ArticlePDF AvailableLiterature Review

Coadaptation: A Unifying Principle in Evolutionary Thermal Biology

Authors:

Abstract and Figures

Over the last 50 yr, thermal biology has shifted from a largely physiological science to a more integrated science of behavior, physiology, ecology, and evolution. Today, the mechanisms that underlie responses to environmental temperature are being scrutinized at levels ranging from genes to organisms. From these investigations, a theory of thermal adaptation has emerged that describes the evolution of thermoregulation, thermal sensitivity, and thermal acclimation. We review and integrate current models to form a conceptual model of coadaptation. We argue that major advances will require a quantitative theory of coadaptation that predicts which strategies should evolve in specific thermal environments. Simply combining current models, however, is insufficient to understand the responses of organisms to thermal heterogeneity; a theory of coadaptation must also consider the biotic interactions that influence the net benefits of behavioral and physiological strategies. Such a theory will be challenging to develop because each organism's perception of and response to thermal heterogeneity depends on its size, mobility, and life span. Despite the challenges facing thermal biologists, we have never been more pressed to explain the diversity of strategies that organisms use to cope with thermal heterogeneity and to predict the consequences of thermal change for the diversity of communities.
Content may be subject to copyright.
282
Coadaptation: A Unifying Principle in Evolutionary Thermal Biology*
Michael J. Angilletta Jr.
1,
Albert F. Bennett
2
Helga Guderley
3
Carlos A. Navas
4
Frank Seebacher
5
Robbie S. Wilson
6
1
Department of Ecology and Organismal Biology, Indiana
State University, Terre Haute, Indiana 47809;
2
Department of
Ecology and Evolutionary Biology, University of California,
Irvine, California 92697;
3
De´partement de Biologie,
Universite´ Laval, Que´bec, Que´bec G1K 7P4, Canada;
4
Departamento de Fisiologia, Instituto de Biocieˆncias,
Universidade de Sa˜o Paulo, Sa˜o Paulo 05508-900, Brasil;
5
School of Biological Sciences, Heydon Laurence Building
A08, University of Sydney, Sydney, New South Wales 2006,
Australia;
6
School of Life Sciences, Goddard Building, St.
Lucia Campus, University of Queensland, Brisbane,
Queensland 4072, Australia
Accepted 4/25/2005; Electronically Published 2/3/2006
ABSTRACT
Over the last 50 yr, thermal biology has shifted from a largely
physiological science to a more integrated science of behavior,
physiology, ecology, and evolution. Today, the mechanisms that
underlie responses to environmental temperature are being
scrutinized at levels ranging from genes to organisms. From
these investigations, a theory of thermal adaptation has
emerged that describes the evolution of thermoregulation, ther-
mal sensitivity, and thermal acclimation. We review and inte-
grate current models to form a conceptual model of coadap-
tation. We argue that major advances will require a quantitative
theory of coadaptation that predicts which strategies should
evolve in specific thermal environments. Simply combining
current models, however, is insufficient to understand the re-
sponses of organisms to thermal heterogeneity; a theory of
coadaptation must also consider the biotic interactions that
influence the net benefits of behavioral and physiological strat-
* This paper was prepared as an overview of a symposium session presented at
“Animals and Environments,” the Third International Conference for Com-
parative Physiology and Biochemistry, Ithala Game Reserve, KwaZulu-Natal,
South Africa, 2004 (http://www.natural-events.com/ithala/default-follow_2.asp).
Corresponding author; e-mail: m-angilletta@indstate.edu.
Physiological and Biochemical Zoology 79(2):282–294. 2006. 2006 by The
University of Chicago. All rights reserved. 1522-2152/2006/7902-4148$15.00
egies. Such a theory will be challenging to develop because each
organism’s perception of and response to thermal heterogeneity
depends on its size, mobility, and life span. Despite the chal-
lenges facing thermal biologists, we have never been more
pressed to explain the diversity of strategies that organisms use
to cope with thermal heterogeneity and to predict the conse-
quences of thermal change for the diversity of communities.
Introduction
Forty years after Heath (1964) placed beer cans in the sun to
caution against a simplistic approach to the study of thermo-
regulation, thermal biology has become a thriving area of re-
search based on mathematical theories, sophisticated technol-
ogies, and model organisms. Much of this research is still
designed to help us understand the causes and consequences
of thermoregulation, but the evolution of thermal reaction
norms (for definitions of this and other thermal biology terms,
see Table 1) has also become a major focus. The core assump-
tions of thermal biology are (1) temperature is one of the most
pervasive state variables affecting biological processes, and (2)
the laws of thermodynamics define the direction and rate of
biochemical processes that underlie the performance of whole
organisms (Haynie 2001; Brown et al. 2004). Given these core
assumptions, thermal biologists have concluded that the en-
vironment exerts strong selective pressures on all organisms
and that knowledge of thermal biology is the key to explaining
many physiological, ecological, and evolutionary patterns.
Adaptive responses to thermal heterogeneity involve all levels
of biological organization from the expression of genes to the
behavior of the organism, but these responses occur on different
temporal scales (Fig. 1). We think of the interactions among
levels of organization that link these responses as a mechanistic
cascade, which flows from the biochemical to the organismal
levels and is nested within a feedback loop. In other words, the
expression of genes shapes the behavior of the organism, but
this behavior causes a change in physiological state that deter-
mines subsequent gene expression. Thus, the relationships
among thermal responses at various levels of organization are
likely to be complex, and thermal responses at biochemical
levels cannot be used to predict changes at higher levels of
organization (Huey and Stevenson 1979; Chaui-Berlinck et al.
2004). The challenge for biologists is to define the mechanistic
links between thermal responses at different levels and to iden-
tify their impact on fitness—in other words, to understand the
evolution of the mechanistic cascade.
Evolutionary Thermal Biology 283
Table 1: Definition of terms in thermal biology
Term Definition
Coadaptation The coevolution of traits within a population via natural selection
Developmental acclimation Irreversible plasticity of a physiological trait in response to an isolated environ-
mental variable, such as temperature, which is experienced during ontogeny
Eurythermy The ability to function over a broad range of body temperatures
Operative temperature The temperature of an inanimate object of zero heat capacity with the same
size, shape, and color as the animal
Seasonal acclimation Reversible plasticity of a physiological trait in response to a seasonal change in
an isolated environmental variable, such as temperature
Stenothermy The restriction of function to a narrow range of body temperatures
Thermal reaction norm A mathematical function relating environmental temperature (or body temper-
ature) to the phenotype expressed by a given genotype; includes thermal per-
formance curves, which describe the thermal sensitivities of behavioral or
physiological rates, and thermal tolerance curves, which describe thermal
sensitivities of fitness
Thermal sensitivity The slope (or derivative) of a thermal reaction norm
Thermoregulation The maintenance of a mean or variance of body temperature that differs from
the mean or variance of operative environmental temperatures, by means of
behavioral, physiological, or morphological strategies
Note. See Mercer (2001) for additional clarification of terminology.
At the organismal level, strategies for coping with thermal
heterogeneity vary along a continuum with two dimensions.
The first dimension describes the strategy of thermoregulation,
from organisms that maintain a nearly constant body temper-
ature despite thermal heterogeneity to those that conform to
environmental temperature. The second dimension describes
the strategy of thermal sensitivity, from organisms whose per-
formance is very sensitive to temperature (stenotherms) to
those whose performance is relatively insensitive to temperature
(eurytherms). All organisms fall somewhere within the contin-
uum bounded by these axes. Many endotherms, such as mam-
mals and birds, are thermoregulators that are stenothermic, but
many ectotherms also regulate their body temperature using
somewhat different mechanisms. Certain fish, such as carp, are
good examples of thermoconformers that are eurythermic. The
position of an organism along this continuum can be altered
by acclimation; in fact, an individual can be stenothermic dur-
ing specific seasons but eurythermic over the course of a year.
Finally, evolution by natural selection alters both acute and
acclimatory responses to temperature. Thus, ecological re-
sponses involve reversible or nonreversible forms of phenotypic
plasticity, and evolutionary responses involve a change in the
reaction norms of genotypes in a population (see Levins 1963
for a generalized discussion of these strategies).
Given that organisms have many options for dealing with
thermal heterogeneity, why does each species adopt certain
strategies over others? To answer this question, we must con-
sider the costs and benefits of each strategy in a range of en-
vironments; such considerations range from purely verbal mod-
els to complicated mathematical ones. Levins (1962, 1963) was
the first to advance models of evolution in heterogeneous en-
vironments, and many models of thermal adaptation were un-
doubtedly inspired by the work described in his classic book
(Levins 1968). Here, we review current models and draw on
recent studies of fish, reptiles, and amphibians to evaluate them.
We then combine the ideas generated from these models to
form a conceptual model of coadaptation that predicts the
optimal suite of strategies given mean, variance, and predict-
ability of environmental temperature. We argue that coadap-
tation is a unifying principle of thermal biology because it
emphasizes the simultaneous evolution of thermoregulation
and thermal sensitivities.
Adaptive Responses to Thermal Heterogeneity:
Theory and Evidence
When the temperature of the environment varies spatially and
temporally, one of four conditions must occur: (1) body tem-
perature is regulated by behavior and physiology; (2) body
temperature fluctuates, but performance is not impaired; (3)
body temperature fluctuates, and performance is initially im-
paired but is restored later by acclimation; or (4) body tem-
perature fluctuates, and performance is impaired and is not
restored by acclimation. As we shall discuss, any of these out-
comes could be adaptive in certain environments. Because each
outcome results from a combination of behavioral and phys-
iological strategies, natural selection should produce organisms
with suites of traits that are coadapted to the environment.
284 M. J. Angilletta Jr., A. F. Bennett, H. Guderley, C. A. Navas, F. Seebacher, and R. S. Wilson
Figure 1. A cascade of mechanisms operating from cellular to organismal
levels is responsible for adaptive responses to thermal heterogeneity.
These mechanisms alter commonly measured properties of organisms,
from gene expression to behavior, on different temporal scales.
Given the complexity of this process, how can we develop a
theory of coadaptation to explain the diversity of strategies that
are manifested in nature? To answer this question, we must
first consider existing theories of thermal adaptation and their
empirical support.
