardyWeinberg equilibrium after Bonferroni correction; A
P
, number of private
alleles.
Population n A
R[14]
H
o
H
e
HWE A
P

Charcos 8 6.95 3.07 0.71 0.42 0.72 0.32 D4, G5 6
Neotropica 8 6.62 3.15 0.75 0.37 0.71 0.32 D4, G5 8
Ventana 7 8.00 3.27 0.71 0.35 0.73 0.32 D4 15
Palmar 7 6.29 2.75 0.55 0.21 0.73 0.14 O1 9
Firestone 8 4.92 2.06 0.66 0.21 0.66 0.16 1
Dominical 8 5.12 1.75 0.61 0.15 0.67 0.13 F1 2
Baru 8 4.83 1.25 0.61 0.25 0.66 0.16 O1
Matapalo 10 5.11 1.99 0.70 0.23 0.65 0.19 3
Portalon 15 5.36 2.15 0.69 0.18 0.73 0.14 O1 1
Savegre 8 5.40 2.17 0.71 0.20 0.68 0.15
Damitas 11 5.24 2.76 0.64 0.29 0.67 0.24 C3 1

Table 6 Differentiation between Oophaga granulifera populations estimated by pairwise F
ST

values. (*) Significant F
ST
values (P < 0.05). Population codes (Pop) as in Table 1.
Pop CH NE VE PA FS DO BA MA PL SA DA
CH
NE 0.05*
VE 0.04* 0.06*
PA 0.14* 0.16* 0.13*
FS 0.17* 0.16* 0.16* 0.11*
DO 0.15* 0.16* 0.14* 0.14* 0.10*
BA 0.15* 0.13* 0.15* 0.10* 0.01 0.09*
MA 0.17* 0.17* 0.17* 0.13* 0.07* 0.11* 0.07*
PL 0.13* 0.12* 0.14* 0.11* 0.07* 0.04* 0.05* 0.04*
SA 0.14* 0.12* 0.12* 0.07* 0.05* 0.06* 0.03 0.06* 0.03*
DA 0.18* 0.16* 0.18* 0.11* 0.06* 0.08* 0.06* 0.06* 0.05* 0.02
53

Figure 7 Assignment probabilities of individuals of Oophaga granulifera from Costa Rica to
putative population clusters at K=2 using the program STRUCTURE. Populations codes (as
in Table 1, font colours as in Fig.5).


3.5 Bioacoustics
Call duration, call rate and dominant frequency were significantly different between southern
and northern lineages, whereas there was no difference in pulse rate (Table 7). In the southern
genetic lineage frogs had longer and slower calls and higher dominant frequency as compared
to the northern lineage (Table 8; Fig. 8). Variability among populations was highest for call
duration while it was lowest for dominant frequency (Table 9). The calls of Charcos,
Neotropica and Firestone showed the lowest variation. The calls of the males from Portalon
and Savegre were the most variable (Table 9). Within the northern lineage, pulse rate
significantly differed between the red and green colour morphs (F = 10.16, P < 0.001).




54

Table 7 ANOVA results testing for differences in call parameters between southern and
northern populations of Oophaga granulifera. Abbreviations for call parameters: PR = pulse
rate, CD = call duration, CR = call rate, DF = dominant frequency. * indicates significant
values.
Variable F P
PR (1s
1
) 3.14 0.082
CD (s) 70.11 <0.0001*
CR (1s
1
) 6.36 0.015*
DF (kHz) 6.36 <0.0001*


Table 8 Male advertisement call parameters in the Oophaga granulifera populations. N =
number of individuals recorded. Abbreviations for call parameters: PR = pulse rate, CD = call
duration, CR = call rate, DF = dominant frequency. All values given as mean SD.

Population N PR (1s
1
) CD (s) CR (1s
1
) DF( kHz)
Charcos 9 174.1417.06 0.530.08 10.370.95 4.210.19
Neotropica 14 165.9914.97 0.570.08 9.931.21 4.200.15
Ventana 7 203.5936.38 0.440.10 11.701.83 4.320.16
Palmar 3 168.7626.73 0.280.07 16.852.07 4.070.28
Firestone 4 172.4918.74 0.330.03 14.601.40 4.150.08
Baru 4 181.5122.85 0.310.1 16.332.12 4.160.18
Matapalo 3 209.1349.36 0.360.09 15.752.36 3.970.11
Portalon 3 211.3724.16 0.260.11 17.672.03 3.980.22
Savegre 5 221.9833.71 0.320.13 14.271.80 4.270.06


55

Figure 8 Variation in male advertisement call parameters among southern and northern
populations of Oophaga granulifera: (A) call duration ; (b) call rate; (c) dominant frequency;
(d) pulse rate
(A)


(B)

56

(C)



(D)


57

Table 9 Coefficients of variation [(mean/standard deviation)*100] for the male advertisement
call parameters. pop = mean across populations. Abbreviations for call parameters: PR =
pulse rate, CD = call duration, CR = call rate, DF = dominant frequency.

CV
PR
CV
CD
CV
CR
CV
DF
pop
Charcos 9.8 14.5 9.1 4.6 9.6
Neotropica 9.0 13.5 12.2 3.7 9.4
Ventana 17.9 23.2 15.6 3.6 14.6
Palmar 15.8 24.3 12.3 6.9 15.1
Firestone 10.9 7.9 9.6 1.8 7.9
Baru 12.6 31.9 13.0 4.3 16.1
Matapalo 23.6 24.2 15.0 2.9 15.7
Portalon 11.4 41.5 11.5 5.5 19.7
Savegre 15.2 40.6 12.6 1.3 19.9

par

14.0

24.6

12.3

3.8

16.3



3.6 Morphology
The parameters weight, body length, femur length, tibia length, humerus length and radio-
ulna length significantly differed between southern and northern genetic lineages (Table 10;
Fig. 9). Frogs belonging to the red southern populations were lighter, smaller and with shorter
limbs than frogs from red, intermediate and green northern populations. Among the northern
populations, head length (F = 7.99; P < 0.0001), head width (F = 24.74; P < 0.0001),
tympanum diameter (F = 9.68; P < 0.0001) and between-nostril distance (F = 6.98; P <
0.0001) together with body weight (F = 15.57; P < 0.0001) and body length (F = 9.34; P <
58

0.0001) were significantly different between red and green colour morphs. In the northern
lineage, red frogs were heavier, longer and had an overall larger head compared to the green
ones.

Table 10 ANOVA comparisons between genetic lineages for ten morphological parameters. *
indicates significant values






Variable F P
weight F= 21.66 P < 0.0001*
body length F = 20.45 P < 0.0001*
femur length F = 28.61 P < 0.0001*
tibia length F = 66.75 P < 0.0001*
humerus length F= 7.15 P < 0.0001*
Radio-ulna
length
F= 0.42 P < 0.0001*
head length F= 0.13 P = 0.71
head width F = 1.19 P = 0.28
tympanum
diameter
F=0.253 P = 0.14
Between-nostril
distance
F = 0.33 P = 0.56
59

Figure 9 Box plots for ten morphological variables comparing southern and northern genetic
lineages in O. granulifera. (A) body weight, (B) body (snout-vent) length, (C) tibia, (D)
femor, (E) humerus, (F) radio-ulna, (G) head-width, (H) head length, (I) tympanum diameter,
(L) between nostril-distance.

(A)






60

(B)


(C)

61

(D)


(E)

62

(F)


(G)

63

(H)


(I)

64

(L)








65

4. Discussion

4.1 General discussion
The fine-scale analysis of genetic and phenotypic variation conducted in the present study
revealed significant geographical divergence in genetics, colouration, advertisement calls and
body size, in Oophaga granulifera populations in Costa Rica. Specifically, the results here
presented show that O. granulifera possesses higher colour diversity than previously
described (Savage, 2002; Wang, 2011). Beside the red, yellow and green colour morphs, the
species shows an array of intermediate morphs and is polymorphic in the northern part of its
distribution. Most of this remarkable variation occurs in a small transition area between
populations monomorphic for red and green phenotypes. To the north of the river Ro
Trraba, first the colour phenotype changes quantitatively following a well defined south-
north cline toward increase of the red body area. Then the colour phenotype differentiates into
an array of colours spanning the chromatic range between red and green. In contrast, all
populations occurring in the Osa peninsula in the south of the distribution are very consistent
in colour characteristics. I provide the first description of colour polymorphic populations of
O. granulifera. The three polymorphic populations consisted of yellow morphs together with
green or red and other novel intermediate phenotypes. Thus, in contrast to previous studies
(Savage, 2002; Wang, 2011), intermediate frogs were always found in polymorphic
populations. The distribution of colour morphs in O. granulifera is asymmetric compared to
the genetic structure resulting from the phylogenetic reconstruction. The southern lineage is
monomorphic red while the northern lineage contains red, green and polymorphic
populations. Within the northern lineage, a monophyletic clade includes all of the northern
populations except Palmar. In this clade, Firestone, Dominical, Portalon, Matapalo, Savegre
and Damitas are closely related while Baru appears to be more differentiated. In contrast to a
66

previous study (Wang, 2011), I did not find a monophyletic yellow-green clade, which
presumably relates to the more comprehensive sampling in the colour transition area
performed in the present study. A missing link between colour and genetic differentiation in
poison frogs has been previously observed in O. pumilio (Hagemann & Prhl, 2007; Wang &
Shaffer, 2008; Brown et al., 2010; Hauswaldt et al, 2011; but see Wang & Summers, 2010).
A recent colour radiation might explain the lack of correspondence between genetic
divergence at mitochondrial loci and colour divergence. According to our phylogeny the
Palmar population belongs to the northern lineage while the TE population from Wang (2011)
belongs to the southern one. Interestingly, the sample location was north of the river Ro
Trraba while Wang (2011) sampled south of it (Ian Wang, pers. comm.; in the cited study the
name of the river was misspelled as Rio Terribe). The River Ro Trraba is the largest of
Costa Rica and has been estimated to have originated in the Quaternary period (Bergoeing, et
al., 1997). The historical presence of such an important physical barrier may well present the
major determinant of the observed inter-lineage divergence. Considering a mutation rate for
mitochondrial DNA of approximately 2% per million years (Avise, 2000), the genetic
distance I observed between lineages (4% for cyt b) is congruent with a role of the river Ro
Trraba in the O. granulifera intra-specific divergence. The analyses of 16S data confirms the
divergence between the southern and northern lineages and show that the intraspecific genetic
distance in O. granulifera is comparable to the distance between different Oophaga species.
The data therefore show, using a much larger sampling than previous studies (nine
populations instead of one), that O. granulifera is highly divergent from all congeneric
species (Hagemann & Prhl; Summers & Clough, 2000). The divergence into two main
lineages is furthermore supported by the haplotype network that defines two clusters
corresponding to the southern and northern lineages. Deep intraspecific genetic divergence
has been found in other dendrobatids. For instance, high divergence in cyt b haplotypes is
reported for Epipedobates femoralis and E. hahneli from Western Amazonia (12% and 7.1%,
67

