of the southwest united states a dissertattion …

121
THE EVOLUTION AND DESIGN OF THE LOCOMOTOR SYSTEM IN LIZARDS OF THE SOUTHWEST UNITED STATES A DISSERTATTION SUBMITTED TO THE GRADUATE DIVISION OF THE UNIVERSITY OF HAWAI‘I AT MĀNOA IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY IN ZOOLOGY AUGUST 2014 By Jeffrey A. Scales Dissertation Committee: Marguerite A. Butler, Chairperson David B. Carlon Kathleen S. Cole Leonard A. Freed Marcelo Kobayashi Keywords: locomotor performance, adaptation, trade-offs, lizards, selective pressures

Upload: others

Post on 29-Dec-2021

0 views

Category:

Documents


0 download

TRANSCRIPT

THE EVOLUTION AND DESIGN OF THE LOCOMOTOR SYSTEM IN LIZARDS

OF THE SOUTHWEST UNITED STATES

A DISSERTATTION SUBMITTED TO THE GRADUATE DIVISION OF THE

UNIVERSITY OF HAWAI‘I AT MĀNOA IN PARTIAL FULFILLMENT OF THE

REQUIREMENTS FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

IN

ZOOLOGY

AUGUST 2014

By

Jeffrey A. Scales

Dissertation Committee:

Marguerite A. Butler, Chairperson

David B. Carlon

Kathleen S. Cole

Leonard A. Freed

Marcelo Kobayashi

Keywords: locomotor performance, adaptation, trade-offs, lizards, selective pressures

ii

ACKNOWLEDGEMENTS

I would like to thank all of the volunteers who helped with fieldwork during this

project including, but not limited to M. Altmann, B. DeRoy, R. Grey, K. Mullins, and

M.J. Truini. I also extend thanks to C. Brong for assistance in video analyses, M.A. Pena

for helping with muscle dissections, and J. Rivera for providing the gene sequences used

in this study. K. Bonine, A.K. Lappin, and W. Sherbrooke provided invaluable assistance

and guidance with lizard collection in the field, for which I am deeply grateful. I express

my gratitude to R. Blob for teaching me the muscle dissection and measurement

techniques used in this study. For logistical support in the field, I thank the AMNH

Southwestern Research Station in Portal, AZ, A.K. Lappin, and J. Scales and R. Scales.

Funding for this research was provided by National Science Foundation (NSF) grants

DDIG 0910400 to M. A. Butler and J. Scales, DEB-0515390 to M.A. Butler, and DEB-

0542360 to A.A. King. Additional funding was supplied by The Society of Integrative

and Comparative Biology (SICB) Grants in Aid of Research, and the University of

Hawaii Colleges of Arts and Sciences Student Research Awards to J. Scales. The Arizona

Fish and Game Department, California Department of Fish and Game, and New Mexico

Department of Game and Fish all provided permits and assistance for lizard collection. I

thank Y. Chan, S. Evers, E. Henry, C. Kokami, M.A. Pena, and J. Rivera constructive

comments that greatly improved this manuscript. Finally, I sincerely thank my advisor,

M.A. Butler and committee members, D.B. Carlon, K.S. Cole, L.A. Freed, and M.

Kobayashi for their mentorship and support throughout my graduate career.

iii

ABSTRACT

Understanding the adaptation and diversification of complex phenotypes is a

central task of evolutionary biology. However, because many phenotypes are complex,

constructed in a in a hierarchy of biological organization that often must perform multiple

functions, studies frequently lack the detail required to fully understand the evolution of

these systems. To gain a deeper understanding regarding the evolution of complex

phenotypes, I integrate evolutionary modeling and functional morphology approaches to

investigate locomotor evolution in lizards of the Southwest United States. I use

evolutionary models based on habitat use, predator escape behavior, and foraging mode

to examine what drives diversity in sprint speed, acceleration, and exertion in these

lizards. I then examine how the locomotor system is designed for specific locomotor

tasks. Because muscle provides the force required for locomotion, I focus on muscle

cross-sectional area (CSA). I first examine the relationship between each performance

and muscle CSA, and then explore how these relationships and CSA distribution vary

with function (flexors vs. extensors) and anatomical location (hip, knee, and ankle).

Finally, because the locomotor system is composed of numerous traits that can vary in

their functional roles, I explore what drives the evolution of each trait individually. I

model the evolution of each trait, and compare the best fitting models across traits to

determine how the locomotor system evolves as a whole. Overall, I demonstrate that

multiple selective pressures act on both locomotor performance and morphology from the

cellular to whole animal level. However, selection is not uniform within the locomotor

system. Performance related selection acts across many locomotor traits, but some are

more constrained or not subject to selection. Behavioral variation is an important driver

of phenotypic diversity as behavioral shifts guide the evolution of performance, muscular

design, and form-function relationships in these lizards. Finally, the complexity of the

locomotor system may promote adaptation and diversification through the mitigation of

potential functional trade-offs and allowing alternative locomotor designs to yield similar

performance capabilities.

iv

TABLE OF CONTENTS

ACKNOWLEDGEMENTS.………………………..………………………………….…ii

ABSTRACT….……………………………………..……………………………...…….iii

LIST OF TABLES….……………………………………..…………………………….viii

LIST OF FIGURES.……...……………………………………………………………....ix

CHAPTER 1. INTRODUCTION…...………………………………………………..…10

Lizards as a study system……………………………………….………..11

CHAPTER 2. ADAPTIVE EVOLUTION IN LOCOMOTOR PERFORMANCE: HOW

SELECTIVE PRESSURES AND FUNCTIONAL RELATIONSHIPS PRODUCE

DIVERSITY.……………………………………………………………………………..15

Abstract.………………………………………………………………………….15

Introduction.………………………………………………………………...……15

Materials and Methods.……………………………………………………..……18

Field-work.………...……………………………………………………..18

Performance trials.……………………………..………………………..18

Phylogeny..……………………………………………………...………..20

Comparative Analyses….…………………………………………..……20

Selective Regimes for OU model…………….…………………………..21

Results and Discussion……….………………………………………………….24

Adaptive evolution of performance.……………………………….......…24

Predator Escape Pressure.……………………………………….24

Foraging strategy.…………………………………………..……25

Habitat use.……………………………………………………....26

v

The reality of performance types.………………………………………..27

Effects of body size on performance.…………………………………….27

Trade-offs in performance...……………………………………………...28

Summary....………………………………………………………………30

CHAPTER 3. HINDLIMB MUSCLE CROSS-SECTIONAL AREA REFLECTS

LOCOMOTOR PERFORMANCE AND MICROHABITAT USE IN LIZARDS.……...37

Abstract.…..……………………………………………………..……………….37

Introduction.…..……………………………………………………….…………38

Materials and Methods…..……………………………………………….………40

Fieldwork…………………………………………………….…………..40

Performance Trials…...…………………………………………….……40

Morphological and Muscle measurements..……………….………….…41

Phylogeny………………………………………………………………...42

Statistical Analyses..………………………………………………….….42

Results.…………………………………………………………………………...43

Scaling of hind limb CSA…..………………………………...……...……43

Hind limb CSA and sprint performance…………………….….…...……44

Hind limb CSA and exertion………...…………………………………...44

Variation in CSA across the hind limb…………………………………...45

Discussion.…………………………………………….…………..……………..45

Muscle CSA and Size…………………………………………..….……..45

CSA, acceleration, and speed...………………………………………….46

Distribution of extensor CSA and locomotor performance...……………48

Summary..…………………………………………….………………….50

vi

CHAPTER 4. RUNNING FOR YOUR LIFE OR RUNNING FOR YOUR DINNER:

WHAT DRIVES FIBER-TYPE EVOLUTION IN LIZARD LOCOMOTOR MUSCLES?

……………………………………………...…………………………………………….60

Abstract..…………………….……………………….…………………………..60

Introduction...…………………………………………………………………….60

Materials & Methods.…………………………..………………………………..63

Muscle fiber-type data and assumptions…………………………………63

Phylogeny.………………………..………………………………………64

Selective Regimes...………………………………………………………65

Analysis...………………………………………………………………...67

Results...…………………………………………….……………………………68

Fast-twitch (FG and FOG)...…………………………………………….68

Slow-twitch oxidative (SO)…………………………...………………….69

Discussion………………………………………...……………………………...69

Selective Hypotheses…………………………………...………………...70

Multiple Evolutionary Models……………………………...……………72

CHAPTER 5. TARGETS OF SELECTION AND POTENTIAL CONSTRAINTS: HOW

THEY SHAPE THE EVOLUTION OF THE LOCOMOTOR SYSTEM IN LIZARDS

……………………………………..…………………….……………………………….78

Abstract……………………………………………….………………………….78

Introduction…………………...………………………………………………….79

Materials and Methods.………………………..…………………………………81

Morphological Data Measurements…………………………...………...81

Morphological traits for evolutionary models………………….………..82

vii

Comparative Analyses……………………………………...……………83

Selective Regimes for evolutionary models……………………....………84

Results…………………………………………….…………………………...…86

Selection on body shape…………………………………………...……..86

Selection on musculature…………………………………………...……86

Selection at the tissue level…………………………………………...….87

Discussion…………………………………………….…………………...……..87

CHAPTER 6. CONCLUSION…..……………………………………………...………95

APPENDIX….………………...………………………….…………...………………...99

REFERENCES…………………………………………….…………….……………..104

viii

LIST OF TABLES

Table 2.1. Performance of alternative models for the evolution of performance………..31

Table 2.2. Model parameters estimated for the predator escape model……………….…31

Table 2.3. Model parameters estimated for the foraging mode model…………………..32

Table 2.4. PC analysis loadings for four performance variables………………………...32

Table 2.5. The relationship between size and performance…………………………...…33

Table 3.1. Relationships between body mass and hind limb CSA……………………….51

Table 3.2. Relationship between hind limb CSA and acceleration and sprint speed…….51

Table 3.3. Relationship between CSA at each joint and speed and acceleration...………52

Table 3.4 Relationship between hind limb CSA and exertion………………….………..53

Table 3.5 PCA loadings of CSA at each joint………………………………………...….54

Table 4.1 Fast-twitch fiber type model selection criteria………………………..…….…74

Table 4.2. Parameters of the best fitting fast-twitch fiber type model…………………...74

Table 4.3. Slow-twitch fiber type model selection criteria……………………………....75

Table 4.4. Parameters of the best fitting slow-twitch fiber type models………..………..75

Table 4.5. Model estimates of fiber type composition compared to raw data……...……76

Table 5.1. Model selection results for locomotor traits………………………………….92

Table 5.2. Comparison of evolutionary models…………………………………………93

ix

LIST OF FIGURES

Figure 2.1. Adaptive evolutionary hypotheses for the evolution of locomotor

performance…………………………………………………...…………………………34

Figure 2.2. Performance phenotypes reflect locomotor demands……………………….35

Figure 2.3. Correlations of PIC between each of the performance variables……………36

Figure 3.1. Phylogenetic relationships of the 20 species of lizards included in this

study……………………………………………………………………………………...55

Figure 3.2. Relationships between log total hind limb and log body mass…………..….56

Figure 3.3. Non-phylogenetic and phylogenetic relationships between hind limb CSA and

performance variables……………………………………………………………………57

Figure 3.4. The proportion of hind limb CSA by function and joint…………………….58

Figure 3.5. PCA of extensor muscle CSA distribution across the hip, knee, and ankle…59

Figure 4.1. Evolutionary hypotheses for the evolution of fiber-type composition in the

iliofibuaris muscle of lizards………………………………………………………..……77

Figure 5.1. Evolutionary hypotheses for the evolution of body shape, musculature, and

individual muscle traits of the locomotor system in lizards……………………………..94

10

CHAPTER 1. INTRODUCTION

Understanding phenotypic adaptation and diversification has been a primary goal

in evolutionary biology, ecology, and physiology. Phenotypic adaptations, features

shaped by natural selection that improve the survival and fitness of an organism, have

played a central role in the study of biology not only because they are a key component

of Darwin's theory of natural selection, but also because adaptations elucidate form-

function relationships, can identify abiotic and biotic factors important to the survival or

reproduction of organisms, and can provide insight into mechanisms of speciation and

determinants of species distributions (Darwin 1859, Schluter 2000, Reznik & Travis

1996). Adaptation to different environments or biotic conditions can also lead to

phenotypic diversification, intimately linking the two (Schluter 2000). As a consequence,

the immense diversity we observe today is frequently attributed to adaptation. However,

adaptation and diversity may be limited because phenotypic traits are commonly subject

to multiple selective pressures and frequently serve multiple functions while forced to

follow design rules (Alfaro et al. 2004, Collar et al. 2006). Thus, to understand the

processes contributing to phenotypic adaptation and diversification requires an

integrative approach, but adaptation is most frequently examined with respect to a single

function or single selective pressure. Here, I combine functional morphology and

evolutionary modeling to examine the evolution of the locomotor system in relation to

multiple selective pressures and locomotor traits to gain a more complete understanding

of the evolution of complex phenotypes.

Locomotion and the locomotor system have served as one of the primary models

for the study of phenotypic adaptation and diversification. First, locomotor performance

is important for numerous fitness related tasks such as escaping predators and obtaining

mates suggesting that it is subject to selection (Alexander 2003, Miles 2004, Lailvaux &

Irschick 2006, Langerhans 2009a). Second, many morphological and physiological traits

have been linked to variation in performance (Bonine et al. 1999, Biewener 2003,

Kohlsdorf & Navas 2012), and variation in performance has been linked to fitness

differences (Arnold 1983, Garland & Losos 1994, O’Steen et al. 2010, Miles 2004,

Langerhans 2009a, Husak et al. 2006, Strobbe et al. 2009) providing evidence for

11

locomotor adaptation. Finally, locomotor performance and morphology is heritable

(Garland 1990) allowing performance to evolve in response to selection.

While it may seem straightforward to examine adaptation and diversification in

locomotion, it can be difficult as the locomotor system is quite complex. Locomotor

systems are organized in a hierarchy of biological levels including cellular, tissue,

system, and whole-animal so that locomotor performance is determined by numerous

morphological and physiological traits, all of which must be integrated to function

properly (Schmidt-Nielsen 1984, Bauwens et al. 1995, Kohlsdorf & Navas 2012).

Locomotor systems must also accomplish a wide array of functions (ie. running, jumping,

climbing, swimming) with the same machinery. If two functions place different demands

on the locomotor system, the two performances may not be optimized simultaneously

resulting in trade-offs, potentially limiting locomotor adaptation and diversity (Lewontin

1978, Arnold 1992, Reznik & Travis 1996). Furthermore, because the locomotor system

must perform so many ecologically relevant tasks, several forms/agents of selection more

than likely act on locomotor performance (Arnold 1992, Reznik & Travis 1996,

Ghalambor et. al. 2004). Despite this complexity, performance is commonly examined

with respect to a single aspect of locomotion such as sprint speed or at a single level of

biological organization, limiting our knowledge of functional roles and integration (e.g.,

Van Damme & Vanhooydonck 2001). Additionally, locomotor morphology and

performance are most often studied in the context of a single selective pressure (e.g.,

Fulton et al. 2003, Verwaijen & Van Damme 2007, Gifford et al. 2008). However, by

examining multiple selective pressures, aspects of performance, and locomotor traits

simultaneously we can gain important inferences into the processes and mechanisms that

promote and/or limit the adaptation and diversification of the locomotor performance and

morphology.

Lizards as a Study system

Lizards have long served as popular models for physiological, ecological, and

evolutionary studies. This popularity largely stems from their enormous diversity. With

roughly 5,900 (and counting) species, lizards make up the bulk of the order squamata

(which also includes snakes and amphisbaenians) and occur widely on all continents

12

except Antarctica (Pough et al. 2004). While many aspects of lizard biology are

conserved (ectothermy, scaled skin, hemipenes in males, etc.), they vary greatly in

lifestyle and life history (Pough et al. 2004). For example, diets range from herbivores,

to omnivores, insectivores and strict carnivores (Pough et al. 2004). Life history varies

even more as sexual and asexual reproduction, and oviparity and viviparity all occur in

lizards (Pough et al. 2004). These diverse lifestyles and life histories have resulted in an

incredible range of morphologies and physiological adaptations.

Diverse lifestyles coupled with a strong reliance on locomotion for ecologically

relevant tasks have made lizards one of the primary models for studies of locomotor

adaptation (Garland & Losos 1994, Aerts et al. 2000, Irschick & Garland 2001).

Thermoregulation, finding mates, defending home ranges and territories, escaping

predators, and obtaining food all depend on locomotion, and lizards have diversified

immensely to accomplish these tasks (Garland & Losos 1994, Pough et al. 2004). Species

specialize in running, climbing, swimming, gliding, and even limbless locomotion, which

has evolved repeatedly for burrowing or serpentine locomotion (Pough et al. 2004). This

high locomotor diversity is associated with several aspects of lifestyle. For example,

lizards occupy a wide array of habitats including hot, dry deserts, rain forests, grasslands,

mountains, and even a marine species (Pough et al. 2004). Within these habitats, they

further vary in microhabitat use, ranging from fossorial species that rarely surface to

arboreal specialists. Lizards also vary greatly in behavior. Predator escape strategies

range along a continuum from species that rely almost exclusively on crypsis to avoid

detection to those that rely heavily on speed to escape attacks (Vitt & Congdon 1978,

Pough et al. 2004). Foraging mode shows a similar spectrum of diversity. Sit-and-wait

predators spend most of their time motionless with occasional short bursts to catch prey,

while at the other end of the spectrum, active foragers spend the majority of time in

motion searching for prey (Reilly et al. 2007). Differences in microhabitat use, predator

escape behavior, and foraging behavior are all commonly invoked to explain variation in

lizard locomotor performance, and recent studies show that all three factors influence

locomotor morphology and performance (Herrel et al. 2002, Melville & Swain 2003,

Miles 2004, Reilly et al. 2007). Thus, lizards vary widely in both their locomotor

demands and the potential selective pressures that act on performance, providing the

13

opportunity to examine how multiple forms of selection interact to guide the evolution of

the locomotor system.

Specifically, I examine locomotor evolution in lizards of the Southwest United

States. I focus mainly on the family Phrynosomatidae, the major radiation of lizards in

North America (Wiens et al. 2010, Wiens et al. 2013). This family includes several

genera, but I focus on the Sceloporus group (fence and spiny lizards), the sand lizards

(Callisaurus, Cophosaurus, Holbrookia, and Uma), and horned lizards (Phrynosoma).

Sceloporines have a “typical” lizard body shape and are more generalist in most behavior,

but vary widely in microhabitat use including terrestrial species as well as arboreal and

rock specialists (Herrel et al. 2002, Bonine et al. 2001, Stebbins 2003). The sand lizards,

on the other hand are largely terrestrial, often inhabiting open areas. Several of these

species are sprint specialist with slender bodies and long hind limbs that use high speeds

to evade predation (Herrel et al. 2002, Bonine et al. 2001, Stebbins 2003). In contrast,

the horned lizards (sister clade to the sand lizards) are cryptic specialists that use

camouflage and morphology to avoid detection and predatory attacks. These lizards are

robust, heavy species with relatively shorter limbs and tails, and ornate scales and

“horns” (Herrel et al. 2002, Bonine et al. 2001, Stebbins 2003, Bergmann & Irschick

2010). Furthermore, most horned lizards feed mainly on ants, which has resulted in

foraging behavior that differs from other Phrynosomatid lizards.

In addition, I include one species of Crotaphytus, two species from the Teiidae

family and a single species from the Anguidae family. Crotaphytus species are large,

quick, sit-and-wait predators that often feed on other lizards within the community (Vitt

& Congdon 1978, Stebbins 2003). Lizards of the genus Aspidoscelis (Teiidae) are

slender, active foragers that rarely stop moving when active, generally use speed to avoid

predation, and occur in a variety of habitats (Vitt & Congdon 1978, Anderson & Karasov

1981, Stebbins 2003). Anguidae is a diverse family of lizards, but most species in North

America are long, slender lizards with relatively small limbs, are active in leaf litter and

other cluttered habitats, and are generally active foragers (Stebbins 2003, Reilly et al.

2007). Thus, the lizards included in this study vary widely in performance abilities,

morphology, habitat use, and foraging and predator escape strategies providing an

excellent opportunity to examine how these factors influence the evolution of locomotor

14

performance and morphology as well as how lizards are designed for specific tasks. I ask

three questions related to the evolution of performance and locomotor system design in

these lizards. 1) What drives the diversity of locomotor performance in these lizards, is it

habitat use, foraging mode, predator escape behavior, or none of these? 2) How are

lizards designed for specific performance capabilities? Which traits are most important

for performance? Does morphological integration allow for multiple pathways to the

same performance? 3) At what level of biological organization does adaptation occur, is it

uniform, or are some traits constrained while others are influenced by selection?

15

CHAPTER 2. ADAPTIVE EVOLUTION IN LOCOMOTOR PERFORMANCE:

HOW SELECTIVE PRESSURES AND FUNCTIONAL RELATIONSHIPS

PRODUCE DIVERSITY

Abstract

The incredible diversity of athletic performances in the animal kingdom has long

fascinated biologists. However, most studies of complex phenotypes take a single

approach, either exploring the role of natural selection in producing sometimes

spectacular athletes or alternatively, focusing on the role of design rules, the material

properties of tissues, or functional interactions in limiting phenotypes. Here we show that

a combination of both approaches is necessary to capture the extent of biodiversity in a

complex phenotype. Adaptation plays a prominent role in explaining the diversity of

locomotor phenotypes in southwestern lizards, but a combination of selective pressures is

required to explain variation in separate aspects of performance. Speed and acceleration

are strongly influenced by selective pressures related to predator escape, whereas exertion

is explained by foraging strategy. The interactions of these different performances also

play a role, resulting in a trade-off between speed and time to exhaustion. Surprisingly,

however, this trade-off is not obligatory and can be overcome by selection for both

performances simultaneously. Therefore, if multiple selective pressures are shaping a

complex phenotype, ignoring the contribution of one will lead to substantial

underestimation of the adaptive component of phenotypic diversification. Alternatively, if

intrinsic and extrinsic factors interact, but intrinsic factors are ignored, selection will

appear inadequate in explaining diversity and we may misinterpret what limits

adaptation.

Introduction

Whether it is for catching dinner, avoiding being someone else's dinner, or

winning contests for mates, an enormous diversity of athletic abilities have evolved

across animals. A variety of selective pressures have been proposed to explain such

16

impressive variation in locomotor capabilities. Behavioral strategies, in particular, play

an important role. For example, predator avoidance is known to select for increases in

speed and acceleration in the fast start of fish (e.g., Langerhans 2009a), swimming speed

of insects (Strobbe et al. 2009), and the sprint speeds of lizards (Miles 2004) and African

herbivores (Bro-Jorgensen 2013). Another commonly involved selective pressure are the

strategies associated with foraging mode. Because the ability to find and acquire prey

determines an animal’s energy budget, foraging strategy can have large fitness

consequences. Interestingly, the different strategies are often quite divergent, with active

foragers relying on frequent but often slow locomotion, whereas ambush predators move

rarely but rapidly (Miles et al. 2007 and references therein), resulting in phenotypes

shaped by selection for energetically efficiency and higher stamina, versus high speed

burst locomotion (Miles et al. 2007 and references therein). Accordingly, variation in

foraging strategy influences locomotor performance and morphology in a wide range of

species (e.g., Norberg and Rayner 1987, Verwaijen & VanDamme 2008, Pruitt 2010).

Furthermore, foraging, predator escape, and other locomotor behaviors all occur in the

context of specific microhabitats, and animals use a wide variety of microhabitats.

Running on narrow branches, for instance, may place a greater premium on stability,

which is not experienced by species that run on flat ground (Losos & Sinervo 1989,

Losos 1990). Therefore, the specific microhabitat context may define the scope of the

functional problem on which selection should act (Losos & Sinervo 1989, Losos 1990,

Fulton et al. 2001, Hodgkison et al. 2004, Gomes et al. 2009).

While natural selection is central in shaping locomotor abilities, phenotypes may

not be able to reach the adaptive peak defined by a single selective pressure. Complex

phenotypes are often limited by functional constraints that place bounds on the action of

adaptive evolution (Arnold 1992, Walker 2007). For example, the same locomotor

machinery is responsible for producing diverse performances such as walking, jumping,

and fighting. As a consequence, if two aspects of performance place conflicting demands

on the same design features, trade-offs can occur so that excellence in one performance

will come at the expense of another. For example, sprinters can achieve the fast running

speeds, but tire quickly, whereas distance runners can run for great lengths of time but

locomote at much lower speeds. The physiological basis for this observed trade-off is

17

thought to occur at the cellular level, as sprinters are enriched with fast-twitch muscle

fibers that produce high power but fatigue quickly, whereas fatigue-resistant muscle

fibers produce lower power outputs (Wilson et al. 2002, Wilson and James 2004). Due to

this negative relationship, we may expect a trade-off between burst speed and sustained

locomotion so that individuals display either high burst speeds coupled with low

endurance or vice versa (Bonine et al. 2005). Therefore, if multiple selective pressures

place antagonistic demands on the locomotor system, the phenotype may reflect a

balance that is suboptimal values for each performance. Alternatively, strong selection an

exclusive strategy may lead to specialization with poor performance for other tasks.

Although numerous empirical studies have documented performance trade-offs in several

functional systems (e.g., Vanhooydonck et al. 2001, Levinton and Allen 2005, Oufiero et

al. 2011), fewer studies have explored to what extent trade-offs may limit adaptation and

diversification (but see Holzman et al. 2011, Vanhooydonck et al. 2014).