Thermoregulation
What factors determine the degree to which an individual
should thermoregulate and the mechanisms it should use to
do so? Huey and Slatkin (1976) modeled the optimal degree
of thermoregulation given energetic costs and benefits. They
assumed that the rate of energy gain was a unimodal function
of temperature. Although they modeled an energetic cost of
thermoregulation, they noted that other costs are certainly plau-
sible; for example, thermoregulators might incur a greater risk
of predation and fewer opportunities for territorial defense or
courtship, reducing fitness via survival and fecundity, respec-
tively. The model makes several predictions: (1) thermoregu-
lation should be greater in patchy environments, where thermal
heterogeneity enables heating and cooling with a low cost of
shuttling; (2) stenotherms should thermoregulate more care-
fully than eurytherms; and (3) individuals in colder environ-
ments should have a lower mean body temperature than in-
dividuals in hotter environments when all other factors are
equal.
Reptiles provide ample evidence that the costs and benefits
of thermoregulation influence the body temperatures of ani-
mals. Many species of diurnal lizards maintain relatively high
and constant body temperatures during activity, which tend to
maximize capacities for performance (Huey 1982; Hertz et al.
1983; Adolph 1990; Angilletta et al. 2002a). Anyone who
watched these lizards could appreciate the tremendous signif-
icance of behavioral mechanisms of thermoregulation, which
include shuttling between microhabitats, perching on objects,
and orienting the body toward solar radiation (Heath 1965;
Hertz and Huey 1981; Christian et al. 1983; Adolph 1990; Bau-
wens et al. 1996). Because solar radiation is the primary source
of heat for ectotherms, thermoregulation is presumably more
costly for species that are nocturnal or inhabit dense forests
(Huey 1974, 1982; Huey and Slatkin 1976). Indeed, body tem-
peratures of nocturnal geckos during activity are significantly
lower than those of diurnal lizards, even though these low
temperatures impede locomotor performance (Huey et al.
1989). For some species of geckos, relatively low body tem-
peratures during activity are undoubtedly caused by costs or
constraints imposed by nocturnal activity because the same
species maintain relatively high and constant body temperatures
in artificial thermal gradients (Huey et al. 1989; Angilletta and
Werner 1998). Some nocturnal geckos thermoregulate as ac-
curately and effectively during the day as diurnal lizards, par-
ticularly in late spring and summer, when considerable thermal
heterogeneity exists within and among sites of retreat (Kearney
and Predavec 2000; Rock et al. 2002). Transplant experiments
also reveal the influence of costs and constraints on thermo-
regulation. For example, Chamaeleo schubotzi, which inhabits
the foggy slopes of Mount Kenya (3,300 m), maintained its
preferred body temperature (33C) for only an hour on clear
days and never approached this body temperature on most days
(Bennett 2005); yet, individuals transplanted to a lower ele-
vation near Nairobi (1,700 m) maintained body temperatures
above 30C throughout the day (Fig. 2). These examples dem-
onstrate the profound flexibility of thermoregulatory behavior
in reptiles.
As predicted by theory, comparative studies over space and
time indicate that mean body temperatures of ectotherms are
lower in colder environments. Within species, this phenome-
non is manifested in one of two ways: individuals in colder
environments (1) maintain lower body temperatures while they
are active (Hertz 1981; Hertz et al. 1983; Van Damme et al.
1989; Adolph 1990; Grant and Dunham 1990; Christian and
Weavers 1996) or (2) maintain elevated body temperatures for
shorter durations each day (Angilletta 2001; Sears and Angilletta
2004). Despite these differences in behavior in natural envi-
ronments, preferred body temperatures measured in artificial
thermal gradients often do not differ among populations when
field body temperatures do (Van Damme et al. 1989; Angilletta
2001). Interspecific comparative analyses have provided mixed
support for the hypothesis that body temperatures will be lower
in colder environments. Among species of Chamaeleo, mean
body temperature decreased with decreasing minimal and max-
imal air temperatures (Bennett 2005). Among tropical species
of Sceloporus and Liolaemus, mean body temperature during
activity decreases by 1C per 1,000 m of elevation (Andrews
1998; Navas 2002). Tropical anurans exhibit more pronounced
Evolutionary Thermal Biology 285
Figure 2. Thermoregulation by (A) native and (B) transplanted cha-
meleons (Chamaeleo schubotzi) in Kenya. Data are mean body tem-
peratures and mean operative temperatures measured in the sun and
shade; operative temperatures in the sun were measured with hollow
copper tubes painted white or black. A, Individuals in the native habitat
of Mount Kenya ( ) never attained body temperatures aboveN p 4
23C. B, Individuals transplanted to Nairobi ( ) maintained bodyN p 6
temperatures above 30C most of the day. Data are from A. F. Bennett
(unpublished manuscript).
elevational clines, with mean body temperature decreasing by
5C per 1,000 m of elevation (Navas 2002). Yet, no elevational
trend in body temperature was evident among temperate spe-
cies of Sceloporus (Andrews 1998); this observation does not
necessarily contradict the prediction of theory because tem-
perate species in colder environments usually maintain pre-
ferred body temperatures for shorter durations per day (Sears
and Angilletta 2004). Thus, lizards in colder environments
maintain lower mean body temperatures over the course of
their lives.
Experimental studies of thermoregulation under perceived
risks of predation also support a model of thermoregulation
based on costs and benefits. Downes (2001) used chemical cues
to alter the perception of risk by garden skinks (Lampropholis
guichenoti) in experimental enclosures. Skinks surrounded by
a high density of scent markings became active later and became
inactive earlier than lizards surrounded by a low density of
scent markings. Consequently, the mean body temperature of
these skinks was lower and their growth was slower. Addition-
ally, velvet geckos (Oedura lesueurii) preferred warm retreats
in the absence of predators but chose cool retreats when warm
retreats were marked by the chemical cues of a predator
(Downes and Shine 1998). Clearly, the risk of predation and
its effect on body temperature depend on the specific mech-
anisms of thermoregulation. Behavioral mechanisms (e.g., shut-
tling and posturing) might make an individual more conspic-
uous to predators, whereas physiological mechanisms (e.g.,
shunting blood and changing color) might not increase risk.
Large reptiles can effectively exploit physiological mechanisms
of thermoregulation, but small reptiles could be forced to rely
largely on behavioral mechanisms (Bartholomew 1982; Dzi-
alowski and O’Connor 2004). Furthermore, the relationship
between thermoregulation and predation risk is complicated
because the ability of prey to escape predators probably depends
on body temperature (reviews in Hertz et al. 1988; Angilletta
et al. 2000b).
Thermal Sensitivity
When body temperature varies, natural selection is expected to
shape the thermal sensitivity of performance. The direction of
evolution should depend greatly on the manner in which an
individual’s performance contributes to its fitness. If perfor-
mance contributes additively to fitness (i.e., the sum of per-
formance over time determines fitness), stenotherms are fa-
vored under most patterns of temporal variation; eurytherms
are more fit than stenotherms only if environmental temper-
ature varies greatly among generations and little within gen-
erations (Gilchrist 1995). A greater degree of eurythermy, how-
ever, would be favored if environmental temperature changed
systematically with time (Huey and Kingsolver 1993). Thus,
the relative magnitude of variation within and among gener-
ations determines the optimal thermal sensitivity when per-
formance contributes additively to fitness. If performance con-
tributes multiplicatively to fitness (i.e., the product of
performance over time determines fitness), optimal thermal
sensitivities are determined more by variation within genera-
tions than by variation among generations; stenothermy is fa-
vored in constant environments, and eurythermy is favored in
variable environments (Lynch and Gabriel 1987). The generality
of these conclusions depends on the validity of a critical as-
sumption: an increase in performance at one or more tem-
peratures necessitates a decrease in performance at other tem-
peratures. In other words, a jack-of-all-temperatures is assumed
to be a master of none (Huey and Hertz 1984). This trade-off
between specialization and generalization, originally proposed
286 M. J. Angilletta Jr., A. F. Bennett, H. Guderley, C. A. Navas, F. Seebacher, and R. S. Wilson
Figure 3. Maximal growth rates (V
max
) of ancestral (filled circles) and
selected (open circles) populations of E. coli. Selected populations were
maintained at 37C for 20,000 generations. Error bars are 95% con-
fidence intervals. Adapted from Cooper et al. (2001).
by Levins (1968), is a common assumption of evolutionary
theories in general.
Both comparative and experimental studies have shown that
eurythermy evolves far more often than predicted by theory.
Thermal sensitivities of growth rate have been compared among
populations of marine fish distributed along latitudinal clines.
If we assume that growth contributes additively to fitness, cur-
rent theory predicts that natural selection should produce
stenotherms that perform best at the mean water temperature.
In most cases, however, genetic divergence along latitudinal
clines has produced eurytherms that perform equally well or
better than stenotherms at any temperature (review in Angil-
letta et al. 2002b). These genetic differences among populations
persist despite considerable gene flow among populations
(Conover 1998), which suggests that natural selection can pro-
duce jacks-of-all-temperatures that are masters of all. Semi-
natural selection of E. coli under various thermal regimes has
provided mixed support for the theory of thermal adaptation.
Selection at 37C during 20,000 generations improved fitness
at 27–39C but reduced fitness at extreme temperatures (Fig.
3). Nevertheless, a series of selection experiments with E. coli
(Bennett and Lenski 1999; Po¨rtner et al. 2006) and a compar-
ative study of Daphnia pulicaria (Palaima and Spitze 2004) have
shown that a trade-off between fitnesses at high and low tem-
peratures is not a necessary outcome of evolution. These com-
parative and experimental results cast doubt on the generality
of current theory.
One possible explanation for the lack of thermal speciali-
zation within natural populations is that physiological perfor-
mances contribute both additively and multiplicatively to
fitness. Consider the number of genes that must underlie a
complex trait such as growth rate. Some fraction of these genes
could also code for proteins that enhance maintenance, repair,
and, hence, survival. Such genes would cause growth rate to
evolve in correlation with thermal tolerance. Because environ-
mental temperature tends to vary more at higher latitudes,
natural selection should produce broader thermal tolerances at
these locations (e.g., van Berkum 1988). Thus, correlated evo-
lution of thermal sensitivities of growth rate and thermal tol-
erance might explain the countergradient variation observed in
fish (Conover and Schultz 1995). Selection experiments can be
used to explore potential genetic correlations between the tol-
erance of extreme temperatures and performance at less ex-
treme temperatures (Mongold et al. 1999). The identification
of mechanistic cascades that produce variation in organismal
performances would shed light on the network of genes shared
between performance and tolerance.