Lougheed et al., 1999; Roberts et al., 2007), and in O. pumilio from Costa Rica and Panama
(7%, Hauswaldt et al., 2011). Nonetheless, the present study is the first providing concordant
lines of evidence for a deep inter-specific divergence, using historical and recent genetic
inference and phenotypic traits. Notably, the two O. granulifera lineages diverged both in
advertisement calls and in morphological traits and present highly reduced gene admixture.
The frogs of the Osa peninsula are smaller and have longer notes (calls), lower call rate and
higher dominant frequency. Thus, the two lineages have undergone divergent evolutionary
trajectories leading to a deep divergence in which reinforcement mechanisms, such as
assortative mating, were presumably involved in impeding hybridization. The opposite
geographical trend has been found in O. pumilio in Costa Rica, in which two southern and
northern genetic lineages roughly correspond to two bioacoustic groups, and the southern
frogs were bigger and had higher call rate (Prhl et al., 2007; Hagemann & Prhl, 2007;
Hauswaldt et al., 2011). This may be explained if stochastic genetic processes, linked for
example to demographic fluctuations during colonization movements, are important factors in
originating intraspecific divergence, secondarily maintained and increased by selection.
The microsatellite STRUCTURE analysis evidences the presence of two well distinguished
southern and northern genetic clusters corresponding to the mitochondrial lineages with the
exception of the population Palmar which has an intermediate genetic structure sharing
genotypes from both clusters (Fig. 7). This is likely to be dependent on the peculiar location
of this population which is in between the southern and the northern populations (Fig. 2). In
particular, the pattern may be explained considering the geographical proximity between the
populations occurring in the Ro Trraba area (Palmar and TE) together with changes in the
water level and position of the river that may have occurred over time. Overall, the genetic
data show a deep intraspecific divergence between populations located south and north of the
river Ro Trraba and indicate the populations of the river basin as historical link between the
two lineages. The differentiation I observed in both calls and morphology between red,
68

intermediate and green colour morphs points to a potential role of mate choice in Oophaga
granulifera colour divergence. Interestingly, in the colour transition area, the majority of the
intermediate phenotypes co-occur with the green phenotype (Portalon, Savegre).
Both genetic and colour geographical variations in O. granulifera do not follow a clear cline,
considering the admixture among populations within the northern cluster and the asymmetric
distribution of intermediate phenotypes in the colour transition area. The lack of a clear colour
cline and a genetically well defined hybrid area may be explained if some of the intermediate
colour phenotypes in O. granulifera are not hybrids between red and green morphs, but
evolutionary steps from the red to the green phenotype. This would explain the fact that we
did not observe any population where pure red and green morphs co-occur. The observed
pattern in colour variation may be the result of complex crossing between red, intermediates
and green morphs over time across the colour transition area.
Wang (2011) stated that the O. granulifera cryptic morph was derived from the bright red
morph. In my opinion Wang (2011) and our phylogeographic reconstruction are congruent
with both a direct evolution of the green morph from a red ancestor and an indirect evolution
via intermediate steps. The green morph might have evolved from the red one and all the
intermediate morphs found in the transition area would be the result of hybridization between
red and green morphs or, alternatively, the intermediate morphs evolved from the red and
subsequently diverged to the green morph. In the second case, the intermediate colours
observed in the transition area would be a mixture of pure phenotypes and hybrids. The
coexistence of bright and dull colour morphs we found in the O. granulifera northern lineage
was also observed in the O. pumilio southern lineage (Hagemann & Prhl, 2007; Wang &
Shaffer, 2008; Hauswaldt et al., 2011). These colour phenotypes have different
conspicuousness against the natural background (Prhl & Ostrowski, 2011; Wang, 2011),
supporting the hypothesis that aposematic and cryptic colour morphs can coexist in systems
primarily evolved as aposematic (Darst et al., 2006; Speed & Ruxton, 2007). Several
69

evolutionary factors have been considered to explain the puzzling colour divergence in poison
frogs: genetic drift, differential predation pressure, toxin availability and sexual selection
(Speed & Ruxton, 2007; Rudh et al., 2007; Summers et al., 1997, 1999; Saporito et al.,
2007a). Interestingly, I found that the polymorphic and green populations show reduced body
size. This is congruent with the evolution of a cryptic antipredator strategy based on lack of
detectability to the predator instead of conspicuous signalling (Hagman & Forsman, 2003). In
addition, the lower genetic variation in the northern populations as revealed by microsatellite
allele diversity points to a possible role of genetic drift in the evolution of colour divergence
in O. granulifera. The effect of stochastic variation on allele frequency was advanced to
explain the evolution of isolated populations (Wright, 1977, Nosil, 2007). Examples have
been reported in O. pumilio and Heliconius butterflies (Rudh et al., 2007; Mallet & Joron,
1999). The presence of highly variable gradual intermediate colour morphs in O. granulifera
and the high population admixture in the northern lineage suggest that a certain amount of
gene flow between colour morphs exists in the colour transition area. Consequently, the
colour divergence within the northern lineage implies the maintenance of selection. Female-
biased parental care as observed in the Oophaga group may have increased female choosiness
and created the conditions for sexual selection through assortative mating by visual and
acoustic cues (Summers et al., 1997). Since male vocalizations are key sexual signals in frogs,
the difference in the pulse rate of the male advertisement call we found between red and green
colour morphs suggest the potential occurrence of female mate choice. The divergence in
body size, may also suggest a role of sexual selection in intraspecific divergence in O.
granulifera. In fact, assortative mating mediated by body size is demonstrated in amphibians
(Berven, 1981; Gabor et al., 2000; Takahashi et al., 2010). Reduced body size may therefore
have originated by natural selection for increased crypsis and be maintained by differential
mating selection in the colour transition area. Empirical tests in the field using mate choice
70

experiments are required to evaluate the actual role of sexual selection in the evolution and
maintenance of colour diversity in O. granulifera.


4.2 Why should a cryptic morph evolve in an aposematic species?
4.2.1 General background
Colour pattern polymorphisms can arise and be maintained by selection favouring multiple
variants within populations (Hoffman & Blouin, 2000; Merilaita, 2006; Vercken et al., 2007;
Dijkstra et al., 2008). Nonetheless, an aposematic species is a peculiar case in which the
evolution of colour diversification is predicted to be highly constrained (Mller, 1879).
Particularly, the possible adaptive role of a cryptic colour variant in an aposematic system is
debated. In fact, this is in constrast with the aposematism theory which predicts the functional
association of toxicity and conspicuous colour to enhance the efficancy of the deterrent signal
to predators. In non toxic species, a cryptic colouration can facilitate predator avoidance
(Cott, 1940; Endler, 1992; Duellman, 2001) therefore being under natural selection
(Robertson, 2008) or can be involved in mate choice (Gomez & Thery, 2007). Laboratory
studies have tested predation rates as a function of colour pattern in polymorphic frogs and
found that predators attacked cryptic (background-matched) morphs less often (Tordoff, 1980;
Morey, 1990). However, for toxically defended anurans such as poison-arrow frogs, bold
colour patterns and overall conspicuousness are thought to protect individuals from predation
attempt and detrimental effects connected (Summers & Clough, 2001; Darst et al., 2006).
Disentangling the selective forces responsible for colour diversification in aposematic
dendrobatids is difficult because several different and potentially interacting processes could
underlie divergent phenotypic expression, including natural (Hairston, 1979; Endler, 1980)
and sexual selection (Endler, 1983; West-Eberhard, 1983; Panhuis et al., 2001; Masta &
71

Maddison, 2002; Maan et al., 2004). I discuss here those among these selective forces that are
considered the most significant and how these may apply to O. granulifera.

4.2.2 Role of genetic drift
Stochastic fluctuations in allelic frequencies have been proposed to play a role in the
evolution of animal populations (Wright, 1933). Recently the actual role of drift has been re-
evaluated and its potential evolutionary effects are not considered of general application.
Nonetheless, under particular conditions such as population reduction and isolation the effect
of drift may be significant and shaping the evolutionary trajectories taken by those
populations (Lande, 1977). This may be the case, for instance, for the colour diversity
observed in the Panamanian populations of O. pumilio, which are small and separated (Rudh
et al., 2007). In O. granulifera, population size reduction may have happened during the
colonization movements toward the northern part of the distribution. The data concerning
microsatellite allelic diversity here presented suggest a decrease in genetic diversity in the
northern intermediate and green populations, concordant with an effect of drift. Drift may
have, for example, determined fluctuations in the frequency of the red aposematic morph
creating the conditions for a colour diversification in combination with other factors. It is
interesting at this regard, considering that theoretical models predict an association between
the frequency of the aposematic morph in the population and the promptness with which
predators learn to avoid it (Speed, 1993; Turner & Speed, 1996; MacDougall & Dawkins,
1998; Servedio, 2000; Speed & Ruxton, 2007).

4.2.3 Role of predation
In aposematic toxic animals, it has been shown that conspicuousness, that is a highly
contrasted appearance to the background, can make individuals more vulnerable to attacks by
nave (e.g., Lindstrm et al., 2001a,b; Riipi et al., 2001; Lindstedt et al., 2008) or specialized
72

predators (Yosef & Whitman, 1992). Moreover, the predators visual and learning abilities and
resistance to prey toxicity, may vary both within and among species determining locally
variable predation risk for a potential aposematic prey (Calvert, 1979; Fink & Brower, 1981;
Yosef & Whitman, 1992; Pinheiro, 1996; Mappes & Alatalo, 1997; Marples et al., 1998;
Exnerov et al., 2003, 2007). Another factor to take into consideration is the ecological
community to which the aposematic animals belong to. Also, the vulnerability to predation
depends on the amount of alternative prey available (Merilaita & Kaitala, 2002; Lindstrm et
al., 2004) and the hunger level of predators (Barnett et al., 2007; Sandre et al., 2010). Overall,
predation could select for intermediate signals in some circumstances (Endler & Mappes,
2004; Tullberg et al., 2005; Ruxton et al., 2009) or allow for signal variation in others (Rowe
et al., 2004; Darst et al., 2006; Ham et al., 2006; Ihalainen et al., 2007). Generally, predators
are thought to learn to avoid unpalatable prey faster if they are conspicuous rather than cryptic
(Gittleman & Harvey, 1980; Roper & Redston, 1987; Alatalo & Mappes, 1996; Lindstrm et
al., 1999b) and if they are clearly visible (highly contrasted) against their background (Roper
& Redston, 1987; Lindstrm et al., 1999b; Gamberale-Stille, 2001; Prudic et al., 2007). Many
typical warning colours (e.g. red, orange or yellow) may also be innately aversive to predators
in comparison with common cryptic colours, such as green and brown, which can increase the
benefits of being warning coloured (Smith, 1975; Schuler & Hesse, 1985). Despite the
advantages of clearly visible warning colours (Guilford, 1986; Gamberale-Stille, 2000;
Schlenoff, 1984; Marples et al., 1998; Thomas et al., 2003; Roper & Redston, 1987; Roper,
1990; and Sherratt & Beatty, 2003; Ham et al., 2006), intermediate warning signals may be
selected instead of conspicuous ones if the predation risk varies (Endler & Mappes, 2004). In
dendrobatids the importance of predation as a driving force in the evolution of colour
diversity has been recently evidenced using empirical tests to assess the actual association of
colour phenotype and predation pressure (Comeault & Noonan, 2011, Saporito et al., 2007b).
The efficacy of an aposematic defense is dependent on the frequency of the conspicuous
73

colour morph characterizing the system (Speed, 1993; Turner & Speed, 1996; MacDougall &
Dawkins, 1998; Servedio, 2000; Lindstrm et al., 2001b; Speed & Ruxton, 2007), because a
high frequency is necessary to elicit the avoidance learning by the predators increasing the
deterrent efficacy of the signal. In O. granulifera, as mentioned, a demographic reduction may
be related to a decrease in genetic variability in the northern green populations. This together
with a colonization movement south-north (Wang, 2011) point to possible demographic
fluctuations linked to the occupancy of new habitats. If, following genetic drift phenomena,
the frequency of the red morph decreased in the O. granulifera northern populations, an
increased predation pressure on the that conspicuous morph may have produce a positive
selection for a cryptic variant. Moreover, the less conspicuous behaviour observed in the
green morph (personal observation) is congruent with an overall cryptic strategy of the green
morphs in O. granulifera. Importantly, to allow a significant evaluation of the role of
predation in the evolution of colour diversity in dendrobatids, the actual level of toxicity of
cryptic and conspicuous morphs has to be taken into consideration (Maan & Cummings,
2012).