While variation in locomotor performances may be governed by ecological

selection and functional interactions, body size can also influence locomotor

performance. Body size has wide-ranging effects on most morphology and physiological

processes, including locomotion (Alexander 1981, Schmidt-Nielsen 1984, Clemente et al.

2012). Thus, locomotor diversity may simply reflect variation in size as opposed to

selection or other design factors.

In this study, we investigate how multiple selective pressures interact to guide the

evolution of locomotor diversity. We focus on the highly diverse lizards of the American

southwest, which include members of the Crotaphytidae, Anguidae and Teiidae as well as

one of the largest radiations of lizards in North America, the Phrynosomatidae (Wiens et

al. 2010). Collectively, these lizards show a broad diversity in locomotor strategies,

differing not only in microhabitat use (Stebbins 2003), but also locomotor behaviors

ranging from extreme cryptic specialists that rely little on speed such as the horned

lizards, to the swift zebra-tailed lizards, who run at impressive speeds, and whiptails that

spend the majority of their time in sustained, slow motion (Anderson & Karasov 1981,

Bonine & Garland 1999). Additionally, their phylogenetic relationships are relatively

well-established (Townsend et al. 2004, Vidal & Hedges 2005, Wiens et al. 2010, Wiens

et al. 2012, Pyron et al. 2013) allowing phylogenetic analyses. Specifically, we test

18

whether selection related to microhabitat use, foraging strategy, and predator escape

behavior act on sprint speed, acceleration, and exertion abilities, and if so, how they

interact to produce the diverse performance types observed in these lizards. We also

examine the relationship between each aspect of performance to determine whether trade-

offs occur as a consequence of multiple locomotor demands, and if so, to what extent

trade-offs limit excellence in multiple aspects of performance. Alternatively, factors

other than selection, such as body size, may explain performance differences. Thus, we

explore the relationship between size and performance as well.

Materials and Methods

Field-work

Three to 10 individuals of 21 species (comprising four families) of lizards were

collected from the wild in California, Arizona and New Mexico during the summers of

2009 and 2010 (Table 1S in Appendix). After capture, lizards were transported to the

Southwestern Research Station or an alternative field lab (a temperature controlled house)

for locomotor trials. Lizards were held overnight for passage of gut contents, with

locomotor performance trials commencing on the day following capture. Lizards were

kept in cloth bags between performance trials with access to water ad libitum. Trials

continued for three days after which time lizards were either euthanized (by sodium

pentobarbital injection) for morphological study, or returned to their site of capture and

released live. All animal procedures followed University of Hawaii IACUC protocol 09-

463.

Performance Trials

Prior to each performance trial, lizards were placed under a heat lamp to achieve

field body temperature (based on previous studies and summarized in Bonine & Garland

1999) for at least an hour. Body temperatures were verified before each trial using a

cloacal thermometer (Miller & Webber Inc.).

For sprint trials, lizards were induced to run down a 3.0 x 0.15m trackway by

clapping or lightly tapping their tail. Each individual was tested five times over a span of

19

two days. Lizards were given a minimum of one hour of rest between each trial and were

run a maximum of three times in one day in order to capture maximum acceleration and

maximum sprint speed. Only runs in which lizards ran straight without jumping, pausing,

or touching the walls were used for analysis.

Sprint speed was measured using a series of infrared photocells placed every

0.25m over the second meter of the trackway. The single fastest time between two

photocells over the five runs was designated as the individual’s maximum sprint speed

under the specified test conditions (Peterson and Husak 2006, Vanhooydonck et al. 2001).

Acceleration was measured by digitally tracking a white marker glued (Elmer’s

glue) dorsally, directly above the pelvis of each lizard. The pelvis is an easily located

landmark near the animal's center of mass and care was taken to minimize handling time

to reduce the lizard’s stress. The first meter of all runs was filmed at 250 frames sec-1

from a dorsal view using a high-speed video camera (Fastec Troubleshooter LE). This

frame rate has shown to be sufficient to accurately measure the speeds and accelerations

observed in these lizards (Walker 1998, Vanhooydonck et al. 2006). The position of the

hip point was tracked and digitized using Kwon 3D (Kwon 1994). The hip position data

was smoothed using a quintic spline in the pspline package (Ramsey & Riley 2010) in the

R computing environment (R development Core Team 2010). Instantaneous acceleration

was calculated by differentiating the smoothed hip position data twice (Walker 1998).

Maximum instantaneous acceleration was considered the maximum acceleration of an

individual over the five runs.

Lizards were given a minimum of 24 hours of rest after their last running trial

before maximal exertion trials. Maximal exertion is defined here as the time and distance

a lizard can run before they reach exhaustion. In maximum exertion trials, lizards were

chased by clapping and tail-tapping around a four meter oval track (marked at 0.25m

intervals) until exhaustion defined here as the point at which an individual refused to run

after 10 consecutive tail taps or loss of righting response (Bennett 1980, Mautz et al.

1992, Vanhooydonck et al. 2014). Maximal exertion was measured in two ways: (i) as

distance covered; and (ii) time to exhaustion (timed with a stopwatch; Bennett 1980,

Mautz et al. 1992, Huyghe et al. 2007, Vanhooydonck et al. 2014).

20

Phylogeny

The desert Southwest lizard species used in our study includes 18 species of

Phrynosomatidae and three additional species. For the 18 species of Phrynosomatidae, we

used the topology and branch lengths from the most comprehensive phylogeny of the

family to date, which included 122 species and used six nuclear and five mitochondrial

genes (Wiens et al. 2010). Retaining branchlengths, we pruned the Wiens topology to

include only the 18 species in this study using the ape package (Paradis et al. 2004) in R.

Our estimate of the relationships of the three species not included in the

Phrynosomatidae phylogeny with respect to the Wiens topology are based on previously

established familial relationships (Townsend et al. 2004, Vidal & Hedges 2005, Wiens et

al. 2012, Pyron et al. 2013). We then combined previously published and new sequences

to estimate branchlengths. Sequences for Elgaria kingii (12s: AY525103.1, 16s: this

study, ND1: AF407538.1, ND2: AF407538.1, ND4: AY605103.1) and A. tigris (BDNF:

EU402619.1, PRLR: HQ130585.1, RAG1: EU402829.1, TRAF6: EU391053.1, 12s:

AY046452.1, 16s: AY046494.1, ND1: HM160771.1, ND2: U71332.1) were downloaded

from Genbank. The 12s, 16s, and ND1 genes for A. uniparens and the 16s sequence for

E. kngii were newly sequenced for the purpose of this study. Sequences for all 21

species were aligned using ClustalX. We used the GTR+ Γ model, selected based on the

Akaike Information Criterion (AIC) using jModelTest, for our analyses. We used

Maximum Likelihood analyses in PAUP 4.0 (Swofford 2003) to estimate the

relationships and branchlengths of all 21 species. We then scaled the branchlengths of

the three additional species so that they were on the same time scale as the Wiens et al.

(2010) tree and grafted these branches to the Wiens et al. (2010) topology and

branchlengths. The resulting tree agrees with previously published phylogenies on the

familial relationships (Townsend et al. 2004, Vidal and Hedges 2005, Wiens et al. 2012,

Pyron et al. 2013) while preserving the topology and branchlengths of Wiens et al.

(2010).

Comparative Analyses

All performance data were log-transformed prior to analysis. Four aspects of

performance, sprint speed, acceleration, and time and distance run to exhaustion were

21

analyzed in a phylogenetic context. The performance data was modeled using Hansen’s

(Hansen 1997) model, an Ornstein-Uhlenbeck (OU) process that models the evolution of

a continuous phenotypic trait subject to the influences of noise and selection (Hansen

1997, Butler & King 2004). The Hansen model describes the change in a trait dX(t), over

time (t), as an increment of a stochastic Brownian motion process (dβ(t)), influenced by

selection (α) toward an optimal trait value (θ) and subject to stochastic changes

proportional to a noise parameter (σ):

dX(t) = α [θ (t) – X(t)] dt + σdβ (t) (1)

Importantly, the Hansen model allows theta to vary along the branches of the phylogeny

to represent shifts in selective regime (or 'adaptive zone' sensu Simpson (1953))

experienced by the evolving lineages. We produced four mappings of selective regimes

on the phylogeny, each representing a distinct evolutionary hypothesis for the evolution

of the trait (sensu Butler & King 2004), see selective regimes below). Each evolutionary

model was fit to the log-transformed data, assuming our phylogeny using the OUCH

software package (Butler & King 2004) in the R statistical computing environment (R

Development Core Team 2010). The fits of each model were compared using the Akaike

information criterion corrected for small sample size (AICc, Butler & King 2004).

Information criteria were used to measure the strength of evidence in support of each

competing model (Burnham & Anderson 2002). Model selection frequencies and model

parameter confidence intervals are based on 2000 bootstrap replicates (Burnham &

Anderson 2002).

Selective Regimes for OU model

We tested four evolutionary scenarios, including Brownian motion and three

selective drivers, for producing the diversity of locomotor abilities observed in extant

species. The Brownian motion (BM) model is a non-adaptive model. This model

assumes that locomotor performance evolves according to random drift, and makes no

assumptions about adaptation with respect to the selective pressures considered here.

22

The first adaptive model is based on microhabitat use and contains five optima:

terrestrial, arboreal, saxicolous, generalist, and litter as defined below (Figure 2.1).

Species were placed into habitat regimes based on published descriptions of microhabitat

use (Table 1S in Appendix). Because most lizards are capable of using multiple

microhabitats, for a species to be placed into a specific regime the literature had to

describe a strong association/preference for a specific microhabitat. Species were

considered terrestrial if the literature described them as spending most of their time on the

ground. Species were placed in the arboreal category if they are strongly associated with

trees or woody bushes, climbing among their branches or trunks. Saxicolous species

were those that showed an association with rocks or vertical rock surfaces. Species were

placed in the clutter category if their microhabitat descriptions specifically included

cluttered areas such as under logs or rocks, or in leaf litter. Finally, species were

considered generalists if the literature didn’t specify a specific preference/association or

the species used multiple habitats.

The second model is based on predator escape behavior and has three optima:

flight, crypsis, and mixed (Figure 2.1; see Table 1S in Appendix for classifications).

Species were classified based on published descriptions of escape behavior (Table 1S in

Appendix). Cryptic escapers (CE) are lizards that rely heavily on crypsis or morphology

to evade predators with little reliance on running. Flight escapers (FE) are species that

rely mainly on bursts of sprinting for escape, whereas mixed escapers (ME) are species

that use a mix of cryptic behavior and short sprints to refugia, or other behaviors (e.g.,

“squirreling”) to avoid predation.

The third adaptive model is based on foraging mode and contains three optima.

Lizard species were classified as sit-and wait predators (SWF), active foragers (AF), or

mixed foragers (MF) based on movement criteria. SWF predators tend to spend long

periods of time motionless, and strike out for prey in very sort and rapid bursts. In

contrast, AF lizards move much more frequently, although at slower speeds (i.e.,

constantly on the go and a moderate pace). Previous authors have demonstrated that

foraging modes can be classified based on percent time moving (PTM) and movements

per minute (MPM) data (Figure 2.1, Table 1S in Appendix), based on the following

criteria: SWF (PTM<10% and MPM<1), MF (MPM>1 and PTM10% < 25%), and AF

23

(MPM >1 and PTM >30%, (Butler 2005, Reilly et al. 2007). Since published PTM and

MPM data were not available for E. kingii, data for the family Anguidae were used,

placing it in the AF category. Most species were readily classified using these criteria,

with some exceptions. The genus Phrynosoma was modeled as MF because its PTM

places it in the AF category, but near the MF category (Reilly et al. 2007), whereas the

MPM designates it as SW (Shaffer & Whitford 1981). Furthermore, the patchy

distribution of Phrynosoma’s food source (largely ants) results in a strategy of actively

searching for food patches followed by a sit-and-wait strategy once a patch is found.

Thus, they exhibit an intermediate foraging strategy between sit-and-wait and active

foraging strategies (Munger 1984, Butler 2005, Perry 2007). Sceloporus graciosus was

modeled as SWF despite an MPM of 1.31, due to its low PTM. Aspidoscelis uniparens

was modeled as an AF despite MPMs less than 1 due to their high MPM (70%<).

The adaptive models of evolution were constructed by assigning adaptive regimes

to individual branches of the phylogeny based on independent ecological data or

descriptions for the extant species (Table 1S in Appendix) with ancestral regimes

reconstructed based on linear parsimony using Mesquite (Maddison & Maddison 2009).

Linear parsimony designates regimes on internal branches that minimize the number of

evolutionary changes across the phylogeny. However, previous study that included more

taxa and specifically more basal species estimated ME to be most basal character state in

this group of lizards (Scales et al. 2009), and so we again used this value (Scales et al.

2009).

We used a principal components analysis (PCA) including all four performance

variables to visualize the performance phenotypes of these lizards in performance space.

We then used linear regression to determine size-performance associations. We also

assessed the correlations between the PICs of each of the four performance variables to

determine whether they display positive correlated evolution, no relationship, or negative

correlated evolution indicative of trade-offs. PIC were performed assuming our

phylogeny and the correlation between each performance trait was determined using

Pearson’s correlation in R. All PIC were performed using the ape package (Paradis et al.

2004) in R.

24

Results and Discussion

Adaptive evolution of performance

We tested four evolutionary hypotheses for the evolution of performance: a

hypothesis of neutral evolution modeled by Brownian motion (BM) and three selective

hypotheses modeled as an Orstein-Uhlenbeck process. These models of adaptation

differed in the source of selective pressure: needs related to predator escape; foraging

strategy; or habitat use (Figure 2.1).

Predator Escape Pressure. We found that sprint speed and acceleration were each

best explained by evolution in response to selection related to predator escape pressures

(Table 2.1). The predator escape model outperformed all others by both AIC.c and SIC

model selection criteria for both sprint speed and acceleration (Table 2.1), with the habitat

use and foraging mode models performing relatively poorly. Bootstrap model selection

frequencies indicate strong support for the predator escape model with regard to speed

and acceleration, as it was selected at least 91% of the time in both cases (Table 2.1).

The evolutionary optima estimated for the predator escape model reveal a trend of

increasing sprint capabilities with greater reliance on speed to escape predators (Table

2.2). Species that use a sprint escape strategy are evolving toward optima for very

effective sprinting, whereas cryptic species evolve toward optima with poor sprinting

capabilities, and generalists are intermediate.

Many studies have established the link between predation intensity and locomotor

performance (as discussed in the introduction), and between sprint performance and

fitness (Watkins 1996, Miles 2004, Husak 2006, Langerhans 2009a) within species.

However, it has rarely been demonstrated that differences in predator escape behavior can

drive performance diversity across a wide range of species (but see Scales et al. 2009).

Our analyses indicate that predation is a strong selective pressure that has resulted

in the diversification of species in terms of speed-related locomotion capabilities. We

tested this hypothesis against strong alternatives. It is not the case that the most complex

model provided the greatest explanatory power, as the habitat model and foraging mode

25

models had as many or more parameters than the predator escape model. Interestingly,

the BM model often outperformed these alternative models (Table 2.1), suggesting that

using the BM model is better than using a misleading ecological model.

Furthermore, an open question is whether the alternative strategies each require

distinct physiological or behavioral adaptations. Interestingly, cryptic and mixed

strategy lizards show substantial overlap in sprint speed and acceleration abilities, while

flight escape species do not (Table 2.2). An ad hoc evolutionary analysis based on a

predator escape model including only two selective regimes, flight specialists and all

other lizards, performed better than our original predator escape model for sprint speed

(ΔAIC.c =-3.30, ΔSIC = -2.84) and acceleration (ΔAIC.c=-1.87, ΔSIC = -1.41). This

suggests that there may be strong selection on sprint specialists above a general minimum

level of performance for all other lizards.

Foraging strategy. While predator escape behavior drives the diversity of speed-

related performance, it does not strongly influence the evolution of maximal exertion.

Instead, the foraging mode hypothesis best explained the evolution of both metrics for

maximal exertion with little support for any other models (Table 2.1). Furthermore, there

was very strong support from bootstrap model selection frequencies as the foraging

model was selected at least 95% of the time according to AIC.c criteria for both exertion

variables (Table 2.1).

The foraging mode model predicts that active foragers should move long

distances and periods of time before exhaustion, whereas sit-and-wait predators should

move shorter distances over a much shorter period of time prior to exhaustion, and mixed

foragers should be intermediate in both (Table 2.3). The estimated optima values are

similar to the measured performance capabilities of lizards in each respective regime

(Table 2.3) and support an evolutionary trend of increased maximal exertion with higher

levels of foraging movement.

Maximal exertion has received less study than speed, but lizards exhibiting

different foraging behaviors are known to differ in energy budgets and expenditure, and

field metabolic rate (Brown & Nagy 2007 and references therein). Furthermore, previous

studies have found that actively foraging species display higher exertion than sit-and-wait

26

species (Bennett 1994 and references therein). Thus, active foragers may require higher

maximal exertion due to high activity levels, whereas sit-and-wait predators may rarely

use their maximal exertion capabilities.

Habitat use. We did not find a strong association between microhabitat use and

locomotor performance. This result is surprising as it is in conflict with previous studies

(e.g., Losos 1990, Melville & Swain 2000, Kholsdorf et al. 2001, Butler and Losos 2002,

Herrel et al. 2002). One potential reason for this difference is that there may be multiple

strategies that are equally effective in a particular microhabitat (e.g., Liolaemus lizards:

Schulte et al. 2004). For example, P. modestum, C. texanus, and A. tigris, share the same

microhabitat and were observed together in a single desert wash during this study, but

they nevertheless exhibit the full range of predator escape and foraging strategies

examined here (Stebbins 2003, J. Scales, pers. obs.). In contrast, the well-studied Anolis

genus shows a strong association between microhabitat and locomotor performance (e.g.,

Losos & Sinervo 1989, Losos 1990). This difference may stem from the fact that Anolis

species are more closely related than the species in this study, and have evolved less

extreme behavioral strategies (e.g., Vanhooydonck et al. 2007, Johnson et al. 2008) such

that shifts in microhabitat use have promoted predictable evolutionary diversification in

locomotion associated morphology and performance (Losos 2009). Here we show that

locomotor performance is not fundamentally linked to microhabitat. Nevertheless,

microhabitat use may still play an indirect role by setting the bounds within which

specialization can occur (Butler, et al. 2000, Mellville & Swain 2003, Collar et al .2010).

In dragon lizards, arboreal specialists showed limited morphological variation while

terrestrial lizards were much more variable (Collar et al. 2010). Other studies have found

that predator escape behavior and locomotor performance were associated with and in

some cases limited by microhabitat use (Melville & Swain, 2003). Thus, some

microhabitats may facilitate behavioral variation whereas other microhabitats may be

more limiting.

27

The reality of performance types

A principal component (PC) analysis showed that variation in performance among

these lizards can be summarized by two principal component axes: a sprint axis (PC1;

acceleration and sprint speed; Table 2.4), and an exertion axis (PC2; distance and time

run to exhaustion; Table 2.4). Therefore, we will hereafter use the term “sprint

parameters” to refer to maximum sprint speed and maximum acceleration collectively,

and “exertion” to refer to time and distance to exhaustion. Three main groups separate in

this performance space: 1) lizards that are slow runners but can run for a long time 2)

lizards with low exertion, but range from slow to fast runners, and 3) lizards that run fast

and can do so for long periods (Figure 2.2). These groupings correspond to a

combination of the two selective pressures favored in our evolutionary analyses above.

These predator avoidance and foraging behaviors are important organizing factors

of locomotor performance. If we map selective regimes onto the PC performance space,

species separate into discrete clusters of exertion abilities according to foraging mode

(Figure 2.2). Similarly, mapping of the predator escape selective regimes reveals

distinctive densities of performance phenotypes by sprint ability, but with some overlap

(Figure 2.2). Interestingly, we find that species can excel along both axes. In conjunction

with the expected fast species with low exertion and slow species with relatively high

exertion, species (e.g., A. tigris) can be active foragers with high exertion and still

accomplish high-speed sprints. Conversely, some species perform poorly in both aspects

showing slow speeds and low exertion. Poor overall performance is an interesting case.

Have these species made behavioral modifications to reduce dependence on locomotion,

or do they live in predator-poor environments?

Effects of body size on performance

Many aspects of the structure, physiology, and function of organisms scale with

size, including locomotor capabilities (e.g., Schmidt-Nielsen 1984, Heglund & Taylor

1988). Surprisingly, we do not observe a strong influence of size on performance here,

with low explanatory power, particularly for sprint speed, acceleration, and time to

exhaustion (r2 values ranging from -0.05 to 0.07 for regression of snout-to-vent length

with performance variables, Table 2.5). Distance to exhaustion showed the strongest

28

relationship with size (r2 = 0.25, P = 0.012). Taken together, these results show the small

and large lizards have similar sprint capabilities, but larger lizards cover greater distances

in the same amount of time before reaching exhaustion. For example, S. graciosus, a

small lizard, and S. clarkii, a relatively large lizard, ran for similar durations, yet S. clarkii

covered almost 10 more meters (Table 2S in Appendix). The scaling of distance to

exhaustion with size may be explained by stride length, which commonly scales with

body size (e.g., Cumming & Cumming 2003, Clemente et al. 2012). Thus, it is possible

that the amount of time animals are able to sprint is governed by physiological limitations

unrelated to size, whereas the distance covered scales with size via limb length and size.

Indeed, time and distance to exhaustion are not correlated (PIC regression: r2= 0.22, p =

0.34, t = 0.97, Figure 2.3).

Trade-offs in performance

Several investigators have proposed a trade-off between speed and acceleration,

either because body size will have opposing effects on the two performance traits (Hill

1950), or because of the force-velocity trade-off predicted by muscle fiber contractile

properties in which increased contraction speed, beneficial for speed, results in low force

production that should be detrimental for acceleration (Vanhooydonck et al. 2006).

However, we observed no such trend. Instead, acceleration and sprint speed show strong

positively correlated evolution (PIC correlation: r2 = 0.84, p < 0.001, t= 6.49, Figure 2.3),

similar to other studies (Huey and Hertz 1984, Vanhooydonck et al. 2006). Since both

traits are important in predator escape, this correlation could be the result of a common

selective pressure (e.g., this study, Huey & Hertz 1984, Walker 2005). Alternatively, it

may be the result of the same design traits improving both aspects of performance.

Vanhooydonck et al. (2006) demonstrated that variation in many of the same

morphological traits were associated with increases in acceleration and sprint speed in

Anolis species. How tight the linkage is between sprint speed and acceleration remains

an open question.

Another commonly invoked trade-off occurs between exertion and sprinting

(Vanhooydonck et al. 2001, Herrel & Bonneaud 2012), and we find evidence for this

trade-off here. However, the relationship is complex and depends on which metric is used

29

for exertion. Faster lizards run farther before exhaustion and show positive correlated

evolution between these traits (PIC correlations speed: r2= 0.53, p= 0.017, t =2.64; and a

marginal correlation with acceleration: r2=0.42, p=0.064, t = 1.97; Figure 2.3). On the

other hand, faster lizards are sooner spent, indicating an evolutionary trade-off between

speed and time to exhaustion (PIC correlations: speed: r2= -0.54, p =0.015, t = -2.69;

acceleration: r2= -0.54, p = 0.014, t=-2.72, Figure 2.3). The physiological characteristics

of muscle fiber types are thought to underlie this trade-off, with fast, powerful muscle

fiber types tiring quickly whereas slower fiber types are much more fatigue resistant (e.g.,

Bonine et al. 2005, Wilson & James 2004). Indeed, we see this performance relationship

in most of the species included in this study. The horned lizards (Phrynosoma) and E.

kingii showed relatively high exertion, but low sprint parameters and have a high

proportion of more fatigue resistant muscle fibers (Bonine et al. 2005). Conversely, the

Sceloporus group and sand lizards have higher sprint parameters and low exertion

accompanied by higher proportions of fast, powerful fibers (Bonine et al. 2005).

Even within this general trend, however, there are interesting exceptions. The

Aspidoscelis species, which have the highest exertion, also exhibit high sprint speeds and

accelerations, especially A. tigris (Figure 2.3). Interestingly, species in the genus

Aspidoscelis are also the only species that experience selection for both performances

according to our model. All other species experience selection for relatively high speeds

coupled with relaxed selection for exertion, or selection for higher exertion accompanied

with relaxed selection for speed. This finding implies that the trade-off is not obligatory

(see Holzman et al. 2011), and selection to perform well at all of these aspects of

locomotion can result in a locomotor design with no apparent trade-off.

How is this trade-off circumvented? It may be that the expected trade-off is not

present because trade-offs at the muscular level do not translate to the whole-organism

level (Wilson et al. 2002, Vanhooydonck et al. 2014). Thus, our expectation for a trade-

off between speed and exertion is unfounded. Additionally, the two performances may be

determined by different traits. Whereas sprint parameters are correlated with traits such

as power output and limb length, exertion may be determined by other physiological

factors such as the circulatory system, blood buffering, or enzymatic activity. In this case

we again would not expect a trade-off. Finally, it may be that there are alternative designs

30

to achieve the same performance ability (many-to-one mapping Wainwright et al. 2005,

Holzman et al. 2011) so that when there is selection for both exertion and speed, the

complexity of the locomotor system can accommodate both (Holzman et al. 2011). How

the Aspidoscelis species accomplish both high speeds and exertion is an interesting

question for future study that may shed light on the limitations of performance.