Acclimation of Thermal Sensitivity
Many ectotherms modify thermal sensitivities of performance
throughout their lifetimes. These modifications can be non-
reversible responses to temperatures experienced during on-
togeny (developmental acclimation) or reversible responses to
gradual changes in temperature among seasons (seasonal ac-
climation). The distinction between developmental and sea-
sonal responses is important when discussing the evolutionary
significance of phenotypic plasticity (Kingsolver and Huey
1998). Both developmental and seasonal acclimation involve a
cascade of processes, which can include the detection of en-
vironmental signals, a signal transduction pathway (or path-
ways), and the activation of the production machinery of the
cell (e.g., structural genes, polymerases, ribosomes) leading to
a change in physiology. During developmental acclimation, the
process involves morphogenesis that is irreversible. Small effects
of developmental temperature early in ontogeny can lead in-
directly to larger effects on the phenotype later in ontogeny.
Thus, developmental acclimation can produce changes that are
distinct from those observed during seasonal acclimation of the
same physiological systems (I. A. Johnston and R. S. Wilson,
unpublished data).
Precht (1958) divided patterns of acclimation into five cat-
egories based on the degree to which the rate of performance
is maintained during environmental change; these categories
are overcompensation, perfect compensation, partial compen-
sation, no compensation, and inverse compensation. All of
these responses have been interpreted as adaptive strategies for
coping with thermal fluctuations (Huey and Berrigan 1996).
For example, when faced with seasonal cooling, an ectotherm
can (1) slow physiological capacity by submitting to Q
10
effects
(e.g., O’Steen and Bennett 2003); (2) depress physiological ca-
pacity by enhancing thermal effects (e.g., St. Pierre and Boutilier
2001); or (3) offset Q
10
effects by maintaining capacities through
partial compensation, perfect compensation, or overcompen-
sation (e.g., Rome et al. 1984; Guderley and St. Pierre 1999).
Evolutionary Thermal Biology 287
Patterns of acclimation are enriched by nature’s complexity;
for example, when environmental warming is accompanied by
a depletion of resources (e.g., food, oxygen), some ectotherms
initiate a metabolic depression that alleviates Q
10
effects (Hand
and Hardewig 1996; de Souza et al. 2004; review in Guppy and
Withers 1999). The presumption that these compensatory re-
sponses enhance the fitness of an organism is known as the
beneficial acclimation hypothesis (Leroi et al. 1994).
Critics of the adaptationist program noted alternatives to
adaptive explanations for acclimatory responses (Garland and
Carter 1994; Feder et al. 2000), and others argued that the
evolutionary significance of such responses has rarely been es-
tablished (Huey and Berrigan 1996; Kingsolver and Huey 1998;
Huey et al. 1999). Huey et al. (1999) encouraged the application
of strong inference (Platt 1964) and advanced four hypotheses
as alternatives to the beneficial acclimation hypothesis. Al-
though we agree that strong inference is a powerful approach,
we contend that current hypotheses have limited power to in-
crease our understanding of the evolution of acclimation. First,
these hypotheses stem from verbal models that consider isolated
mechanisms rather than quantitative models that consider mul-
tiple causality; for example, the hypothesis that “colder is
better” is based on the common observation that organisms
reach larger body sizes at lower temperatures, whereas the hy-
pothesis of developmental buffering is based on a mechanism
that is independent of body size (Huey et al. 1999). Because
these hypotheses are not mutually exclusive and neither leads
to quantitative predictions about the magnitude of acclimation,
one cannot easily design crucial experiments to distinguish be-
tween them (see Quinn and Dunham 1983). More importantly,
these hypotheses do not stem from an explicit consideration
of the variance and predictability of environmental tempera-
ture. A quantitative model, however, would provide the explicit
analysis of costs and benefits that is needed to predict the
optimal capacity for acclimation given the mean, variance, and
predictability of environmental temperature. Furthermore,
such a model would enable investigators to make a priori pre-
dictions about the acclimation of thermal sensitivity when mul-
tiple mechanisms (or constraints) are involved.
Currently, only one quantitative model can be used to infer
the optimal acclimation of thermal sensitivity. This model, con-
structed by Gabriel and Lynch (1992), describes the evolution
of an environmental tolerance curve, which can be thought of
as the relationship between temperature and fitness. As with a
performance curve, a tolerance curve can be characterized by
an optimum and a breadth. The model predicts that the net
benefit of acclimation increases with increasing variation in
body temperature among generations and decreasing variation
in body temperature within generations. This result, however,
follows from the fact that fitness is zero if body temperature
falls outside of the tolerance curve at any point during the life
of an organism (i.e., once you’re dead, you’re dead). The op-
timal acclimation of performance curves would differ quali-
tatively from that of tolerance curves if performance contributes
additively to fitness rather than multiplicatively (see “Thermal
Sensitivity”).
Despite the limitations of current theory, experimental stud-
ies of acclimation have generally rejected the beneficial accli-
mation hypothesis (Leroi et al. 1994; Zamudio et al. 1995; Huey
and Berrigan 1996; Bennett and Lenski 1997; Huey et al. 1999;
Gibert et al. 2001). Given that these studies used rigorous ex-
perimental designs to test and reject a long-held assumption
within evolutionary physiology, they confirm the need to crit-
ically evaluate adaptive explanations for phenotypic plasticity.
Much of the rigorous nature of these experiments was afforded
by the careful selection of model organisms, such as E. coli and
Drosophila melanogaster. Huge numbers of replicates, short
generation times, and comparatively simple procedures of
maintenance were all important advantages of using these spe-
cies. In the majority of these studies, however, organisms were
raised at different temperatures for one or more generations.
A consequence of this experimental design was that reversible
and irreversible forms of acclimation might have been con-
founded (Wilson and Franklin 2002).
Our understanding of acclimation comes mostly from hun-
dreds of descriptions of seasonal acclimation and its underlying
mechanisms (Prosser 1979). Although recent experiments have
focused on developmental acclimation in organisms with short
generations, the benefits of seasonal acclimation still remain to
be investigated (Guderley and St. Pierre 2002; Wilson and
Franklin 2002). This gap in our understanding of acclimation
is surprising when one considers that many model organisms
in thermal biology have generations that span one or more
years and are therefore routinely exposed to seasonal changes
in temperature. One key to better understanding the evolution
of either developmental or seasonal acclimation will be the
careful selection of a trait that has a clear connection to survival
or reproduction. For example, Wilson and Johnston (2005)
studied the thermal acclimation of mating success of males of
Gambusia holbrooki. Males of this species copulate entirely
through sneaky encounters because females always resist mat-
ing. When tested at 18C, males acclimated to this temperature
swam faster, maneuvered better, and had greater mating success
than males acclimated to 30C; the converse was true for tests
conducted at 30C (Fig. 4). Because mating success is likely to
correlate strongly with fitness, this experiment provides some
evidence for the benefits of seasonal acclimation.
Even when compensation would seem beneficial, failure to
compensate perfectly could reflect costs of acclimation (Hoff-
mann 1995; Guderley 2004). Acclimation of thermal perfor-
mance curves would seem beneficial if (1) fluctuations in tem-
perature within generations are predictable, (2) the time lag
required for compensation is short relative to the rate of change
in temperature, and (3) the energetic or survival cost of com-
pensation is less than the loss of energy or reduction of sur-
vivorship that is caused by a failure to compensate (see DeWitt
288 M. J. Angilletta Jr., A. F. Bennett, H. Guderley, C. A. Navas, F. Seebacher, and R. S. Wilson
Figure 4. Effect of thermal acclimation on the number of copulations
obtained by males of Gambusia holbrooki at 18 and 30C. Copulations
were recorded during 5 min of exposure to two females in an aquarium.
Significant differences were detected between groups at both test tem-
peratures ( for each group). Data are . AdaptedN p 20 means SE
from Wilson and Johnston (2005).
et al. 1998). Perfect compensation of performance is costly
because of the time and energy required to synthesize new
enzymes, modify membranes, and redistribute resources. For
example, compensation of muscular function during exposure
to low temperature can involve increases in the concentrations
of enzymes (Sa¨nger 1993), the expression of myosin isoforms
(Johnston and Temple 2002), the fluidity of cellular membranes
(Hazel 1995), and the oxidative capacity or volume of mito-
chondria (Guderley and St. Pierre 1999). These changes require
time as well as energy; for example, the compensatory response
of rainbow trout (Oncorhynchus mykiss) is biphasic and requires
a minimum of 6 wk to stabilize (Bouchard and Guderley 2003).
Beyond the obvious costs of producing these changes, the en-
ergetic cost of “primed” mitochondria and the deleterious ef-
fects of the reactive oxygen species generated during electron
transport are additional costs of compensation (review in Gud-
erley 2004). Furthermore, natural covariation between tem-
perature and other resources can enhance the costs of com-
pensation. A seasonal decrease in temperature is usually
accompanied by a reduction in the availability of food, which
could favor reverse compensation rather than partial or perfect
compensation (Clarke 1993). Indeed, the nutritional state of
an animal has marked effects on the aerobic and glycolytic
capacities of its muscles (Pelletier et al. 1993a, 1993b; Dutil et
al. 1998). More effort must be expended to determine the net
benefits of acclimation in natural environments (Huey and Ber-
rigan 1996), and cooperation between physiologists and ecol-
ogists will be invaluable in this undertaking (Kingsolver and
Huey 1998).
A Theory of Coadaptation
Although the coadaptation of thermoregulation and thermal
reaction norms has been appreciated for over 20 yr (Huey and
Slatkin 1976; Huey and Bennett 1987), this appreciation has
not led to an understanding of why some organisms modify
behavior, physiology, and biochemistry in certain combina-
tions. Consider the two most commonly tested hypotheses re-
garding thermal coadaptation: (1) the mean (or preferred) body
temperature of an organism should match the thermal opti-
mum for performance, and (2) the variance in body temper-
ature should correlate with the thermal performance breadth.
These predictions deal with covariation between phenotypes
but say nothing about which phenotypes should be observed
in a given environment. Do some organisms, such as tropical
anurans (Navas 1996), use mainly biochemical modifications
of thermal sensitivity because the cost of thermoregulation is
too high? Are other organisms, such as temperate lizards (Sears
and Angilletta 2004), more likely to use thermoregulation be-
cause its costs are lower in their environments? And what are
the causes of variation along this continuum within major
groups of ectotherms?