4.2.4 Role of toxicity
The presence of chemical defences is a fundamental trait of aposematic defences as the
predators learn to avoid a conspicuously colours animals due to the unpalatability and adverse
effect provoked by toxins (Muller, 1879). Therefore, the possible variation of the actual
toxicity in a population may influence the selection pressure at the local level and create
potential for the diversification in aposematic signals (Wang, 2011). It has been proposed that
once aposematism is established in a system, then colouration can become decoupled and
diversifies following local variation in toxicity. There is evidence that in O. pumilio, the dull
coloured less constrasted (cryptic) morphs are significantly less toxic than the bright coloured
morphs (Maan & Cummings, 2012). This finding is in contrast with Wang (2011) for O.
74

granulifera that showed an overall higher toxicity in green and intermediate morphs. These
two apparently contrasting findings may be explained considering a crucial point in
dendrobatid chemical defences that is the dependence on the diet as a source of toxins. In fact,
spatial and temporal variation in the prey communities from which dendrobatids derive their
alkaloids must be taken into consideration before depicting any related evolutionary scenario.
In the two mentioned studies these longterm observational data are missing therefore any
deduction has to be careful (see Saporito et al., 2007c). Another factor that has to be
thoroughly weighed is the methodology that has been followed up to date to evaluate the
toxicity of dendrobatids, which is an intradermal injection of frog skin extract in laboratory
mice. Both the species used and the way of assumption may not be optimal representation of
the real processes as they happen in nature. Nonetheless, the difference in toxicity was
quantitative (toxicity essays; both studies) and qualitative (number of identified alkaloids;
Wang, 2011) and is consistent across different cryptic morphs therefore it may represent an
important factor in the evolution of colour diversity in aposematic species. The possible
reason explaining the constrasting pattern observed in O. pumilio and O. granulifera may be
dependent on the different biogeographic conditions in which the two systems evolved.
Indeed, most of O. pumilio colour diversity is found in the Bocas del Toro archipelago in
Panama, formed by many different small islands of recent origin while O. granulifera is
found in mainland Costa Rica in a potentially continuous distribution. In O. granulifera, the
effect of an increased toxicity may have reduced the need for a bright colouration over time.
This is because, predators can learn to avoid a highly toxic not profitable prey even if dully
coloured. The related evolutionary mechanism is compatible with individual selection because
any predation attempt may be neutralized if powerful toxins cover the body surface, such as in
the case of dendrobatids, determining immediate repellent and noxious effects to the predator.
Evolving a dull colouration, the potential costs involved in the bright pigments production are
avoided. Several factors will have to be clarified to test the feasibility of this proposed
75

scenario. Other than the already mentioned possible fluctuations in toxicity and the
questionable ecological significance of the toxicity essay used, information on the pigments
involved in dendrobatids pigmentation are necessary to evaluate the actual cost of pigment
production.

4.2.5 Role of sexual selection
As discussed before, together with the role of natural selection mediated by predators, sexual
selection has been suggested to play an important role in the diversification of aposematic
signals in poison-dart frogs. Females show preference for brighter (Maan & Cummings, 2009)
and more colourful males (Maan & Cummings, 2008) as well for males of their own
population in O. pumilio (Summers et al., 1999; Reynolds & Fitzpatrick, 2007; Maan &
Cummings, 2008). I showed here, that aposematic red and cryptic green morphs have a
different pulse rate in the male advertisement call in the northern populations. A significant
differentiation in this key sexual signal indicates that the females may potentially select males
of their own colour, similarly with what has been observed in the congeneric species O.
pumilio.






76

5. Conclusion

The present work provides evidence for deep intraspecific divergence in the polymorphic
aposematic poison-dart frog Oophaga granulifera. I described the patterns of geographical
variation in historical and recent genetic structure and in key phenotypic traits such as
morphology and advertisement call. The data here presented all congruently show that the
southern and northern lineages identified are clearly diverging and may well represent
different species. The phylogeographic reconstruction and the recent population structure
indicate that the river Ro Trraba is likely to be the geographic barrier between the lineages
that are, therefore, geographically, genetically and phenotypically separated.
As concerned to the recent evolution of colour diversity in O. granulifera, several factors may
potentially be involved in the origin and maintenance of this notable differentiation.
Alongside correlation data such as those here presented future research efforts will be
necessary to empirically evaluate the effective strength in mating preferences shaping the
observed variation in colouration. Moreover, many aspects of dendrobatid ecology, such as
actual predator community, prey community and related spatial and temporal variation in the
acquired toxicity, will have to be clarified to define a clearer evolutionary scenario explaining
the reason of the remarkable colour variation in this aposematic vertebrate.




77

6. References

Aichinger, M. (1991) Tadpole transport in relation to rainfall, fecundity and body size in five
species of poison-dart frogs from Amazonian Peru. Amphibia-Reptilia, 12, 4955.
Alatalo, R.V. & Mappes, J. (1996) Tracking the evolution of warning signals. Nature, 382,
708709.
Andersson, S., rnborg, J. & Andersson, M. (1998) Ultraviolet sexual dimorphism and
assortative mating in blue tits. Proceedings of the Royal Society of London series B, 265,
445450.
Arak, A. (1983) Sexual selection by male-male competition in natterjack toad choruses.
Nature, 306, 261262.
Arntzen, J.W. & Wallis, G.P. (1999) Geographic variation and taxonomy of crested newts
(Triturus cristatus superspecies): morphological and mitochondrial DNA data.
Contributions to Zoology, 68, 181203.
Auber, L. (1957) The distribution of structural colours and unusual pigments in the class
Aves. Ibis, 99, 463476.
Baeta, R., Faivre, B., Motreuil, S., Gaillard, M. & Moreau, J. (2008) Carotenoid trade-off
between parasitic resistance and sexual display: an experimental study in the blackbird
(Turdus merula). Proceedings of the Royal Society series B, 275, 427434.
Bailey, R.I., Thomas, C.D. & Butlin, R.K. (2004) Premating barriers to gene exchange and
their implications for the structure of a mosaic hybrid zone between Chorthippus brunneus
and C. jacobsi (Orthoptera: Acrididae). Journal of Evolutionary Biology, 17, 108119.
Barlow, G.W. (1961) Causes and significance of morphological variation in fishes.
Systematic Zoology, 10, 105117.
78

Barluenga, M., Stlting, K.N., Salzburger, W., Muschick, M. & Meyer, A. (2006) Sympatric
speciation in Nicaraguan crater lake cichlid fish. Nature, 439, 719723.
Barnett, C.A., Bateson, M. & Rowe, C. (2007) State-dependent decision making: educated
predators strategically trade off the costs and benefits of consuming aposematic prey.
Behavioural Ecology, 18, 645651.
Barton, N. (2004) Speciation: Why, how, where and when? Current Biology, 15, R603R604.
Beatty, C.D., Beirinck, K. & Sherratt, T.N. (2004) The evolution of Mllerian mimicry in
multispecies communities. Nature, 431, 6367.
Benson, W. (1972) Natural selection for Mullerian mimicry in Heliconius erato in Costa Rica.
Science, 176, 936939.
Berven, K.A. (1981) Mate choice in the wood frog, Rana sylvatica. Evolution, 35, 707722.
Berven, K.A. (1982) The genetic basis of altitudinal variation in the wood frog Rana
sylvatica. An experimental analysis of life history traits. Evolution, 36, 962983.
Bezzerides, A.L., McGraw, K.J., Parker, R.S. & Husseini, J. (2007) Elytra color as a signal of
chemical defense in the Asian ladybird beetle Harmonia axyridis. Behavioral Ecology and
Sociobiology, 61, 14011408.
Blair, W.F. (1958) Mating call in the speciation of anuran amphibians. American Naturalist,
92, 2751.
Blount, J.D., Metcalfe, N.B, Birkhead, T.R & Surai, P.F. (2003) Carotenoid modulation of
immune function and sexual attractiveness in zebra finches. Science, 300, 125127.
Blount, J.D., Speed, M.P, Ruxton, G.D. & Stephens, P.A. (2009) Warning displays may
function as honest signals of toxicity. Proceedings of the Royal Society series B, 276, 871
877.
Bolaos, F. (1990) Actividad de canto y territorialidad en Dendrobates granuliferus. Tesi de
maestria, Universidad de Costa Rica.
79

Brakefield, P.M. & Liebert, T.G. (1985) Studies of colour polymorphism in some marginal
populations of the aposematic jersey tiger moth Callimorpha quadripunctaria. Biological
Journal of the Linnean Society, 26, 225241.
Bossuyt, F. & Milinkovitch, M.C. (2000) Convergent adaptive radiations in Madagascan and
Asian ranid frogs reveal covariation between larval and adult traits. Proceedings of the
National Academy of Sciences USA, 97, 65856590.
Bourne, G.R., Collins, A.C., Holder, A.M. & McCarthy, C.L. (2001) Vocal communication
and reproductive behavior of the frog Colostethus beebei in Guyana. Journal of
Herpetology, 35, 272281.
Brown, J., Maan, M.E., Cummings, M.E. & Summers, K. (2010) Evidence for selection on
coloration in a Panamanian poison frog: a coalescent-based approach. Journal of
Biogeography, 37, 891901.
Caldwell, J.P. (1997) Pair bonding in spotted poison frogs. Nature, 385, 211214.
Caldwell, J.P., & de Oliveira, V.R.L. (1999) Determinants of biparental care in the spotted
poison frog, Dendrobates vanzolinii (Anura: Dendrobatidae). Copeia, 3, 565575.
Calvert, W.H. (1979) Mortality of the monarch butterfly (Danaus plexippus L.): avian
predation at five overwintering sites in Mexico. Science, 204, 847851.
Camp, D.C., Marshall, J.L., Austin, R.M. Jr. (2000) The evolution of adult body size in black-
bellied salamanders (Desmognathus quadramaculatus complex). Canadian Journal of
Zoology, 78, 17121722.
Campbell, J. (1999) Distribution patterns of amphibians in Middle America. In: Patterns of
Distribution of Amphibians (ed. Duellman). The John Hopkins University Press,
Baltimore, Maryland.
Campbell, J.A. & Frost, D.R. (1993) Anguid lizards of the genus Abronia: revisionary notes,
descriptions of 4 new species, a phylogenetic analysis, and key. Bulletin of the American
Museum of Natural History, 1121.
80

Carson, H.L. (1975) The genetics of speciation at the diploid level. American Naturalist, 109,
7392.
Carson, H.L. & Templeton, A.R.A. (1984) Genetic revolutions in relation to speciation
phenomena: the founding of new populations. Annual Review of Ecology and Systematics,
15, 97131.
Castellano, S. & Giacoma, C. (1998) Morphological variation of the green toad, Bufo viridis,
in Italy: a test of causation. Journal of Herpetology, 32, 540550.
Cherry, L.M., Case, S.M. & Wilson, A.C. (1978) Frog perspective on the morphological
difference between humans and chimpanzees. Science, 200, 209211.
Clarke, R.D. (1972) The effect of toe clipping on survival in Fowler's toad (Bufo woodhousei
fowleri). Copeia, 1, 182185.
Clement, M., Posada, D. & Crandall, K.A. (2000) TCS: a computer program to estimate gene
genealogies. Molecular Ecology, 9, 16571659.
Clough, M. & Summers, K. (2000) Phylogenetic systematics and biogeography of the poison
frogs: evidence from mitochondrial DNA sequences. Biological Journal of the Linnean
Society, 70, 515540.
Coates, A.C. & Obando, J.A. (1996) The geological evaluation of the Central American
Isthmus. In: Evolution and Environment in Tropical America (eds. Jackson, B.C., Budd,
A.F. & Coates, A.G.). University of Chicago Press, Chicago. Illinois, 2156.
Colangelo, P., Castiglia, R., Franchini, P. & Solano E. (2010) Pattern of shape variation in the
eastern African gerbils of the genus Gerbilliscus (Rodentia, Muridae): environmental
correlations and implication for taxonomy and systematics. Mammalian Biology, 75, 302
310.
Comeault, A. A. & Noonan, B.P. (2011) Spatial variation in the fitness of divergent
aposematic phenotypes of the poison frog, Dendrobates tinctorius. Journal of Evolutionary
Biology, 24, 13741379.
81

Cott, H.B. (1940) Adaptive Coloration in Animals. Methuen, London, UK.
Couldridge, V.C.K & Van Staaden, M.J. (2004) Habitat-dependent transmission of male
advertisement calls in bladder grasshoppers (Orthoptera; Pneumoridae). Journal of
Experimental Biology, 207, 27772786.
Coyne, J. A. (1992) Genetics and speciation. Nature, 355, 511515.
Crawford, A.J. (2003) Huge populations and old species of Costa Rican and Panamanian dirt
frogs inferred from mitochondrial and nuclear gene sequences. Molecular Ecology, 12,
25252540.
Crawford, A.J., Bermingham, E. & Polania, C. (2007) The role of tropical dry forest as a
longterm barrier to dispersal: a comparative phylogeographical analysis of dry forest
tolerant and intolerant frogs. Molecular Ecology, 16, 47894807.
Crump, M.L. (1972) Territoriality and mating behaviour in Dendrobates granuliferus (Anura:
Dendrobatidae). Herpetologica, 28, 195198.
Daly, J.W. & Myers, C.W. (1967) Toxicity of Panamanian poison frogs (Dendrobates): some
biological and chemical aspects. Science, 156, 970973.
Daly, J.W., Myers, C.W. & Whittaker, N. (1987) Further classification of skin alkaloids from
neotropical poison frogs (Dendrobatidae), with a general survey of toxic/noxious
substances in the Amphibia. Toxicon, 10, 10231095.
Daly, J.W., Garraffo, H.M., Spande, T.F., Jaramillo, C. & Rand, A.S. (1994a) Dietary source
for skin alkaloids of poison frogs (Dendrobatidae)? Journal of Chemical Ecology, 20, 943
955.
Daly, J.W., Gusovsky, F., Myers, C.W., Yotsu-Yamashita, M. & T. Yasumoto (1994b) First
occurrence of tetrodotoxin in a dendrobatid frog (Colostethus inguinalis), with further
reports for the bufonid genus Atelopus. Toxicon, 32, 279285.
Daly, J.W., Garraffo, H.M. & Myers, C.W. (1997) The origin of frog skin alkaloids: an
enigma. Pharmaceutical News, 4, 914.
82