Summary

Here we demonstrate that multiple approaches are essential not only in

understanding phenotypic diversity, but also for understanding what factors are

responsible for the evolutionary diversification of phenotypes. Our results indicate that

multiple sources of selection are responsible for the locomotor diversity in lizards. Both

the need to flee predators as well as the efficient food acquisition are powerful selective

factors, but they act on different aspects of the locomotor system. Behavioral variation

can therefore play a significant role in shaping phenotypic diversity. Furthermore, the

functional integration of the locomotor system does not necessarily limit the potential

combinations of locomotor performances. We generally observe an expected locomotor

trade-off between speed and exertion, but our analyses show that it is possible to

circumvent this trade-off when selection is strong and acting on multiple performance

axes.

31

Table 2.1. Performance of alternative models for the evolution of locomotor performance.

Model fit statistics for sprint speed, acceleration, time and distance to exhaustion (Ex).

The model with the best fit is in bold and listed first. Bootstrap selection frequencies are

listed in parentheses. The Best model was predator escape for both sprint speed and

acceleration and Foraging Mode for both time and distance to exhaustion

Sprint Speed

Acceleration

Time to Ex

Distance to Ex

Model AIC.c AIC.c AIC.c AIC.c

Predator Escape 0 (92%) 0 (97%) 2.4 (2%) 18.3 (0%)

Brownian Motion 2.8 (1%) 12.5 (0%) 7.7 (1%) 19.8 (0%)

Foraging Mode 4.0 (6%) 6.9 (2%) 0 (95%) 0 (100%)

Habitat 6.8 (1%) 16.3 (0%) 12.4 (2%) 33.9 (0%)

Table 2.2. Model parameters estimated for the predator escape model, the best fitting

sprint speed and acceleration model. Estimated optimal values () for sprint speed (m/s)

and acceleration (m/s2) data given in columns. The predator escape model posits that

lizards are in selective regimes for cryptic escapers (CE), mixed escapers (ME), and

flight escapers (FE). Bootstrap 95% confidence intervals are in parentheses. Measured

data are means of performance data for species grouped by selective regime.

Parameter

Sprint Speed

(m/s)

Measured Sprint

Speed (m/s)

Acceleration

(m/s2)

Measured

Acceleration

(m/s2)

35.4 (13.0,

1284) 56.57 (18.37,

1170)

6.16 (1.80,

201)

3.72 (0.86,

81.5)

CE 1.94 (1.33,

2.92) 2.21 (0.60)

25.96 (20.27,

31.80)

27.92

(6.47)

ME 2.18 (1.76,

2.73) 2.41 (0.79)

30.29 (26.77,

34.25)

30.43

(5.49)

FE 3.78 (2.85,

5.11) 3.69 (0.42)

43.15 (36.53,

51.44)

43.00

(10.04)

32

Table 2.3. Parameters estimated for the foraging mode model, the best fitting maximal

exertion model, and actual measured performance. Estimated optimal values (q) for the

distance (m) and time (s) run to exhaustion data are presented. The foraging mode model

posits that lizards are in selective regimes for active foragers (AF), mixed foragers (MF),

and sit-and-wait foragers (SWF). Bootstrap 95% confidence intervals are in parentheses.

Measured data (standard deviations in parentheses) are means of performance data for

species grouped by selective regime.

Parameter Distance θ (s)

Measured

Distance

(m) Time θ (s)

Measured

Time (s)

α 418.5 (58.07, 612.3) 197.1 (26.32, 637.2)

σ 56.51 (8.09, 92.10) 11.64 (1.16, 44.2)

θAF 38.69 (29.09, 52.54)

40.16

(12.37) 239.41 (199.1, 291.1)

239.72

(14.92)

θMF 29.76 (22.27, 40.28)

30.06

(5.35) 126.70 (104.3, 172.9)

127.12

(12.40)

θSWF 20.52 (18.05, 23.47)

21.30

(6.12) 59.75 (54.81, 65.40)

60.92

(12.16)

Table 2.4. PC analysis loadings and the percent of variation explained by each PC axes

for four performance variables. PC 1 can be interpreted as primarily a linear function of

speed and acceleration, while PC 2 consists of the maximal exertion variables. Together

the two PC’s explain 92% of the variation in performance. Only correlation values greater

than 0.200 are shown.

Performance PC1 PC2

Acceleration 0.662

Sprint Speed 0.669

Distance to Exhaustion 0.309 0.665

Time to Exhaustion 0.737

Variation explained 51% 41%

33

Table 2.5. The relationship between size and performance. Linear regressions of

performance on size (snout-to-vent length) show no influence of size on sprint speed,

acceleration and time to exhaustion. Only distance to exhaustion is significantly related to

size.

Performance Variable r2 F P-value

Sprint Speed 0.01 1.21 0.28

Acceleration 0.07 2.6 0.12

Time to Exhaustion -0.02 0.65 0.43

Distance to Exhaustion 0.25 7.78 0.012

34

Figure 2.1. Adaptive evolutionary hypotheses for the evolution of locomotor

performance abilities. Each selective regime is represented by a different color, and color

codes are located below each hypothesis. Branchlengths are not to scale.

35

Figure 2.2. Performance phenotypes reflect the locomotor demands of foraging mode and predator escape. High scores on PC1

indicate fast sprinters while high scores on PC2 species with high exertion. A) Exertion abilities separate into discrete clusters by

foraging mode (green = sit-and-wait, purple=mixed, active= orange) with active foragers having high exertion and sit-and-wait

predators have low exertion. B) Sprints vary by predator escape behavior with sprint specialists reaching higher speeds (blue= cryptic

and mixed species, red= sprint specialists). Ellipses represent kernel density estimates for 30, 60, and 90% levels of each group.

36

Figure 2.3. Correlations of phylogenetic independent contrasts (PIC) between each of the

performance variables. (A) Sprint speed and acceleration show positive correlated

evolution, as do (B) sprint speed/acceleration and distance run to exhaustion. However,

there is negative correlated evolution between (C) sprint speed/acceleration and time run

to exhaustion, but no relationship between time and (D) distance run to exhaustion. Plots

for acceleration are excluded because they closely resemble those of sprint speed.

37

CHAPTER 3. HINDLIMB MUSCLE CROSS-SECTIONAL AREA REFLECTS

LOCOMOTOR PERFORMANCE AND MICROHABITAT USE IN LIZARDS

Abstract

Locomotor performance is essential to the fitness of many terrestrial vertebrates.

Therefore, determining what defines locomotor performance is an integral part of

understanding the evolution of morphology, physiology, and behavior of vertebrates. The

majority of vertebrate movement is powered by muscle indicating that muscle design

should impact performance abilities. Specifically, the cross-sectional area (CSA) of a

muscle is a significant determinant of force production capabilities, suggesting that CSA

may influence performance abilities. However, few studies examine how muscle CSA

actually impacts performance, especially in a comparative context. Here I examine how

the CSA of muscles of the hind limb are related to acceleration, sprint speed, and exertion

in 21 species of lizards. I find that speed and acceleration, but not exertion increase with

CSA, and the CSA of extensor muscles better predicts speed and acceleration abilities

than flexors. Furthermore, muscle CSA scales isometrically in these lizards, but sprint

specialists tend to have more CSA for their size, while horned lizards have less. PCA

also shows that fast lizards have increased extensor CSA, but the distribution varies.

Fast, terrestrial species have muscle distributed more at the ankle, whereas as climbers

have more muscle at the hip. Interestingly, species with similar speed and acceleration

abilities can have very different muscle distributions, suggesting that there is more than

one way to achieve high speeds and acceleration. I conclude that muscle CSA plays an

important role in defining some, but not all performance capabilities, and that the amount

and distribution of hind limb musculature reflects the locomotor demands of these

lizards.

38

Introduction

Locomotor performance is critical to the survival and reproduction of most

vertebrates and this locomotion is chiefly driven by muscle. In particular, muscle cross-

sectional area (CSA) is likely a key contributor to locomotor performance, as it is tightly

linked with the amount of force a muscle can generate (Alexander 2003). In terrestrial

runners force production is an important determinant of locomotor abilities such as speed

and acceleration (Weyand 2000, Roberts & Scales 2002, Curtin et al. 2005, Weyand

2010). Thus, one may expect the amount of CSA to be a significant determinant of

running performance (e.g., Crook et al. 2008). However, how muscle CSA relates to

performance and is designed for specific tasks is rarely tested, even though relationships

between muscle CSA, body size and how musculature is distributed across the limb

potentially play a significant role in governing locomotor abilities.

Body size is intimately linked to locomotor design and performance (e.g.,

Schmidt-Nielsen 1973, Alexander 2003, Iriarte-Diaz 2012, Clemente et al. 2012),

especially in the muscular system. In fact, the majority of variation in muscle function

observed in vertebrates can be explained by scaling effects (Medler 2002). Thus,

variation in size should influence the amount and distribution of muscle CSA. In animals

that experience similar locomotor demands and have comparable lifestyles, one may

expect muscle CSA to scale isometrically (i.e., in a geometrically similar manner) with

size. However, CSA seldom scales isometrically, but instead, frequently varies with

locomotor mode (Bennett 1996, Kikuchi 2010), posture (Alexander et al. 1981, Kikuchi

2010), behavior (Alexander et al. 1981, Pollock & Shadwick 1994, Bennett 1996) or

some combination thereof, suggesting that the CSA-size relationship plays a prominent

role in locomotion and may be an important design trait dictating the limits of locomotor

performance.

Along with size, the arrangement of muscle CSA should also impact locomotor

performance. Muscles are arranged into functional groups that vary in their role during

locomotion. For example, in running, a stride is composed of the stance phase, when the

limb is contacting the ground providing propulsion, and swing phase, when the limb is

moving through the air to be repositioned (Biewener 2003, Higham et al. 2011).

39

Individual muscles may serve a support and propulsion function, while others function to

swing the limb (Reilly 1995, Higham 2011). Muscle function can further vary by joint so

that musculature across joints has different purposes during locomotion (Alexander et al.

1981, Pollock & Shadwick 1994, Zaaf 1999, Reilly 1995). Accordingly, the distribution

of muscle CSA across joints commonly reflects performance specialties (Pasi & Carrier

2003), mode of locomotion (Kikuchi 2010, Bennett 1996, Michilsens et al. 2009), or

differences in lifestyle (Alexander et al. 1981, Pollock & Shadwick 1994, Myatt et al.

2011).

The divergence in function of the muscular groups suggests that although CSA is

important, not all functional groups may influence performance in the same manner (Zaaf

et al. 1999, Higham et al. 2011). Furthermore, because of the complexity in structure and

function of the locomotor system, it is possible that similar abilities may be achieved with

different designs (Alfaro et al. 2004, Wainwright et al. 2005). This many-to-one mapping

of form and function is thought to be an emergent property of complex systems such as

the locomotor system (Wainwright et al. 2005). However, how muscle CSA of different

functional groups influence performance abilities is rarely explicitly tested leaving gaps

in our understanding of how the locomotor system is constructed. From a functional

perspective, researchers may fail to recognize how overall musculature is designed to

achieve a given performance. From an evolutionary perspective, researchers may miss

out on how the muscle system is adjusted over time to meet the different locomotor needs

of organisms.

Here I examine the relationship between hind limb muscle CSA and locomotor

performance in lizards. Lizards serve as an exceptional model for locomotor studies due

to their tremendous diversity in locomotor behaviors and performance abilities. I

concentrate on lizards of the Southwest United States as they display a wide breadth of

locomotor abilities ranging from slow, cryptic specialists to fast, sprint specialists

(Bonine & Garland 1999, Bonine et al. 2001, Chapter 2). I include only hind limb

musculature as the hind limbs provide the propulsion for running in lizards (Curtin et al.

2005). First I test whether hind limb CSA influences three aspects of performance, sprint

speed, acceleration, and exertion. While acceleration and sprint speed are force related

performances and should improve with increased CSA, exertion is not. Thus, I expect

40

that the relationship between CSA and locomotor abilities will vary with aspect of

performance. Second, I examine whether the relationship between CSA and performance

varies by functional group (stance and swing) or by joint (hip, knee, and ankle). Finally, I

explore how the muscular design of the hind limb varies across species. Specifically, is

CSA a function of body size or is it related to locomotor demands so that variation in

CSA distribution is associated with specific performance abilities or lifestyle?

Materials and Methods

Fieldwork

Twenty species of lizards comprising three families were used in this study.

Three to 10 individuals of each species were collected from the wild in California,

Arizona, and New Mexico from June to August in each of 2009 and 2010 (Table 1S in

Appendix). After capture, all lizards were transported to the Southwestern Research

Station or a house (Sceloporus occidentalis) for locomotor trials. Lizards were kept

overnight to allow gut passage, and locomotor trials were performed on the day following

capture. Lizards were kept in cloth bags between performance trials with access to water

ad libitum. Trials were performed for three sequential days after capture. At the

completion of all performance trials lizards were humanely euthanized for morphological

study, or returned to their site of capture. All animal procedures followed University of

Hawaii IACUC protocol 09-463.

Performance Trials

Lizards were placed under a heat lamp for a minimum of one hour prior to all

performance trials to achieve field body temperature. Body temperatures were verified

before each trial using a cloacal thermometer (Miller & Webber Inc.).

For sprint trials, lizards were induced to run down a 3 x 0.15m wide trackway by

clapping, or lightly tapping the tail. Each individual was raced five times over a span of

two days. Lizards were given a minimum of one hour of rest between each trial and were

run a maximum of three times in one day in order to capture maximum acceleration and

41

maximum sprint speed. Only runs in which the lizard ran straight without jumping or

touching the wall were used for analysis.

Sprint speed was measured using a series of four infrared photocells placed every

0.25 m over the second meter of the trackway. The single fastest time between two

photocells over the five runs was used as an estimate of the individual’s maximum sprint

speed (Huey & Pianka 1981, Miles & Smith 1987, Vanhooydonck et al. 2001).

Acceleration was measured using high-speed video. The first meter of each run

was filmed at 250 frames sec-1 from a dorsal view using a high-speed video camera

(Fastec Troubleshooter LE). A white marker was glued dorsally, directly above the pelvis

of each lizard for digitizing position coordinates. The pelvis was chosen as it is an easily

located landmark near the animal's center of mass. The position of the hip point was

tracked and digitized using Kwon 3D (Kwon 1994). The hip position data was smoothed

using a quintic spline in the pspline package (v. 1.0-14) in the R computing environment

(R Development Core Team 2010). Instantaneous acceleration was calculated by

differentiating the smoothed hip position data twice. Maximum instantaneous

acceleration was calculated from the maximum acceleration value of an individual over

the five runs.

Lizards were given a minimum of 24 hours of rest after their last running trial

before maximal exertion trials. In maximum exertion trials, lizards were chased by

clapping and tail-tapping around a four meter oval track (marked every 0.25m) until

exhaustion, defined here as the point at which an individual refused to run after 10

consecutive tail taps or loss of righting response (Bennett 1980, Mautz et al. 1992,

Huyghe et al. 2005, Vanhooydonck et al. 2014). Maximal exertion was measured in two

ways: (i) as distance covered; and (ii) time to exhaustion (timed with a stopwatch;

Bennett 1980, Mautz et al. 1992, Huyghe et al. 2005).

Morphological and Muscle measurements

Snout-vent length (SVL, distance from the tip of the snout to the cloaca) was

measured using digital calipers (Mitutoyo Digimatic D6-6CX) and body mass was

measured shortly after capture with Pesola scales to the nearest 0.1 g. Three to six

individuals of each species were then used for muscle morphology measurements. The

42

majority of the hind limb muscles were removed from the right hind limb. Each muscle

was blotted dry and measured to the nearest 0.0001g using an Acculab Sartorius ACL-

80.4 scale. The fascicle length (the length of muscle fiber bundles, used here to estimate

fiber length) of a muscle was measured at mid muscle (the midpoint of the muscle length)

under a Zeiss DV4 stemi spot dissecting scope to the nearest 0.01mm using Mitoyo

Digimatic D6-6CX calipers (Biewener & Full 1992). Pennation angle () was estimated

by bisecting the muscle and measuring the angle of the fibers relative the tendon, bone, or

aponeurosis at mid-muscle (Biewener & Full 1992). These methods were applied

uniformly across species, and should provide adequate data for comparison across

species. The cross-sectional area (CSA) of a muscle was estimated as the muscle mass

(g) divided by fascicle length (cm), assuming a muscle density of 1.06g/cm (Pasi &

Carrier 2003). The effective CSA of pennate muscles was then estimated as CSA

multiplied by the cosine of the pennation angle (cos).

Phylogeny

To construct a phylogeny of the 20 species included in this study, I used the same

phylogeny constructed by Scales and Butler (see Chapter 1 Materials and Methods), but

pruned Elgaria kingii from the tree because of its overwhelming influence on analyses.

Thus, I provide only a brief description of its construction here. For the 18 species of

Phrynosomatidae, I used the topology and branchlengths from the Wiens et al. (2010)

phylogeny, which is the most comprehensive phylogeny of Phrynosomatidae to date

including 122 species and is based on 6 nuclear and 5 mitochondrial genes. I then

estimated the phylogenetic relationships of the remaining species to the Wiens et al.

(2010) tree based on previously established familial relationships. The resulting tree

(Figure 3.1) agrees with previously published phylogenies on the familial relationships

(Townsend et al. 2004, Vidal & Hedges 2005) while retaining the topology and

branchlengths of Wiens et al. (2010) for a majority of species included in this study.

Statistical Analyses

Log transformed species means for performance and morphology variables were

used for all analyses. To determine how hind limb scales with body mass, linear

43

regressions were performed on body mass and hind limb CSA. To correct for the effects

of size, I regressed each performance and CSA on body mass. I then used the residuals

from these regressions for further analyses.

Muscle data were analyzed with respect to functional group. Muscles were first

grouped into extensor and flexor muscles. The muscle groups were then further divided

by joint into hip retractors, knee extensors, ankle extensors, hip protractors, knee flexors,

and ankle flexors. The CSA of a functional group was the sum of the CSA of each muscle

included in that functional group.

Pearson’s product moment correlation was used to test CSA correlations between

functional groups. Linear regressions were also used to determine the relationship

between each performance and the muscle CSA of each functional group. The

distribution of extensor musculature was compared across species using principal

components analysis (PCA) on size corrected data.

Phylogenetic independent contrasts (PIC) were used for phylogenetic analyses of

the relationships between each performance variable and CSA (Felsenstein 1985). PICs

were calculated using the ape package in R, assuming our phylogeny (Paradis 2004).

Linear regressions forced through the origin (Garland et al. 1992) were performed on the

PIC of body mass and PICs of hind limb CSA to determine the relationship between

muscle CSA and body mass. To correct for the effects of size, I regressed the PIC of each

performance and CSA on the PIC of body mass. I then regressed the residual PIC of each

performance on the residual PIC of muscle CSA to determine the relationships between

CSA and each performance.

Results

Scaling of hind limb CSA

Hind limb CSA scales isometrically with body mass in these lizards (slope=0.675,

Figure 3.2, Table 3.1). While there is an overall trend of isometry, there are also

intriguing deviations from this trend. Phrynosoma species fall well below the general

trend line (Figure 3.2) suggesting that they have less CSA for their body mass.

Conversely, two sand lizards (C.texanus and C. draconoides), the Aspidoscelis species,

44

and C. collaris tend to have more muscle CSA for their body mass (Figure 3.2). From an

evolutionary perspective, hind limb CSA and body size show correlated evolution (Table

3.1).

Hind limb CSA and sprint performance

Speed and acceleration are positively correlated with total hind limb CSA (Figure

3.3, Table 3.2), such that lizards with relatively large hind limb CSA are fast runners.

However, these relationships become marginal when phylogeny is taken into account

(Table 3.2) and are primarily associated with the evolutionary split of the sand and

horned lizards (Figure 3.3). In fact, the contrast value at the node of the horned-sand

lizard split (3.66) is far more than a standard deviation (sd=1.23) from the mean

standardized contrast (0.22).

The relationship between CSA and the sprint performances varies with functional

group. Extensor and flexor CSA are highly correlated (r=0.93, t=10.92, p<0.001). Yet,

extensor CSA is more strongly related to speed and acceleration than flexor CSA (Table

3.2). Among the joints of the hind limb, ankle CSA is most strongly associated with

speed and acceleration (Table 3.3). Phylogenetic analyses also indicate that flexor and

extensor CSA show correlated evolution (r=0.94, t=11.58, p<0.01), and although the

relationships between muscle CSA and performance are now marginal, extensor CSA

shows a stronger evolutionary association with speed and acceleration than flexor CSA

(Table 3.2). Similar to total hind limb CSA, the evolutionary CSA-performance

relationship is correlated with the evolutionary divergence of the horned and sand lizards

(Figure 3.3).

Hind limb CSA and exertion

In contrast to the sprint performances, distance and time run to exhaustion are

unrelated to hind limb CSA (Figure 3.3, Table 3.4). However, according to phylogenetic

analyses, time run to exhaustion does show marginal negative correlated evolution with

hind limb CSA (Figure 3.3, Table 3.4). This negative relationship remains marginal for

extensor CSA, but is significant for flexor CSA (Table 3.4). Again, these evolutionary

trends are related to variation between the sand and horned lizard clades (Figure 3.3).

45

Horned lizards have relatively low CSA, but still run for a relatively long time, whereas

sand lizards have high CSA, but low times to exhaustion (Figure 3.3). The two

Aspidoscelis species, however, do not follow this trend as they have very long times to

exhaustion and high hind limb CSA (Figure 3.3).

Variation in CSA across the hind limb

Because extensor CSA best explains speed and acceleration performance, I

explore its variation across the hind limb in more depth. Extensor CSA generally makes

up more than half of hind limb CSA (Figure 3.4) and is commonly largest at the ankle,

but distribution varies widely across species (Figure 3.4). A PCA shows that variation

across the hip, knee, and ankle can be summarized by two axes: an axis of general

amount of CSA (PC1; ankle loading most heavily followed by the knee, Table 3.5), and

an axis of distribution between the hip and knee (PC2; hip contrasting with the knee,

Table 3.5). Together, the axes explain 97.5% of the variation in extensor CSA

distribution.

Several Sceloporines and the Urosaurus species have similar CSA distributions

with higher scores on PC2 suggesting that more CSA is at the hip (Figure 3.5 A). On the

other hand, the C. draconoides, C. texanus, and Aspidscelis species score high on PC1,

but relatively low on PC2 and C. collaris, has the lowest score on PC2 suggesting that

these species have high overall CSA but more of it is distributed at the ankle and knee

(Figure 3.5 A). Phrynosoma species score low on PC1 and intermediate to low on PC2

indicating they have low CSA in general (Figure 3.5 A).

Discussion

Muscle CSA and Size

The relationship between muscle CSA and body size can often reflect variation in

locomotor mode, behavior, or posture (Alexander et al. 1981, Pollack & Shadwick 1994,

Bennett 1996, Kikuchi 2009). Given the variety of behaviors and habitats used by the

lizards studied here, it is somewhat surprising that I find hind limb CSA scales

isometrically with body mass. One reason I may observe isometry is that these lizards all

46

use a sprawling posture typical of most lizards. Posture likely plays an important role in

the CSA-size relationship as hind limb CSA scaled close to isometry in birds with similar

postures (Bennett 1996), but scaled allometrically in mammals where posture and

locomotor habits differed (Alexander et al. 1981, Pollock & Shadwick 1994). Thus, the

sprawling posture of lizards may require a common CSA-mass relationship so that there

may be a fundamental “lizard” muscle design across species.

In spite of the overall isometric scaling of CSA and mass, there are several

deviations from this trend that are reflective of differences in lifestyle. Callisaurus

draconoides, Cophosaurus texanus, Crotaphytus collaris and the Aspidoscelis species all

rely heavily on sprinting to escape predators and experience selection for high speeds to

do so (Chapter 2). These lizards also have high hind limb CSA for their size.

Conversely, the horned lizards are cryptic specialists that rely little on locomotion to

escape predators, and experience relaxed selection for sprinting (Chapter 2).

Accordingly, they have reduced CSA for their size. Thus, while there maybe a typical

“lizard design”, the relative amount of hind limb CSA appears to be shifted up or down

according to the locomotor needs of a species.

CSA, acceleration, and speed

Hind limb CSA appears to be an important determinant of sprint performance in

these lizards as it was positively associated with both acceleration and sprint speed. This

correlation is likely because muscular force production is directly related to CSA

(Alexander 2003), and both acceleration and speed are linked to force. Acceleration is a

direct function of force (a= f/m) and runners only accelerate when a foot is in contact

with the ground, while sprint speed has been linked to the amount and rate at which force

is applied to the ground in runners (Weyand 2000, 2010). Extensors are generally active

when the foot is in contact with the ground providing propulsion (Reilly 1995, Nelson &

Jayne 2001) which explains why extensor CSA showed the strongest relationship with the

two sprint performances, a trend observed in at least one other species of lizard

(Sceloporus woodi, Higham et al. 2011). Furthermore, ankle extensors showing the

strongest association with sprint performances reflects that in lizards, propulsion is

47

powered by ankle extension in level running (Aerts et al. 2000, Fieler & Jayne 1998, Zaaf

et al. 1999).

Increases in hind limb CSA, especially in the extensors, are likely required to

meet the elevated force demands of both high speed and accelerations (Curtin et al. 2005)

indicating that the same design traits underlie the two performances. Vanhooydonck et al.

(2006) also observed a partial overlap in traits that determine sprint speed and

acceleration in Anolis lizards. Thus, speed and acceleration appear to be functionally

linked at the muscular level so that fast lizards also accelerate well, a trend frequently

observed in lizards (Huey & Hertz 1982, 1984, Vanhooydonck et al. 2006, Chapter 2).