Currently, no theory of coadaptation links phenotypic strat-
egies to thermal environments because models of thermoreg-
ulation assume a thermal reaction norm and models of thermal
reaction norms assume a profile of body temperatures. Optimal
thermal reaction norms depend on temporal variation in body
temperature (Huey and Kingsolver 1993; Gilchrist 1995). The
benefit of thermoregulation depends on thermal reaction
norms for all phenotypes that contribute to fitness; benefits are
high for thermal specialists and low for thermal generalists
(Huey and Slatkin 1976; Huey 1982). Each body of theory
makes assumptions about the other; therefore, one must view
the evolution of thermoregulation and thermal sensitivities of
performance as a coadaptive process (Huey and Bennett 1987).
Current theories enable us to understand the covariation be-
tween thermoregulation and thermal physiology, but they do
not enable us to predict which suites of traits will evolve in
particular environments.
How can one unify current theories of thermal adaptation
to produce a more powerful theory of coadaptation? We start
by offering a conceptual model, which can be further developed
quantitatively. Two assumptions pervade current theories
(Huey and Kingsolver 1989): (1) a jack-of-all-temperatures is
a master of none (i.e., a trade-off exists between performance
at a given temperature and the thermal breadth of perfor-
mance), and (2) hotter is better (i.e., performance at the ther-
mal optimum is greater when the thermal optimum is higher).
If the first assumption is valid, a stenotherm with a constant
body temperature that is equal to its thermal optimum should
have a higher fitness than a eurytherm with any distribution
of body temperatures (Gilchrist 1995). If the second assump-
tion is valid, a stenotherm with a high and constant body
Evolutionary Thermal Biology 289
Figure 5. Conceptual model illustrating how thermoregulation and the thermal sensitivity of performance coevolve in ectotherms.
temperature should have a higher fitness than a stenotherm
with a low and constant body temperature. Using these two
assumptions, we can predict the strategies produced by coad-
aptation. In the simplest case, an organism in a stable envi-
ronment obviously must thermoconform and should be a spe-
cialist for the modal temperature (see Janzen 1967). But what
about organisms in variable environments? Here, the cost of
thermoregulation might be the primary driver of coadaptation
(Fig. 5). Environments in which the cost of thermoregulation
is low will favor specialists that thermoregulate precisely. The
preferred body temperature (or preferred range of body tem-
peratures) should depend on the cost-benefit ratio of ther-
moregulation, the effect of temperature on absolute perfor-
mance (i.e., whether hotter truly is better), and physical
constraints on the stability of cellular and molecular structures.
Alternatively, environments in which the cost of thermoregu-
lation is high will favor either specialists or generalists that
thermoconform; organisms whose generation is long should be
specialists, whereas organisms whose generation is short should
be generalists. This conclusion follows from the fact that the
optimal thermal sensitivity is determined by the relative mag-
nitude of within-generation versus among-generation variation
in temperature (see “Adaptive Responses to Thermal Hetero-
geneity”). Note that eurythermy on a generational scale might
be achieved by developmental plasticity or acclimatization,
which would appear to be stenothermy on a shorter temporal
scale. If a trade-off between specialization and generalization
is not universal, the process of coadaptation might be more
complex. Yet, our current understanding of biochemical ad-
aptation (Hochachka and Somero 2002; Guderley 2004) leads
us to conclude that this trade-off must occur at the biochemical
level, even if it can be masked at the organismal level (see
Angilletta et al. 2003).
Putting the Game into Theories of Thermal Adaptation
Evolutionary thermal biologists have made great strides in dis-
covering the processes that operate at or below the organismal
level but have generally underappreciated the biotic arena in
which organisms employ their behavioral and physiological
strategies. This oversight is hardly punishable, considering that
ecologists have only just begun to appreciate the role of en-
vironmental temperature in shaping the dynamics of com-
munities (Panikov 1999; Rooney and Kalff 2000; Al-Rabai’ah
et al. 2002; Garvey et al. 2003; McClanahan and Maina 2003;
Genner et al. 2004). In nature, temperature is likely to influence
the quantity and quality of competitors, predators, and mu-
tualists. These other players must also find strategies for coping
with thermal heterogeneity while dealing with their own set of
competitors, predators, and mutualists. The result is a rich web
of interactions among species that must ultimately shape the
process of thermal coadaptation within species (see Magnuson
et al. 1979).
This game-theoretic perspective offers more than a macro-
scopic view of thermal adaptation. Indeed, the interactions
among species are likely to produce emergent phenomena that
differ qualitatively from the predictions of current theories (e.g.,
Davis et al. 1998). In other words, one cannot understand
thermal adaptation in a community of species by studying the
evolution of each species in isolation. We draw this conclusion
for two reasons: (1) the net benefits of thermoregulation and
thermal acclimation depend on the strategies adopted by other
290 M. J. Angilletta Jr., A. F. Bennett, H. Guderley, C. A. Navas, F. Seebacher, and R. S. Wilson
Figure 6. Variation in environmental temperature is perceived differently by each species in a community. Species that live one or more years
experience more variation within generations than they do among generations, whereas those that live only a few weeks experience more
variation among generations than within generations. This point is illustrated by data from Oak Ridge, Tennessee (360005N, 841456 W;
elevation 266 m), recorded by the National Oceanic and Atmospheric Administration. The effect of generation time is shown by expanding
several periods of the year 1999, ranging from 6 wk (oval frame) to the entire year (rectangular frame).
organisms, and (2) species within a community are expected
to have different thermal sensitivities of performance. The first
point should be obvious from studies of thermoregulation un-
der perceived risks of predation (see “Adaptive Responses to
Thermal Heterogeneity”). Whether predators are capable of
activity under a thermal condition affects the cost of ther-
moregulation by their prey (Huey and Slatkin 1976; Huey
1982). The same coevolutionary dynamic applies to seasonal
acclimation; whether a species should compensate for seasonal
changes or submit to a decrease in physiological function de-
pends on whether their predators and prey are doing so because
the net benefit of compensation depends on the availability of
resources (Clarke 1993). The expectation that species in a com-
munity will have different thermal sensitivities of performance
follows from the recognition that all species perceive the ther-
mal environment differently. For example, consider an envi-
ronment in which temperature cycles daily and annually (Fig.
6). Species with generations of a few weeks (or less) will ex-
perience less variation within generations and more variation
among generations; moreover, variation among generations will
be unpredictable because temperature fluctuates stochastically
within seasons (hence the terms “cold snap” and “heat wave”).
Species with generations of a few months will experience mod-
erate variation within and among generations; however, vari-
ation among generations will be caused by predictable changes
in season. Finally, species with generations of a year or more
will see more variation within generations than among gen-
erations; much of the variation within generations will be
caused by predictable changes in season. Spatial variation is
also perceived differently because small and large species at the
same location can have very different body temperatures (Ste-
venson 1985); this problem is exacerbated by differences in
mobility between small and large species (or between plants
and animals; Bradshaw 1972; Huey et al. 2002). These differ-
ences in the perception of temporal and spatial variation among
species should influence the evolution of thermal sensitivities
and acclimatory responses and hence should have consequences
for the ecological and evolutionary dynamics of communities.
The Need for Immediate Progress in Evolutionary
Thermal Biology
A theory of thermal coadaptation should have impacts beyond
the realm of basic science. Over 99% of all organisms are ec-
tothermic, and global climate change has and will continue to
pose thermal challenges to these organisms. On a regional basis,
some environments have warmed by as much as 2Cinthe
last 30 yr, while other environments have cooled by a similar
degree (Walther et al. 2002). These thermal changes have altered
the growth, phenology, and distribution of many plants and
animals (Atrill and Power 2002; Beaugrand et al. 2002; Walther
et al. 2002; Parmesan and Yohe 2003). Because temperature
Evolutionary Thermal Biology 291
affects many traits that determine the interactions among spe-
cies (Garvey et al. 2003; Jiang and Morin 2004), the structure
and dynamics of communities will likely be perturbed by global
climate change. These perturbations will be particularly prob-
lematic when responses to thermal change are asynchronous
among species whose interactions are normally in a dynamic
equilibrium (e.g., Fitter and Fitter 2002).
The development of a game theory of thermal coadaptation
could offer a major advantage to applied ecologists. Although
great effort has been expended to understand the ecological
consequences of global climate change, the evolutionary con-
sequences remain less clear. Yet, evolutionary responses will
determine the interactions among organisms in future ecosys-
tems. This interplay between ecological and evolutionary re-
sponses will be complex because certain species will evolve more
rapidly than others. Knowledge of the rates and magnitudes of
potential evolutionary responses could help us to assess the
long-term consequences of global climate change. Thus, a game
theory of thermal coadaptation might ultimately lead to insights
that will enable us to manage biodiversity successfully in the
coming decades.
Acknowledgments
We thank S. Morris and A. Vosloo for organizing the Third
International Congress of Comparative Physiology and Bio-
chemistry. Participation by M.J.A. was made possible by an
international travel grant from Indiana State University. A.F.B.,
H.G., and C.A.N. were supported by grants from the National
Science Foundation (IBN0091308), Natural Sciences and En-
gineering Research Council, and Fundac¸a˜o de Amparo a`
Pesquisa do Estado de Sa˜o Paulo (03-01577-8), respectively.
R.S.W. was supported by a grant from the University of Queens-
land and a discovery grant from the Australian Research Coun-
cil. We thank R. Huey for providing thoughtful comments on
a previous version of the manuscript.
Literature Cited
Adolph S.C. 1990. Influence of behavioral thermoregulation on
microhabitat use by two Sceloporus lizards. Ecology 71:315
327.
Al-Rabai’ah H.A., H.L. Koh, D. DeAngelis, and H.L. Lee. 2002.
Modeling fish community dynamics in the Florida Ever-
glades: role of temperature variation. Water Sci Technol 46:
71–78.
Andrews R.M. 1998. Geographic variation in field body tem-
perature of Sceloporus lizards. J Therm Biol 23:329–334.
Angilletta M.J. 2001. Thermal and physiological constraints on
energy assimilation in a widespread lizard (Sceloporus un-
dulatus). Ecology 82:3044–3056.