Daly, J.W. (1998) Thirty years of discovering arthropod alkaloids in amphibian skin. Journal
of Natural Products, 61, 162172.
Daly, J.W., Garraffo, H.M. & Spande, T.F. (1999) Alkaloids from amphibian skins. In
Alkaloids: chemical and biological perspectives (ed. Pelletier, S.W.). Pergamon, New
York, 1161.
Daly, J.W., Garraffo, H.M. Spande, T.F. Decker, M.W. Sullivan, J.P. & Williams, M. (2000)
Alkaloids from frog skins: the discovery of epibatidine and the potential for developing
novel non-opioid analgesics. Natural Products Report, 17, 131135.
Darst, C.R., Cummings, M.E. & Cannatella, D.C. (2006) A mechanism for diversity in
warning signals: conspicuousness versus toxicity in poison frogs. Proceedings of the
National Academy of Sciences of the USA, 103, 58525857.
Darwin C (1859) On the Origin of Species by Means of Natural Selection (John Murray,
London).
Dijkstra, P.D., Seehausen, O. & Groothuis, T.G.G. (2008) Intra-sexual competition among
females and the stabilization of a conspicuous colour polymorphism in a Lake Victoria
cichlid fish. Proceedings of the Royal Society of London Series B, 275, 519526.
Dobzhansky T. (1940) Speciation as a stage in evolutionary divergence. American Naturalist
74, 312321.
Doebeli, M. & Dieckmann, U. (2004) Adaptive dynamics of speciation: spatial structure. In:
Adaptive Speciation (eds. Dieckmann, U., Doebeli, M., Metz, J.A.J., Tautz, D.).
Cambridge University Press, Cambridge, UK, 160167.
Dopman, E.B., Robbins, P.S. & Seaman, A. (2010) Components of reproductive isolation
between North American pheromone strains of the European corn borer. Evolution, 64,
881902.
Dosen, L.D. & Montgomerie, R. (2005) Female size influences mate preferences of male
guppies. Ethology, 110, 245255.
83

Dover, G. (1995) A species definition: a functional approach. Trends in Ecology &
Evolution, 12, 489490.
Driscoll, D.A. (1998) Genetic structure, metapopulation processes and evolution influence the
conservation strategies for two endangered frog species. Biological Conservation, 83, 43
54.
Duellman, W.E. (2001) The Hylid Frogs of Middle America. Society for the Study of
Amphibians and Reptiles, Ithaca, New York.
Dumbacher, J.P., Wako, A., Derrickson, S.R., Samuelson, A., Spande, T.F. & Daly., J.W.
(2004). Melyrid beetles (Choresine): a putative source for the batrachotoxin alkaloids
found in poison-dart frogs and toxic passerine birds. Proceedings of the National Academy
of Science USA, 101, 1585715860.
Edmunds, M. (1974) Defence in Animals. A Survey of Anti-Predator Defences. Longman,
New York.
Elwood, R., Gibson, J., Neil, S. (1987) The amorous Gammarus: size assortative mating in G.
pulex. Animal Behaviour, 35, 16.
Endler, J.A. (1977) Geographic Variation, Speciation and the Origin of Species. Princeton,
University Press, Princeton, New Jersey.
Endler, J.A. (1980) Natural selection on color patterns in Poecilia reticulata. Evolution, 34,
7691.
Endler, J.A. (1983) Natural and sexual selection on colour pattern in poeciliid fishes.
Environmenatl. Biology of Fish, 9, 173190.
Endler, J.A. (1992) Signals, signal conditions, and the direction of evolution. American
Naturalist, 139, 125153.
Endler, J. & Mappes, J. (2004) Predator mixes and the conspicuousness of aposematic signals.
American Naturalist, 163, 532547.
84

Ereshefsky, M. (1992) The Units of Evolution: Essays on the Nature of Species, MIT Press,
Cambridge, Massachussetts.
Evanno, G., Regnaut, S. & Goudet, J. (2005) Detecting the number of clusters of individuals
using the software STRUCTURE: a simulation study. Molecular Ecology, 14, 26112620.
Excoffier, L., Smouse, P.E. & Quattro, J.M. (1992) Analysis of molecular variance inferred
from metric distances among DNA haplotypes: application to human mitochondrlal DNA
restriction data. Genetics, 131, 479491.
Excoffier, L., Laval, G. & Schneider, S. (2005) Arlequin version 3.5: an integrated software
package for population genetics data analysis. Evolutionary Bioinformatics Online, 1, 47
60.
Exnerov, A., Landov, E., Stys, P., Fuchs, R., Prokopov, M. & Cehlarikov, P. (2003)
Reactions of passerine birds to aposematic and non-aposematic firebugs
(Pyrrhocorisapterus, Heteroptera). Biological Journal of Linnean Society, 78, 517525.
Exnerov, A., Stys, P., Fucikov, E., Vesel, S., Svadov, K., Prokopov, M., Jarosk, V.,
Fuchs, R. & Landov, E. (2007) Avoidance of aposematic prey in European tits (Paridae):
learned or innate? Behavioural Ecology, 18, 148156.
Fandino, M.C., H. Luddecke & A. Amezquita. (1997) Vocalisation and larval transportation
of male Colostethus subpunctatus (Anura: Dendrobatidae). Amphibia-Reptilia, 18, 3948.
Faulkner, D. J. (1992) Chemical defenses of marine molluscs. In: Ecological Roles of Marine
Natural Products (ed. Paul, V. J.). Comstock Publishing Associates, Ithaca, New York,
119163.
Feder, J.L., Berlocher, S.H., Roethele, J.B., Dambroski, H., Smith, J.J., Perry, W.L.,
Gavrilovic, V., Filchak, K.E., Rull, J. & Aluja, M. (2003) Allopatric origins for sympatric
host-plant shifts and race formation in Rhagoletis. Proceedings of the National Academy of
Science USA, 100, 1031410319.
85

Feder, J.L., Xie, X., Rull, J., Velez, S., Forbes, A., Dambroski, H., Filchak, K. & Aluja., M.
(2005) Mayr, Dobzhansky, Bush and the complexities of sympatric speciation in
Rhagoletis. Proceedings of the National Academy of Science USA, 102, 65736580.
Feder, J. L. & Nosil, P. (2009) Chromosomal inversions and species differences: when are
genes affecting adaptive divergence and reproductive isolation expected to reside within
inversions? Evolution, 63, 30613075.
Fink, L.S. & Brower, L.P. (1981) Birds can overcome the cardenolide defense of monarch
butterflies in Mexico. Nature, 291, 6770.
Fisher, D.O. & Cockburn, A. (2006) The large-male advantage in brown antechinuses: female
choice, male dominance, and delayed male death. Behavioral Ecology, 17, 164171.
Fitzpatrick, B.M., Fordyce, J.A. & Gavrilets, S. (2008) What, if anything, is sympatric
speciation? Journal of Evolutionary Biology, 21, 14521459.
Fitzpatrick, B.M., Fordyce, J.A. & Gavrilets, S. (2009) Pattern, process and geographic modes
of speciation. Journal of Evolutionary Biology, 22, 23422347.
Forester, D.C. & Czarnowsky, R. (1985) Sexual selection in the spring peeper Hyla crucifer
(Amphibia, Anura): role of the advertisement call. Behaviour, 92, 112128.
Foster, S.A & Endler, J.A. (1999) Geographic variation in behavior: perspectives on
evolutionary mechanisms. Oxford University Press, New York.
Frazer, J.F.D. & Rothschild, M. (1960), Defence mechanisms in warningly-coloured moths
and other insects Proceedings of the 11th International Congress of Entomology, Vienna,
1959, 249256.
Friedl, T.W.P. & Klump, G.M. (1997) Some aspects of population biology in the European
treefrog, Hyla arborea. Herpetologica, 53, 321330.
Gabor, C.R., Krenz, J.D. & Jaeger, R.G. (2000) Female choice, male interference, and sperm
precedence in the red-spotted newt. Behavioral Ecology, 11, 115124.
86

Gabor, C. & Page, R. (2003) Female preference for large males in sailfin mollies, Poecilia
latipinna: the importance of predation pressure and reproductive status. Acta ethologica, 6,
712.
Gamberale-Stille, G. (2000) Decision time and prey gregariousness influence attack
probability in nave and experienced predators. Animal Behavior, 60, 9599.
Gamberale-Stille, G. (2001) Benefit by contrast: an experiment with live aposematic prey.
Behavioural Ecology, 12, 768772.
Garca-Pars, M., Good, D.A., Parra-Olea., G. & Wake, D.B. (2000) Biodiversity of Costa
Rican salamanders: implications of high levels of genetic differentiation and
phylogeographic structure for species formation. Proceedings of the National Academy of
Sciences, USA, 97, 16401647.
Gavrilets, S. (1997) Evolution and speciation on holey adaptive landscapes. Trends in
Ecology and Evolution, 12, 307312.
Gavrilets, S., Li, H. & Vose, M.D. (2000) Patterns of parapatric speciation. Evolution,
54, 11261134.
Gavrilovic, V., Filchak, K.E., Rull, J. & Aluja, M. (2003) Allopatric origins for sympatric
hostplant shifts and race formation inRhagoletis. Proceedings of the National Academy of
Science USA, 100, 1031410319.
Gerhardt, H.C. & Huber, F. (2002) Acoustic Communication in Insects and Anurans.
University of Chicago Press, Chicago.
Gittleman, J.L. & Harvey, P.H. (1980) Why are distasteful prey not cryptic? Nature, 286,
149150.
Golay, N. & Durrer, H. (1994) Inflammation due to toe-clipping in natterjack toads (Bufo
calamita). Amphibia-Reptilia, 15, 8196.
87

Gomez, D. & Thery, M. (2007) Simultaneous crypsis and conspicuousness in color patterns:
comparative analysis of a neotropical rainforest bird community. American Naturalist, 169,
4261.
Goodman, D.E. (1971) Territorial behaviour in a neotropical frog, Dendrobates granuliferus.
Copeia, 2, 365370.
Goudet, J. (1995) FSTAT version 1.2: a computer program to calculate F-statistics. Journal of
Heredity, 86, 485486.
Grant, B.R & Grant, P.R (1989) Evolutionary dynamics of a natural population: the large
cactus finch of the Galpagos University of Chicago Press, Chicago, Illinois.
Grant, T., Frost, D.R., Caldwell, J.P., Gagliardo, R., Haddad, C.L., Kok, P.J., Means, D.B.,
Noonan, B.P., Schargel, W.E. & Wheeler, W.C. (2006) Phylogenetics systematics of
dartpoison frogs and their relatives (Amphibia: Athesphatanura: Dendrobatidae).
Bulletin of the American Museum of Natural History, 299, 1268.
Gray, S.M. & McKinnon, J.S. (2006) Linking color polymorphism maintenance and
speciation. Trends in Ecology and Evolution, 22, 7179.
Green, D.M. (1993) Evolutionary relationships of Eastern Palearctic brown frogs, genus
Rana: paraphyly of the 24-chromosome species group and the significance of chromosome
number change. Zoological Journal of the Linnean Society, 109, 125.
Greenwood, J.J.D., Wood, E.M. & Batchelor, S. (1981) Apostatic selection of distasteful
prey. Heredity, 47, 2734.
Grether, G.F., Hudon, J. & Endler, J.A. (2001) Carotenoid scarcity, synthetic pteridine
pigments and the evolution of sexual coloration in guppies (Poecilia reticulata).
Proceedings of the Royal Society series B, 268, 12451253.
Griffith, S.C., T.H. Parker & Valrie, A. Olson (2006) Melanin- versus carotenoid-based
sexual signals: is the difference really so black and red? Animal Behaviour, 71, 749763.
88