Flexor CSA was also correlated with speed and acceleration although not as well

as extensor musculature. Given that extensors are important for providing propulsion for

acceleration and sprint speed, why do we see a similar increase in flexor CSA? Flexors

are generally active during running as they swing the hind limb through the air to

reposition the leg for the next ground contact (Snyder 1954, Reilly 1995, Nelson & Jayne

2001). As a result, increases in flexor musculature may be required to effectively swing

the increased extensor musculature of fast runners. Furthermore, faster running requires

increased recruitment of swing-phase muscles (Jayne et al. 1990, Marsh et al. 2004),

which are generally flexors (Snyder 1954, Reilly 1995). Therefore, both flexors and

extensors likely play a role in determining speed related performances. While flexor

CSA may not influence speed as strongly as extensor CSA, or limit performance within

species (Higham et al. 2011), it may partially contribute to the large-scale performance

differences observed between the species studied here.

The increases in hind limb CSA associated with sprint abilities may be especially

effective for improved performance if they are due to fast-glycolytic (FG) fibers. FG

fibers have larger diameters, force production, and speed of contraction, all of which

would improve acceleration and speed (Rome et al. 1990, Bonine 2001, Biewener 2003).

This association may be the case as the larger diameters of FG fibers would contribute to

muscle CSA, and many of the lizards with high CSA in this study also have high

proportions of FG fibers (Bonine et al. 2001, Bonine et al. 2005).

Contrary to the sprint parameters, the relationship between muscle CSA and

exertion is unclear, but CSA does not appear to be an important determinant of exertion

48

capabilities. Instead, distance may be determined by other traits such as limb and stride

length (Christiansen & Bonde 2002, Cumming & Cumming 2003, Irschick & Jayne

2000, White & Anderson 1994, Clemente et al. 2012) while time to exhaustion may be

the result of physiological factors such as blood buffering, heart size, and enzymatic

activities (Bennett et al. 1984, Bennett 1994, Bonine 2007). Thus, exertion and

speed/acceleration may be largely determined by different attributes. Separate design

traits underlying these performances would allow for the independent evolution of speed

and exertion, and may explain the ambiguous relationship between these performances

not only in these lizards, but in other species as well (Chapter 2, Tsuji et al. 1989, Wilson

et al. 2002, Herrell & Bonneaud 2012).

Our analyses indicate that the emergence of the sand and horned lizard clades was

an important event in the diversification of hind limb CSA and performance, a trend also

observed in the evolution of muscle fiber types (Bonine et al. 2001). The horned and

sand lizards occupy similar habitats but have diverged significantly in behavior. Sand

lizards have become sprint specialists that rely on speed for many aspects of life, whereas

horn lizards have become cryptic, feed mainly on ants and rely little on speed for most of

their behaviors (Vitt & Congdon 1978, Stebbins 2003, Bonine et al. 2001, Sherbrooke

2003). Thus, substantial divergence in lifestyle can be accompanied by large phenotypic

divergences, and these events may be times of considerable innovation. Furthermore,

that a single evolutionary event can result in such a significant divergence indicates that

evolution does not necessarily happen in small increments, but may occur in larger

evolutionary steps.

Distribution of extensor CSA and locomotor performance

Our PC analysis suggests that the amount of extensor CSA is likely related to

sprint performance, which corroborates the findings of the regression analyses. Slower

species tend to have lower hind limb CSA overall, indicating a lower reliance on speed

(Figure 3.5 B). Conversely, faster species generally have high overall muscle CSA,

especially at the ankle (Figure 3.5 B). Thus, the variation across this first PC axis likely

reflects a performance gradient.

49

Interestingly, most of the fast sprint specialists are also largely terrestrial

compared to the generalist and climbing species studied here, and have more ankle and

knee extensors relative to hip retractors (Figure 3.5 B, C). In level running, knee

extensors must support the body mass, and propulsion is powered by ankle extension

(Fieler & Jayne 1998, Zaaf et al. 1999, Aerts et al. 2000). Hence, higher extensor CSA at

the knee and ankle compared to the hip may be indicative of a more terrestrial lifestyle

(Figure 3.5 C). Accordingly, Zaaf et al. (1999) found that a terrestrial gecko species

exhibited much higher moments at the knee and ankle compared to an arboreal species.

On the other hand, a few Sceloporus (S. jarovii, S. clarkii, S. poinsettii) and

Urosaurus species tend to be climbers, either arboreal or saxicolous (utilizing vertical

rock faces, Stebbins 2003, Chapter 2 and references therein). These species also have

relatively more muscle CSA at the hip (Figure 3.5 C). S. occidentalis a more generalist

species, but an adept climber, was frequently observed climbing trees and rocks also

showed relatively high hip CSA (Figure 3.5 C). Hip musculature is important for

climbing as it supplies the propulsive forces to overcome inertia and counteract gravity

(Zaaf et al. 1999, Aerts et al. 2000). Zaaf et al. (1999) found higher hip moments in a

specialized climber compared to a terrestrial species of gecko, and Higham and Jayne

(2004) found increases in EMG amplitude at the hip (and knee) with increased inclines in

Chameleo calyptratus both suggesting that musculature at the hip is important for

climbing. Only, C. collaris, a rock dweller, does not show high extensor CSA at the hip.

However, this species often require open ground for running, and are not considered

exceptional climbers (Smith 1995). Thus, the distribution of CSA along the limb likely

reflects mode of locomotion suggesting that microhabitat use also has an important

influence on muscular design.

Our analyses also reveal that all of the performances studied here can be achieved

through multiple muscular designs, a trend called many-to-one mapping. Sprint speed

and acceleration both increase with hind limb CSA, but the muscle can be distributed

very differently along the limb and still achieve relatively high speeds and accelerations

(Figure 3.5 B). For example, C. texanus, which has high ankle CSA, and U. graciosus,

which has high hip CSA show similar accelerations (Figure 3.5 B). Horned lizards and

Aspidoscelis species both display relatively high exertion, but Aspidoscelis have high

50

hind limb CSA and proportions of FG fibers, whereas horned lizards have low CSA and

proportions of FG fibers (this study, Bonine et al. 2005). This many-to-one mapping

suggests that the locomotor system can accommodate multiple demands placed on it by

different tasks. Moreover, finding that many-to-one mapping occurs for the three

performances studied here indicates that it may be widespread in locomotor systems.

Summary

The relative amount and distribution of hind limb CSA is linked to locomotor

performance and reveals important information about locomotor modes and lifestyles.

While CSA increases with body mass, the relative amount shifts up or down according to

the locomotor demands of species. Hind limb CSA plays a role in determining speed and

acceleration, but not exertion, suggesting that different traits underlie these performances.

Large behavioral shifts appear to be especially important in the evolution of these CSA-

performance relationships. With respect to speed and acceleration, extensor CSA best

explains variation in performance, but CSA can be distributed in different ways.

Climbers, such as arboreal and saxicolous species tend to have more CSA at the hip,

while in terrestrial species CSA is largely concentrated at the knee and ankle. Thus, CSA

reflects variation in mode of locomotion and microhabitat use.

51

Table 3.1. Relationships between body mass and hind limb CSA. Results of non-

phylogenetic and PIC regressions of hind limb CSA on body mass. Italics indicate results

are from phylogenetic analyses.

Muscle CSA Slope r2 p F

Total Hind limb 0.665 0.80 p<0.001 78.03

0.73 0.88 p<0.001 146.80

Table 3.2. Relationship between hind limb CSA and acceleration and sprint speed.

Results of non-phylogenetic and PIC regressions of sprint performances on hind limb

CSA. Italics indicate results from phylogenetic analyses.

Muscle CSA r2 p F

Sprint Speed

Total Hind limb 0.42 0.001 14.84

0.12 0.079 3.47

Extensor 0.46 <0.001 17.36

0.12 0.076 3.55

Flexor 0.33 0.005 10.42

0.10 0.096 3.08

Acceleration

Total Hind limb 0.38 0.002 12.51

0.12 0.077 3.52

Extensor 0.41 0.001 14.08

0.13 0.067 3.80

Flexor 0.30 0.007 9.29

0.09 0.109 2.85

52

Table 3.3. Relationship between CSA at each joint and speed and acceleration. Results

multiple regression analyses of sprint performance and hind limb CSA by joint. Overall

models were significant. Speed: F= 5.626, p =0.008, r2= 0.42, Acceleration: F= 4.61, p =

0.017, r2 = 0.36.

Joint CSA Sum Sq Df F p

Sprint speed

Hip 0.00401 1 0.0770 0.7849

Knee 0.01635 1 0.3142 0.5829

Ankle 0.15229 1 2.9259 0.1065

Acceleration

Hip 0.0177 1 0.4484 0.5126

Knee 0.0058 1 0.1470 0.7065

Ankle 0.0674 1 1.7072 0.2098

53

Table 3.4. Relationship between hindlimb CSA and exertion variables. Results of non-

phylogenetic and PIC regressions of exertion performances on hind limb CSA. Italics

indicate results from PIC.

Muscle CSA r2 p F

Distance to exhaustion

Total Hindlimb -0.04 0.623 0.25

-0.05 0.858 0.03

Extensor - - -

- - -

Flexor - - -

- - -

Time to exhaustion

Total Hindlimb -0.01 0.370 0.85

0.14 0.059 4.07

Extensor -0.04 0.584 0.31

0.16 0.082 3.41

Flexor 0.05 0.170 2.05

0.16 0.044 4.70

54

Table 3.5. PCA loadings of CSA at each joint. The two PC axes explain 97.5% of the

variation in hind limb CSA. PC1 explains 81.3% of the variation and is largely related to

amount of ankle CSA, while PC2 explains 16.2% and is a contrast between muscle CSA

at the hip and knee. Only loadings greater than 0.2 are shown.

Joint PC1 PC2

Hip 0.38 -0.88

Knee 0.47 0.43

Ankle 0.80

81.3% 16.2%

55

Figure 3.1. Phylogenetic relationships of the 20 species of lizards included in this study.

The tree is scaled to one and branches are proportional to estimated branchlengths.

56

Figure 3.2. Relationship between log total hindlimb CSA and log of body mass for 20

species of lizards. Total hind limb CSA scales isometrically with body mass in these

lizards (slope=0.675, p<0.001, F= 81.35, r2= 0.819). However, several lizards that rely

heavily on speed (Crotaphytus collaris, Callisaurus draconoides, Cophosaurus texanus,

Aspidoscelis tigris, and A uniparens) have more CSA for their size (red), while horned

lizards (Phrynosoma cornutum, P. solare, and P, modestum), cryptic specialists that rely

little on speed, have less CSA for their mass (green).

57

Figure 3.3. Non-phylogenetic and phylogenetic relationships between hind limb CSA and

acceleration and the exertion variables (distance and time run to exhaustion). Sprint

speed is not shown as it shows trends very similar to acceleration. Acceleration and

sprint speed are positively correlated with hind limb CSA, but this relationship is

marginal when phylogeny is taken into account. Distance and time run to exhaustion are

unrelated to CSA, but time to exhaustion exhibits a marginal negative relationship

according to PIC. The CSA-performance relationships are associated with the

evolutionary split between horned and sand lizards.

58

Figure 3.4. The proportion of hind limb muscle CSA composed of extensors (A) and the

proportion of extensor CSA comprised of hip (dark gray), knee (light gray), and ankle

(white) muscles (B). Species are ordered from slowest relative sprint speeds

(Phrynosoma cornutum) to fastest (Aspidoscelis tigris). Extensor musculature makes up

at least half of hind limb CSA in all species. CSA at the knee or ankle make up the largest

proportion of extensor CSA, but distribution varies widely.

59

Figure 3.5. PCA of extensor muscle CSA distribution across the hip, knee, and ankle. PC1 represents the general amount of CSA, and

the ankle has the highest loading. PC2 reflects a contrast between hip CSA and knee CSA. B) PCA in relation to speed and

acceleration (black- species with fast speed or acceleration, red- species with intermediate speed and acceleration, green – species with

the slowest speed or acceleration). PC1 reflects performance abilities, with fast species generally having high amounts of CSA.

However, the CSA can be distributed largely at the ankle, knee or hip. C) PC2 reflects mode of locomotion related microhabitat use

(Climbing- black, generalist- red, terrestrial – green). Climbing species generally have higher CSA at the hip compared to the knee,

whereas terrestrial species often have more CSA distributed at the ankle or knee.

60

CHAPTER 4. RUNNING FOR YOUR LIFE OR RUNNING FOR YOUR

DINNER: WHAT DRIVES FIBER-TYPE EVOLUTION IN LIZARD

LOCOMOTOR MUSCLES?

Abstract

Despite its role in whole-animal performance, the adaptation of muscle

physiology remains little explored regarding terrestrial locomotion. We tested

evolutionary models based on predator escape and foraging strategies of lizards to test

whether fiber-type composition of a leg muscle is adaptive for behavior. The best-fitting

model for fast-twitch fiber-type evolution was one based on predator escape strategy,

while the foraging-mode model fared poorly (Akaike Information Criterion with small

sample size correction [hereafter, AICc], AICc = 29.7). According to the predator

escape model, lizards relying on sprints to avoid predators are predicted to have higher

relative proportions of fast glycolytic (FG) fibers (70%) while cryptic lizards are

predicted to have higher relative fast oxidative glycolytic (FOG) fiber proportions (77%).

This pattern suggests an evolutionary trend toward greater FG (respectively FOG) fiber

composition among lizards that specialize in sprinting (respectively crypsis). The best-

fitting model for slow-twitch fibers was a single optimum suggesting a common selective

pressure across these lizards. The second-best model explaining slow-twitch fiber-type

evolution was Brownian motion (AICc = 0.80), indicating some support for neutral

evolution. We find evidence suggesting that different fiber types occurring in the same

muscle can evolve under different evolutionary pressures.

Introduction

The morphological evolution of organisms is intimately linked to adaptation via

whole animal performance and behavior (Arnold 1983, Garland & Losos 1994). Lizards

have long served as a model system for locomotor adaptation because they exhibit a wide

variety of locomotor behaviors and abilities. Among ecologically differentiated lizards,

variation in gross morphology, especially external limb length, has evolved in association

61

with differences in sprint performance and environmental locomotor needs, thus

providing compelling evidence for adaptation (Losos & Sinervo 1989, Losos 1990,

Garland & Losos 1994, Bonine & Garland 1999, Irschick & Jayne 1999, Irschick 2002).

Two principal rationales are commonly invoked to explain the observed variation

in lizard locomotor behavior: selection associated with foraging strategy or with predator

escape behavior (Snell et al. 1988, Cooper 1994, Garland & Losos 1994, Miles et al.

2007). Lizards are particularly diverse in foraging strategy. Sit-and-wait (or ambush)

foragers remain motionless for much of the time, and rely on short bursts of locomotion

to capture prey that moves within close proximity. Alternatively, active foragers spend

much of the time moving in search of prey. Differences in foraging mode have been

shown to influence the evolution of life-history strategies covering such wide-ranging

aspects as behavior, diet, morphology, and physiology of lizards (Reilly et al. 2007). Sit-

and-wait predators often rely on short bursts of speed to capture prey, but because of their

brief durations of continuous movement, they often do not require high endurance

(Anderson & Karasov 1981). Conversely, the high activity levels of active foragers may

put a premium on locomotor endurance (Anderson & Karasov 1981). These different

demands imply that selection for acceleration and speed for overtaking prey or endurance

for continuous searching may influence the evolution of muscle fiber type in locomotion.

Predator escape behavior is commonly evoked as a selective force acting on the

morphology and performance of lizards (Snell et al. 1988, Garland & Losos 1994,

Vervust et al. 2007). While some species rely on crypsis to avoid detection by predators

(with some remaining extremely still), other species use sprinting to out-run predators.

Thus, lizards that escape by sprinting are presumably under stronger selection for high

accelerations and maximum speed than lizards that use a generalized strategy, while

lizards that rely on crypsis may experience relaxed selection. Relaxed selection can

increase the variation in a locomotor trait, or may increase the value of a given trait, if the

alternative, unused trait has a deleterious effect.

Locomotion is achieved using skeletal muscles. In vertebrates, muscles are

composite structures made up mostly of three common fiber-types: fast-twitch glycolytic

fibers, fast-twitch oxidative-glycolytic fibers, and slow oxidative fibers (Gleeson et al.

1980, Gleeson & Johnston 1987). Each of these fiber types differ in force generated,

62

speed of contraction, and fatigue resistance. Fast-twitch glycolytic (FG) fibers have fast

shortening velocities (V), and high maximum shortening velocities (Vmax, Rome et al.

1990). FG fibers can produce the high force and power needed for sprinting, but fatigue

quickly (Peter et al. 1972). Many animals use FG fibers for high-speed burst locomotion

(fish: Rome et al. 1988, frogs: Lutz et al. 1998; savannah monitor lizard, Varanus

exanthematicus: Jayne et al. 1990). By contrast, slow oxidative (SO) fibers have low V

and Vmax values, produce low force and power, but fatigue slowly (Peter et al. 1972,

Gleeson et al. 1980, Gleeson & Johnston 1987, Rome et al. 1990). Finally, fast-twitch

oxidative-glycolytic (FOG) fibers have intermediate V and Vmax values, generally

producing intermediate force and power, and show intermediate fatigue-resistance (Peter

et al. 1972, Gleeson et al. 1980, Gleeson & Johnston 1987). Given that there are limits

on total muscle mass (and hence, the number of muscle fibers), the fiber type proportion

of a muscle should reflect its functional requirements.

Bonine et al. (2001; 2005) found a large variation in muscle fiber type

composition among lizard species, and that lizards generally have a negative relationship

between FG and FOG fibers. In particular, Bonine et al. (2005) found that two small

clades of lizards (sand lizards which includes the genera Uma, Callisaurus, Cophosaurus,

and Holbrookia, and horned lizards which consist of the genus Phrynosoma) have

experienced an unusually large amount of evolutionary change in muscle fiber-type

composition over a short period of time. Based on these finding, the authors

hypothesized that natural selection may shape the iliofibularis fiber type composition of

these lizards. However, their Brownian motion (BM) model could not test whether

changes in fiber-type composition of the iliofibularis are under selection within the

contexts set by the ecology and behavior of the species of lizards studied.

The fiber-type composition of a muscle has been proposed to evolve for a specific

whole muscle function, but to date no study has provided strong evidence for such

adaptation. Here, we use the Bonine et al. (2005) data and the direct modeling approach

of Hansen (1997) and Butler and King (2004) to test whether fiber-type composition of a

locomotor muscle evolves under the influence of drift and selection. We explicitly model

different selective pressures (namely, foraging or predator escape strategies, a single

global optimum for all lizards, or random drift) and compare reasonable alternative

63

models that may be responsible for the evolution of fiber type composition in locomotor

muscles.

Materials & Methods

Muscle fiber-type data and assumptions

We obtained iliofibularis muscle fiber-type composition data from Bonine et al.

(2005), the largest interspecific dataset available for any locomotor muscle. The

iliofibularis is a hindlimb muscle that is active during burst locomotion and its muscle

fiber recruitment increases with running speed (Jayne et al. 1990). The iliofibularis is a

swing-phase muscle, and although it is not active during stance phase, it may still

contribute to faster running speed and endurance. At high running speeds, limbs are

cycled quickly, (Fieler & Jayne 1998), and some lizards use stride frequency to modulate

their sprint speed (Vanhooydonck et al. 2002). Additionally, although stance-phase

muscles consume most of the energy during running, a substantial amount of energy is

also used in swing phase muscles (Marsh et al. 2004, Marsh et al. 2006). Consequently,

the contractile properties of the iliofibularis may influence stride frequency, sprint speed,

and fatigue-resistance during continuous locomotion in lizards (Bonine et al. 2005).

Bonine et al. (2005) used independent contrasts and ancestral character state

reconstruction to study fiber type composition of the iliofibularis muscle in lizards. We

analyzed proportional cross sectional areas of each fiber-type (FG, FOG, SO) derived

from proportions reported in Bonine et al. (2005) for the iliofibularis. The evolution of

fiber-type proportions was modeled using the Hansen model, an Ornstein-Uhlenbeck

(OU) process with discrete switches of adaptive regime, applied to the analysis of

phenotypic evolution along phylogenies (Hansen 1997, Butler & King 2004). Hansen’s

(1997) OU process models both stabilizing selection and drift and allows for multiple

trait optima. The OU model (Equation 1) expresses the change in a quantitative

character, dX(t), over time, t, based on stabilizing selection toward an optimal trait value

and random drift.

dX(t) = α [θ (t) – X(t)] dt + σdβ (t) (1)

64

In this model, (t) is the optimal trait value, which can vary along any lineage; ,

measures the strength of selection, and the strength of drift. The Butler and King

(2004) approach uses the OU process to test alternative evolutionary hypotheses

concerning the location of different selective regimes (regimes are the combination of

environmental and organismal characters associated with exhibiting a particular behavior,

Baum & Larson 1991) on the phylogeny. Their method also allows the estimation of

optimal trait values for each selective regime. The model assumes that the trait of interest

remains close to a fitness optimum by selection and that the strength of selection can be

interpreted in a similar manner across different regimes (Hansen 1997). This method, like

all other comparative methods, assumes that the phylogeny accurately reflects the

evolutionary history of the organisms.

Phylogeny

We began with the phylogenetic hypothesis published in Bonine et al. (2005),

which provides a topology, but lacks the branch lengths that are required for applying the

OU model sensu Butler and King (2004). The Bonine phylogenetic hypothesis is a

composite of available phylogenies, each of which is well supported (see references in

Bonine et al. 2005). We revised the placement of the scincid and anguid branches of the

Bonine topology in light of the more recent phylogenetic hypothesis of Vidal and Hedges

(2005), which places the Scincidae more basal than Anguidae. We call this revised

topology the BVH topology.

We estimated the branch lengths on the BVH topology based on nucleotide

sequence data collected from the NCBI Entrez nucleotide database. We used

mitochondrial 12s rRNA gene sequences because this locus has data available for the

most species. For Laudakia stellio we substituted the con-generic L. stoliczkana, as these

species are a similar phylogenetic distance from Phrynosomatidae (Macey et al. 1997,

Macey et al. 2000). For two species (D. dorsalis, and P. sicula) suitable sequences could

not be located: we excluded these species from the analysis. All sequences were aligned

using ClustalX, resulting in a sequence dataset of 248 basepairs. The tree topology was

constrained to the BVH topology and the branch lengths were estimated using maximum

likelihood in PAUP. However, the mitochondrial 12s rRNA gene sequences could not

65

resolve the basal branch lengths between the Scincidae (Carlia and Eumeces) and

Lacertidae/Teiidae (Aspidoscelis and Acanthodactylus) split, or between the

Lacertidae/Teiidae (Aspidoscelis and Acanthodactylus) and Anguidae (Elgaria) split. In

these cases we used the branch lengths reported by Vidal and Hedges (2005).

Selective Regimes

We constructed five evolutionary scenarios for the proportion of fiber-types in

lizard iliofibularis muscles by assigning adaptive regimes to particular branches of our

phylogeny. The first adaptive model contains three optima and is based on the foraging

modes of each species. Debate continues as to whether foraging mode is a discrete or

continuous trait (e.g., Perry 1999, Cooper 2007, Reilly et al. 2007). Nevertheless, based

on multivariate movement data, several studies have demonstrated that sit-and-wait and

active foragers form distinct clusters that are validated statistically (Cooper 1994, Butler

2005, Cooper 2005, Cooper 2007). Thus, even if foraging mode is not strictly discrete, it

can usefully be described as modal.

Lizard species are classified as sit-and-wait foragers (SWF), active foragers (AF),

or mixed foragers (MF) based on published data on percent time moving (PTM) and

movements per minute (MPM, Table S3 in Appendix). Several studies have found

support for threshold values of 1 MPM and 10% PTM for sit-and-wait foragers (Butler

2005; Cooper 2007; Miles et al. 2007). Additional studies suggest that active foragers

have PTM greater than 30%, and species that fall between 10% and 30% PTM may be

mixed foragers (Reilly & McBrayer 2007). Miles et al. (2007) additionally found

significant morphological differences among foraging modes. We used these criteria to

classify lizards: SWF, (PTM <10% and MPM <= 1), MF (MPM > 1 and 10% < PTM <

30%), and AF (MPM > 1 and PTM > 30%). If published data on PTM or MPM were not

available for a species we used data for a related species that exhibits similar behavior

(eg. Eumeces laticeps for E. fasciatus), or for the family if only one foraging mode is

known to occur in the family (e.g., Anguidae, Miles et al. 2007). The genus Phrynosoma

was modeled as both SWF and AF because its PTM and MPM are from different studies

and would place them in different categories (Table S3 in Appendix). Sceloporus

graciosus was also modeled as both SWF and AF based on PTM and MPM, while E.

66

fasciatus was only modeled as AF due to the very high value of PTM used for this

species (Table S3 in Appendix). Our results and conclusions based on the different

models were qualitatively the same, so models from only one set of foraging mode

classifications are reported.

The second model also has three optima, but is based on predator escape mode

(flight, crypsis, and mixed). Lizard species were classified according to published

descriptions of escape behavior, because little quantitative data are available (Table S3 in

Appendix). Species were classified as flight escapers (FE) if their main mode of escape

relies heavily on bursts of sprinting, as cryptic escapers (CE) if they rely heavily on

crypsis to remain undetected with lower reliance on running, or mixed escapers (ME) if

they rely on both cryptic behavior, and short sprints to evade predators.

The third model of evolution assumed that both foraging and predator-escape

behaviors were important determinants of muscle fiber-type composition. Under this

model, each unique combination of foraging mode and predator escape behaviors

corresponds to a distinct adaptive regime. For instance, Holbrookia maculata is a sit-

and-wait predator that relies on crypsis so it is in the sit-and-wait/crypsis category

(SWF/CE), while Callisaurus draconoides is a sit-and wait predator that relies mainly on

sprinting from predators so it is placed in the sit-and-wait/flight category (SWF/FE).