Angilletta M.J., T. Hill, and M.A. Robson. 2002a. Is physio-
logical performance optimized by thermoregulatory behav-
ior? a case study of the eastern fence lizard, Sceloporus un-
dulatus. J Therm Biol 27:199–204.
Angilletta M.J., P.H. Niewiarowski, and C.A. Navas. 2002b. The
evolution of thermal physiology in ectotherms. J Therm Biol
27:249–268.
Angilletta M.J. and Y.L. Werner. 1998. Australian geckos do not
display diel variation in thermoregulatory behavior. Copeia
1998:736–742.
Angilletta M.J., R.S. Wilson, C.A. Navas, and R.S. James. 2003.
Tradeoffs and the evolution of thermal reaction norms.
Trends Ecol Evol 18:234–240.
Atrill M.J. and M. Power. 2002. Climatic influence on a marine
fish assemblage. Nature 417:275–278.
Bartholomew G.A. 1982. Physiological control of body tem-
perature. Pp. 167–211 in C. Gans and F.H. Pough, eds. Bi-
ology of the Reptilia. Vol. 12. Academic Press, New York.
Bauwens D., P.E. Hertz, and A.M. Castilla. 1996. Thermoreg-
ulation in a lacertid lizard: the relative contributions of dis-
tinct behavioral mechanisms. Ecology 77:1818–1830.
Beaugrand, G., P.C. Reid, F. Iban˜ez, J.A. Lindley, and M. Edward.
2002. Reorganization of North Atlantic marine copepod bio-
diversity and climate. Science 296:1692–1694.
Bennett A.F. 2005. Thermoregulation in African chameleons.
Pp. 234–241 in S. Morris and A. Vosloo, eds. Animals and
Environments: Proceedings of the Third International Con-
ference of Comparative Physiology and Biochemistry. Elsev-
ier, Amsterdam.
Bennett A.F. and R.E. Lenski. 1997. Evolutionary adaptation to
temperature. VI. Phenotypic acclimation and its evolution
in Escherichia coli. Evolution 51:36–44.
———. 1999. Experimental evolution and its role in evolu-
tionary physiology. Am Zool 39:346–362.
Bouchard P. and H. Guderley. 2003. Time course of the re-
sponse of mitochondria from oxidative muscle during ther-
mal acclimation of rainbow trout, Oncorhynchus mykiss. J
Exp Biol 206:3455–3465.
Bradshaw A.D. 1972. Some of the evolutionary consequences
of being a plant. Evol Biol 5:25–47.
Brown J.H., J.F. Gillooly, A.P. Allen, V.M. Savage, and G.B. West.
2004. Toward a metabolic theory of ecology. Ecology 85:
1771–1789.
Chaui-Berlinck J.G., C.A. Navas, L.H. Monteiro, and J.E. Bi-
cudo. 2004. Temperature effects on a whole metabolic re-
action cannot be inferred from its components. Proc R Soc
Lond B 271:1415–1419.
Christian K.A., C.R. Tracy, and W.P. Porter. 1983. Seasonal shifts
in body temperature and use of microhabitats by Galapagos
land iguanas (Conolophus pallidus). Ecology 64:463–468.
Christian K.A. and B.W. Weavers. 1996. Thermoregulation of
monitor lizards in Australia: an evaluation of methods in
thermal biology. Ecol Monogr 66:139–157.
Clarke A. 1993. Seasonal acclimatization and latitudinal com-
292 M. J. Angilletta Jr., A. F. Bennett, H. Guderley, C. A. Navas, F. Seebacher, and R. S. Wilson
pensation in metabolism: do they exist? Funct Ecol 7:139–
149.
Conover D.O. 1998. Local adaptation in marine fishes: evidence
and implications for stock enhancement. Bull Mar Sci 62:
477–493.
Conover D.O. and E.T. Schultz. 1995. Phenotypic similarity
and the evolutionary significance of countergradient varia-
tion. Trends Ecol Evol 10:248–252.
Cooper V.S., A.F. Bennett, and R.E. Lenski. 2001. Evolution of
thermal dependence of growth rate of Escherichia coli pop-
ulations during 20,000 generations in a constant environ-
ment. Evolution 55:889–896.
Davis A.J., J.H. Lawton, B. Shorrocks, and L.S. Jenkinson. 1998.
Individualistic species responses invalidate simple physio-
logical models of community dynamics under global envi-
ronmental change. J Anim Ecol 67:600–612.
de Souza S.C., J.E. de Carvalho, A.S. Abe, J.E. Bicudo, and M.S.
Bianconcini. 2004. Seasonal metabolic depression, substrate
utilisation and changes in scaling patterns during the first
year cycle of tegu lizards (Tupinambis merianae). J Exp Biol
207:307–318.
DeWitt T.J., A. Sih, and D.S. Wilson. 1998. Costs and limits to
phenotypic plasticity. Trends Ecol Evol 13:77–81.
Downes S. 2001. Trading heat and food for safety: costs of
predator avoidance in a lizard. Ecology 82:2870–2881.
Downes S. and R. Shine. 1998. Heat, safety or solitude? using
habitat selection experiments to identify a lizard’s priorities.
Anim Behav 55:1387–1396.
Dutil J.-D., Y. Lambert, H. Guderley, P.U. Blier, D. Pelletier,
and M. Desroches. 1998. Nucleic acids and enzymes in At-
lantic cod (Gadus morhua) differing in condition and growth
rate trajectories. Can J Fish Aquat Sci 55:788–795.
Dzialowski E.M. and M.P. O’Connor. 2004. Importance of the
limbs in the physiological control of heat exchange in Iguana
Iguana and Sceloporus undulatus. J Therm Biol 29:299–305.
Feder M.E., A.F. Bennett, and R.B. Huey. 2000. Evolutionary
physiology. Annu Rev Ecol Syst 31:315–341.
Fitter A.H. and R.S.R. Fitter. 2002. Rapid changes in flowering
time in British plants. Science 296:1689–1691.
Gabriel W. and M. Lynch. 1992. The selective advantage of
reaction norms for environmental tolerance. J Evol Biol 5:
41–59.
Garland T., Jr., and P.A. Carter. 1994. Evolutionary physiology.
Annu Rev Ecol Syst 56:579–621.
Garvey J.E., D.R. Devries, R.A. Wright, and J.G. Miner. 2003.
Energetic adaptations along a broad latitudinal gradient: im-
plications for widely distributed assemblages. Bioscience 53:
141–150.
Genner M.J., D.W. Sims, V.J. Wearmouth, E.J. Southall, A.J.
Southward, P.A. Henderson, and S.J. Hawkins. 2004. Re-
gional climatic warming drives long-term community
changes of British marine fish. Proc R Soc Lond B 271:655–
661.
Gibert P., R.B. Huey, and G.W. Gilchrist. 2001. Locomotor
performance of Drosophila melanogaster: interactions among
developmental and adult temperatures, age, and geography.
Evolution 55:205–209.
Gilchrist G.W. 1995. Specialists and generalists in changing en-
vironments. I. Fitness landscapes of thermal sensitivity. Am
Nat 146:252–270.
Grant B.W. and A.E. Dunham. 1990. Elevational covariation
in environmental constraints and life histories of the desert
lizard, Sceloporus merriami. Ecology 71:1765–1776.
Guderley H. 2004. Metabolic responses to low temperature in
fish muscle. Biol Rev 79:409–427.
Guderley H. and J. St. Pierre. 1999. Seasonal cycles of mito-
chondrial ADP sensitivity and oxidative capacities. J Comp
Physiol B 169:474–480.
———. 2002. Going with the flow or life in the fast lane:
contrasting mitochondrial responses to thermal change. J
Exp Biol 205:2237–2249.
Guppy M. and P. Withers. 1999. Metabolic depression in an-
imals: physiological perspectives and biochemical generali-
zations. Biol Rev 74:1–40.
Hand S.C. and I. Hardewig. 1996. Downregulation of cellular
metabolism during environmental stress: mechanisms and
implications. Annu Rev Physiol 58:539–563.
Haynie D.T. 2001. Biological Thermodynamics. Cambridge
University Press, Cambridge.
Hazel J.R. 1995. Thermal adaptation in biological membranes:
is homeoviscous adaptation the explanation? Annu Rev
Physiol 57:19–42.
Heath J.E. 1964. Reptilian thermoregulation: evaluation of field
studies. Science 146:784–785.
———. 1965. Temperature regulation and diurnal activity in
horned lizards. Univ Calif Publ Zool 64:97–136.
Hertz P.E. 1981. Adaptation to altitude in two West Indian
anoles (Reptilia: Iguanidae): field thermal biology and phys-
iological ecology. J Zool (Lond) 195:25–37.
Hertz P.E. and R.B. Huey. 1981. Compensation for altitudinal
changes in the thermal environment by some Anolis lizards
on Hispaniola. Ecology 62:515–521.
Hertz P.E., R.B. Huey, and T. Garland Jr. 1988. Time budgets,
thermoregulation, and maximal locomotor performance: are
reptiles Olympians or body scouts? Am Zool 28:927–938.
Hertz P.E., R.B. Huey, and E. Nevo. 1983. Homage to Santa
Anita: thermal sensitivity of sprint speed in agamid lizards.
Evolution 37:1075–1084.
Hochachka P.W. and G.N. Somero. 2002. Biochemical Adap-
tation. Oxford University Press, Oxford.
Hoffmann A.A. 1995. Acclimation: increasing survival at a cost.
Trends Ecol Evol 10:1–2.
Huey R.B. 1974. Behavioral thermoregulation in lizards: im-
portance of associated costs. Science 184:1001–1003.
———. 1982. Temperature, physiology, and the ecology of rep-
Evolutionary Thermal Biology 293
tiles. Pp. 25–74 in C. Gans and F.H. Pough, eds. Biology of
the Reptilia. Vol. 12. Academic Press, New York.
Huey R.B. and A.F. Bennett. 1987. Phylogenetic studies of coad-
aptation: preferred temperatures versus optimal performance
temperatures of lizards. Evolution 41:1098–1115.
Huey R.B. and D.A. Berrigan. 1996. Testing evolutionary hy-
potheses of acclimation. Pp. 205–237 in I.A. Johnston and
A.F. Bennett, eds. Animals and Temperature: Phenotypic and
Evolutionary Adaptation. Society for Experimental Biology
Seminar Series. Cambridge University Press, Cambridge.