Gbitz, T., Thorpe, R.S. & Malhotra, A. (2005) The dynamics of genetic and morphological
variation on volcanic islands. Proceedings of the Royal Society of London Series B, 272,
751757.
Guilford, T. (1986) How do warning colors work? Conspicuousness may reduce
recognition errors in experienced predators. Animal Behaviour, 34, 286288.
Guilford, T. (1990) Evolutionary pathways to aposematism. Acta Oecologica, 11, 835842.
Hagemann, S. & Prhl, H. (2007) Mitochondrial paraphyly in a polymorphic poison frog
species (Dendrobatidae; D. pumilio). Molecular Phylogenetics and Evolution, 45, 740747.
Hagman, M. & Forsman, A. (2003) Correlated evolution of conspicuous coloration and body
size in poison frogs (Dendrobatidae). Evolution, 57, 29042910.
Hairston, N.G. (1979) Adaptive significance of color polymorphism in 2 Species of
Diaptomus (Copepoda). Limnological Oceanography, 24, 1537.
Halliday, T. (1996) Amphibians Ecological Census Techniques, A Handbook (ed. Sutherland,
W.J.). Cambridge University Press, Cambridge, UK.
Ham, A.D., Ihalainen, E., Lindstrm, L. & Mappes, J. (2006) Does color matter? The
importance of color in avoidance learning, memorability and generalisation. Behavioural
Ecology and Sociobiology, 60, 482491.
Harari, A.R., Handler, A.M. & Landolt, P.J. (1999) Size-assortative mating, male choice and
female choice in the curculionid beetle Diaprepes abbreviatus. Animal Behaviour, 58,
11911200.
Hartl, D.L. & Clark, A.G. (2007) Principles of Population Genetics, Sinauer, Sunderland,
Massachussetts.
Hausdorf, B. (2011) Progress toward a general species concept. Evolution, 65, 923931.
Hauswaldt, J.S., Ludewig, A.K., Vences, M. & Prhl, H. (2011) Widespread co-occurrence of
divergent mitochondrial haplotype lineages in a Central American species of poison frog
(Oophaga pumilio). Journal of Biogeography, 38, 711726.
89

Hero, J.M. (1989) A simple code for toe clipping anurans. Herpetological Review, 20, 6667.
Hill, G.E (1990) Females house finch prefer colourful males: sexual selection for a condition-
dependant trait. Animal Behaviour, 40, 563572.
Hill, G.E. (1993) Geographic variation in the carotenoid plumage pigmentation of male house
finches (Carpodacus mexicanus). Biological Journal of the Linnean Society, 49, 6386.
Hill, G.E. & McGraw, K.J. (2006) Bird Coloration. Vol II. Function and Evolution. Harvard
University Press, Cambridge, Massachussetts.
Hodek, I. & Honk, A. (1996) Ecology of Coccinellidae. Kluwer Academic Publishers,
Dordrecht, The Netherlands.
Hoffman, E.A. & Blouin, M.S. (2000) A review of colour and pattern polymorphisms in
anurans. Biological Journal of the Linnean Society, 70, 633665.
Hoffman, E.A., Schueler, F.W., Jones, A.G. & Blouin, M.S. (2006) An analysis of selection
on colour polymorphism in the northern leopard frog. Molecular Ecology, 15, 26272641.
Hoekstra, H.E., Drumm, K.E. & Nachman, M.W. (2004) Ecological genetics of adaptive
color polymorphism in pocket mice: geographic variation in selected and neutral genes.
Evolution, 58, 13291341.
Hoekstra, H.E., Krenz, J.G. & Nachman, M.W. (2005) Local adaptation in the rock pocket
mouse (Chaetodipus intermedius): natural selection and phylogenetic history of
populations. Nature Heredity, 94, 217228.
Holdridge, L.R. (1947) Determination of world plant formations from simple climatic data.
Science, 105, 367368.
Humphries, R.B. (1979) Dynamics of a breeding frog community. PhD Thesis. The
Australian National University, Canberra, Australia.
Ihalainen, E., Lindstrm, L. & Mappes, J. (2007) Investigating Mllerian mimicry: predator
learning and variation in prey defenses. Journal of Evolutionary Biology, 20, 780791.
90

Joron, M. & Mallet, J.L.B. (1998) Diversity in mimicry: paradox or paradigm? Trends in
Ecology and Evolution, 13, 461466.
Junc, F.A., Altig, R. & Gascon, C. (1994) Breeding biology of Colostethus stepheni, a
dendrobatid with a nontrasnported nidicolous tadpole. Copeia, 747750.
Kapan, D.D. (2001) Three-butterfly system provides a field test of mllerian mimicry. Nature,
409, 338340.
Kingston, J.L., Rosenthal, G.G. & Ryan, M.J. (2003) The role of sexual selection in
maintaining a colour polymorphism in the pygmy swordtail, Xiphophorus pygmaeus.
Animal Behavior, 65, 735743.
Kirkpatrick, M. & Ravign, V. (2002) Speciation by natural and sexual selection: models and
experiments. American Naturalist, 159, 2235.
Kirkpatrick, M. & Barton, N.H. (2006) Chromosome inversions, local adaptation and
speciation. Genetics, 173, 419434.
Kohda, M., Yanagisawa, Y. Sato, T. Nakaya, K. Niimura, Y. Matsumoto, K. & Ochi, H.
(1996) Geographical colour variation in cichlid fishes at the southern end of Lake
Tanganyika. Environmental Biology of Fish, 45, 237248.
Kohlmann, B., Wilkinson, J. & Lulla, K. (2002) Costa Rica desde el espacio/Costa Rica from
Space, 1st edn. Fundacon Neotrpica, San Jos.
Kohlmann, B., Solis, A., Elle, O., Soto, X. & Russo, R. (2007) Biodiversity, conservation,
and hotspot atlas of Costa Rica: a dung beetle perspective (Coleoptera: Scarabaeidae:
Scarabaeinae). Zootaxa, 1457, 134.
Lande, R. (1976) Natural selection and random genetic drift in phenotypic evolution.
Evolution, 30, 314334.
Lande, R. (1981) Models of speciation by sexual selection on polygenic traits. Proceedings of
the National Academy of Sciences, USA, 78, 37213725.
91

Larsson, K. & Forslund, P. (1991) Environmentally induced morphological variation in the
Barnacle Goose, Branta leucopsis. Journal of Evolutionary Biology, 4, 619636.
Lee, T.J., Speed, M.P. & Stephens, P.A. (2011) Honest Signaling and the Uses of Prey
Coloration. The American Naturalist , 178, 19.
Lemckert, F. (1996) Effects of toe-clipping on the survival and behaviour of the Australian
frog Crinia signifera. Amphibia-Reptilia, 17, 287290.
Lindstedt, C., Lindstrm, L. & Mappes, J. (2008) Hairiness and warning colors as
components of antipredator defense: additive or interactive benefits? Animal Behaviour,
75, 17031713.
Lindstedt, C., Eagerb, H., Ihalainena, E., Kahilainena, A., Stevens, M. & Mappes J. (2011)
Direction and strength of selection by predators for the color of the aposematic wood tiger
moth. Behavioral Ecology, 22, 580587.
Lindstrm, L., Alatalo, R.V. & Mappes, J. (1999a) Inherited avoidance towards color,
gregariousness and conspicuousness: experiments with naive and experienced predators.
Behavioural Ecology, 10, 317322.
Lindstrm, L., Alatalo, R., Mappes, J., Riipi, M. & Vertainen, L. (1999b) Can aposematic
signals evolve by gradual change? Nature, 397, 249251.
Lindstrm, L., Alatalo, R., Lyytinen, A. & Mappes, J. (2001a) Predator experience on cryptic
prey affects the survival of conspicuous aposematic prey. Proceedings of the Royal Society
of London series B, 268, 357361.
Lindstrm, L., Alatalo, R.V., Lyytinen, A. & Mappes, J. (2001b) Strong antiapostatic
selection against novel rare aposematic. Proceedings of the National Academy of Sciences,
98 91819184.
Lindstrm, L., Alatalo, R.V., Lyytinen, A. & Mappes, J. (2004) The effect of alternative prey
on the dynamics of Batesian and Mullerian mimicries. Evolution, 58, 12941302.
92

Lougheed, S.C., Austin, J.D., Bogart, J.P., Boag, P.T. & Chek, A.A. (2006) Multi-character
perspectives on the evolution of intraspecific differentiation in a neotropical hylid frog.
BMC Evolutionary Biology, 6, 23.
Lu, G. & Bernatchez, L. (1999) Correlated trophic specialization and genetic divergence in
sympatric lake whitefish ecotypes (Coregonus clupeaformis) support for the ecological
speciation hypothesis. Evolution, 53, 14911505.
Luddecke, H. (2000) Behavioral aspects of the reproductive biology of the Andean frog
Colostethus palmatus (Amphibia: Dendrobatidae). Revista de la Academia Colombiana de
Ciencias Exactas, Fisicas y Naturales, 23, 303316.
Maan, M.E., Seehausen, O., Soderberg, L., Johnson, L., Ripmeester, E.A.P., Mrosso, H.D.J.,
Taylor, M.I., van Dooren, T.J.M. & van Alphen, J.J.M. (2004) Intraspecific sexual
selection on a speciation trait, male coloration, in the Lake Victoria cichlid Pundamilia
nyererei. Proceedings of the Royal Society of London Series B, 271, 24452452.
Maan, M.E. & Cummings, M.E. (2008) Female preferences for aposematic signal components
in a polymorphic poison frog. Evolution, 62, 23342345.
Maan, M.E. & Cummings, M.E. (2009) Sexual dimorphism and directional sexual selection
on aposematic signals in a poison frog. Proceedings of the National Academy of Sciences
of the USA, 106, 1907219077.
Maan, M.E. & Seehausen, O. (2011) Ecology, sexual selection and speciation. Ecology
Letters, 14, 591602.
MacDougall, A. & Dawkins, M. S. (1998) Predator discrimination error and the benefits of
Mullerian mimicry. Animal Behaviour, 55, 12811288.
MacDougall, A.K. & Montgomerie, R. (2003) Assortative mating by carotenoid-based
plumage colour: a quality indicator in American goldfinches, Carduelis tristis.
Naturwissenschaften 90, 464467.
93

Malhotra, A. & Thorpe, R.S. (1997) Size and shape variation in a Lesser Antillean anole,
Anolis oculatus (Sauria: Iguanidae) in relation to habitat. Biological Journal of the Linnean
Society, 60, 5372.
Mallet, J. (1986) Hybrid zones in Heliconius butterflies in Panama, and the stability and
movement of warning colour clines. Heredity, 56, 191202.
Mallet, J. & Singer, M.C. (1987) Individual selection, kin selection and the shifting balance in
the evolution of warning colours: the evidence from butterflies. Biological Journal of the
Linnean Society, 32, 337350.
Mallet, J.L.B. & Barton, N.H. (1989) Strong natural selection in a warning color Mllerian
mimicry. Evolutionary Ecology, 13, 777806.
Mallet, J. & Joron, M. (1999) Evolution of diversity in warning color and mimicry:
polymorphism, shifting balance and speciation. Annual Review of Ecology and
Evolutionary Sytematics, 30, 201233.
Mallet, J., Meyer, A., Nosil, P. & Feder, J.L. (2009) Space, sympatry and speciation. Journal
of Evolutionary Biology.
Marroig, G. & Cheverud, J.M. (2001) A comparison of phenotypic variation and covariation
patterns and the role of phylogeny, ecology, and ontogeny during cranial evolution of new
world monkeys. Evolution, 55, 25762600.
Mappes, J. & Alatalo, R.V. (1997) Effect of novelty and gregariousness in survival of
aposematic prey. Behavioural Ecology, 8, 174177.
Masta, S.E. & Maddison, W.P. (2002) Sexual selection driving diversification in jumping
spiders. Proceedings of the National Academy of Science USA, 99, 44424447.
Marchan-Rivadanera, M.R., Larsen, P.A., Phillips, C.J., Strauss, R.E. & Baker, R.J. (2012),
On the association between environmental gradients and skull size variation in the great
fruit-eating bat, Artibeus lituratus (Chiroptera: Phyllostomidae). Biological Journal of the
Linnean Society, 105, 623634.
94