Although there are nine logically possible combinations, the lizards studied here exhibit

only six of these combinations (Figure 4.1).

Ancestral regimes were reconstructed using linear parsimony. This algorithm

chooses regimes on internal branches to minimize the number of evolutionary changes

along the phylogeny. Linear parsimony resulted in a few ambiguities. For example, the

branch immediately ancestral to the Phrynosomatidae family could be assigned to either

mixed or flight escape with equal parsimony. In these cases, all possibilities were tested

and results compared (however, these variations did not affect the conclusions).

The final two models make no assumptions about adaptation. The simplest model

is Brownian motion, or random drift. The last model assumes stabilizing selection with a

single global optimum fiber-type composition for all species. It implies that all lizards

share the same adaptive regime, varying little outside of the optimal range of fiber type

67

compositions. This situation would be expected to occur if the functional demands of

locomotion are identical among all lizards irrespective of behavior or ecology.

Analysis

The fiber-type data in Bonine et al. (2005) were proportions that sum to one and

hence only have two degrees of freedom. The proportional nature of the fiber-type data

result in non-independence, and no manipulation will make them truly independent.

However, to minimize some of the problems associated with non-independence in our

analysis, we calculated the proportion of SO amongst the total (SO/(SO+FG+FOG)), and

that of FG only amongst the fast (FG/(FG+FOG). Thus, both of the proportions we

analyzed could range from 0 to 1 within a single lizard species, whereas the traditionally

calculated proportions of SO and FG would be constrained to range between 1-FG and 1

– SO, respectively. Consequently, the fiber type proportions we used allowed us to

capture the essential biological information in this system while reflecting the two

degrees of freedom present in the data. We note that had we modeled all three fiber-type

proportions separately, interpretation would have been more complicated because of

interdependency among the models. We logit transformed relative proportions of the

fiber-types to convert the data to a scale not bounded by zero and one. We used quantile-

quantile plots to diagnose deviations from normality.

We fit each of our evolutionary models to the muscle fiber-type data assuming the

BVH phylogeny. The fits of each of the models were compared using Akaike

Information criterion corrected for small sample size (AICc) and the more conservative

Schwartz information criterion (SIC). We use information criteria to assess the strength

of evidence in favor of the competing models (Burnham & Anderson, 2002). We fit

models of evolution using the OUCH package (version 1.2-4, Butler & King 2004) in the

R statistical computing environment (R Development Core Team 2007). Model selection

frequencies and 95% confidence intervals for model parameters were calculated based on

2000 bootstrap replicates (Burnham & Anderson 2002). All statistical analyses were

conducted in the R statistical computing environment (R Development Core Team 2007).

Once the optima for each model were calculated, these parameters were back-

transformed into proportions to facilitate interpretation.

68

Results

Fast-twitch (FG and FOG)

Predator escape mode was the best predictor of fast-twitch fiber type evolution.

In the best fitting model, evolution of fast-twitch (FG and FOG) fiber-type composition is

predicted by adaptive regimes related to predator escape strategy (Table 4.1). The

predator-escape model out-performed all other models by both model selection criteria

and showed both selection and drift to be important (Tables 4.1 & 4.2). Under the

predator-escape model, the fast-twitch optimum indicates that lizards that sprint to escape

should have very high proportions of FG fibers, while lizards that rely on crypsis should

have mostly FOG fibers (Table 4.2 & 4.5). Lastly, lizards that use a mixed strategy

should have an intermediate proportion with slightly more FG than FOG fibers (Tables

4.2 & 4.5).

Bootstrap model selection frequencies demonstrated that there was little support

for the Brownian motion, single-optimum, and combination models, while the foraging

mode model fared the worst (Table 4.1). The predator escape model was selected the

majority of the time using both the AIC and SIC (98.1% and 88.3% respectively). The

poor performance of the foraging mode and combination models indicates that foraging

mode does not substantially contribute to the evolution of fast-twitch fiber types.

Bootstrap values show overlap of 95% CI’s of optimal FG proportions for the flight and

mixed categories, but no overlap between the mixed or flight and the cryptic category

(Table 4.2).

The data presented here support an evolutionary trend toward muscles composed

mostly of one fast-twitch fiber-type, which is associated with increased behavioral

specialization. Two of the three behavioral strategies (flight and crypsis) exhibited

selection for a single fiber type, while the third behavioral strategy (mixed) showed

intermediate values of the two fast fiber types. The sand lizards, which rely heavily on

speed for predator escape (Vitt & Price 1982, Dial 1986, Bulova 1994), have high FG

fiber-type proportions (0.68), while the horned lizards that rely heavily on crypsis

(Sherbrooke 1987; Stebbins 2003), have very low FG fiber-type proportions (0.29). The

Sceloporus group (which consists of the genera Sceloporus, Uta, and Urosaurus), more

69

generalist in their behaviors (Sites et al. 1992; Stebbins 2003), had intermediate FG fiber-

type proportions (0.51). These fiber compositions associate well with sprint

performance, the sand lizards tend to reach the fastest sprint speeds, the Sceloporines

more intermediate, and horned lizards reached slower speeds (see Bonine & Garland

1999).

Slow-twitch oxidative (SO)

A model with a single optimum was the best fitting for the evolution of the

proportion of SO fibers in the iliofibularis, out performing all other models by both

model selection criteria (Table 4.3). The single optimum model predicts that only 8% of

the iliofibularis will be composed of SO fibers (Tables 4.4 & 4.5). However, the

proportion of SO fibers vary little in the lizards examined here (Bonine et al. 2005, SO

proportions do show more variation in other lizards, see Gleeson and Harrison 1986, and

Mutungi, 1990). The single optimum model was selected more frequently than any other

model (72.1% and 66.6%) based on bootstrap model selection frequencies for both model

selection criteria (Table 4.3). Brownian motion showed similar AICc and SIC values to

the single-optimum model as did the predator escape, but the Brownian motion model

was selected at less than half the frequency of the single optimum model and the predator

escape model was very rarely selected (Table 4.3). On the other hand, the foraging mode,

and combination models received little support based on the bootstrap model selection

frequencies (Table 4.3).

Discussion

Predator escape behavior appears to have driven the evolution of fast-twitch fiber-

type composition. Why is predator escape the driving force behind the fast-twitch muscle

fiber-type composition as opposed to foraging mode or a combination of the two?

Selection for survival against predatory attacks is probably stronger than selection for

incremental increases in foraging efficiency (Abrams 1986, Brodie & Brodie 1999).

Dawkins and Krebs (1979) referred to this as the “life-dinner principle”. While failure to

escape a predator ends all reproduction, poor foraging efficiency only marginally reduces

70

fitness by lowering energy intake and reducing mate acquisition, number of offspring, or

offspring survival (Frey-Ross et al. 1995, Morse & Stephens 1996, Kaspi et al. 2000).

Therefore, foraging behavior is likely a relatively weaker selective pressure on locomotor

musculature when compared to the strong selection of predation. Alternatively, selection

may be operating on speed. Lizards use near-maximal speed while fleeing predators, but

use lower speeds in prey capture attempts (these attempts were aimed at artificial lures;

Irschick & Losos 1998, Husak 2006). Furthermore, sit-and-wait lizards use a variety of

foraging speeds. For example, Crotaphytus collaris runs to capture prey, whereas P.

platyrhinos captures ants at slow speeds, yet both are considered sit-and-wait predators

(Vitt & Congdon, 1978). Examining the variation in FG and FOG fiber composition, sit-

and-wait species span the entire range of variation (Table S3 in Appendix).

Selective Hypotheses

What are the possible mechanisms of selection acting within the predator escape

regime? According to the predator escape model, lizards relying on sprint speed to

escape predators should have fast-twitch muscle composed mostly of FG fibers. Several

lines of evidence suggest an association between FG proportions and running

performance. Recruitment of FG fibers for acceleration and speed is found in organisms

such as fish, frogs, and lizards (e.g., Rome et al. 1988; Jayne et al. 1990; Lutz et al.

1998). Jayne et al. (1990) showed that, in the lizard Varanus exanthematicus, as speed

increased, additional white fibers in the iliofibularis were recruited (80% of the white

fibers are FG; Mutungi 1990). Within Phrynosomatid lizards, sprint speed is associated

with higher FG fiber type proportions (Bonine et al. 1999, 2001, 2005). Furthermore,

sprint speed has been linked to higher survival in lizards (Urosaurus ornatus, Miles 2004,

Crotaphytus collaris, Husak 2006; this latter study measured sprint speed at predator

escape). Thus, having high numbers of FG fibers likely helps achieve fast accelerations

and high running speeds, both of which are integral for escaping predators.

Alternatively, cryptic lizards have reduced FG and higher FOG fiber proportions

in comparison to flight escapers. There are multiple plausible mechanisms that may lead

to high FOG fiber proportions. First, there need not be direct selection for this

phenotype. Rather, it may reflect a lack of selection for FG fibers. Second, because

71

cryptic species escape predators largely by avoiding detection they have little need for

rapid accelerations. However, they still require locomotion for other activities (e.g., mate

acquisition, thermoregulation, feeding). The majority of these movements are well below

their maximum capabilities (Irschick & Losos, 1998, Jayne & Irschick 2000, Mattingly &

Jayne 2004, Husak 2006, Cooper 2007), and thus they may experience little selection for

FG fibers. Instead, FOG fibers may serve as the best general-purpose fiber type. FOG

fibers still allow animals to sprint (for catching prey, which often occurs at lower speeds

than predator escape [Irschick & Losos 1998, Husak 2006]), but have good contractile

properties at lower running speeds, and show greater fatigue resistance while running

(Peter et al. 1972; Gleeson et al. 1980; Gleeson & Johnston 1987). Third, if endurance

improves fitness (Barbosa & Moreno, 1999, Irschick, 2003), selection for fatigue

resistance may be acting on fast-twitch fiber-type composition, favoring FOG fibers.

There is very little known about endurance in cryptic species, but one study found that

horned lizards do not have high endurance at relatively fast speed (Garland 1994).

Whether the same is true at slower speeds is unknown. Finally, as there are differences in

the cellular composition of the different fiber types, they may differ in maintenance costs

(as far as we know, this idea has not been studied), which may play a role in selection for

energetically inexpensive muscle maintenance or contractions. Lack of certainty about

the precise selective mechanism does not diminish the significance of our results, which

clearly show an evolutionary pattern in fiber type evolution in which locomotor muscle is

composed largely of FOG fibers in cryptic species.

Further research is needed to determine the precise selective mechanism acting on

muscle fiber-type composition. Data are needed on fiber type involvement at different

speeds, their relationship with endurance, as well as more data on speeds used in nature.

Furthermore, we have examined only one of many muscles involved in lizard locomotion

and different muscles in the limbs differ in fiber type composition (Putnam et al. 1980;

Gleeson 1983). It would be interesting to see whether analyses of muscles involved in

the propulsive phase of running suggest the action of different selective pressures.

Finally, a systems-level consideration of all locomotor muscles may illuminate the

frequently unclear connection between individual muscle performance and whole animal

performance (e.g., Gleeson & Harrison 1988, Wilson et al. 2002). Locomotion is a

72

complex behavior involving many muscles. A full understanding of the mechanisms of

selection acting on locomotor muscles will require a more intimate knowledge of

interaction between muscles of the locomotor system, and links between muscle, whole-

animal performance, and fitness.

Multiple Evolutionary Models

Why is there no single best model for both fast and slow twitch fiber type

evolution? The answer may lie in their different functional roles. SO fiber-types provide

high endurance across a wide range of lower-speed activities such as walking/slow

running (Jayne et al. 1990), and are possibly involved in joint stabilizing, body support,

or posturing (Putnam et al. 1980 and references therein); behaviors which are shared

across lizard taxa. Consistent with this hypothesis, SO fiber composition remains

relatively constant in this dataset, whereas FG and FOG vary widely (Bonine et al. 2005).

It is intriguing that phenotypes which are components of the same functional unit

are predicted by very different evolutionary models. Whereas fast-twitch are best

explained by a strongly adaptive model driven by behavioral strategy, slow-twitch is best

explained by a global optimum model, or even a model of pure drift with no selection.

Perhaps this should not be so surprising, since locomotor muscles must perform such a

wide variety of functions throughout the life of a lizard. Functional variety here is

accomplished by different functional subunits (i.e., fiber types) comprising the muscle,

whose proportions can be differentially molded by evolution. Interestingly, one appears to

be highly conserved whereas the others are more free to vary.

Muscle physiology is often assumed to be under selection for whole animal

performance, but our study is one the first to provide direct evidence for an adaptive link

between muscle physiology and the behavior of an organism. Because Bonine et al.

(2005) were limited to the use of Brownian motion (BM) as their underlying evolutionary

model they were unable to test for selection within the contexts set by the ecology and

behavior of the species of lizards. Furthermore, determining whether foraging or

predator escape behavior exerts a stronger selective pressure on the muscle physiology of

these lizards would be unfeasible unless the two alternatives are compared directly. This

73

illustrates the importance of posing our best biologically-informed models and allowing

them to compete for the best explanation of the data.

74

Table 4.1. Fast-twitch fiber type model selection criteria. Model fit statistics for fast-

twitch fiber-type composition of the iliofibularis muscle. The Model with the best fit

(AICc = 24.67, SIC = 25.61) is in bold and bootstrap model selection frequencies (%)

based on 2000 replicates are included in parentheses. See the “Selective Regimes”

portion of the Materials and Methods section for a full explanation of each model.

Model AICc SIC

Predator Escape

0.00 (98.10%) 0.00 (88.30%)

Combination

12.96 (1.85%) 6.84 (11.70%)

Brownian Motion

18.13 (0.05%) 18.74 (0.00%)

Single Optima

20.26 (0.00%) 21.25 (0.00%)

Foraging Strategy 29.65 (0.00%) 29.65 (0.00%)

Table 4.2. Parameters of the best fitting fast-twitch fiber type model. Model

parameter estimates (dX(t) = α [(t) - X(t)] dt + dβ(t)) including selection (), drift (),

and optima () of the best -fitting fast twitch fiber-type model, predator escape.

Estimated optima values are cross-sectional proportions of FG relative to fast twitch

fibers (FG/(FG+FOG)). Escape strategy abbreviations are: FE = flight escape, ME =

mixed escape, and CE = cryptic escape. Bootstrap 95% confidence intervals are included

in parentheses.

Predator

Escape

Parameters

2.05 (0.44,12.58)

0.62 (0.36,1.30)

FE

0.76 (0.65,0.98)

ME

0.53 (0.46,0.77)

CE 0.16 (<0.01,0.39)

75

Table 4.3. Slow-Twitch fiber type model selection criteria. Model fit statistics for slow-

twitch fiber-type composition of the iliofibularis muscle. The Model with the best fit

(AICc = 56.52, SIC = 58.46) is in bold and bootstrap model selection frequencies (%)

based on 2000 replicates are included in parentheses.

Model AICc SIC

Single Optima

0.00 (72.10%) 0.00 (66.60%)

Brownian Motion

0.80 (25.55%) 0.41 (29.15%)

Predator Escape

1.40 (1.45%) 0.41 (1.90%)

Foraging Strategy

10.12 (0.85%) 9.13 (1.55%)

Combination 12.94 (0.05%) 7.76 (0.80%)

Table 4.4. Parameters of the best fitting slow-twitch fiber type models. Model

parameter estimates (dX(t) = α [(t) - X(t)] dt + dβ(t)) including selection (α), drift (σ),

and optimum (; for Brownian motion indicates the estimated ancestral state) for slow-

twitch fiber-type composition of the iliofibularis muscle. The estimated optimum value is

the cross-sectional proportion of SO fibers relative to all fibers (SO/(FG+FOG+SO)).

Bootstrap 95% confidence intervals are included in parentheses.

Model Brownian

Motion

Single

Optimum

Parameters

2.96 (1.02,50.62)

1.14 (0.78,1.46) 1.87 (1.12,7.01)

0.09 (0.04,0.23) 0.08 (0.05,0.13)

76

Table 4.5. Model estimates of fiber type composition compared to raw data. ‘s indicate

the fiber type proportions relative to total muscle volume estimated by the best models

for fast- (predator escape) and slow-twitch (single optimum) fiber types. FG, FOG, and

SO, are mean fiber type proportions of each regime based on the raw data presented in

Bonine et al. (2005). Because SO is 0.08, FG estimates were calculated as 0.92*

(estimated by the predator escape model), and FOG was calculated as 1-(FG+ SO).

See Table 2 for list of abbreviations.

Predator

Escape

Regime

Fiber

Types

FG

FG FOG FOG SO SO

FE

0.70 0.64 0.22 0.30 0.08 0.06

ME

0.53 0.46 0.39 0.42 0.08 0.12

CE 0.15 0.34 0.77 0.58 0.08 0.07

77

Figure 4.1: Evolutionary hypotheses for the evolution of fiber-type composition in the

iliofibularis muscle of lizards. Each color represents a different selective regime and

color codes are indicated below each hypothesis. The time scale of the phylogeny is

standardized to 1.0 from the basal node to terminal species.

78

CHAPTER 5. TARGETS OF SELECTION AND POTENTIAL CONSTRAINTS:

HOW THEY SHAPE THE EVOLUTION OF THE LOCOMOTOR SYSTEM IN

LIZARDS

Abstract

Adaptation arises in response to selective pressures acting on organismal

performance. Performance is generally a product of a complex phenotype constructed in

a hierarchy of biological organization from the cellular to whole animal level. Traits

across these levels can serve different functions and are involved in multiple aspects of

performance. Therefore, all phenotypic traits may not evolve in a uniform manner.

Phenotypic traits may be shaped by different selective pressures, result from a

compromise between function, or be limited in their adaptation due to constraints.

However, studies have largely examined the evolution of the many traits composing a

complex phenotype individually. To further our understanding of complex phenotypic

evolution, I test evolutionary models based on the behavior and ecology to assess how

selection on performance influences the evolution of traits from the cellular to the whole

animal level in the locomotor system of lizards. I find that performance based selection

does not act uniformly across the locomotor system. Selection related to both predator

escape and foraging behaviors acts on several traits spanning the cellular to whole-

organism level. However, the evolution of many traits are explained by very different

evolutionary models. A global optimum model best explains the evolution of several

muscular traits. This model indicates selection is similar on musculature across lizards.

However, this finding may also be a consequence of functional constraints in the

muscular system. Interestingly, a Brownian motion best explained femur and tibia

evolution. While this finding suggests portions of the limb may not be subject to

selection, it may also result from compromises between competing functions. Overall, I

find traits show distinct patterns of evolution within a complex phenotype, with some

reflecting selection related to performance, while others are more constrained in their

evolution.

79

Introduction

Organismal performance often results from the integration of a sizeable suite of

morphological and physiological traits (e.g., Liem 1973, Kohlsdorf et al. 2012).

Therefore, one may expect selection acting directly upon variation in performance to

translate into selection across the suite of underlying traits (Arnold 1983, Garland &

Losos 1994, Irschick et al. 2008). However, morphological and physiological variation

may not always reach the evolutionary optima set by performance-based selection as

genetic, developmental, or functional trade-offs may constrain their evolutionary

potential (Wake & Larson 1987, Wake 1991, Arnold 1992, Endler 1995, Walker 2007).

Identifying where selection or constraints dominate the evolutionary dynamics of a

complex, integrated phenotype can provide valuable insight into the processes that guide

the evolution of both performance and morphology/physiology, but multilevel selection is

rarely examined.

Locomotion is one of the primary models used for studying evolution in

performance and morphology, largely because it plays a critical role in the survival and

reproduction of many animals (e.g., Irschick and Garland 2001, Irschick et al. 2008). A

plethora of studies have identified functional links between form (morphology or

physiology) and performance as well as demonstrating that variation in performance can

be associated with fitness differences. However, the majority of these functional studies

do not take into account how selection acts across multiple levels of the biological

hierarchy. The locomotor system consists of a suite of traits integrated across a hierarchy

of biological levels, including the cellular (e.g., muscle fiber type), tissue (e.g., muscle

pennation, muscle cross-sectional area), and whole body (e.g., limb length, tail length)

levels. Yet, how variation at these different levels is shaped by evolution remains little

explored. Furthermore, the same musculoskeletal machinery must perform several tasks

including running, climbing, jumping, and walking for extended periods of time

(Vanhooydonck et al. 2001, Russell & Higham 2009, Oufiero et al. 2011, Herrel &

Bonneaud 2012,). However, trait evolution is most commonly examined in isolation with

very few studies looking at how selection for multiple functions influences the evolution

80

of all the traits involved. Researchers therefore, know little about how multiple forms of

selection act across complex phenotypes such as the locomotor system.

Lizards have long served as a model for selection and adaptation in locomotor

performance and morphology. Across lizards, studies have successfully linked

morphology and performance (e.g., Bonine & Garland 1999, Losos 1990, Melville &

Swain 2000, Goodman et al. 2008), and performance and fitness (Husak et al. 2006,

Miles 2004, Irschick et al. 2008) providing strong evidence for locomotor adaptation.

Furthermore, variation in performance and morphology has been linked to numerous

agents of selection including habitat use (Losos & Sinervo 1989, Losos 1990, Melville &

Swain 2000, Herrel et al. 2002, Kohlsdorf et al. 2004, Goodman et al. 2008), predator

escape behavior (Melville & Swain 2003, Miles 2004, Husak 2006, Vervust et al. 2007,

Gifford et al. 2008, Scales et al. 2009), and foraging mode (Huey et al. 1984, Reilly et al.

2007, McElroy et al. 2008, Verwaijen & VanDamme 2008) indicating that multiple

selective agents likely act on the locomotor system. Finally, morphological and

physiological traits across multiple levels of biological organization show variation with

performance (limb length: Bonine & Garland 1999, musculature: Zaaf et al. 1999,

Vanhooydonck et al. 2006 fiber type: Gleeson and Harrison 1988, Bonine et al. 2001,

2005, Kohlsdorf et al. 2004) suggesting that selection acts across the system.

Previously I used a group of lizards from the Southwest United States to

demonstrate that selection related to predator escape behavior drives the evolution of

sprint speed and acceleration, while selection related to foraging mode drives the

evolution of maximal exertion. However, as multiple traits across the locomotor system

influence performance in these lizards (Bonine et al. 2001, Herrel et al. 2002, Chapter 3),

how selection acts on this complex system remains unclear: Does performance related

selection act uniformly across the locomotor system or is it concentrated at one or more

levels? Does selection at one level of the biological hierarchy cascade down to lower

mechanistic levels? Can different selective agents target specific traits?

81

Materials and Methods

Here, I assess whether and where performance based selection acts across the

locomotor system by modeling the evolution of morphological traits across multiple

biological levels of the locomotor system. I use the same evolutionary models that were

previously used to model the evolution of performance in these same lizards (Chapter 2).

I then compare the models that best explain the evolution of morphological traits to the

models that best explained the evolution of performance. If selection acts uniformly

across the locomotor system, I expect all morphological traits to evolve according to the

same selection as performance. Alternatively, if selection is concentrated at one

biological level, I expect to see morphological traits at that level experience selection

similar to performance, but not traits at other levels. Finally, selection may occur across

levels of biological organization, but concentrated only at certain traits. If this is the case,

I expect some, but not all traits distributed across levels, to experience the same selection

as performance.

Morphological Data Measurements

To address the question of whether selection that guides the evolution of

locomotor performance also acts across multiple levels of organization of the locomotor

system, I use the performance data and models presented in Chapter 2 and the hind limb

muscle morphology data from Chapter 3. In addition to these data, I also include

morphological data for an individual muscle, the gastrocnemius, and body shape traits

based on gross morphology. Morphological data for the gastrocnemius was collected in

the same manner used in Chapter 3. Therefore, I supply only a limited explanation of

how the data was collected here. The gastrocnemius was removed, blotted dry and

weighed to the nearest 0.0001g using an Acculab Sartorius ACL-80.4 scale. The fascicle

length was measured at mid-muscle under a Zeiss DV4 stemi spot dissecting scope to the

nearest 0.01mm using Mitoyo Digimatic D6-6CX calipers (Biewener & Full 1992).

Pennation angle was estimated by bisecting the muscle and measuring the angle of the

fibers relative to the tendon or aponeurosis at mid-muscle (Biewener & Full 1992).

82

Snout-vent length (SVL, distance from the tip of the snout to the cloaca), tail

length (measured from the cloaca to the tip of the tail), fore limb length (measured from

the shoulder to the tip of the longest toe), and pelvic width (distance between the 2 hip

joints) were measured using digital calipers (Mitutoyo Digimatic D6-6CX) on all

individuals. Femur length (the greater trochanter to medial condyle), tibia length (top of

the medial condyle to medial malleolus), and foot length (base of the heel to the tip of the

longest toe) were all measured directly on the bone(s) of interest once muscles were

removed. Bone measurements were made on the same subset of individuals as muscle

measurements (see Chapter 3). Body mass was measured on live lizards with Pesola

scales to the nearest 0.1 g.

Morphological traits for evolutionary models

I modeled the evolution of several morphological traits and compare the best

performing of these models to the models that explain the evolution of performance from

Chapter 2. Traits selected for evolutionary analyses were those that potentially affect

locomotor performance in lizards, including several body shape traits including pelvic

width, tail length, fore and hind limb lengths. Previous studies of locomotor performance

in these lizards show that while hind limb length is important for sprint performance, the

different elements of the hind limb can vary in their relationships with performance (Zaaf

& VanDamme 2001). Thus, I also model the evolution of femur, tibia, and foot lengths

separately.

I similarly modeled the evolution of the musculature of the hind limb. I focus on

musculature involved in propulsion during running, as these muscles are more tightly

associated with running performance in the lizards studied here (Chapter 3, Higham et al.