Huey R.B., D. Berrigan, G.W. Gilchrist, and J.C. Herron. 1999.
Testing the adaptive significance of acclimation: a strong in-
ference approach. Am Zool 39:323–336.
Huey R.B., M. Carlson, L. Crozier, M. Frazier, H. Hamilton,
C. Harley, A. Hoang, and J.G. Kingsolver. 2002. Plants versus
animals: do they deal with stress in different ways? Integr
Comp Biol 42:415–423.
Huey R.B. and P.E. Hertz. 1984. Is a jack-of-all-temperatures
a master of none? Evolution 38:441–444.
Huey R.B. and J.G. Kingsolver. 1989. Evolution of thermal
sensitivity of ectotherm performance. Trends Ecol Evol 4:
131–135.
———. 1993. Evolution of resistance to high temperature in
ectotherms. Am Nat 142(suppl.):S21–S46.
Huey R.B., P.H. Niewiarowski, J. Kaufmann, and J.C. Herron.
1989. Thermal biology of nocturnal ectotherms: is sprint
performance of geckos maximal at low body temperatures?
Physiol Zool 62:488–504.
Huey R.B. and M. Slatkin. 1976. Cost and benefits of lizard
thermoregulation. Q Rev Biol 51:363–384.
Huey R.B. and R.D. Stevenson. 1979. Integrating thermal phys-
iology and ecology of ectotherms: discussion of approaches.
Am Zool 19:357–366.
Janzen D.H. 1967. Why mountain passes are higher in the
tropics. Am Nat 101:233–249.
Jiang L. and P.J. Morin. 2004. Temperature-dependent inter-
actions explain unexpected responses to environmental
warming in communities of competitors. J Anim Ecol 73:
569–576.
Johnston I.A. and G.K. Temple. 2002. Thermal plasticity of
skeletal muscle phenotype in ectothermic vertebrates and its
significance for locomotory behaviour. J Exp Biol 205:2305–
2322.
Kearney M. and M. Predavec. 2000. Do nocturnal ectotherms
thermoregulate? a study of the temperate gecko Christinus
marmoratus. Ecology 81:2984–2996.
Kingsolver J.G. and R.B. Huey. 1998. Evolutionary analyses of
morphological and physiological plasticity in thermally var-
iable environments. Am Zool 38:545–560.
Leroi A.M., A.F. Bennett, and R.E. Lenski. 1994. Temperature
acclimation and competitive fitness: an experimental test of
the beneficial acclimation assumption. Proc Natl Acad Sci
USA 91:1917–1921.
Levins R. 1962. Theory of fitness in a heterogenous environ-
ment. I. The fitness set and adaptive functions. Am Nat 96:
361–373.
———. 1963. Theory of fitness in a heterogenous environment.
II. Developmental flexibility and niche selection. Am Nat 97:
75–90.
———. 1968. Evolution in Changing Environments. Princeton
University Press, Princeton, NJ.
Lynch M. and W. Gabriel 1987. Environmental tolerance. Am
Nat 129:283–303.
Magnuson J.J., L.B. Crowder, and P.A. Medvick. 1979. Tem-
perature as an ecological resource. Am Zool 19:331–343.
McClanahan T.R. and J. Maina. 2003. Response of coral as-
semblages to the interaction between natural temperature
variation and rare warm-water events. Ecosystems 6:551–563.
Mercer J. 2001. Glossary of terms for thermal physiology. Jpn
J Physiol 51:245–280.
Mongold J.A., A.F. Bennett, and R.E. Lenski. 1999. Evolutionary
adaptation to temperature. VII. Extension of the upper ther-
mal limit of Escherichia coli. Evolution 53:386–394.
Navas C.A. 1996. Metabolic physiology, locomotor perfor-
mance, and thermal niche breadth in Neotropical anurans.
Physiol Zool 69:1481–1501.
———. 2002. Herpetological diversity along Andean eleva-
tional gradients: links with physiological ecology and evo-
lutionary physiology. Comp Biochem Physiol 133A:469–485.
O’Steen S. and A.F. Bennett. 2003. Thermal acclimation effects
differ between voluntary, maximum, and critical swimming
velocities in two cyprinid fishes. Physiol Biochem Zool 76:
484–496.
Palaima A. and K. Spitze. 2004. Is a jack-of-all-temperatures a
master of none? an experimental test with Daphnia pulicaria
(Crustacea: Cladocera). Evol Ecol Res 6:215–225.
Panikov N.S. 1999. Understanding and prediction of soil mi-
crobial community dynamics under global change. Appl Soil
Ecol 11:161–176.
Parmesan C. and G. Yohe. 2003. A globally coherent fingerprint
of climate change impacts across natural systems. Nature 421:
37–42.
Pelletier D., H. Guderley, and J.-D. Dutil. 1993a. Does the
aerobic capacity of fish muscle change with growth rate? Fish
Physiol Biochem 12:83–93.
———. 1993b. Effects of growth rate, temperature, season and
body size on glycolytic enzyme activities in the white muscle
of Atlantic cod (Gadus morhua). J Exp Zool 265:477–487.
Platt J.R. 1964. Strong inference. Science 146:347–353.
Po¨rtner H.O., A.F. Bennett, F. Bozinovic, A. Clarke, M.A. Lar-
dies, M. Lucassen, B. Pelster, F. Schiemer, and J.H. Stillman.
2006. Trade-offs in thermal adaptation: the need for a mo-
lecular to ecological integration. Physiol Biochem Zool 79:
295–313.
Precht H. 1958. Concepts of the temperature adaptation of
unchanging reaction systems of cold-blooded animals. Pp.
294 M. J. Angilletta Jr., A. F. Bennett, H. Guderley, C. A. Navas, F. Seebacher, and R. S. Wilson
50–78 in C.L. Prosser, ed. Physiological Adaptations. Amer-
ican Physiological Society, Washington, DC.
Prosser C.L. 1979. Comparative Animal Physiology. Wiley, New
York.
Quinn J. and A.E. Dunham. 1983. On hypothesis testing in
ecology and evolution. Am Nat 122:602–617.
Rock J., A. Cree, and R.M. Andrews. 2002. The effect of re-
productive condition on thermoregulation in a viviparous
gecko from a cool climate. J Therm Biol 27:17–27.
Rome L.C, P.T. Loughna, and G. Goldspink. 1984. Temperature
acclimation: improved sustained swimming performance in
carp at low temperatures. Science 228:194–196.
Rooney N. and J. Kalff. 2000. Inter-annual variation in sub-
merged macrophyte community biomass and distribution:
the influence of temperature and lake morphometry. Aquat
Bot 68:321–335.
Sa¨nger A.M. 1993. Limits to the acclimation of fish muscle.
Rev Fish Biol Fish 3:1–15.
Sears M.W. and M.J. Angilletta. 2004. Body size clines in Sce-
loporus lizards: proximate mechanisms and demographic
constraints. Integr Comp Biol 44:433–442.
Stevenson R.D. 1985. Body size and limits to the daily range
of body temperatures in terrestrial ectotherms. Am Nat 125:
102–117.
St. Pierre J. and R.G. Boutilier. 2001. Aerobic capacity of frog
skeletal muscle during hibernation. Physiol Biochem Zool
74:390–397.
van Berkum F.H. 1988. Latitudinal patterns of the thermal sen-
sitivity of sprint speed in lizards. Am Nat 132:327–343.
Van Damme R., D. Bauwens, A.M. Castilla, and R.F. Verheyen.
1989. Altitudinal variation of the thermal biology and run-
ning performance in the lizard Podarcis tiliguerta. Oecologia
80:516–524.
Walther G.-R., E. Post, P. Convey, A. Menzel, C. Parmesan,
T.J. Beebee, J.-M. Fromentin, O. Hoegh-Guldberg, and F.
Bairlein. 2002. Ecological responses to recent climate change.
Nature 416:389–395.
Wilson R.S. and C.E. Franklin. 2002. Testing the beneficial ac-
climation hypothesis. Trends Ecol Evol 17:66–70.
Wilson R.S. and I.A. Johnston. 2005. Combining studies of
comparative physiology and behavioural ecology to test the
adaptive benefits of thermal acclimation. Pp. 201–208 in S.
Morris and A. Vosloo, eds. Animals and Environments: Pro-
ceedings of the Third International Conference of Compar-
ative Physiology and Biochemistry. Elsevier, Amsterdam.
Zamudio K.R., R.B. Huey, and W.D. Crill. 1995. Bigger isn’t
always better: body size, developmental and parental tem-
perature and male territorial success in Drosophila melano-
gaster. Anim Behav 49:671–677.
... From a global warming perspective, particularly in the realm of environmental influences on marine ecosystems, attention has been directed towards physiological responses in ectotherms relating to thermal tolerance and suitability of habitats (Farrell, 2009;Habary et al., 2017;McKenzie et al., 2016;Pörtner & Knust, 2007). Body temperature is a fundamental controlling factor of all activities in ectothermic animals (Angilletta et al., 2002;Angilletta Jr et al., 2006;Fry, 1971;Huey & Stevenson, 1979), and species have evolved their performances and tolerances within their local habitat's temperature regimes (Pörtner & Farrell, 2008). For example, species which have evolved in relatively thermally stable habitats typically have narrow thermal tolerance envelopes (Angilletta Jr et al., 2006;Habary et al., 2017;Huey & Kingsolver, 1993) and are therefore less resilient to temperature increases and thermal fluctuations, such as those caused by global warming (Angilletta Jr et al., 2006;Pörtner, 2001;Pörtner & Farrell, 2008). ...
... Body temperature is a fundamental controlling factor of all activities in ectothermic animals (Angilletta et al., 2002;Angilletta Jr et al., 2006;Fry, 1971;Huey & Stevenson, 1979), and species have evolved their performances and tolerances within their local habitat's temperature regimes (Pörtner & Farrell, 2008). For example, species which have evolved in relatively thermally stable habitats typically have narrow thermal tolerance envelopes (Angilletta Jr et al., 2006;Habary et al., 2017;Huey & Kingsolver, 1993) and are therefore less resilient to temperature increases and thermal fluctuations, such as those caused by global warming (Angilletta Jr et al., 2006;Pörtner, 2001;Pörtner & Farrell, 2008). Unfortunately, there is a general lack of knowledge on the thermal performance of Arctic fishes and whether they will be influenced as much as tropical fishes in future warmer waters (Nati et al., 2021). ...