Marples, N.M., Roper, T.J. & Harper, D.G.C. (1998) Response of wild birds to novel prey:
evidence of dietary conservatism. Oikos, 83, 161165.
Mayden, R.L. (1997) A hierarchy of species concepts: the denouement in the saga of the
species problem. In: Species: the units of biodiversity (eds. Claridge, M.F., Dawah, H.A. &
Wilson, M.R.). Chapman and Hall, London, UK, 381424.
Mayr, E, (1947) Ecological factors in speciation. Evolution, 1, 263288.
Mayr, E. (1963) Animal species and evolution. Museum of Comparative Zoology, Harvard
University, Cambridge, Massachussetts.
Mayr, E. (1970) Populations, species and evolution. Harvard University Press, Cambridge,
Massachussetts.
McCarthy, M.A. & Parris, K.M. (2004) Clarifying the effect of toe clipping on frogs with
Bayesian statistics. Journal of Applied Ecology, 41, 780786.
McGuigan, K., McDonald, K., Parris, K. & Moritz, C. (1998) Mitochondrial DNA diversity
and historical biogeography of a wet forest-restricted frog (Litoria pearsoniana) from mid-
east Australia. Molecular Ecology, 7, 175186.
McKay, J.K. & Latta, R.G. (2002) Adaptive population divergence: markers, QTL and traits.
Trends in Ecology and Evolution, 17, 285291.
Merilaita, S & Kaitala V. (2002) Community structure and the evolution of aposematic
coloration. Ecology Letters, 5, 495501.
Merilaita, S. (2006) Frequency-dependent predation and maintenance of prey polymorphism.
Journal of Evolutionary Biology, 19, 20222030.
Meyer, E. (1992) Erfolgreiche Nachzucht von Dendrobates granuliferus Taylor, 1958.
Herpetofauna, 14, 1121.
Meyer, E. (1993) Fortplanzung und Brutpflegeverhalten von Dendrobates granuliferus
Taylor, 1958. Verffentlichungen Naturhistorisches Museum Schleusingen, 7/8, 113142.
95

Morey, S.R. (1990) Microhabitat selection and predation in the pacific treefrog, Pseudacris
regilla. Journal of Herpetology, 24, 292296.
Morrison, C & Hero, J. (2003) Geographic variation in lifehistory characteristics of
amphibians: a review. Journal of Animal Ecology, 72, 270279.
Mller, F. (1879) Ituna and Thyridia: a remarkable case of mimicry in butterflies. Proceedings
of the Entomological Society of London, 2029.
Myers, C.W., Daly, J.W. & Malkin, B. (1978) A dangerously toxic new frog (Phyllobates)
used by Ember Indians of western Colombia, with discussion of blowgun fabrication and
dart poisoning. Bulletin of the American Museum of Natural History, 161, 307366.
Myers, C.W., Daly, J.W. & Martinez, V. (1984) An arboreal poison frog (Dendrobates) from
western Panama. American Museum Novitates, 2783, 120.
Narins, P.M., H, W. & Grabul., D.S. (2003) Bimodal signal requisite for agonistic behaviour
in a dart-poison frog, Epipedobates femoralis. Proceedings of the National Academy of
Science USA, 100, 577580.
Narins, P.M., Grabul, D.S., Soma, K.K., Gaucher, P. & Hdl, W. (2005) Cross-modal
integration in a dart-poison frog. Proceedings of the National Academy of Science USA,
102, 24252429.
Noonan, B.P. & Gaucher, P. (2006) Refugial isolation and secondary contact in the dyeing
poison frog Dendrobates tinctorius. Molecular Ecology, 15, 44254435.
Noonan, B.P., & Comeault, A.A. (2009) The role of predator selection on polymorphic
aposematic poison frogs. Biology Letters 5, 5154.
Noor, M.A.F. (1999) Reinforcement and other consequences of sympatry. Heredity, 83, 503
508.
Nosil, P., Vines, T.H. & Funk, D.J. (2005) Reproductive isolation caused by natural selection
against immigrants from divergent habitats. Evolution, 59, 705719.
96

Nosil, P & Crespi, B.J. (2006) Experimental evidence that predation promotes divergence in
adaptive radiation. Proceedings of the National Academy of Sciences of the United States
of America, 103, 90909095.
Obika, M. & Bagnara, J.T. (1964) Pteridines as pigments in amphibians. Science, 143, 485
487.
Ohmer, M.E., Robertson, J.M. & Zamudio, K.R. (2009) Discordance in body size, colour
pattern, and advertisement call across genetically distinct populations in a Neotropical
anuran (Dendropsophus ebraccatus). Biological Journal of the Linnean Society, 97, 298
313.
Olsson, M. (1993) Male preference for larger females and assortative mating for body size in
the sand lizard (Lacerta agilis). Behavioral Ecology and Sociobiology, 32, 337341.
van Oppen, M.J.H., Turner, G.F., Rico, C., Robinson, R.L., Deutsch, J.C., Genner, M.J. &
Hewitt, G.M. (1998) Assortative mating among rock-dwelling cichlid fishes supports high
estimates of species richness from Lake Malawi. Molecular Ecology, 7, 9911001.
Orr, H.A. & Turelli, M. (2001) The evolution of postzygotic isolation: Accumulating
Dobzhansky-Muller incompatibilities. Evolution, 55, 10851094.
Palumbi, S.R., Martin, A., Romano, S., McMillan, W.O., Stice, L. & Grabowski, G. (1991) A
simple fools guide to PCR. Version 2.0. University of Hawaii, Honolulu, HI.
Panhuis, T.M., Butlin, R., Zuk, M. & Tregenza, T. (2001) Sexual selection and speciation,
Trends in Ecology and Evolution, 16, 364371.
Parris, K.M. & McCarthy, M.A. (2001) Identifying effects of toe clipping on anuran return
rates: the importance of statistical power, Amphibia-Reptilia, 22, 275-289.
Parsons, Y.M. & Shaw, K.L. (2001) Species boundaries and genetic diversity among
Hawaiian crickets of the genus Laupala identified using amplified fragment length
polymorphism. Molecular Ecology, 10, 17651772.
97

Pinheiro, C.E.G. (1996) Palatability and escaping ability in Neotropical butterflies: tests with
wild kingbirds (Tyrannus melancholicus, Tyrannidae). Biological Journal of the Linnean
Society, 59, 351365.
Peakall, R. & Smouse, P.E. (2006) GENALEX 6: genetic analysis in Excel. Population genetic
software for teaching and research. Molecular Ecology Notes, 6, 288295.
Posada, D. (2008) jModelTest: phylogenetic model averaging. Molecular Biology and
Evolution, 25, 12531256.
Poulton, E.B. (1890) The Colour of Animals: their Meaning and Use. Kegan Paul, Trench,
Trubner, London, UK.
Povoln D. (1999) Mendelian dihybridism in central European polymorphic populations of
aposematic burnet-moth Zygaena ephialtes (Linnaeus, 1767) and its Mllerian-Batesian
mimicry (Lepidoptera, Zygaenidae). Acta Universitatis Agriculturae et Silviculturae
Mendelianae Brunensia, 47, 4166.
Preston, B.T., Stevenson, I.R., Pemberton, J.M., Coltman, D.W. & Wilson, K. (2005) Male
mate choice influences female promiscuity in Soay sheep. Proceedings of the Royal
Society of London Series B, Biological Sciences, 272, 365373.
Price, J.J., Friedman, N.R. & Omland, K.E. (2007) Song and plumage evolution the new
world orioles (Icterus) show similar lability and convergence in patterns. Evolution, 61,
850863.
Pritchard, J.K., Stephens, M. & Donnelly, P. (2000) Inference of population structure using
multilocus genotype data. Genetics, 155, 945959.
Prhl, H. (1997) Territorial behavior of the strawberry poisondart frog, Dendrobates
pumilio. Amphibia-Reptilia, 18, 437442.
Prhl, H. & Hdl, W. (1999) Parental investment, potential reproductive rates and mating
system in the strawberry poison-dart frog Dendrobates pumilio. Behavioural Ecology and
Sociobiology, 46, 215220.
98

Prhl, H. & Berke, O. (2001) Spatial distribution of male and female strawberry poison frogs
and their relation to female reproductive resources. Oecologia, 129, 534542.
Prhl, H. (2003) Variation in male calling behaviour and relation to male mating success in
the strawberry poison frog (Dendrobates pumilio). Ethology, 109, 273290.
Prhl, H. (2005) Territorial behavior in dendrobatid frogs. Journal of Herpetology, 39, 354
365.
Prhl, H., Koshy, R.A., Mueller, U., Rand, A.S. & Ryan, M.J. (2006) Geographic variation of
genetic and behavioral traits in northern and southern tngara frogs. Evolution 60, 1669
1679.
Prudic, K., Skemp, A.K. & Papaj, D.R. (2007) Aposematic coloration, luminance contrast,
and the benefits of conspicuousness. Behavioural Ecology, 18, 4146.
de Queiroz, A. & Wimberger, P.H. (1993) The usefulness of behavior for phylogeny
estimation: levels of homoplasy in behavioral and morphological characters. Evolution 47,
4660.
Ratcliffe, J.M. & Nydam, M.L. (2008) Multimodal warning signals for a multiple predator
world. Nature 455, 9699.
Raymond, M. & Rousset, F. (1995a) GENEPOP (version 1.2): a population genetics software
for exact test and ecumenicism. Journal of Heredity, 86, 248249.
Raymond, M. & Rousset, F. (1995b) An exact test of population differentiation. Evolution,
49, 12801283.
Reaser, J.K. & Dexter, R.E. (1996) Rana pretiosa (spotted frog). Toe clipping effects.
Herpetological Review, 27, 195196.
Reynolds, R.G. & Fitzpatrick, B.M. (2007) Assortative mating in poison-dart frogs based on
an ecologically important trait. Evolution, 61, 22532259.
99

Richards, C.L. & Knowles, L.L. (2007) Tests of phenotypic and genetic concordance and
their application to the conservation of Panamanian golden frogs (Anura, Bufonidae).
Molecular Ecology, 16, 31193133.
Richards-Zawacki, C.L. & Cummings, M.E. (2011) Intraspecific reproductive character
displacement in a polymorphic poison dart frog, Dendrobates pumilio. Evolution, 65, 259
267.
Riipi, M., Alatalo, R., Lindstrm, L. & Mappes, J. (2001) Multiple benefits of gregariousness
cover detectability costs in aposematic aggregations. Nature, 413, 512514.
Roberts, J.D. & Wardell-Johnson, G. (1995) Call differences between peripheral isolates of
the Geocrinia rosea complex (Anura: Myobatrachidae) in Southwestern Australia. Copeia,
4, 899906.
Robertson, J.M. (2008) Genetic and phenotypic diversity patterns in two polymorphic,
neotropical anurans: biogeography, gene flow and selection. PhD thesis, Cornell
University, Ithaca, New York.
Robertson, J.M. & Robertson, A.D. (2008) Spatial and temporal patterns of phenotypic
variation in a Neotropical frog. Journal of Biogeography, 35, 830843.
Robertson, J.M., Duryea, M.C. & Zamudio, K.R. (2009), Discordant patterns of evolutionary
differentiation in two Neotropical treefrogs. Molecular Ecology, 18, 13751395.
Ronquist, F. & Huelsenbeck, J.P. (2003) MrBayes 3: Bayesian phylogenetic inference under
mixed models. Bioinformatics, 19, 15721574.
Roper, T.J. & Redstone, S. (1987) Conspicuousness of distasteful prey affects the strength
and durability of one-trial avoidance learning. Animal Behaviour, 35, 739747.
Roper, T.J. (1990) Responses of domestic chicks to artificially colored insect prey: effects of
previous experience and background color. Animal Behaviour, 39, 466473.
Rosenblum, E.B., Hoekstra, H. E. & Nachman, M.W. (2004) Adaptive reptile color variation
and the evolution of the MCIR gene. Evolution, 58, 17941808.
100