2011). I model the evolution of stance phase muscle cross-sectional area (CSA) of the

hind limb as a whole, but also independently model functional groups, including the

major hip retractor, knee extensors, and ankle extensors as they can vary in function

during locomotion (Snyder 1954, Vanhooydonck et al. 2006, Zaaf et al. 1999).

Additionally, to examine how selection acts at the tissue level, I modeled the evolution of

muscle mass, fascicle length, and pennation angle in the gastrocnemius, a muscle

important for force production and propulsion running performance (Reilly 1995, Roberts

83

& Scales 2002, Higham et al. 2011). Although I previously modeled performance in

absolute terms, I chose to model size-corrected morphological data for two reasons. 1)

Relative morphology is a strong predictor of relative performance suggesting that relative

traits are still tightly linked to performance abilities. 2) The evolution of body size is

influenced by a wide range of life history and ecological factors (Blackburn et al. 1999,

Shine 2005, Faribairn 1997), however body shape likely reflects the locomotor demands

of an organism (Losos 1990, Vanhooydonck & VanDamme 1999, Melville & Swain

2000, Bergmann & Irschick 2010). I corrected body morphology traits for SVL to

highlight differences in shape. I corrected all muscle traits for body mass because

muscles provide the force for locomotion, and the effectiveness of force production is

directly related to the mass the force must move (a=f/m). For example, if two lizards

have the same muscle CSA, but one has twice the mass, I expect the lighter lizard to

achieve greater acceleration all else being equal.

Comparative Analyses

For our comparative analyses, I used the same phylogeny constructed Chapter 2,

which contains 21 species from the families Phrynosomatidae (17 species, 6 genera),

Teiidae (Aspidoscelis tigris and A. uniparens), Anguidae (Elgaria kingii) and

Crotaphytidae (Crotaphytus collaris).

All morphological data were log-transformed prior to analysis except for the

gastrocnemius pennation angle, which was arcsine transformed. The data were modeled

using Hansen’s (Hansen 1997) model, an Ornstein-Uhlenbeck (OU) process that models

the evolution of a continuous phenotypic trait subject to the influences of stabilizing

selection and noise (Hansen 1997, Butler & King 2004). The Hansen model describes

the change in a trait dX(t), over time (t), as an increment of a stochastic Brownian motion

process (dβ(t)), influenced by selection (α) toward an optimal trait value (θ) and subject

to stochastic changes proportional to a drift parameter (σ):

dX(t) = α [θ (t) – X(t)] dt + σdβ (t) (1)

84

Notably, the Hansen model allows optimal trait values to vary along the branches of the

phylogeny to represent shifts in selective regime (or 'adaptive zone' sensu Simpson 1953)

experienced by the evolving lineages. I compared 5 mappings of selective regimes on the

phylogeny, each representing a distinct evolutionary hypothesis for the evolution of the

trait (sensu Butler & King 2004, see selective regimes below). Each evolutionary model

was fit to the transformed data, assuming our phylogeny using the OUCH software

package (Butler & King 2004) in the R statistical computing environment (R

development Core Team 2010). The fits of each model were compared using the Akaike

information criterion corrected for small sample size (AICc). Information criteria were

used to measure the strength of evidence in support of each competing models (Burnham

& Anderson 2002). Model selection frequencies and model parameter confidence

intervals are based on 2000 bootstrap replicates (Burnham and Anderson 2002).

Selective Regimes for evolutionary models

I tested six evolutionary scenarios, Brownian motion and five adaptive models,

for producing variation in the locomotor morphology of these lizards. The Brownian

motion (BM) model is a stochastic model with no terms for selection. This model

assumes that locomotor performance evolves according to random noise or drift, and

makes no assumptions about adaptation with respect to the selective pressures considered

here.

Three of the five adaptive models are those previously used by Scales and Butler

(in review) to model locomotor performance, so they are covered only very briefly here

(see Scales and Butler in review for a detailed construction of evolutionary models). All

adaptive models of evolution were constructed by assigning adaptive regimes to

individual branches of the phylogeny (Butler & King 2004) and ancestral regimes were

reconstructed based on linear parsimony using Mesquite (Maddison & Maddison 2009).

The first of these adaptive models is based on microhabitat use and contains five

optima. Species were classified as terrestrial if they spend the majority of their time on

the ground. Arboreal species are those strongly associated with trees or woody bushes

whereas saxicolous species show a strong association with rocks and their vertical

surfaces. Litter species are those that are found in leaf litter and cluttered areas and

85

species were considered generalist if they don’t have a specific microhabitat associations

or frequent multiple microhabitat types (Figure 5.1). The second adaptive model is based

on foraging mode and contains three optima. Lizard species were classified as sit-and

wait predators (SWF), active foragers (AF), or mixed foragers (MF) based on movement

criteria (Figure 5.1). The third model is based on predator escape behavior and also has

three optima: flight, crypsis, and mixed (Figure 5.1). In this model cryptic escapers (CE)

are lizards that rely heavily on crypsis or morphology to evade predators with little

reliance on running. Flight escapers (FE) are species that rely mainly on bursts of

sprinting for escape, whereas mixed escapers (ME) are species that use a mix of cryptic

behavior and short sprints to refugia, or other behaviors (e.g., “squirreling” the act of

continually running to the opposite side of trees or rocks to avoid predators) to avoid

predation.

I also incorporated two additional models that were not initially included in the

performance study of Chapter 2. The first new model is a second predator escape model

with only two selective regimes. In Chapter 2, a post hoc analysis showed that a predator

escape model in which lizards were separated into sprint specialists versus all other

lizards (Figure 5.1), outperformed the other models in explaining sprint performance, but

was still within 2 AICc units of the 3 regime model. In this two-regime model, species

classified as flight escapers (see classifications of Chapter 2) experience selection on

performance and morphology that differs from all other predator escape behaviors used

by the lizards here (Figure 5.1).

Finally, I also added a global optimum model that assumes stabilizing selection

with a single optimum for a trait value for all species. It implies that all lizards should

share a single adaptive regime for a given trait with little variation out side of the optimal

range. This model was included because some morphological traits may have an optimal

value for functional purposes such as force production.

86

Results

Selection on body shape

The best predictor of body shape evolution varied across traits. Predator escape

behavior provides the top explanation for the evolution of hind limb length (Table 5.1).

While the three-regime predator model showed the lowest AICc score, the two-regime

predator model also received substantial bootstrap support (Table 5.1). The BM model

was also within it 2 AICc units of the predator escape models, but received almost no

bootstrap support. Interestingly, two different models best explain the evolution of the

separate hind limb segments. The BM model best explained the evolution of both femur

and tibia length, but was not strongly supported as there was modest bootstrap support for

multiple other models (Table 5.1). Conversely, the evolution of foot length was best

explained by the three-regime predator escape model, which had relatively strong

bootstrap support with moderate support for the two-regime model (Table 5.1). Thus, the

evolutionary trend observed for the hind limb appears to be driven by selection on the

foot, suggesting that foot length is a strong target of selection for level running. These

results also show that selection on performance for sprinting does not translate into

selection acting uniformly across the hind limb (Table 5.2).

Other body shape traits also show variation in evolutionary trends. For example,

forelimb length evolution is best explained by the BM model, but multiple models also

showed modest bootstrap support (Table 5.1). On the other hand, the evolution of tail

length and pelvic width are both driven by evolution in response to selection related to

foraging mode. The foraging mode model had strong support for both traits, with all

other models receiving almost no bootstrap support (Table 5.1). Overall, performance

based selection does not guide the evolution of all body shape traits (Table 5.2), but the

two selective pressures that drive sprinting and exertion abilities also target traits such as

foot and tail length, and pelvic width (Table 5.2).

Selection on musculature

Selection acting on musculature also did not always match performance related

selection (Table 5.2). The global optimum model best explained the evolution of overall

87

extensor CSA, as well as muscle CSA at the hip and knee (Table 5.1). This model was

well supported in all three cases (Table 5.1). Alternatively, the two-regime predator

escape model best explained the evolution of CSA at the ankle. While the predator

escape model was reasonably well supported, the global optimum model also received

moderate bootstrap support (Table 5.1). These findings indicate that selection drives the

evolution of hind limb musculature (Table 5.2), but only the more distal musculature at

the ankle is the focus of performance related selection, at least for level sprinting (Table

5.2).

Selection at the tissue level

Even the evolution of traits within an individual muscle are explained by different

evolutionary models. Relative muscle mass and fascicle length of the gastrocnemius both

evolve according to selection related to predator escape behavior (Table 5.1).

Interestingly, the three-regime model outperformed all other models for the evolution of

fascicle length, while the two-regime model best explained muscle mass with minimal

support for any other mode (Table 5.1). The evolution of pennation angle is best

explained by the global optimum model, which was strongly supported (Table 5.1).

These findings suggest that even within a given tissue the forms of selection can vary, but

the evolution of at least some traits are matched to that of performance.

Discussion

The locomotor system is organized in a hierarchy of biological levels that must be

integrated to function properly, but selection may not act uniformly across the system. I

tested whether morphological traits experience the same selection as the locomotor

performances they help to determine. By comparing six evolutionary hypotheses, I find

that the selective pressures that act on performance also act on morphological traits

across multiple levels of biological organization, but not in a uniform fashion.

According to Arnold’s paradigm (1983), natural selection acts directly on

organismal performance (or behavior, Garland & Losos 1994) and secondarily on

morphology resulting in the evolution of suites of coadapted morphological and

88

physiological characters. I find strong evidence for such a scenario as traits at the tissue

level (the gastrocnemius), the muscular system, and body shape, all evolve according to

selection associated with predator escape behavior, the same selective pressure that drives

the evolution of sprint performances. Additionally, Scales et al. (2009) showed that

predator escape behavior also drives the evolution of fast-glycolitic (FG) fiber type

proportions in a group of lizards that largely overlap with the species in this study (Table

5.2). Combined, these studies suggest that selection on the locomotor system spans the

cellular to the whole organism levels resulting in a coadpated suite of traits (Table 5.2).

A combination of several possibilities may underlie the non-uniform selection

across the locomotor system. One probable explanation is that traits in the locomotor

system experience differential selection. Differential selection may be a consequence of

varying selection strength among traits, or be the result of separate selective agents

targeting distinct traits. For example, some traits are stronger determinants of locomotor

performance than others (Scales et al. 2009, Kohlsdorf & Navas 2012, Vanhooydonck et

al. 2006, Herrel et al. 2002), and may be the targets of selection. In the lizards studied

here, foot length, ankle extensor CSA, and FG fiber-type proportions show tight

relationships with sprint performance compared to other traits (Bonine et al. 2001, 2005,

Herrel et al. 2002, Chapter 3). These traits, critical for performance, also evolve in

association with the sprint performances, suggesting that the most crucial determinants of

performance are also the strongest targets of selection. On the other hand, traits including

forelimb length and hip musculature are less associated with horizontal sprint

performances (Vanhooydonck et al. 2006), and may experience weak, or no selection

related to sprint and exertion performance.

Specific targets of selection may also explain why so few of the traits included in

this study evolve in association with foraging mode. Many of the traits studied here are

strong determinants of sprint performance, while others are associated with maximal

exertion. Morphological traits varying in their degree of association with a given aspect

of performance is common across many species (Hildebrand 1985, Alexander 2003,

Vanhooydonck et al. 2006, Kohlsdorf & Navas 2012) suggesting that selection targeting

specific traits may be widespread in complex phenotypes.

89

Locomotor traits may also differ in their evolution because different agents of

selection act on them. The evolution of traits such as foot length and ankle musculature is

driven by predator escape behavior due to their association with speed (Zaaf and

VanDamme 2001, Herrel et al. 2002, Vanhooydonck et al. 2006, Chapter 3). In contrast,

tail length and pelvic width evolve in association with foraging mode. A slender body

form can help reduce the cost of transport and improve maneuverability in lizards

(Garland & Losos 1994, Vanhooydonck & VanDamme 1999), both of which are

important for active foragers. The tail can also play an important role in locomotor

performance in lizards (e.g., Jusufi et al. 2008, Bateman & Flemming 2009 and

references therein) and may facilitate movement through some microhabitats (Wiens &

Slingluff 2001). The tail is also used as an anti predator device in many species (e.g.,

Cooper 1998, Bateman & Flemming 2009 and references therein). Active foragers are

likely more exposed to predation due to their increased movements (Lima & Bednekoff

1999, Hawlena 2009) and may benefit from longer tails to improve escape success (Huey

& Pianka 1981, Vitt 1983). Therefore, foraging mode driving the evolution of pelvic

width and tail length follows biomechanical and ecological predictions and provides

evidence that multiple forms of selection can act on distinct traits within a complex

phenotype.

Separate forms of selection acting on distinct traits may explain additional

variation in our results. Musculature at the hip is more essential for climbing, whereas the

knee and ankle musculature play a more significant role in level running (Zaaf et al.

1999, Zaaf & Van Damme 2001). The more arboreal and saxicolous species in this study

tend to have hind limb musculature distributed more towards the hip, whereas horizontal

runners have more musculature concentrated at the ankle (Chapter 3). Accordingly,

selection related to level sprinting acts strongly on ankle musculature, but not on hip

musculature. This situation may be reversed if I examined selection related to climbing

speed, where we can expect selection at the hip, but not the ankle. Again, such a scenario

lends further support to the idea that the many functional roles of the locomotor system

lead to different selective agents acting on specific traits resulting in distinct evolutionary

patterns.

90

An alternative, but by no means mutually exclusive explanation for our findings is

that some traits may be more limited in their evolution than others. Because locomotor

traits must perform so many functional roles, some may be limited in adaptation due to

functional constraints. Functional constraints may lead to compromises or trade-offs so

that traits cannot be simultaneously optimized for two functional roles (Vanhooydonck et

al. 2001, Ghalambor et al. 2003, Levinton & Allen 2005, Langerhans 2009b, Herrel &

Bonneaud 2012). The tibia and femur must perform level running, climbing, and

jumping all while maintaining posture and supporting body mass (Russell & Bels 2001).

Selection to perform all of these tasks may not allow optimization for a single

performance. Thus, while BM may reflect that traits like tibia and femur length are not

targets of selection, it may also be a consequence of selective pressures interacting so that

these traits are not optimized for a single function, but are compromises between several.

Conversely, traits that serve a single function may be easily adjusted to locomotor

demands. FG fibers are good for power production regardless of the function, be it

climbing, jumping, or level running (Biewener 2003, Bonine et al. 2001). Therefore, the

proportion of FG fibers in the hind limb muscles may be readily adjusted by selection for

specific power production needs regardless of the activity.

The evolution of some locomotor traits may be limited by their multiple

functions, but others may also be restricted by the need to perform as single function

well. A prime example may occur within the gastrocnemius muscle. While

gastrocnemius muscle mass and fascicle length evolve according to the same selective

pressure as sprint performances, a single optimum best explained the evolution of

pennation angle. The gastrocnemius plays an important role in terrestrial running,

providing force for propulsion during the stance phase (Zaaf et al. 2001, Roberts &

Scales 2002, Vanhooydonck et al. 2006, Higham et al. 2011). Pennate muscles are

generally capable of higher force production than parallel fibered muscles (Biewener

2003). Therefore, selection for a pennation angle that maximizes force production for

running may guide the evolution of pennation in the gastrocnemius muscle, limiting

variation.

The single optimum model may not reflect selection at all, but instead, pennation

angle may be constrained resulting in conserved muscle architecture across species. The

91

demands of force production for propulsion may functionally constrain pennation angles,

limiting the range of angles possible in the gastrocnemius. The muscle architecture of the

iliofibularis, another hind limb muscle, is also highly conserved (Bonine et al. 2001, Zaaf

et al. 2001). The iliofibularis, a swing phase muscle that may not require high force

production, but high contraction velocity is parallel-fibered in all species studied here

(unpublished data). Thus, muscle function may constrain muscle architecture, so that

within a muscle, pennation angle is constrained, but other traits are free to vary according

to locomotor needs. However, I cannot rule out developmental or genetic constraints as

pennation angles for a given muscle are frequently limited throughout lizards (Snyder

1954, Zaaf et al. 2001, Higham et al. 2011). Regardless of the mechanism, constraints

may inhibit selection on some traits in the locomotor system, while selection acts to fine

tune those traits that are free to vary.

Understanding the evolution and adaptation of complex phenotypes such as the

locomotor system is one of the main goals of evolutionary biology. Here we show that

the selective agents that act on locomotor performance in lizards also act across multiple

levels of organization in the locomotor system, but not in a uniform fashion. Instead,

adaptation of the locomotor system appears to be the result of a complex balance of

selection and constraint. Traits important for performance are likely targeted by

selection, but different agents of selection can act on individual traits. Conversely, the

adaptation of some traits may be functionally constrained due to compromises between

multiple functions or canalization for a specific function, while other traits may not even

experience selection. Thus, the evolution of the locomotor system seems to consist of

selection tuning evolutionarily labile traits to performance needs, while other traits

remain fixed. Using methods such as this one, helps tease apart the targets of selection

from constraints, providing a deeper understanding of how selection results in the

adaptation of complex, integrated phenotypes.

92

Table 5.1. Model selection criteria for locomotor traits. Model fit statistics for the

evolution of morphological traits across multiple levels of biological organization.

Models with the best fit (Akaike Information Criteria corrected for small sample size,

AICc) are listed in bold (with a value of zero), while ∆AICc values are provided for all

other models. Bootstrap model selection frequencies based 2,000 replicates are included

in parentheses. See “Selective Regimes” for explanations of each model.

Models

BM OU1

Predator

3-regime

Predator

2-regime

Foraging

Mode

Habitat

Use Body shape

Hind limb

0.80

(5.50%)

4.6

(0.35%) 0

(54.15%)

1.9

(31.70%)

2.8

(8.05%)

5.9

(0.25%)

Femur 0

(46.80%)

3.9

(11.80%)

6.4

(3.60%)

4.7

(13.40%)

5.8

(18.10%)

3.2

(6.30%)

Tibia

0

(42.15%)

1.40

(13.10%)

1.88

(4.50%)

0.96

(14.95%)

1.65

(18.60%)

8.87

(6.70%)

Foot

6.4

(0.75%)

8.1

(0.00%) 0

(69.45%)

0.2

(25.35%)

4.5

(4.30%)

15.6

(0.10%)

Pelvis

19.1

(0.00%)

19.4

(0.00%)

11.4

(0.30%)

21.1

(0.00%) 0

(99.70%)

29.9

(0.00%)

Forelimb

0

(45.30%)

3.3

(13.30%)

7.5

(3.90%)

5.1

(13.20%)

6.1

(18.30%)

5.6

(6.00%)

Tail

15.7

(0.00%)

14.3

(0.00%)

11.799

(1.30%)

15.099

(0.00%) 0

(98.70%)

21.2

(0.00%)

Musculature

Extensors

7.4

(0.10%) 0

(81.50%)

3.2

(2.70%)

0.1

(9.05%)

5.7

(5.50%)

7.3

(1.15%)

Hip

12.4

(0.35%) 0

(82.05%)

5.9

(2.80%)

2.7

(8.00%)

5.3

(5.35%)

10.6

(1.45%)

Knee

5.0

(0.01%) 0

(78.90%)

4.4

(2.90%)

0.200

(2.90%)

4.9

(6.55%)

6.0

(1.25%)

Ankle

8.0

(0.10%)

1.5

(27.50%)

2.7

(6.10%) 0

(62.85%)

7.40

(2.80%)

9.2

(0.65%)

Gastrocnemius

Gast mass

4.2

(0.20%)

4.3

(1.30%)

0.70

(12.15%) 0

(84.35%)

3.0

(1.80%)

12.5

(0.20%)

Gast length

5.9

(0.90%)

6.2

(0.50%) 0

(70.80%)

0.5

(16.65%)

6.9

(1.80%)

12.4

(0.20%)

Gast pennation

10.8

(0.40%) 0

(82.05%)

4.0

(3.25%)

2.6

(8.00%)

1.8

(5.20%)

13.1

(1.10%)

93

Table 5.2. Comparison of evolutionary models. The best fitting models for the evolution

of locomotor performance and traits within the locomotor system. Models in bold reflect

morphological traits that experience the same form of selection as performance.

Best Model

*Performance

Speed Predator Escape Acceleration Predator Escape

Exertion Foraging Mode

Whole animal

Hind limb Predator Escape Femur BM Tibia BM Foot Predator Escape

Pelvis Foraging Mode Forelimb Global

Tail Foraging Mode

Musculature

Extensors Global Hip Global

Knee Global Ankle Predator Escape

Tissue (Gastrocnemius)

Mass Predator Escape Fascicle length Predator Escape Pennation angle Global

+Cellular

FG fiber type Predator Escape SO fiber type Global

* Data from J. Scales Chapter 2.

+ Data from Scales et al. (2009).

94

Figure 5.1: Evolutionary hypotheses for the evolution of body shape, musculature, and

individual muscle traits of the locomotor system in lizards. Each color represents a

different selective regime and color codes are indicated below each hypotheses. The

global optimum model is not shown, but would be represented by a single color across

the entire phylogeny. The time scale of the phylogeny is standardized to 1.0 from the

basal node to the terminal species.

95

CHAPTER 6. CONCLUSIONS

Understanding the adaptation and diversification of phenotypic traits is one of the

primary goals of evolutionary biology, physiology and ecology. However, this task can

be difficult as many phenotypes are complex, consisting of numerous traits integrated

across multiple levels of biological organization that must perform a variety of functions,

and experience several forms of natural selection. Using the locomotor system as a

model, I have elucidated several principles regarding phenotypic evolution that are likely

broadly applicable across complex phenotypes. First, multiple selective pressures act on

both performance and morphology. At least two selective pressures act on locomotor

performance and morphology in lizards of the Southwest United States. However, I only

examined a few aspects of performance. Had I included more performance types such as

jumping, climbing, or maneuverability, I would undoubtedly discover the action of even

more selective pressures. Even if they do not act directly, selective pressures such as

habitat use may set the bounds in which specialization may occur (Butler et al. 2000,

Collar et al. 2010). Thus, without considering multiple aspects of performance and

selective pressures, we likely underestimate important selective forces as well as

functional diversity.

Second, variation in behavior is important in promoting phenotypic diversity. In

the lizards studied here, behavioral shifts are associated with the diversification of both

locomotor performance and morphology. Variation in foraging behavior leads to

differences in exertion abilities, while variation in predator escape behavior drives

differences in sprint capabilities. Furthermore, behavioral shifts also guide the evolution

of several morphological traits (muscle fiber type, foot length, etc.) and form-function

relationships between traits such as muscle CSA and speed. Although behavior can have

diverse effects on phenotypic evolution (Huey et al. 2003), we find support for the idea

that behavior is an important driver of evolutionary change. This idea has received

increasing support, especially in terms of locomotion, and has led to a modification of

Arnold’s (1983) morphology-performance-fitness paradigm (Garland & Losos 1994).

Third, performance based selection does not act uniformly across complex

phenotypes. Selection acts directly on variation in performance and only indirectly on

96

the underlying morphology (Arnold 1983). However, we find that selection varies across

traits within the locomotor system. Some traits, the most important determinants of

performance, are targets of selection. For example, foot length, ankle CSA, and muscle

fiber type, all evolve according to the same selective pressure as the locomotor

performances they help determine, sprint speed and acceleration. Interestingly, specific

traits can be subject to distinct selective pressures. Thus, traits serving separate

functional roles may experience distinct evolutionary influences. In contrast, other traits

may be unaffected by selection or limited in their evolutionary capacity. Such limitation

may result from strong selection to perform a single function or functional constraints

restricting variation. Therefore, examining multiple traits across a complex phenotype is

important for understanding functional relationships and integration as well as the factors

that promote or limit the evolution and adaptation of complex phenotypes.

Fourth, the actions of selective pressures can result in distinct phenotypes

integrated from the cellular to whole-organism level designed for specific tasks. For

example, the amount of hind limb muscle cross-sectional area (CSA) and proportion of

fast glycolitic fibers are both linked with sprinting behavior related to predator escape

(Chap.3, Bonine et al. 2001). When the data from this study is combined with other

studies, we see that species that rely heavily on sprints to evade predators should have

long hind limbs, especially the foot, with more muscle CSA, especially at the ankle, and

high proportions of FG fibers. On the other hand, species that rely more on crypsis, but

little on sprints should have shorter limbs with less CSA and higher proportions of FOG

fibers. Thus, selection results in a suite of coadapted traits spanning the locomotor

system, a phenomenon observed in other lizards and systems (Holzman et al. 2011,

Kohlsdorf & Navas 2012) suggesting it is a common theme throughout complex

phenotypes.

Finally, many phenotypes are very complex, and this complexity appears to allow

the mitigation of potential constraints and promote diversification. The locomotor system

is composed of numerous traits and must perform many functions. I find that different

traits are important for specific functions. Many traits help determine speed and

acceleration, but only two are linked with foraging strategy and these are not necessarily

the same traits that govern speed. Thus, exertion abilities and sprint abilities are likely

97

determined by different traits, which may allow them to evolve independently so that

lizards can be well suited for both activities (Arnold 1992). Furthermore, although I find

an expected speed-exertion trade-off, this trade-off is linked to the presence or absence of

selection. Selection for only a single performance, speed or exertion, results in a trade-

off. Conversely, when there is selection for both performances, species excelled at both,

suggesting the presence of mechanisms to alleviate this commonly invoked trade-off.

The ability to mitigate potential trade-offs may be an emergent property of complex

phenotypes such as the locomotor system (Alfaro et al. 2005, Wainwright et al. 2005,

Holzman et al. 2011) as high complexity within a system can allow the evolution of

performance combinations that defy theoretical trade-offs (Holzman et al. 2011).