... Body temperature is a fundamental controlling factor of all activities in ectothermic animals (Angilletta et al., 2002;Angilletta Jr et al., 2006;Fry, 1971;Huey & Stevenson, 1979), and species have evolved their performances and tolerances within their local habitat's temperature regimes (Pörtner & Farrell, 2008). For example, species which have evolved in relatively thermally stable habitats typically have narrow thermal tolerance envelopes (Angilletta Jr et al., 2006;Habary et al., 2017;Huey & Kingsolver, 1993) and are therefore less resilient to temperature increases and thermal fluctuations, such as those caused by global warming (Angilletta Jr et al., 2006;Pörtner, 2001;Pörtner & Farrell, 2008). Unfortunately, there is a general lack of knowledge on the thermal performance of Arctic fishes and whether they will be influenced as much as tropical fishes in future warmer waters (Nati et al., 2021). ...
Article
Full-text available
Global warming affects the metabolism of ectothermic aquatic breathers forcing them to migrate and undergo high‐latitudinal distribution shifts to circumvent the temperature‐induced mismatch between increased metabolic demand and reduced water oxygen availability. Here the authors examined the effects of temperature on oxygen consumption rates in an Arctic stenotherm, the Greenland halibut Reinhardtius hippoglossoides, and calculated the optimal temperature for maximum aerobic scope, AS(Topt,AS), which was found to be 2.44°C. They also investigated cardiac performance as limiting the oxygen transport chain at high temperatures by measuring maximum heart rate (fHmax) over acute temperature increases and found various metrics related to fHmax to be at least 3.2°C higher than Topt,AS. The authors’ measured Topt,AS closely reflected in situ temperature occurrences of Greenland halibut from long‐term tagging studies, showing that AS of the species is adapted to its habitat temperature, and is thus a good proxy for the species' sensitivity to environmental warming. The authors did not find a close connection between fHmax and Topt,AS, suggesting that cardiac performance is not limiting for the oxygen transport chain at high temperatures in this particular Arctic stenotherm. The authors’ estimate of the thermal envelope for AS of Greenland halibut was from −1.89 to 8.07°C, which is exceptionally narrow compared to most other species of fish. As ocean temperatures increase most rapidly in the Arctic in response to climate change, and species in these areas have limited possibility for further poleward‐range shifts, these results suggest potential severe effects of global warming on Arctic stenotherms, such as the Greenland halibut. The considerable economic importance of the species raises concerns for future fisheries and species conservation of Arctic stenotherms in the Northern Hemisphere.
... In this regard, thermal performance curves (TPCs) have provided a general framework to assess how temperature affects different aspects of life, from biochemical and physiological rates up to ecologically relevant variables such as predation rates and population growth [8][9][10][11]. These curves describe how these attributes vary as a function of temperature, indicating the range of temperatures in which performance can be sustained (CT min and CT max ), the optimal temperature (T opt ) in which performance is maximized, among others [11][12][13][14]. They are pervasive across evolutionary lineages and levels of organization, and suggest that thermodynamic constraints at lower levels scale up and somehow contribute to macroecological patterns such as distribution ranges [11,[14][15][16]. ...
... In other words, do patterns of thermal adaptation observed at a given level of organization hold at different levels or do responses vary depending on the studied trait? A correspondence across traits would indicate that they are co-adapted [12], as thermal co-adaptation would imply that natural selection in Darwinian fitness should result in correlated responses in the physiological machinery. Under this scenario, one should expect congruent variation in TPCs of warm-versus cold-adapted populations or lineages across levels of biologic organization. ...
... We now discuss correlated responses between metabolism and egg-to-adult survival. The TPC descriptors obtained for these traits were always positively correlated (figure 4), indicating that their thermal responses are indeed co-adapted [12]. That is not to say, however, that descriptors are quantitatively identical because these traits are expressed at different levels of organization and temporal windows, which are known to affect the overall shape of performance curves [11,30,31,49]. ...
Article
Full-text available
Understanding how species adapt to different temperatures is crucial to predict their response to global warming, and thermal performance curves (TPCs) have been employed recurrently to study this topic. Nevertheless, fundamental questions regarding how thermodynamic constraints and evolution interact to shape TPCs in lineages inhabiting different environments remain unanswered. Here, we study Drosophila simulans along a latitudinal gradient spanning 3000 km to test opposing hypotheses based on thermodynamic constrains (hotter-is-better) versus biochemical adaptation (jack-of-all-temperatures) as primary determinants of TPCs variation across populations. We compare thermal responses in metabolic rate and the egg-to-adult survival as descriptors of organismal performance and fitness, respectively, and show that different descriptors of TPCs vary in tandem with mean environmental temperatures, providing strong support to hotter-is-better. Thermodynamic constraints also resulted in a strong negative association between maximum performance and thermal breadth. Lastly, we show that descriptors of TPCs for metabolism and egg-to-adult survival are highly correlated, providing evidence of co-adaptation, and that curves for egg-to-adult survival are systematically narrower and displaced toward lower temperatures. Taken together, our results support the pervasive role of thermodynamics constraining thermal responses in Drosophila populations along a latitudinal gradient, that are only partly compensated by evolutionary adaptation.
... Thus, lower juvenile mass in less heterothermic animals does not necessarily have to be correlated with parents' physiologies but could be constrained by higher litter size itself. The modern theoretical model for adaptive thermoregulation suggests the existence of a continuum of thermoregulatory strategies (Angilletta et al., 2006;. Since the overall area under the theoretical performance curve is expected to be maximized (but also finite), thermoregulatory specialists (e.g. ...
... Since the overall area under the theoretical performance curve is expected to be maximized (but also finite), thermoregulatory specialists (e.g. strictly homeothermic animals) operating at a narrow T b range should perform better at optimal temperatures than thermoregulatory generalistsheterothermic animals characterized by a wide T b range (Angilletta et al., 2006(Angilletta et al., , 2010. This varying thermosensitivity can underlie/cause (Boyles and Warne, 2013;Seebacher and Little, 2017) intraindividual variation in the performance of skeletal muscles (Choi et al., 1998;Rojas et al., 2012;James et al., 2015), immune function (Butler et al., 2013), and generally complex molecular networks (Schwartz and Bronikowski, 2013). ...
Article
Full-text available
Heterothermy is considered to be the most effective energy-saving strategy improving survival under natural conditions. Interspecific studies suggest that this strategy is also associated with reduced reproductive output. Yet little is known about the reproductive consequences of heterothermy use at the intraspecific level and thus its repercussions for microevolutionary processes. Moreover, as yet no study has aimed to test if litter size and juvenile mass are affected by torpor use in wild captured animals under undemanding laboratory conditions. Here we tested the hypothesis that intraspecific variation in heterothermy use is associated with different reproductive successes, being the result of the evolution of distinct life histories. We predicted that heterothermy use in winter negatively correlates with litter size and juvenile body mass during the subsequent breeding season. To test this prediction, we used yellow-necked mice from a population in which individuals consistently differ in their use of heterothermy in winter. We measured body size (head width) and body mass, basal metabolic rate, as well as metabolism and body temperature during fasting-induced torpor in wild caught mice in winter. Phenotyped mice were bred in the subsequent summer selectively – males and females with similar heterothermy characteristics were paired, the most to the least heterothermic. Dam body size, but not basal metabolism, was positively correlated with litter size (but not juvenile mass). However, when accounting for this relationship, litter size was negatively while juvenile mass was positively correlated with the average heterothermy use of a given couple. Our study indicates that heterothermy use correlates with specific life-history strategies arising from a fundamental evolutionary trade-off between survival and reproduction.
... However, temperature variations across elevations are the major driver for the environmental selection of ecological strategies (Jump et al., 2009;Körner & Paulsen, 2004;Midolo et al., 2019) by exerting a substantial influence on species' fitness (i.e. growth, productivity and survival) while shaping the kinetic energy of molecules and biological processes (Angilletta et al., 2006;Madeira et al., 2012;Nievola et al., 2017), such as enzyme activity, protein folding and particularly membrane fluidity (Hasanuzzaman, 2020;Keller & Seehausen, 2012;Kingsolver, 2009;Lamers et al., 2020). ...
Article
1. The effect of environmental gradients on the remarkable diversity of mountain- associated plants and on the species' abilities to cope with climate change tran-scends species-specific strategies. For instance, our understanding of the impact of thermal gradients on ecological divergences in populations of widely distrib-uted species is limited, although it could provide important insights regarding spe-cies' response to climate change. 2. Here, we investigated whether populations of an endemic species broadly distrib-uted across an elevation gradient employ unique or multiple divergent ecological strategies according to specific environmental conditions. We hypothesised that populations employ distinct strategies, producing a tolerance-avoidance trade-off related to the thermal conditions they experience across elevations. 3. We conducted our research with 125 individuals of Pitcairnia flammea(Bromeliaceae) sampled from various elevations spanning from sea level to ~2200 m and cultivated under the same conditions. To assess specific ecological strategies of P. flammea populations across elevations, we examined leaf temper-ature, heat and cold tolerances, as well as other structural/morphological, optical, physiological and biochemical leaf traits. 4. We majorly observed that water-saving traits diminish as elevation increases while membrane fluidity, majorly associated with unsaturated and very- long- chain lipids, enhances. Low- elevation individuals of P. flammea invest in water storage tissues, which likely prevent excessive water loss through the intense transpiration rates under warming periods. Conversely, high-elevation plants ex-hibit increased membrane fluidity, a possible response to the stiffening induced by low temperature. 5. Our results revealed a tolerance-avoidance trade-off related to thermal strate-gies of populations distributed across an elevation gradient. Low-elevation plants avoid excessive leaf temperature by investing in water-saving traits to maintain transpiration rates. High-elevation individuals, in turn, tend to invest in mem-brane properties to tolerate thermal variations, particularly cold events. 6. Our findings challenge the conventional notion that plants' vulnerability to warm-ing depends on species-specific thermal tolerance by showing diverse thermal strategies on populations across an elevation gradient.
... Thus, one would expect seasonal acclimation and behavioural strategies to be coordinated, thereby maximizing both survival during unfavourable seasons and performance during the rest of the year (e.g. [35,36]). Nevertheless, there is limited information available on how seasonal acclimation and behaviour interact in nature (e.g. ...