Rosso, A., Castellano, S. & Giacoma, C. (2004) Ecogeographic analysis of morphological and
life-history variation in the Italian treefrog. Evolutionary Biology 18, 303321.
Roulin, A. (2004) The evolution, maintenance and adaptive function of genetic colour
polymorphism in birds. Biological Reviews, 79, 815848.
Rousset, F. (2008) GENEPOP007: a complete re-implementation of the GENEPOP software for
Windows and Linux. Molecular Ecology Resources, 8, 103106.
Rowe, C., Lindstrm, L. & Lyytinen, A. (2004) The importance of pattern similarity between
Mllerian mimics in predator avoidance learning. Proceedings of the Royal Society of
London series B, 271, 407413.
Rowland, H.M., Ihalainen, E., Lindstrm, L., Mappes, J. & Speed, M.P. (2007) Co-mimics
have a mutualistic relationship despite unequal defense levels. Nature, 448, 6466.
Rudh, A., Rogell, B. & Hglund, J. (2007) Non-gradual variation in colour morphs of the
strawberry poison frog Dendrobates pumilio: genetic and geographical isolation suggest a
role for selection in maintaining polymorphism. Molecular Ecology, 16, 42844294.
Rueffler, C., Van Dooren, T.J.M., Leimar, O. & Abrams, P.A. (2006) Disruptive selection and
then what? Trends in Ecology and Evolution, 21, 238245.
Rutowski, R.L., Macedonia, J.M., Morehouse, N.I. & Taylor-Taft, L. (2005) Pterin pigments
amplify iridescent ultraviolet signal in males of the orange sulphur butterfly, Colias
eurytheme . Proceedings of the Royal Society series B, 272, 23292335.
Ruxton, G.D., Sherratt, T.N. & Speed, M.P. (2004) Avoiding attack. Evolutionary ecology of
crypsis, warning signals and mimicry. Oxford University Press, New York.
Ruxton, G.D., Speed, M. & Broom, M. (2007) The importance of initial protection of
conspicuous mutants for the coevolution of defense and aposematic signaling of the
defense: a modeling study. Evolution, 61, 21652174.
101

Ryan, M.J. & Rand, A.S. (1990) The sensory basis of sexual selection for complex calls in the
tngara frog, Physalaemus pustulosus (sexual selection for sensory exploitation).
Evolution, 44, 305314.
Ryan, M.J. & Wilczynski, W. (1991) Evolution of intraspecific variation in the advertisement
call of a cricket frog (Acris crepitans, Hylidae). Biological Journal of the Linnean Society,
44, 249271.
Ryan, M.J., Rand, A.S. & Weigt, L.A. (1996) Allozyme and advertisement call variation in
the tngara frog, Physalaemus pustulosus. Evolution, 50, 24352453.
Ryan, M.J. (1999) Sexual selection, receiver biases, and the evolution of sex differences.
Science, 28, 19992003.
Sanderson, N., Szymura, J.M. & Barton, N.H. (1992) Variation in mating call across the
hybrid zone between fire-bellied toads Bombina bombina and B. variegata. Evolution, 46,
595607.
Sandre, S.L., Stevens, M. & Mappes, J. (2010) The effect of predator appetite, prey warning
coloration and luminance on predator foraging decisions. Behaviour, 147, 11211143.
Santos, J. C., Coloma, L. A. & Cannatella, D. C. (2003) Multiple, recurring origins of
aposematism and diet specialization in poison frogs. Proceedings of the National Academy
of Science USA, 100, 1279212797.
Santos, J.C., Coloma, L.A., Summers, K., Caldwell, J.P., Ree, C. & Cannatella, R.D.
(2009) Amazonian amphibian diversity is primarily derived from late miocene
Andean lineages. PLoS Biology, 7, e56.
Saporito, R.A., Garraffo, H.M., Donnelly, M.A., Edwards, A.L., Longino, J.T. & J.W. Daly.
(2004) Formicine ants: an arthropod source for the pumiliotoxin alkaloids of dendrobatid
poison frogs. Proceedings of the National Academy of Science USA, 101, 80458050.
102

Saporito, R.A., Norton, R.A., Garraffo, H.M., Spande, T.F. & Daly, W.D. (2007a) Oribatid
mites as a major dietary source for alkaloids in poison frogs. Proceedings of the National
Academy of Science USA, 104, 88858890.
Saporito, R.A, Zuercher, R., Roberts, M., Gerow, K.G. &. Donnelly, M.A (2007b)
Experimental Evidence for Aposematism in the Dendrobatid Poison Frog Oophaga
pumilio.
Saporito, R.A., Donnelly, M.A., P. Jain, Garraffo, H.M., Spande, T. F. & Daly, J. W. (2007c)
Spatial and temporal patterns of alkaloid variation in the poison frog Oophaga pumilio in
Costa Rica and Panama over 30 years. Toxicon, 50, 757778.
Savage, J.M. (1968) The dendrobatid frogs of Central America. Copeia, 4, 745776.
Savage, J.M. (2002) The Amphibians and Reptiles of Costa Rica: A Herpetofauna Between
Two Continents, Between Two Seas. The University of Chicago Press, Chicago, Illinois.
Schuble, C.S. (2004) Variation in body size and sexual dimorphism across geographical and
environmental space in the frogs Limnodynastes tasmaniensis and L. peronii. Biological
Journal of the Linnean Society, 82, 3956.
Schlenoff, D.H. (1984) Novelty: a basis for generalisation in prey selection. Animal
Behaviour, 32, 919921.
Schliewen, U., Kocher, T., McKaye, K.R., Seehausen, O. & Tautz, D. (2006) Evidence for
sympatric speciation? Nature, 444, E12E13.
Schluter, D. (2009) Evidence for ecological speciation and its alternative. Science, 323, 737
741.
Schuler, W. & Hesse, E. (1985) On the function of warning coloration, a black and yellow
pattern inhibits prey attack by nave domestic chicks. Behavioural Ecology and
Sociobiology, 16, 245255.
103

Seehausen, O. & van Alphen, J.J.M. (1998) The effect of male coloration on female mate
choice in closely related Lake Victoria cichlids (Haplochromis nyererei complex).
Behavioral Ecology and Sociobiology , 42, 18.
Servedio, M.R. (2000) The effects of predator learning, forgetting, and recognition errors on
the evolution of warning coloration. Evolution, 54, 751763.
Sherratt, T.N. & Beatty, C.D. (2003) The evolution of warning signals as reliable indicators of
prey defense. American Naturalist, 162, 377389.
Siddiqi, A., Cronin, T.W., Loew, E.R., Vorobyev, M. & Summers, K. (2004) Interspecific
and intraspecific views of color signals in the strawberry poison frog Dendrobates pumilio.
Journal of Experimental Biology, 207, 24712485.
Siefferman, L. & Hill, G. E. (2003) Structural and melanin coloration indicate parental effort
and reproductive success in male eastern bluebirds. Behavioral Ecology, 14, 855861.
Silverstone, P.A. (1973) Observations on the behavior and ecology of a Colombian poison-
arrow frog, the kokoep (Dendrobates histrionicus Berthold). Herpetologica, 29, 295301.
Silverstone, P.A. (1975) A revision of the poison-arrow frogs of the genus Dendrobates.
Scientific Bulletin Of The Natural History Museum, Los Angeles County, 21, 155.
Slabbekoorn, H. & Smith, T.B. (2002) Bird song, ecology and speciation. Philosophical
Transactions of the Royal Society of London Series B, 357, 493503.
Smith, S.M. (1977) Coral-snake pattern recognition and stimulus generalisation by naive great
kiskadees (Aves: Tyrannidae). Nature, 265, 535536.
Smith, S.M. (1975) Innate recognition of coral snake pattern by a possible avian predator.
Science, 187, 759760.
Smith, M.J., Roberts, J.D., Hammond, T.J. & Davis, R.A. (2003) Intraspecific variation in the
advertisement call of the sunset frog Spicospina flammocaerula (Anura: Myobatrachidae):
a frog with a limited geographic distribution. Journal of Herpetology, 37, 285291.
104

Speed, M.P. (1993) Muellerian mimicry and the psychology of predation. Animal Behaviour,
45, 571580.
Speed, M.P. & Ruxton, G.D. (2007) How bright and how nasty: explaining diversity in
warning signal strength. Evolution, 61, 623635.
Stalker, H.D. & Carson, H.L. (1947) Morphological variation in natural populations of
Drosophila robusta Sturtevant. Evolution, 1, 237248.
Steffen, J.E. & McGraw, K.J. (2007) Contributions of pterin and carotenoid pigments to
dewlap coloration in two anole species. Comparative Biochemistry and Physiology B, 146,
4246.
Stimson, J. & Berman, M. (1990) Predator induced colour polymorphism in Danaus
plexippus L. Lepidoptera: Nymphalidae) in Hawaii. Heredity, 65, 401406.
tys, P. (2004) Phylogenetic constraints on occurence of warning coloration in the European
Heteroptera. In: Abstract Book ISBE 2004, 211. Jyvskyla, Finland: University of
Jyvskyla.
Summers, K. (1989) Sexual selection and intra-female competition in the green poison-dart
frog, Dendrobates auratus. Animal Behaviour, 37, 797805.
Summers, K., Bermingham, E., Weigt, L., McCafferty, S., & Dahlstrom, L. (1997)
Phenotypic and genetic divergence in three species of dart-poison frogs with contrasting
parental behavior. Journal of Heredity, 88, 813.
Summers, K., Symula, R., Clough, M. & Cronin, T. (1999) Visual mate choice in poison
frogs. Proceedings of the Royal Society of London Series B, 266, 21412145.
Summers, K. & Clough, M. (2001) The evolution of coloration and toxicity in the poison frog
family (Dendrobatidae). Proceedings of the National Academy of Sciences of the United
States of America, 98, 62276232.
105

Summers, K. & McKeon, C.S. (2004) The evolutionary ecology of phytotelmata use in
Neotropical poison frogs. Miscellaneous Publications, Museum of Zoology, University of
Michigan, 193, 5573.
Swofford, D.L. (2002) PAUP*: phylogenetic analysis using parsimony, version 4b10. Sinauer
Associates, Sunderland, MA.
Takada, W., Sakata, T., Shimano, S., Enami, Y., Mori, N., Nishida, R. & Kuwahara. Y.
(2005) Scheloribatid mites as the source of pumiliotoxins in dendrobatid frogs. Journal of
Chemical, Ecology, 31, 24032415.
Tamura, K. Dudley, J., Nei, M. & Kumar, S. (2007) MEGA4: Molecular Evolutionary
Genetics Analysis (MEGA) software version 4.0. Molecular Biology and Evolution, 24,
15961599.
Templeton, A.R. (1980) The theory of speciation via the founder principle. Genetics, 94,
10111038.
Templeton, A.R. (1981) Mechanisms of speciation: a population genetic approach. Annual
Review of Ecology and Systematics, 12, 23 48.
Thomas, R.J., Marples, N.M., Cuthill, I.C., Takahashi, M. & Gibson, E.A (2003) Dietary
conservatism may facilitate the initial evolution of aposematism. Oikos, 101, 458466.
Thorpe, R.S., Malhotra, A., Black, H., Daltry, J.C. & Wster, W. (1995) Relating geographic
pattern to phylogenetic process. Proceedings of the Royal Society of London Series B, 349,
6168.
Thorpe, R.S & Stenson, A.G. (2003) Phylogeny, paraphyly and ecological adaptation of the
colour and pattern in the Anolis roquet complex on Martinique. Molecular Ecology, 12,
117132.
Tordoff, W. (1980) Selective predation of gray jays, Perisoreus canadensis, upon boreal
chorus frogs, Pseudacris triseriata. Evolution, 24, 10041008.
106