Mechanisms such as many-to-one mapping, the idea that alternative designs can produce

the same performance, may be especially important in overcoming trade-offs (Alfaro et

al. 2005, Wainwright et al. 2005). In fact, I find evidence of many to one mapping with

respect to hind limb muscle CSA and speed. Species with similar sprint performance can

have muscle distributed more towards the ankle or hip (Chap. 2). Thus, the complexity

across the locomotor system may be what allows it to accommodate multiple functions

(Alfaro et al. 2005, Wainwright et al. 2005, Holzman et al. 2011 Strobbe et al. 2009) and

promote functional and morphological diversification. This phenomenon is almost

certainly widespread across complex phenotypes that underlie ecologically important

performances (Wainwright et al. 2005, Holzman et al. 2011).

Overall, the evolution of the locomotor system and locomotor performance is a

complex interaction of selective pressures acting on multiple aspects of performance

along with a suite of traits across several levels of biological organization. The effects of

selection are coupled with traits potentially unaffected by selection or limited in their

evolution due to functional canalization and constraints. The high level of morphological

and functional complexity likely helps mitigate trade-offs so that functional integration

does not limit possible performance combinations, promoting morphological and

performance diversification. Large behavioral shifts provide the fuel for such

diversification. However, this complexity also means understanding the evolution of

complex phenotypes can be very challenging. Here I show that integrating several

approaches is required to adequately describe performance and morphological diversity,

98

and the evolutionary process that underlie them. Applying an integrated approach to

other systems such as feeding or mating should provide further insights into the

adaptation and diversification of complex phenotypes.

99

APPENDIX

Table 1S. Selective Regime Classifications. Foraging mode classifications are based on references in Miles et al. (2007) and Scales et

al. (2009). Foraging mode variables: MPM = movements per minute, PTM = percent time moving. Foraging mode clasifications:

SW = sit & wait, MF = mixed forager, AF = active forager. Predator escape classifications: CE = cryptic escape, ME = mixed escape,

FE = flight escape. Predator escape classifications are based on Pianka & Parker 1975, Vitt & Congdon 1978, and references in Scales

et al. (2009). Predator escape classifications: CE = cryptic escape, ME = mixed escape, FE = flight escape. Habitat use classifications

are based on Stebbins 2003, and Brennan & Holycross 2006.

Species Family (Clade)

Foraging

mode MPM PTM

Predator

Escape Habitat use

Callisaurus

draconoides Phrynosomatidae

(Sand) SW - 0.02 FE

Terrestrial

Cophosaurus

texanus Phrynosomatidae

(Sand) SW 0.46 0.02 FE

Terrestrial

Holbrookia

maculata Phrynosomatidae

(Sand) SW 0.44 0.02 CE

Terrestrial

Phrynosoma

modestum Phrynosomatidae

(Horned) MF* 0.12 - CE

Terrestrial

Phrynosoma

cornutum Phrynosomatidae

(Horned) MF* - - CE

Terrestrial

Phrynosoma solare Phrynosomatidae

(Horned) MF* - - CE

Terrestrial

Sceloporus

graciosus Phrynosomatidae

(Sceloporus) SW 1.31 5.80 ME

Generalist

Sceloporus

magister Phrynosomatidae

(Sceloporus) SW 0.05 0.20 ME

Generalist

Sceloporus

clarkii Phrynosomatidae

(Sceloporus) SW 0.20 0.80 ME

Arboreal

Sceloporus

poinsettii Phrynosomatidae

(Sceloporus) SW 0.08 0.20 ME Rock-

dwelling

100

Sceloporus

occidentalis Phrynosomatidae

(Sceloporus) SW - - ME

Generalist

Sceloporus

tristichus Phrynosomatidae

(Sceloporus) SW† 0.31† 1.00† ME†

Generalist

Sceloporus

virgatus Phrynosomatidae

(Sceloporus) SW 0.40 0.80 ME

Generalist

Sceloporus

jarrovii Phrynosomatidae

(Sceloporus) SW 0.30 0.90 ME Rock-

dwelling

Urosaurus

graciosus Phrynosomatidae

(Sceloporus) SW* - - CE

Arboreal

Urosaurus ornatus Phrynosomatidae

(Sceloporus) SW 0.66 2.30 ME

Arboreal

Uta stansburiana Phrynosomatidae

(Sceloporus) SW 0.20 0.60 ME

Generalist

Crotaphytus

collaris Crotaphytidae SW 0.09 0.00 FE

Rock-

dwelling

Elgaria kingii Anguidae

AF* - - ME Litter

Aspidoscelis

uniparens Teiidae AF 0.79 78.70 FE

Terrestrial

Aspidoscelis

tigris Teiidae AF 1.62 87.00 FE

Terrestrial

* indicates classifications based on data of closely related species or family designation.

† data based on S. undulatus as it was formerly a subspecies of the S. undulatus complex

101

Table 2S. Locomotor performance and snout-vent length for 21 species of lizards. Lizards varied widely in locomotor performance.

Species means and standard deviations are presented for 4 performance variables: sprint speed, the maximum speed over a 0.25m

distance; acceleration, the maximum instantaneous acceleration; distance, the distance moved prior to exhaustion; time, time run prior

to exhaustion; SVL, distance from tip of the snout to the vent.

Species Sprint Speed

(m/s) Acceleration

(m/s2) Distance (m) Time (s) SVL (mm)

E. kingii 0.95 (0.22) 24.22 (8.37) 25.93 (3.97) 230.00 (58.57) 81.84 (5.67)

U. stansburiana 2.32 (0.37) 30.11 (4.63) 17.21 (5.49) 61.78 (22.29) 48.78 (1.70)

S. graciosus 2.18 (0.18) 24.92 (4.65) 17.88 (4.06) 58.33 (14.73) 51.76 (1.80)

S. magister 3.22 (0.48) 33.26 (7.75) 32.91 (10.05) 78.30 (17.20) 104.62 (9.73)

S. virgatus 1.57 (0.22) 22.51 (4.55) 14.44 (4.42) 64.20 (27.20) 52.79 (3.87)

S. tristichus 2.37 (0.30) 29.54 (7.61) 18.93 (3.56) 62.33 (12.25) 59.58 (5.86)

S. occidentalis 2.75 (0.36) 31.87 (8.68) 26.46 (3.46) 55.50 (9.32) 65.78 (4.46)

S. poinsettii 3.37 (0.46) 37.54 (9.66) 18.68 (1.41) 42.80 (6.83) 98.87 (14.31)

S. jarrovii 1.71 (0.13) 29.57 (5.83) 15.98 (2.92) 71.90 (14.19) 81.45 (5.11)

S. clarkii 3.49 (0.35) 40.8 (17.08) 27.61 (1.56) 55.50 (14.77) 99.60 (3.75)

U. ornatus 2.53 (0.48) 30.44 (8.18) 12.50 (2.37) 67.62 (19.32) 49.58 (3.56)

U. graciosus 2.82 (0.28) 37.79 (1.94) 19.00 (3.38) 62.67 (8.96) 54.70 (2.04)

P. solare 1.95 (0.30) 22.00 (7.17) 36.17 (5.18) 134.43 (20.65) 85.84 (7.01)

P. modestum 1.63 (0.20) 25.10 (5.93) 26.26 (5.64) 112.80 (35.96) 53.10 (5.28)

P. cornutum 1.76 (0.37) 23.73 (3.94) 27.75 (3.05) 134.13 (42.21) 82.49 (1.89)

H. maculata 2.87 (0.23) 31.00 (12.87) 19.06 (3.12) 55.90 (7.17) 52.33 (3.78)

C. texanus 3.50 (0.47) 35.56 (4.85) 20.94 (5.94) 37.20 (9.08) 66.26 (4.66)

C. draconoides 3.94 (0.43) 55.85 (17.12) 31.01 (5.81) 52.57 (11.42) 82.13 (3.43)

C. collaris 3.73 (0.84) 40.36 (7.02) 26.24 (4.60) 85.57 (22.88) 88.97 (1.67)

A. uniparens 3.05 (0.39) 32.35 (6.29) 46.19 (12.45) 256.90 (131.28) 66.01 (4.84)

A. tigris 4.15 (1.03) 50.88 (17.55) 48.35 (9.73) 232.25 (59.11) 81.79 (2.90)

102

Table 3S. Selective regime classifications for fiber-type data. Designations for foraging mode are based on references in Miles et al.

(2007) and Scales et al. (2009). Foraging mode variables: MPM = movements per minute, PTM = percent time moving. Foraging

mode classifications: SW = sit & wait, MF = mixed forager, AF = active forager. Predator escape classifications are based on Pianka

& Parker 1975, Vitt & Congdon 1978, and references in Scales et al. (2009). Predator escape classifications: CE = cryptic escape, ME

= mixed escape, FE = flight escape. Fiber-type proportion data for each species are from Bonine et al. (2005). FG/(FOG+FG) values

were calculated for use in analyses.

Species Family FG FOG SO FG/

(FG+FOG) Foraging

mode MPM PTM Predator

Escape Callisaurus

draconoides Phrynosomatidae 0.70 0.29 0.01 0.71 SW - 0.02 FE

Cophosaurus texanus Phrynosomatidae 0.65 0.33 0.03 0.66 SW 0.46 0.02 FE

Holbrookia maculata Phrynosomatidae 0.65 0.32 0.03 0.67 SW 0.44 0.02 CE

Uma notata Phrynosomatidae 0.64 0.25 0.11 0.72 SW* - 2.40 FE

Phrynosoma modestum Phrynosomatidae 0.30 0.64 0.06 0.32 SW*† 0.12 - CE

Phrynosoma cornutum Phrynosomatidae 0.25 0.66 0.08 0.27 SW*† - - CE

Phrynosoma mcallii Phrynosomatidae 0.31 0.56 0.13 0.36 SW*† - 32.00 CE

Phrynosoma platyrhinos Phrynosomatidae 0.23 0.72 0.06 0.24 SW*† - - CE

Sceloporus graciosus Phrynosomatidae 0.49 0.41 0.11 0.54 SW† 1.31 5.80 ME

Sceloporus magister Phrynosomatidae 0.41 0.42 0.17 0.49 SW 0.05 0.20 ME

Sceloporus undulatus Phrynosomatidae 0.43 0.43 0.15 0.50 SW 0.31 0.40 ME

Sceloporus virgatus Phrynosomatidae 0.44 0.45 0.11 0.49 SW 0.40 0.80 ME

Urosaurus ornatus Phrynosomatidae 0.45 0.41 0.14 0.52 SW 0.66 2.30 ME

Uta stansburiana Phrynosomatidae 0.48 0.46 0.07 0.51 SW 0.20 0.60 ME

Crotaphytus collaris Crotaphytidae 0.58 0.37 0.06 0.61 SW 0.09 0.00 FE

103

Gambelia wislizenii Crotaphytidae 0.58 0.38 0.04 0.60 SW* - - FE

Laudakia stellio Agamidae 0.37 0.46 0.17 0.45 SW* - - FE

Elgaria kingii Anguidae 0.43 0.41 0.16 0.51 AF* - - ME

Acanthodactylus

scutellatus Lacertidae 0.60 0.25 0.15 0.71 SW 1.00 7.70 FE

Aspidoscelis tigris Teiidae 0.76 0.22 0.03 0.78 AF 1.62 87.00 FE

Carlia fusca Scincidae 0.54 0.39 0.07 0.58 MF 2.00 17.70 ME*

Eumeces fasciatus Scincidae 0.49 0.38 0.13 0.56 AF* 0.60 72.40 ME

* indicates classifications based on data of closely related species or family designation.

† indicates a species modeled as both AF and SW due to differences in MPM and PTM.

104

REFERENCES

Abrams, P. A. 1986. Adaptive responses of predators to prey and prey to predators: the

failure of the arms race analogy. Evolution 40:1229-1247.

Aerts, P., R. Van Damme, B. Van Hooydonck, A. Zaaf, and A. Herrel. 2000. Lizard

locomotion: How morphology meets ecology. Neth. J. Zool. 50:261-277.

Alexander, R. M. 2003. Principles of animal locomotion. Princeton University Press,

Princeton, NJ.

Alexander, R. M., A.S. Jayes, G.M.O. Maloiy, and E.M. Wathuta 1981. Allometry of the

leg muscles of mammals. Proc. Zool. Soc. Lond. 194:539-552.

Alfaro, M. E., Bolnick, D.I., Wainwright, P.C., Wiens, J. 2004. Evolutionary dynamics of

complex biomechanical systems: an example using the four-bar mechanism.

Evolution 58:495-503.

Anderson, R. A., and W. H. Karasov. 1981. Contrasts in energy intake and expenditure in

sit-and-wait and widely foraging lizards. Oecologia 49:67-72.

Arnold, S. J. 1983. Morphology, performance and fitness. Am. Zool. 23:347 - 361.

Arnold, S. J. 1992. Constraints on phenotypic evolution. Am. Nat. 140 (suppl.):S85-S107.

Barbosa, A., and E. Moreno. 1999. Hindlimb morphology and locomotor performance in

waders: an evolutionary approach. Biol. J. Linn. Soc. 67:313-330.

Bateman, P. W., and P.A. Flemming. 2009. To cut a long tail short: a review of lizard

caudal autotomy studies carried out over the last 20 years. J. Zool. 277:1-14.

Baum, D. A., and A. Larson. 1991. Adaptation reviewed - A phylogenetic methodology

for studying character macroevolution. Syst. Zool. 40:1-18.

Bauwens, D., T. Garland Jr., A.M. Castilla, and R. Van Damme. 1995. Evolution of sprint

speed in Lacertid lizards: morphological, physiological, and behavioral

covariation. Evolution 49:848-863.

Bennett, A. F. 1980. The thermal dependence of lizard behavior. Anim. Behav. 28:752 -

762.

Bennett, A. F. 1994. Exercise performance of reptiles. Adv. Vet. Sci. Comp. Med.

38B:113-138.

105

Bennett, A. F., R. B. Huey, and H. John-Alder. 1984. Physiological correlates of natural

activity and locomotor capacity in two species of lacertid lizards. J. Comp.

Physiol. B 154:113-118.

Bennett, M. B. 1996. Allometry of the leg muscles of birds. Proc. Zool. Soc. Lond.

238:435-443.

Bergmann, P. J., and D. J. Irschick. 2010. Alternate pathways of body shape evolution

translate into common patterns of locomotor evolution in two clades of lizards.

Evolution 64:1569-1582.

Biewener, A. A. 2003. Animal Locomotion. Oxford University Press, New York, NY.

Biewener, A. A., and R. J. Full. 1992. Force platform and kinematic analysis.

Biomechanics- Structures and Systems a Practical Approach. University of

Chicago Press, Chicago, IL.

Blackburn, T. M., T. M. Gaston, and N. Loder. 1999. Geographic gradients in body size: a

clarification of Bergmann's rule. Divers. Distrib. 5:165-174.

Bonine, K. E. 2007. Physiological correlates of lizard foraging mode. pp. 94-119 in S. M.

Reilly, L. B. McBrayer, and D. B. Miles eds. Lizard Ecology, Cambridge

University Press, Cambridge, U.K.

Bonine, K. E., and T. Garland Jr. 1999. Sprint performance of phrynosomatid lizards,

measured on a high-speed treadmill, correlates with hindlimb length. J. Zool.,

Lond. 248:255 - 265.

Bonine, K. E., Gleeson, T. T., and T. Garland Jr. 2005. Muscle fiber-type variation in

lizards (Squamata) and phylogenetic reconstruction of hypothesized ancestral

states. J. Exp. Biol. 208:4529 - 4547.

Bonine, K. E., T. T. Gleeson, and T. Garland Jr. 2001. Comparative analysis of fiber-type

composition in the iliofibularis muscle of phrynosomatid lizards (Squamata). J.

Zool., Lond. 248:255-265.

Brennan, T. C., and A. T. Holycross. 2006. A field guide to amphibians and reptiles in

Arizona. Arizona Game and Fish Department, Pheonix, AZ.

Bro-Jorgensen, J. 2013. Evolution of sprint speed in African savannah herbivores in

relation to predation. Evolution 67:3371-3376.

106

Brodie, E. D., and E. D. Brodie Jr. 1999. Predator-prey arms races. Bioscience 49:557-

568.

Brown, T. K., and K. A. Nagy. 2007. Lizard energetics and the sit-and-wait vs. wide-

foraging paradigm. pp. 120-140 in S. M. Reilly, L. B. McBrayer, and D. B. Miles

eds. Lizard Ecology, Cambridge University Press, Cambridge, U.K.

Bulova, S. J. 1994. Ecological correlates of population and individual variation in

antipredator behavior of two species of desert lizards. Copeia 1994:980-992.

Burnham, K. P., and D. R. Anderson. 2002. Model selection and multimodel inference.

Springer, New York, NY.

Butler, M. A. 2005. Foraging mode of the chameleon, Bradypodion pumilum: a challenge

to the sit - and - wait versus active forager paradigm? Biol. J. Linn. Soc. 84:797 -

808.

Butler, M. A., and A. A. King. 2004. Phylogenetic comparative analysis: A modeling

approach for adaptive evolution. Am. Nat. 164:683 - 695.

Butler, M. A., and J. B. Losos. 2002. Multivariate sexual dimorphism, sexual selection,

and adaptation in Greater Antillean Anolis lizards. Ecol. Monogr. 72:541 - 559.

Butler, M. A., T. W. Schoener, and J. B. Losos. 2000. The relationship between sexual

size dimorphism and habitat type in Greater Antillean Anolis. Evolution 54:259 -

272.

Christiansen, P., and N. Bonde. 2002. Limb proportions and avian terrestrial locomotion.

J. Ornithol. 143:356-371.

Clemente, C. J., P. C. Withers, and G. Thompson. 2012. Optimal body size with respect to

maximal speed for the yellow-spotted monitor lizard (Varanus panoptes;

Varanidae). Physiol. Biochem. Zool. 85:265-273.

Collar, D., J. A. Schulte II, B. C. O'Meara, and J. B. Losos. 2010. Habitat use affects

morphological diversification in dragon lizards. J. Evol. Biol. 23:1033-1049.

Collar, D. C., and P. C. Wainwright. 2006. Discordance between morphological and

mechanical diversity in the feeding mechanism of centrarchid fishes. Evolution

60:2575-5284.

Cooper Jr., W. E. 1998. Reactive and anticipatory display to deflect predatory attack to an

autotomous lizard tail. Can. J. Zool. 76:1507-1510.

107

Cooper, W. E. 1994. Prey chemical discrimination, foraging mode, and phylogeny. pp 95-

116 in L.J. Vitt, and E.R. Pianka eds. Lizard Ecology: Historical and experimental

perspectives. Princeton University Press, Princeton, NJ.

Cooper, W. E. 2005. The foraging mode controversy: both continuous variation and

clustering of foraging movements occur. J. Zool. 267:179 - 190.

Cooper, W. E. 2007. Foraging modes as suites of coadapted movement traits. J. Zool.

272:45-55.

Crook, T. C., S. E. Cruickshank, C. M. McGowan, N. Stubbs, J. M. Wakeling, A. M.

Wilson, and R. C. Payne. 2008. Comparative anatomy and muscle architecture of

selected hind limb muscles in the Quarter Horse and Arab. J. Anat. 212:144-152.

Cumming, D. H., and G. S. Cumming. 2003. Ungulate community structure and

ecological processes: body size, hoof area and trampling in African savannas.

Oecologia 134:560-568.

Curtin, N. A., R. C. Woledge, and P. Aerts. 2005. Muscle directly meets the vast power

demands in agile lizards. Proc. R. Soc. B 272:581 - 584.

Darwin, C. 1859. On the origin of species by means of natural selection, or the

preservation of favoured races in the struggle for life. 1st ed.

Dawkins, R., and J. R. Krebs. 1979. Arms races between and within species. Proc. R.

Soc. B 205:489-511.

Dial, B. E. 1986. Tail display in two species of iguanid lizards: a test of the 'predator

signal' hypothesis. Am. Nat. 127:103-111.

Endler, J. A. 1995. Multiple-trait coevolution and environmental gradients in guppies.

Trends Ecol. Evol. 10:22-29.

Fairbairn, D. J. 1997. Allometry for sexual size dimorphism: Pattern and process in the

coevolution of body size in males and females. Annu. Rev. Ecol. Evol. Syst.

28:659 - 687.

Felsenstein, J. 1985. Phylogenies and the comparative method. Am. Nat. 125:1-15.

Fieler, C. L., and B. C. Jayne. 1998. Effects of speed on the hindlimb kinematics of the

lizard Dipsosaurus dorsalis. J. Exp. Biol. 201:609-622.

108

Frey-Ross, F., P.A. Brodmann, and H. Reyer. 1995. Relationship between food resources,

foraging patterns, and reproductive success in the water pipit, Anthus sp.

spinoletta. Behav. Ecol. 6:287 -295.

Fulton, C. J., D. R. Bellwood, and P. C. Wainwright. 2001. The relationship between

swimming ability and habitat us in wrasses (Labridae). Mar. Biol. 139:25-33.

Garland, T. Jr. 1994. Phylogenetic analyses of lizard endurance capacity in relation to

body size and body temperature. pp. 237-259 in L. J. Vitt and E. R. Pianka eds.

Lizard ecology: Historical and experimental perspectives, Princeton University

Press, Princeton, NJ.

Garland, T. Jr., and J. B. Losos. 1994. Ecological morphology of locomotor performance

in squamate reptiles. pp. 240-302 in P. C. Wainwright and S. M. Reilly eds.

Ecological Morphology. The University of Chicago Press, Chicago, IL.

Garland, T. Jr., A. F. Bennett, and C. B. Daniels. 1990. Heritability of locomotor

performance and its correlates in a natural population. Experientia 46:530-533.

Garland, T. Jr., P. H. Harvey, and A. R. Ives. 1992. Procedures for the analysis of

comparative data using phylogenetically independent contrasts. Syst. Biol. 41:18-

32.

Ghalambor, C. K., D. N. Reznick, and J. A. Walker. 2004. Constraints on adaptive

evolution: The functional trade-off between reproduction and fast-start swimming

performance in the Trinidadian guppy (Poecilia reticulata). Am. Nat. 164:38-50.

Ghalambor, C.K., J.A. Walker, and D.N. Reznik. 2003. Multi-trait selection, adaptation,

and constraints on the evolution of burst swimming performances Integr. Comp.

Biol. 43:431-438.

Gifford, M. E., A. Herrel, and D. L. Harmon. 2008. The evolution of locomotor

morphology, performance, and anti-predator behavior among populations of

Leiocephalus lizards from the Dominican Republic. Biol. J. Linn. Soc. 93:445-

456.

Gleeson, T. T. 1983. A histochemical and enzymatic study of the muscle fiber types in the

water monitor, Varanus salvator. J. Exp. Zool. 227:191-201.

109

Gleeson, T. T., and I. A. Johnston. 1987. Reptilian skeletal muscle: contractile properties

of identified, single fast-twitch and slow fibers from the lizard Dipsosaurs

dorsalis. J. Exp. Zool. 242:283-290.

Gleeson, T. T., and J. M. Harrison. 1986. Reptilian skeletal muscle: Fiber-type

composition and enzymatic profile in the lizard, Iguana iguana. Copeia 1986:324-

332.

Gleeson, T. T., and J. M. Harrison. 1988. Muscle composition and its relation to sprint

running in the lizard Dipsosaurus dorsalis. Am. J. Phys. 255:R470 - R477.

Gleeson, T. T., R. W. Putnam, and A. F. Bennett 1980. Histochemical, enzymatic, and

contractile properties of skeletal muscle fibers in the lizard Dipsosaurus dorsalis.

J. Exp. Biol. 214:293-302.

Gomes, F. R., E. L. Rezende, M. B. Grizante, and C. A. Navas. 2009. The evolution of

jumping performance in anurans: morphological correlated and ecological

implications. J. Evol. Biol. 22:1088-1097.

Goodman, B. A., D. B. Miles, and L. Schwarkopf. 2008. Life on the rocks: habitat us

drives morphological and performance evolution in lizards. Ecology 89:3462-

3471.

Hansen, T. F. 1997. Stabilizing selection and the comparative analysis of adaptation.

Evolution 51:1341 - 1351.

Hawlena, D. 2009. Colorful tails fade when lizards adopt less risky behaviors. Behav.

Ecol. Sociobiol. 64:205-213.

Heglund, N. C., and C. R. Taylor. 1988. Speed, stride frequency and energy cost per

stride: how do they change with body size and gait? J. Exp. Biol. 138:301-318.

Herrel, A., and C. Bonneaud. 2012. Trade-offs between burst performance and maximal

exertion in a wild amphibian, Xenopus tropicalis. J. Exp. Biol. 215:3106-3111.

Herrel, A., J. J. Meyers, and B. Vanhooydonck. 2002. Relations between microhabitat use

and limb shape in phrynosomatid lizards. Biol. J. Linn. Soc. 77:149 - 163.

Higham, T. E., and B. C. Jayne. 2004. In vivo muscle activity in the hindlimb of the

arboreal lizard, Chameleo calyptratus: general patterns and the effects of incline.

J. Exp. Biol. 207:249-261.

110

Higham, T. E., P. G. Korchari, and L. D. McBrayer. 2011. How muscles define maximum

running performance in lizards: an analysis using swing- and stance-phase

muscles. J. Exp. Biol. 214:1685-1691.

Hildebrand, M. 1985. Walking and running. pp. 38-57 in M. Hildebrand, D.M. Bramble,

K.F. Liem, and D.B. Wake eds. Functional Vertebrate Morphology, Belknap Press

of Harvard University Press, Cambridge, MA.

Hill, A. V. 1950. The dimensions of animals and their muscular dynamics. Sci. Prog.

38:209-230.

Hodgkison, R., S. T. Balding, Z. Akbar, and T. H. Kunz. 2004. Habitat structure, wing

morphology, and the vertical stratification of Malaysisan fruit bats

(Megachiroptera: Pteropodidae). J. Trop. Ecol. 20:667-673.

Holzman, R., D. C. Collar, R. S. Mehta, and P. C. Wainwright. 2011. Functional

complexity can mitigate performance trade-offs. Am. Nat. 177:E69-E83.

Huey, R. B., A. F. Bennett, H. John-Alder, and K. Nagy. 1984. Locomotor capacity and

foraging behaviour of Kalahari lacertid lizards. Anim. Behav. 32:41-50.

Huey, R. B., and P. E. Hertz. 1982. Effects of body size and slope on sprint speed of a

lizard (Stellio (Agama) stellio). J. Exp. Biol. 97:401-409.

Huey, R. B., and P. E. Hertz. 1984. Effects of body size and slope on acceleration of a

lizard (Stellio stellio). J. Exp. Biol. 110:113-123.

Huey, R. B., and E. R. Pianka. 1981. Ecological consequences of foraging mode. Ecology

62:991-999.

Huey, R. B., P. E. Hertz, and B. Sinervo 2003. Behavioral drive versus inertia in

evolution: a null model approach. Am. Nat. 161:357-366.

Husak, J. F. 2006. Does survival depend on how fast you can run or how fast you do run?

Funct. Ecol. 20:1080 - 1096.

Husak, J. F., S. F. Fox, M. B. Lovern, and R. A. Van Den Bussche. 2006. Faster lizards

sire more offspring: sexual selection on whole-animal performance. Evolution

60:2122 - 2130.

Huyghe, K., B. Vanhooydonck, A. Herrel, Z. Tadic. 2007. Morphology, performance,

behavior, and ecology of three color morphs in males of the lizard Podarcis

melisellensis. Integr. Comp. Biol. 47:211-220.

111

Iriarte-Diaz, J. 2012. Differential scaling of locomotor performance in small and large

terrestrial mammals. J. Exp. Biol. 205:2897-2908.

Irschick, D. J. 2002. Evolutionary approaches for studying functional morphology. Integr.

Comp. Biol. 42:278 - 290.

Irschick, D. J. 2003. Measuring performance in nature: implications for studies of fitness

within populations. Integr. Comp. Biol. 43:396-407.

Irschick, D. J., and B. C. Jayne. 2000. Size matters: Ontogenetic variation in the three-

dimensional kinematics of steady-speed locomotion in the lizards Dipsosaurus

dorsalis. J. Exp. Biol. 203:2133-2148.

Irschick, D. J., and B. C. Jayne 1999. Comparative three-dimensional kinematics of the

hindlimb for high-speed bipedal and quadrupedal locomotion of lizards. J. Exp.

Biol. 202:1047 - 1065.

Irschick, D. J., and J. B. Losos. 1998. A comparative analysis of the ecological

significance of maximal locomotor performance in Caribbean Anolis lizards.

Evolution 52:219 - 226.

Irschick, D. J., and T. Garland Jr. 2001. Integrating function and ecology in studies of

adaptation: Invesigations of locomotor capacity as a model system. Annu. Rev.

Ecol. Evol. Syst. 32:367-396.

Irschick, D. J., J. J. Meyers, J. F. Husak, and J. F. LeGalliard. 2008. How does selection

operate on whole-organism functional performance capacities? A review and

synthesis. Evol. Ecol. Res. 10:177-196.

Jayne, B. C., A. F. Bennet, and G. V. Lauder. 1990. Muscle recruitment during terrestrial

locomotion: How speed and temperature affect fiber type use in a lizard. J. Exp.

Biol. 152:101 - 128.

Johnson, M. A., M. Leal, L. R., Schettino, A. C. Lara, L. J. Revell, and J. B. Losos. 2008.

A phylogenetic perspective on foraging mode evolution and habitat use in West

Indian Anolis lizards. Anim. Behav. 75:555-563.

Jusufi, A., D. I. Goldman, S. Revzen, and R. J. Full. 2008. Active tails enhance arboreal

acrobatics in geckos. Proc. Natl. Acad. Sci. U.S.A. 105:4215-4219.

112

Kaspi, R., P. W. Taylor, and B. Yuval. 2000. Diet and size influence sexual advertisement

and copulatory success of males in Mediterranean fruit fly leks. Ecol. Entomol.

25:279 - 284.

Kikuchi, Y. 2010. Comparative analysis of muscle architecture in primate arm and

forearm. Anat. Histol. Embryol. 39:93-106.

Kohlsdorf, T., and C. A. Navas. 2012. Evolution of form and function:

morphophysiological relationships and locomotor performance in tropidurine

lizards. J. Zool. 288:41-49.

Kohlsdorf, T., R. S. James, J. E. Carvalho, R. S. Wilson, M. D. Pai-Silva, and C. A.

Navas. 2004. Locomotor performance of closely related Tropidurus species:

relationships with physiological parameters and ecological divergence. J. Exp.

Biol. 207:1183-1192.

Kohlsdorf, T., T. Garland Jr., and C. A. Navas. 2001. Limb and tail lengths in relation to

substrate usage in Tropidurus lizards. J. Morphol. 248:151-164.

Kwon, Y. H. 1994. KWON3D Motion analysis package 2.1 User's reference manual. V-

TEK Corporation Anyang, Korea.

Lailvaux, S. P., and D. J. Irschick. 2006. A functional perspective on sexual selection:

insights and future prospects. Anim. Behav. 72:263-273.

Langerhans, R. B. 2009a. Morphology, performance, fitness: functional insight into a

post-Pleistocene radiation of mosquitofish. Biol. Lett. 5:488-491.

Langerhans, R. B. 2009b. Trade-off between steady and unsteady swimming underlies

predator-driven divergence in Gambusia affinis. J. Evol. Biol. 22:1057-1075.

Levinton, J. S., and B. J. Allen 2005. The paradox of the weakening combatant: trade-off

between closing force and gripping speed in a sexually selected combat structure.

Funct. Ecol. 19:159-165.

Lewontin, R. C. 1978. Adaptation. Sci. Am. 239:212-228.

Liem, K. F. 1973. Evolutionary strategies and morphological innovations: Cichlid

pharyngeal jaws. Syst. Zool. 22:425 - 441.

Lima, S. L., and P. A. Bednekoff. 1999. Temporal variation in danger drives antipredator

behavior: the predation risk allocation hypothesis. Am. Nat. 153:649-659.

113

Losos, J. B. 1990. The evolution of form and function: morphology and locomotor

performance in West Indian Anolis lizards. Evolution 44:1189 - 1203.

Losos, J. B. 2009. Lizards in an evolutionary tree: ecology and adaptive radiation of

Anoles. University of California Press, Berkeley, CA.

Losos, J. B., and B. Sinervo. 1989. The effects of morphology and perch diameter on

sprint performance of Anolis lizards. J. Exp. Biol. 145:23 - 30.

Lutz, G. J., S. Bremner, N. Lajevardi, R. L. Lieber, and L. C. Rome. 1998. Quantitative

analysis of muscle fiber type and myosin heavy chain distribution in the frog

hindlimb: implications for locomotor design. J. Muscle Res. Cell M. 19:717 - 731.

Macey, J. R., A. Larson, N. B. Ananjeva, and T. J. Papenfuss. 1997. Evolutionary shifts in

three major structural features of the mitochondrial genome among iguanian

lizards. J. Mol. Evol. 44:660-674.

Macey, J. R., J. A. Schulte II, A. Larson, N. B. Ananjeva, Y. Wang, R. Pethiyagoda, N.

Rastegar-Pouyani, and T. J. Papenfuss. 2000. Evaluating trans-tethys migration:

An example using acrodont lizard phylogenies. Syst. Biol. 49:233-256.

Maddison, W. P., and D. R. Maddison. 2009. Mesquite: a modular system for

evolutionary analysis. Version 2.72 http://mesquiteproject.org.

Marsh, R. L., D. J. Ellerby, J. A. Carr, H. T. Henry, and C. I. Buchanan. 2004. Partitioning

the energetics of walking and running: Swinging the limbs is expensive. Science

303:80-83.

Mautz, W. J., C. B. Daniels, and A. F. Bennett. 1992. Thermal dependence of locomotion

and aggression in a Xantusiid lizard. Herpetologica 48:271 - 279.

McElroy, E. J., K. L. Hickey, and S. M. Reilly. 2008. The correlated evolution of

biomechanics, gait and foraging mode in lizards. J. Exp. Biol. 211:1029-1040.

Medler, S. 2002. Comparative trends in shortening velocity and force production in

skeletal muscles. Am. J. of Physiol.- Reg. I. 283:R368-R378.

Melville, J., and R. Swain. 2003. Evolutionary correlations between escape behavior and

performance ability in eight species of snow skinks (Niveoscincus: Lygosominae)

from Tasmania. J. Zool. 261:79-89.

114

Melville, J., and R. Swain. 2000. Evolutionary relationships between morphology,

performance and habitat openess in the lizard genus Niveoscincus (Scincidae:

Lygosominae). Biol. J. Linn. Soc. 70:667-683.

Miles, D. B. 2004. The race goes to the swift: fitness consequences of variation in sprint

performance in juvenile lizards. Evol. Ecol. Res. 6:63 - 75.

Miles, D. B., and R. G. Smith. 1987. A microcomputer-based timer and Data Acquisition

Device for measuring sprint speed and acceleration in cursorial animals. Funct.

Ecol. 1:281-286.

Miles, D. B., J. B. Losos, and D. J. Irschick. 2007. Morphology, performance, and

foraging mode. pp. 49-93 in S.M. Reilly, L.B. McBrayer, and D.B. Miles eds.

Lizard Ecology. Cambridge University Press, Cambridge, U.K.

Morse, D. H., and E. G. Stephens. 1996. The consequences of adult foraging success on

the components of lifetime fitness in a semelparous, sit and wait predator. Evol.

Ecol. 10:361 - 373.

Munger, J. C. 1984. Optimal foraging? Patch use by horned lizards (Iguanidae:

Phrynosoma). Am. Nat. 123:654-680.

Mutungi, G. 1990. Histochemistry, innervation, capillary density, and mitochoindrial

volume of red and white muscle fibers isolated from a lizard, Varanus

exanthimaticus. Can. J. Zool. 68:476-481.

Myatt, J. P., R. H. Crompton, and S. K. S. Thorpe. 2011. Hindlimb muscle architecture in

non-human great apes and a comparison of methods for analyzing inter-species

variation. J. Anat. 219:150-166.

Nelson, F. E., and B. C. Jayne. 2001. The effects of speed on the in vivo activity and

length of a limb muscle during the locomotion of the iguanian lizard Dipsosaurus

dorsalis. J. Exp. Biol. 204:3507-3522.

Norberg, U. M., and J. M. V. Rayner. 1987. Ecological morphology and flight in bats

(Mammalia; Chiroptera): Wing adaptations, flight performance, foraging strategy

and echolocation. Phil. Trans. R. Soc. B 316:335 - 427.

O'Steen, S., A. J. Cullum, and A. F. Bennett. 2002. Rapid evolution of escape ability in

Trinidadian guppies (Poecilia reticulata). Evolution 56:776 -784.

115

Oufiero, C. E., M. R. Walsh, D. N. Reznick, and T. Garland Jr. 2011. Swimming

performance trade-offs across a gradient in community composition in

Trinidadian killifish (Rivulus hartii). Ecology 92:170-179.

Paradis, E., J. Claude, and K. Strimmer. 2004. APE: analyses of phylogenetics and

evolution in R language. Bioinformatics 20:289 - 290.

Pasi, B. M., and D. R. Carrier. 2003. Functional trade-offs in the limb muscles of dogs

selected for running vs. fighting. J. Evol. Biol. 16:324-332.

Perry, G. 1999. The evolution of search modes: Ecological versus phylogenetic

perspectives. Am. Nat. 153:98-109.

Perry, G. 2007. Movement patterns in lizards: measurement, modality, and behavioral

correlates. pp. 13-48 in S.M. Reilly, L.B. McBrayer, and D.B. Miles eds. Lizard

Ecology. Cambrdige University Press, New York, NY.

Peter, J. B., R. J. Barnard, V. R. Edgerton, C. A. Gillespie, and K. E. Stemple. 1972.

Metabolic profiles of three fiber types of skeletal muscle in guinea pigs and

rabbits. Biochemistry 11:2627-2633.

Peterson, C. C., and J. F. Husak. 2006. Locomotor performance and sexual selection:

Individual variation in sprint speed of collard lizards (Crotaphytus collaris).

Copeia 2006:216-224.

Pianka, E. R., and W. S. Parker. 1975. Ecology of horned lizards: a review with special

reference to Phrynosoma platyrhinos. Copeia 1975:141-162.

Pollock, C. M., and R. E. Shadwick. 1994. Allometry of muscle, tendon, and elastic

energy storage capacity in mammals. Am. J. Physiol.-Reg. I. 266:R1022-R1031.

Pough, F. H., R. M. Andrews, J. E. Cadle, M. L. Crump, A. H. Savitzky, and K. D. Wells

2004. Herpetology. Pearson Prentice Hall, Upper Saddle River, NJ.

Pruitt, J. N., and J. F. Husak. 2010. Context-dependent running speed in funnel-web

spiders from divergent populations. Funct. Ecol. 24:165-171.

Putnam, R. W., T. T. Gleeson, and A. F. Bennett. 1980. Histochemical determination of

the fiber composition of locomotory muscles in a lizard, Dipsosaurus dorsalis. J.

Exp. Zool. 214:303-309.

116

Pyron, R. A., F. T. Burbrink, and J. J. Wiens. 2013. A phylogeny and revised classification

of Squamata, including 1416 species of lizards and snakes. BMC Evol. Biol.

13:93.

R development core team. 2007. R: A language and environment for statistical

computing. R foundation for statistical computing Vienna, Austria:URL

http://www.R-project.org/.

R development core team. 2010. R: A language and environment for statistical

computing. R Foundation for statistical computing Vienna, Austria: URL

http://www.R-project.org/.

Ramsey, J., and B. Riley. 2010. pspline: Penalized smoothing splines. R package version

1.0-14. http://CRAN.R-project.org/package=pspline.

Reilly, S. M. 1995. Quantitative electromyography and muscle function of the hind limb

during quadrupedal running in the lizard Sceloporus clarki. Zoology 98:263-277.

Reilly, S. M., and L. D. McBrayer. 2007. Prey capture and prey processing behavior and

the evolution of lingual and sensory characteristics: divergences and convergences

in lizard feeding biology. pp. 302-333 in S. M. Reilly, L.B. McBrayer, and D.B.

Miles eds. Lizard ecology, Cambrdige University Press, New York, NY.

Reilly, S. M., L. D. McBrayer, and D. B. Miles. 2007. Lizard Ecology: The evolutionary

consequences of foraging mode. Cambridge University Press, Cambridge, U.K.

Reznik, D. N., and J. Travis. 1996. The empirical study of adaptation in natural

populations. pp. 243-290 in M.R. Rose and G.V. Lauder eds. Adaptation.

Academic Press, San Diego, CA.

Roberts, T. J., and J. A. Scales. 2002. Mechanical power output during running

accelerations in wild turkeys. J. Exp. Biol. 205:1485 - 1494.

Rome, L. C., R. P. Funke, R. M. Alexander, G. Lutz, H. Aldridge, F. Scott, and M.

Freadman. 1988. Why animals have different muscle fiber types. Nature 335:824

- 827.

Rome, L. C., A. A. Sosnicki, and D. O. Goble. 1990. Maximum velocity of shortening of

three fiber types from horse soleus muscle: implications for scaling with body

size. J. Physiol. 431:173 - 185.

117

Russell, A. P., and T. E. Higham. 2009. A new angle on clinging in geckos: incline, not

substrate, triggers the deployment of the adhesive system. Proc. R. Soc. B

276:3705-3709.

Russell, A. P., and V. Bels. 2001. Biomechanics and kinematics of limb-based locomotion

in lizards: review, synthesis and prospectus. Comp. Biochem. Physiol. A Physiol.

131:89-112.

Scales, J. A., A. A. King, and M. A. Butler. 2009. Running for your life or running for

your dinner: What drives fiber-type evolution in lizard locomotor muscles? Am.

Nat. 173:543 - 553.

Schluter, D. 2000. The ecology of adaptive radiation. Oxford University Press, New

York, NY.

Schmidt-Nielsen, K. 1984. Scaling: Why is animal size so important? Cambridge

University Press, Cambridge, UK.

Schulte II, J. A., J. B. Losos, F. B. Cruz, and H. Nunez. 2004. The relationship between

morphology, escape behaviour and microhabitat occupation in the lizard clade

Liolaemus (Iguanidae: Tropidurinae*: Liolaemini). J. Evol. Biol. 17:408-420.

Shaffer, D. T. J., and W. G. Whitford. 1981. Behavioral responses of a predator, the

round-tailed horned lizard. Phrynosoma modestum and its prey, honey pot ants,

Myrmecocystus spp. Am. Midl. Nat. 105:209-216.

Sherbrooke, W. C. 1987. Defensive head posture in horned lizards (Phrynosoma: Sauria:

Iguanidae). Southwest. Nat. 32:512-515.

Sherbrooke, W.C. 2003. Introduction to horned lizards of North America. University of

California Press, Berkeley, California.

Shine, R. 2005. Life-history evolution in reptiles. Annu. Rev. Ecol. Evol. Syst. 36:23-46.

Simpson, G. G. 1953. The major features of evolution. Columbia University Press, New

York, NY.

Sites, J. W. J., J. W. Archie, C. J. Cole, and O. F. Villela. 1992. A review of phylogenetic

hypotheses for the lizard genus Sceloporus (Phrynosomatidae): implications for

ecological and evolutionary studies. Bull. Am. Mus. Nat. Hist. 213.

Smith, H. M. 1995. Handbook of lizards: Lizards of the United States and Canada.

Cornell University Press Ithica, New York:557.

118

Snell, H. L., R. D. Jennings, H. M. Snell, and S. Harcourt. 1988. Intrapopulation variation

in predator-avoidance performance of Galapagos lava lizards: the interaction of

sexual and natural selection. Evol. Ecol. 2:353 - 369.

Snyder, R. C. 1954. The anatomy and function of the pelvic girdle and hindlimb in lizard

locomotion. Am. J. Anat. 95:1-45.

Stebbins, R. C. 2003. A field guide to western reptiles and amphibians. Houghton Mifflin

Company, New York, NY.

Strobbe, F., M. A. McPeek, M. de Block, L. de Meester, and R. Stoks. 2009. Survival

selection on escape performance and its underlying phenotypic traits: a case of

many-to-one mapping. J. Evol. Biol. 22:1172-1182.

Swofford, D. L. 2003. PAUP*. Pylogenetic analyss using parsimony (*and other

methods) Version 4b10. Sinauer Associates, Sunderland, MA.

Townsend, T. M., A. Larson, E. Louis, and R. Macey. 2004. Molecular phylogenetics of

squamata: The position of snakes, amphisbaenians, and dibamids, and the root of

the squamate tree. Syst. Biol. 53:735-757.

Tsuji, J. S., R. B. Huey, F. H. Van Berkum, T. Garland Jr., and R. G. Shaw. 1989.

Locomotor performance of hatchling fence lizards (Sceloporus occidentalis):

quantitative genetics and morphometric correlates. Evol. Ecol. 3:2240-2252.

Van Damme, R., and B. Vanhooydonck. 2001. Origins of interspecific variation in lizard

sprint capacity. Funct. Ecol. 15:186-202.

Vanhooydonck, B., A. Herrel, and D. J. Irschick. 2007. Determinants of sexual

differences in escape behavior in lizards of the genus Anolis: a comparative

approach. Integr. Comp. Biol. 47:200-210.

Vanhooydonck, B., A. Herrel, R. Van Damme, and D. J. Irschick. 2006. The quick and the

fast: the evolution of acceleration capacity in Anolis lizards. Evolution 60:2137-

2147.

Vanhooydonck, B., and R. Van Damme. 1999. Evolutionary relationships between body

shape and habitat use in lacertid lizards. Evol. Ecol. Res. 1:785-805.

Vanhooydonck, B., R. Van Damme, and P. Aerts. 2001. Speed and stamina trade-off in

lacertid lizards. Evolution 55:1040 - 1048.

119

Vanhooydonck, B., R. S. James, J. Tallis, P. Aerts, Z. Tadic, K. A. Tolley, G. J. Measey,

and A. Harrel. 2014. Is the whole more than the sum of its parts? Evolutionary

trade-offs between burst and sustained locomotion in lacertid lizards. Proc. R.

Soc. B. 281:20132677.

Vanhooydonck, B., R. Van Damme, and P. Aerts. 2002. Variation in speed, gait

characteristics and microhabitat use in lacertid lizards. J. Exp. Biol. 205:1037-

1046.

Vervust, B., I. Grbac, R. Van Damme. 2007. Differences in morphology, performance and

behaviour between recently diverged populations of Podarcis sicula mirror

differences in predation pressure. Oikos 116:1343-1352.

Verwaijen, D., and. R. Van Damme. 2008. Foraging mode and locomotor capacities in

Lacertidae. Amphibia-Reptilia 29:197-206.

Verwaijen, D., and R. Van Damme. 2007. Does foraging mode mould morphology in

lacertid lizards? J. Evol. Biol. 20:1950-1961.

Vidal, N., and S. B. Hedges. 2005. The phylogeny of squamate reptiles (lizards, snakes,

and amphibaenians) inferred from nine nuclear protein - coding genes. C.R.

Biologies 328:1000 - 1008.

Vitt, L. J. 1983. Tail loss in lizards: the significance of foraging and predator escape

modes. Herpetologica 39:151-162.

Vitt, L. J., and J. D. Congdon. 1978. Body shape, reproductive effort, and relative clutch

mass in lizards: Resolution of a paradox. Am. Nat. 112:595 - 608.

Vitt, L. J., and H. J. Price. 1982. Ecological and evolutionary determinants of relative

clutch mass in lizards. Herpetologica 38:237-255.

Wainwright, P. C., M. E. Alfaro, D. I. Bolnick, and C. D. Hulsey. 2005. Many-to-one

mapping of form to function: A general principle in organismal design? Integr.

Comp. Biol. 45:256-262.

Wake, D. B. 1991. Homoplasy: The result of natural selection, or evidence of design

limitations? Am. Nat. 138:543-567.

Wake, D. B., and A. Larson. 1987. Multidimensional analysis of an evolving lineage.

Science 238:42-48.

120

Walker, J. A. 1998. Estimating velocities and accelerations of animal locomotion: a

simulation experiment comparing numerical differentiation algorithms. J. Exp.

Biol. 201:981 - 995.

Walker, J. A. 2007. A general model of functional constraints on phenotypic evolution.

Am. Nat. 170:681-689.

Walker, J. A., C. K. Ghalambor, O. L. Griset, D. McKenney, and D. N. Reznick. 2005. Do

faster starts increase the probability of evading predators? Funct. Ecol. 19:808-

815.

Watkins, T. B. 1996. Predator-mediated selection on burst swimming performance in

tadpoles of the pacific tree frog, Pseudacris regilla. Physiol. Zool. 69:154-167.

Weyand, P. G., D. B. Sternlight, M. J. Bellizzi, and S. Wright. 2000. Faster top running

speeds are achieved with greater ground forces not more rapid leg movements. J.

Appl. Physiol. 89:1991-1999.

Weyand, P. G., R. F. Sandell, D. N. L. Prime, and M. W. Bundle 2010. The biological

limits to running speed are imposed from the ground up. J. Appl. Physiol.

108:950-961.

White, T. D., and R. A. Anderson. 1994. Locomotor patterns and costs as related to body-

size and form in Teiid lizards. J. Zool. 233:107-128.

Wiens, J. J., and J. L. Slingluff. 2001. How lizards turn into snakes: a phylogenetic

analysis of body-form evolution in anguid lizards. Evolution 55:2303-2318.

Wiens, J. J., C. A. Kuczynski, A. Saad, and T. W. Reeder. 2010. Phylogenetic

relationships of phrynosomatid lizards based on nuclear and mitochondrial data,

and a revised phylogeny for Sceloporus. Mol. Phylogen. Evol. 54:150 -161.

Wiens, J. J., C. R. Hutter, D. G. Mulcahy, B. P. Noonan, T. M. Townsend, J. W. Sites Jr.,

and T. W. Reeder. 2012. Resolving the phylogeny of lizards and snakes

(Squamata) with extensive sampling of genes and species. Biol. Lett. 8:1043-

1046.

Wiens, J. J., K. H. Kozak, and N. Silva. 2013. Diversity and niche evolution along aridity

gradients in North American lizards (Phrynosomatidae). Evolution 67:1715-1728.

Wilson, R. S., and R. S. James. 2004. Constraints on muscular performance: trade-offs

between power output and fatigue resistance. Proc. R. Soc. B 271:S222-S225.

121

Wilson, R. S., R. S. James, and R. Van Damme. 2002. Trade-offs between speed and

endurance in the frog Xenopus laevis: a multi-level approach. J. Exp. Biol.

205:1145-1152.

Zaaf, A., and R. VanDamme. 2001. Limb proportions in climbing and ground-dwelling

geckos (Lepidosauria, Gekkonidae): a phylogenetically informed analysis.

Zoomorphology 121:45-53.

Zaaf, A., A. Herrel, P. Aerts, and F. DeVree. 1999. Morphology and morphometrics of the

appendicular musculature in geckoes with different locomotor habits

(Lepidosauria). Zoomorphology 119:9-22.