Tullberg, B.S., Merilaita, S. & Wiklund, C. (2005) Aposematism and crypsis combined as a
result of distance dependence: functional versatility of the colour pattern in the swallowtail
butterfly larva. Proceedings of the Royal Society series B, 272, 13151321.
Turelli, M., Barton, N.H. & Coyne, J.A. (2001) Theory and speciation. Trends in Ecology &
Evolution, 16, 330343.
Turner, J.R.G. & Speed, M.P. (1996) Learning and memory in mimicry. Simulations of
laboratory experiments. Philosophical Transactions of the Royal Society of London Series
B, 351, 11571170.
Turner, J.R.G. & Mallet, J.L.B. (1996) Did forest islands drive the diversity of warningly
coloured butterflies? Biotic drift and the shifting balance. Phylosophical Transactions of
the Royal Society of London Series B, 351, 835845.
Vences, M., Kosuch, J. Boistel, R. Haddad, C.F.B., La Marca, E., Lotters S. & Veith. M.
(2003) Convergent evolution of aposematic coloration in neotropical poison-dart frogs: a
molecular phylogenetic perspective. Organismal Diversity and Evolution, 3, 215226.
Vercken, E., Massot, M., Sinervo, B. & Clobert, J. (2007) Colour variation and alternative
reproductive strategies in females of the common lizard Lacerta vivipara. Journal of
Evolutionary Biology, 20, 221232.
Verrell, P.A. (1985) Male mate choice for large, fecund females in the red-spotted newt,
Notophthalmus viridescens: how is size assessed? Herpetologica, 41, 382386.
Waichman, A.V. (1992) An alphanumeric code for toe clipping amphibians and reptiles.
Herpetological Review, 23, 1921.
Wallace, A. R. (1867) Mimicry and other protective resemblances among animals.
Wang, I.J. & Shaffer, H.B. (2008) Rapid color evolution in an aposematic species: a
phylogenetic analysis of color variation in the strikingly polymorphic strawberry poison-
dart-frog. Evolution, 62, 27422759.
107

Wang, I.J. & Summers, K. (2010) Genetic structure is correlated with phenotypic divergence
rather than geographic isolation in the highly polymorphic strawberry poison-dart frog.
Molecular Ecology, 19, 447458.
Wedell, N. & Sandberg, T. (1995) Female preference for large males in the bushcricket
Requena sp. 5 (Orthoptera: Tettigoniidae). Journal of Insect Behavior, 8, 513522.
Weigt, L.A., Crawford, A.J., Rand, A.S. & Ryan, M.J. (2005) Biogeography of the tungara
frog, Physalaemus pustulosus: a molecular perspective. Molecular Ecology, 14, 3857
3876.
Weir, B.S. & Cockerham, C.C. (1984) Estimating F-statistics for the analysis of population
structure. Evolution, 38, 13581370.
Welch, J.J. (2004) Accumulating Dobzhansky-Muller incompatibilities: Reconciling theory
and data. Evolution, 58, 11451156.
Wellborn, G.A. & Broughton, R.E. (2008) Diversification on an ecologically constrained
adaptive landscape. Molecular Ecology, 17, 29272936.
Wells, K.D. (1978) Courtship and parental behavior in a Panamanian poison-arrow frog
(Dendrobates auratus). Herpetologica, 34, 148155.
Wells, K.D. (1980a) Behavioral ecology and social organization of a dendrobatid frog
(Colostethus inguinalis). Behavioral Ecology and Sociobiology, 6, 199209.
Wells, K.D. (1980b) Evidence for growth of tadpoles during parental transport in Colostethus
inguinalis. Journal of Herpetology, 14, 428430.
Wells, K.D. (1980c) Social behavior and communication of a dendrobatid frog (Colostethus
trinitatis). Herpetologica, 36, 189199.
West-Eberhard, M.J. (1983) Sexual selection, social competition, and speciation. Quarterly
Review of Biology, 58, 155183.
108

Weygoldt, P. (1980) Complex brood care and reproductive behavior in captive poison-arrow
frogs, Dendrobates pumilio, O. Schmidt. Behavioral Ecology and Sociobiology, 7, 329
332.
Weygoldt, P. (1987) Evolution of parental care in dart poison frogs (Amphibia: Anura:
Dendrobatidae). Zeitschrift fr Zoologische Systematik und Evolutionsforschung, 25, 51
7.
White, M.J.D. (1968) Models of speciation. Science, 159, 10651070.
Wiens, J.J. & Penkrot, T.A. (2002) Delimiting species using DNA and morphological
variation and discordant species limits in spiny lizards (Sceloporus). Systematic Biology,
51, 69 91.
van Wijngaarden, R. (1985) Some remarks on Dendrobates granuliferus. Lacerta, 43, 181
183.
van Wijngaarden, R. & Bolaos, F. (1992) Parental care in Dendrobates granuliferus (Anura:
Dendrobatidae) with a description of the tadpole. Journal of Herpetology, 26, 102105.
van Wijngaarden, R. & van Gool. S. (1994) Site fidelity and territoriality in the dendrobatid
frog Dendrobates granuliferus. Amphibia-Reptilia, 15, 171181.
Williamson, I. & Bull, C.M. (1996) Population ecology of the Australian frog Crinia
signifera: adults and juveniles. Wildlife Research, 23, 249266.
Wollenberg, K., Noonan, B.P. & Ltters, S. (2006) Polymorphism versus species richness-
systematics of large Dendrobates from the eastern Guiana Shield (Amphibia:
Dendrobatidae). Copeia. 4, 623629.
Wright, S. (1931) Evolution in Mendelian populations. Genetics, 16, 97159.
Yosef, R. & Whitman, D.W. (1992) Predator exaptations and defensive adaptations in
evolutionary balance: no defense is perfect. Evolutionary Ecology, 6, 527536.
Zamudio, K.R. & Greene, H.W. (1997) Phylogeography of the bushmaster (Lachesis muta:
109

Viperidae): implications for Neotropical biogeography, systematics, and conservation.
Biological Journal of the Linnean Society, 62, 421442.
Zimmermann, H., & E. Zimmermann (1988) Etho-Taxonomie und zoogeographische
artengruppenbildung bei pfeilgiftfrschen (Anura: Dendrobatidae). Salamandra, 24, 125
160.












110

Additional Information

1. Sequences accession numbers

List of all sequences used in the study. Samples specifically sequenced for the present work
are marked with *. Table reports, sampled animal, sampling site (and code) and the haplotype
code (with relative accession number). For samples sharing the same haplotype identified in
the present study, the accession number is reported only once.
ID Sampling site [code] Cyt b (AccN) 16S (AccN)
O. granulifera 1-CH* Charcos [CH] HE775251 (H1) HE804764 (1)
O. granulifera 2-CH* Charcos [CH] HE775251 (H1) HE804764 (1)
O. granulifera 3-CH* Charcos [CH] HE775252 (H2) HE804765 (2)
O. granulifera 4-VE* Ventana [VE] HE775253 (H3)
O. granulifera 5-VE* Ventana [VE] HE775253 (H3)
O. granulifera 6-PA* Palmar [PA] HE775254 (H4) HE804766 (3)
O. granulifera 7-PA* Palmar [PA] HE775255 (H5) HE804766 (3)
O. granulifera 8-PA* Palmar [PA] HE775256 (H6) HE804766 (3)
O. granulifera 9-FS* Firestone [FS] HE775257 (H7) HE804767 (4)
O. granulifera 10-FS* Firestone [FS] HE775257 (H7)
O. granulifera 11-FS* Firestone [FS] HE775258 (H8) HE804767 (4)
O. granulifera 12-BA* Baru [BA] HE775258 (H8) HE804767 (4)
O. granulifera 13-BA* Baru [BA] HE775259 (H9)
O. granulifera 14-BA* Baru [BA] HE775260 (H10)
O. granulifera 15-DO* Dominical [DO] HE775257 (H7) HE804767 (4)
O. granulifera 16-DO* Dominical [DO] HE775257 (H7) HE804767 (4)
O. granulifera 17-DO* Dominical [DO] HE775257 (H7)
111

O. granulifera 18-MA* Matapalo [MA] HE775257 (H7)
O. granulifera 19-MA* Matapalo [MA] HE775257 (H7)
O. granulifera 20-MA* Matapalo [MA] HE775258 (H8) HE804767 (4)
O. granulifera 21-MA* Matapalo [MA] HE775262 (H12)
O. granulifera 22-PL* Portalon [PL] HE775257 (H7) HE804767 (4)
O. granulifera 23-PL* Portalon [PL] HE775257 (H7)
O. granulifera 24-PL* Portalon [PL] HE775261 (H11) HE804767 (4)
O. granulifera 25-SA* Savegre [SV] HE775257 (H7) HE804767 (4)
O. granulifera 26-DA* Damitas [DA] HE775257 (H7) HE804768 (5)
O. granulifera 27-DA* Damitas [DA] HE775257 (H7)
O. granulifera 28-DA* Damitas [DA] HE775257 (H7)
O. granulifera Wang0925-FC Fila Chonta [FC] HQ841077 (H7)
O. granulifera Wang0927-FC Fila Chonta [FC] HQ841078 (H16)
O. granulifera Wang0928-FC Fila Chonta [FC] HQ841079 (H15)
O. granulifera Wang0929-CN Cerro Nara [CN] HQ841080 (H15)
O. granulifera Wang0930-CN Cerro Nara [CN] HQ841081 (H7)
O. granulifera Wang0931-CN Cerro Nara [CN] HQ841082 (H15)
O. granulifera Wang0932-DM Rio Damitas [DM] HQ841083 (H15)
O. granulifera Wang0933-DM Rio Damitas [DM] HQ841084 (H7)
O. granulifera Wang0934-DM Rio Damitas [DM] HQ841085 (H15)
O. granulifera Wang0939-SV Rio Savegre [SV] HQ841086 (H7)
O. granulifera Wang0940-SV Rio Savegre [SV] HQ841087 (H7)
O. granulifera Wang0941-SV Rio Savegre [SV] HQ841088 (H7)
O. granulifera Wang1007-BR Rio Baru [BR] HQ841089 (H10)
O. granulifera Wang1009-BR Rio Baru [BR] HQ841090 (H8)
O. granulifera Wang1010-BR Rio Baru [BR] HQ841091 (H10)
O. granulifera Wang1012-PB Piedras Blancas [PB] HQ841092 (H14)
O. granulifera Wang1014-PB Piedras Blancas [PB] HQ841093 (H13)
112








O. granulifera Wang1015-PB Piedras Blancas [PB] HQ841094 (H14)
O. granulifera Wang1016-TE Ro Trraba [TE] HQ841095 (H14)
O. granulifera Wang1017-TE Ro Trraba [TE] HQ841096 (H14)
O. granulifera Wang1018-TE Ro Trraba [TE] HQ841097 (H14)
O. granulifera Wang1019-DB Drake Bay [DB] HQ841098 (H14)
O. granulifera Wang1020-DB Drake Bay [DB] HQ841099 (H14)
O. granulifera Wang1021-DB Drake Bay [DB] HQ841100 (H14)
O. granulifera Wang1022-CO Corcovado [CO] HQ841101 (H14)
O. granulifera Wang1024-CO Corcovado [CO] HQ841102 (H14)
O. granulifera Wang1026-CO Corcovado [CO] HQ841103 (H14)
Oophaga histrionica _ HQ841104 DQ502032
Oophaga sylvatica _ HQ290567 AY364569
Oophaga arborea _ DQ502036
Oophaga lehmanni _ DQ502034
Oophaga pumilio-N Nicaragua DQ502235
Oophaga pumilio-S Panama EU342663
Oophaga speciosa _ DQ502037
Oophaga vicentei _ DQ502167
Dendrobates leucomelas
Dendrobates tinctorius
_ AF124119
DQ768797
113


2. Population structure K results









114

Ackowledgments
Thanks to the Volkswagen foundation for the very generous resources provided.
Thanks to the MINAE for releasing research permits.
Thanks to Heike (Prhl) for giving an opportunity to an unknown student from Italy.
Thanks to Nick (Mundy) for letting me enter in his lab at the University of Cambridge.
Thanks to Eberhard Meyer for providing information on some frog populations locations.
Thanks to Adriana Bellati for advising on molecular data analysis.
Thanks to Sabine Sippel for her collaboration in the lab.
Thanks to Snke von den Berg for helping with graphics.
Thanks to Cristina who has been an important support during several years.
Thanks to my parents who are still doing an effort to understand where I want to go.
Thanks to all people who helped me out during the crazy time in the field.
Thanks to the tiny granular poison frogs for existing.
Thanks to the lethal snakes of Costa Rica for leaving me alive and able to conclude this work.
Thanks to Costa Rica cause I realized there is not just one way to live.
Thanks to Germany and England cause I understood cultural differences may be funny.
Thanks to Julia for showing me one does not need many things to be happy.

Das könnte Ihnen auch gefallen