Article
Full-text available
Many ectothermic organisms counter harsh abiotic conditions by seeking refuge in underground retreats. Variations in soil hydrothermal properties within these retreats may impact their energy budget, survival and population dynamics. This makes retreat site choice a critical yet understudied component of their strategies for coping with climate change. We used a mechanistic modelling approach to explore the implications of behavioural adjustments and seasonal acclimation of metabolic rate on retreat depth and the energy budget of ectotherms, considering both current and future climate conditions. We used a temperate amphibian, the alpine newt (Ichthyosaura alpestris), as a model species. Our simulations predict an interactive influence of different thermo- and hydroregulatory strategies on the vertical positioning of individuals in underground refuges. The adoption of a particular strategy largely determines the impact of climate change on retreat site choice. Additionally, we found that, given the behavioural thermoregulation/hydroregulation and metabolic acclimation patterns considered, behaviour within the retreat has a greater impact on ectotherm energetics than acclimation of metabolic rate under different climate change scenarios. We conclude that further empirical research aimed at determining ectotherm behavioural strategies during both surface activity and inactivity is needed to understand their population dynamics and species viability under climate change.
... Recently, however, L. peronii larvae reared and exposed to UVBR at cool temperatures accumulated less DNA damage than larvae reared in warm temperatures and acutely exposed to UVBR at cool temperatures (Hird, Cramp, et al., 2023). This suggests that thermal acclimation, which is defined as a phenotypic change in response to the thermal environment within the lifetime of an organism (Angilletta et al., 2006;Huey et al., 1999a;Kingsolver & Huey, 1998), can improve the resilience of larvae to UVBR-associated DNA damage. The degree to which amphibians demonstrate plasticity to changing abiotic factors in nature likely reflects their ability to cope with rapid global change (Angilletta, 2009;Chevin et al., 2010;Huey et al., 1999b;Kingsolver & Huey, 1998;Seebacher et al., 2014;Woods and Harrison, 2002). ...
Article
Full-text available
Amphibian declines are sometimes correlated with increasing levels of ultraviolet radiation (UVR). While disease is often implicated in declines, environmental factors such as temperature and UVR play an important role in disease epidemiology. The mutagenic effects of UVR exposure on amphibians are worse at low temperatures. Amphibians from cold environments may be more susceptible to increasing UVR. However, larvae of some species demonstrate cold acclimation, reducing UV-induced DNA damage at low temperatures. Understanding of the mechanisms underpinning this response is lacking. We reared Limnodynastes peronii larvae in cool (15°C) or warm (25°C) waters before acutely exposing them to 1.5 h of high intensity (80 µW cm −2) UVBR. We measured the color of larvae and mRNA levels of a DNA repair enzyme. We reared larvae at 25°C in black or white containers to elicit a skin color response, and then measured DNA damage levels in the skin and remaining carcass following UVBR exposure. Cold-acclimated larvae were darker and displayed lower levels of DNA damage than warm-acclimated larvae. There was no difference in CPD-photolyase mRNA levels between cold-and warm-acclimated larvae. Skin darkening in larvae did not reduce their accumulation of DNA damage following UVR exposure. Our results showed that skin darkening does not explain cold-induced reductions in UV-associated DNA damage in L. peronii larvae. Beneficial cold-acclimation is more likely underpinned by increased CPD-photolyase abundance and/or increased photolyase activity at low temperatures. K E Y W O R D S amphibian declines, DNA repair, ecophysiology, melanin, photolyase, plasticity, temperature
... Yet recent advancements have demonstrated that cost-benefit ratios are determined not by an average of the thermal environment, but by its finer scale landscape characteristics, including variability, configuration, and connectivity (Sears et al., 2016;Sears & Angilletta, 2015). These developments show that a landscape-level perspective of physiological and behavioral "coadaptation" to temperature (Angilletta et al., 2006) is needed to predict which ectotherms will tolerate or avoid climate change threats. However, empirical studies supporting this perspective are overwhelmingly focused on lizards, leaving implications for aquatic ectotherms unclear (Donelson et al., 2019;Huey et al., 2012). ...
Article
Full-text available
Abstract Climate change is a global phenomenon, but natural selection occurs within landscapes. Many global analyses predict how climate change will shape behavior and physiology, but few incorporate information from the landscape scales at which animals actually respond to selective pressure. We compared cold‐water fish (redband trout Oncorhynchus mykiss newberrii) from neighboring habitats in a naturally warm, recently fragmented basin to understand how different responses to warming may arise from landscape constraints. Trout in warm, hydrologically connected Upper Klamath Lake fled summer temperatures and sought refuge in cool tributaries, while trout in an equally warm but fragmented reach of the Klamath River endured summer conditions. Trout in the river were more physiologically tolerant of high temperatures than trout in the lake across multiple metrics, including capacity for aerobic activity, recovery from exertion, and loss of equilibrium. Two independent metrics of energetic condition indicated that the behavioral strategy of trout in the lake came at a substantial energetic cost, while the physiological strategy of trout in the river was able to mitigate most energetic consequences of high temperatures. No clear genetic basis for increased tolerance was found in trout from the river, which may suggest tolerance was derived from plasticity, although our analysis could not rule out genetic adaptation. Our results show that landscape processes such as fragmentation can cause different climate survival strategies to emerge in neighboring populations. Connecting the mechanisms that favor similar survival strategies among related organisms at broad scales with mechanisms that drive landscape‐scale variability within taxa should be a major goal for future predictions of biological responses to climate change.
... Previous studies have indicated that the thermal preference and thermal physiology of an organism would show a co-adaptation, keeping the preferred and optimal temperatures very close (Huey and Bennett 1987;Gilchrist 1995;Angilletta et al. 2006;Schwerdt et al. 2020a). However, we found that the optimal temperature observed in M. thorelli (close to 26 • C) was significantly higher than the preferred temperature (close to 15 • C). ...
Article
Thermal preference and thermal performance are used to describe the thermal biology of an ectothermic organism through parameters, i.e., estimating locomotor performance by maximum running speed. In this study, we assessed the thermal preference and locomotor performance of the spider Mecicobothrium thorelli Holmberg, 1882, a wintry mygalomorph spider endemic to the native mountainous grasslands of central Argentina and Uruguay. The preferred temperatures of the 72.4% of the individuals were in the range of 10-20°C. The highest frequencies of preferred temperatures were 10-15 °C in males and 15-20°C in females. The sprint speed showed significant differences between all the temperatures evaluated and showed the highest speeds at 25 °C and the lowest at 3 °C. The optimal temperature was 26.09 °C which was significantly higher than the preferred temperature in both males and females. We concluded that M. thorelli selects a wide range of temperatures and prefers to stay in medium and low temperatures, which are correlated with winter activity in the wild. However, the species showed maximum speed at higher temperatures which implies that spiders would perform even better in nature and maximize their locomotion by being active during a warmer period.
Article
Temperature is an important environmental factor that affects how organisms allocate metabolic resources to physiological processes. Laboratory experiments that determine absolute thermal limits for representative species are important for understanding how fishes are affected by climate change. Critical Thermal Methodology (CTM) and Chronic Lethal Methodology (CLM) experiments were utilized to construct a complete thermal tolerance polygon for the South American fish species, Mottled catfish (Corydoras paleatus). Mottled catfish showed Chronic Lethal Maxima (CLMax) of 34.9 ± 0.52 °C and Chronic Lethal Minima (CLMin) of 3.8 ± 0.08 °C. Fish were chronically acclimated (∼2 weeks) to 6 temperatures ranging from 7.2 ± 0.05 °C →32.2 ± 0.16 °C (7 °C, 12 °C, 17 °C, 22 °C, 27 °C, and 32 °C), and CTM used to estimate upper and lower acute temperature tolerance. Linear regressions of Critical Thermal Maxima (CTMax) and Minima (CTMin) data with each acclimation temperature were used along with CLMax and CLMin to create a complete thermal tolerance polygon. The highest CTMax was 38.4 ± 0.60 °C for fish acclimated to 32.2 ± 0.16 °C, and the lowest CTMin was 3.36 ± 1.84 °C for fish acclimated to 7.2 ± 0.05 °C. Mottled catfish have a polygon measuring 785.7°C2, and the slope of the linear regressions showed the species gained 0.55 °C and 0.32 °C of upper and lower tolerance per degree of acclimation temperature, respectively. We compared slopes of CTMax or CTMin regression lines to each other using a set of comparisons between 3, 4, 5, or 6 acclimation temperatures. Our data demonstrated that 3 acclimation temperatures were as sufficient as 4 → 6 to pair with estimates of chronic upper and lower thermal limits for accurately determining a complete thermal tolerance polygon. Construction of this species' complete thermal tolerance polygon provides a template for other researchers. The following is sufficient to generate a complete thermal tolerance polygon: Three chronic acclimation temperatures that are spread somewhat evenly across a species' thermal range, include an estimation of CLMax and CLMin, and are followed by CTMax and CTMin measurements.
Article
Full-text available
Abstrak. Ikan badut (A. ocellaris) adalah salah satu ikan model yang digunakan sebagai objek penelitian perubahan lingkungan. Penelitian ini bertujuan untuk mengetahui dampak perubahan iklim khususnya peningkatan suhu terhadap pertumbuhan dan kelangsungan hidup ikan badut (A. ocellaris). Penelitian dilakukan di dalam laboratorium menggunakan rancangan acak lengkap. Perlakuannya adalah suhu 28OC (control), 29OC, 30OC dan 31OC, untuk mengamati pertumbuhan panjang dan bobot, laju pertumbuhan relatif, kelangsungan hidup dan perubahan warna ikan badut selama 8 minggu pemeliharaan. Pertumbuhan panjang, bobot, laju pertumbuhan spesifik dan kelangsungan hidup terbaik adalah perlakuan suhu 29OC (P<0,05) dengan nilai masing-masing 1,81 ± 0,09 cm, 1,06 ± 0,15g, 0,78 ± 0,09% dan 100±00%. Pertumbuhan dan kelangsungan hidup pada suhu 31OC adalah 1,02 ± 0,04cm, 0,18 ± 0,16 g, 0,30 ± 0,04% dan 66,67 ± 5,44%. Ikan badut mengalami perubahan warna merah, hijau dan biru pada perlakuan suhu 31OC (<,0,05). Peningkatan suhu dapat memicu tingkat mortalitas dan menghambat pertumbuhan ikan badut. Abstract. One of the model fish used in climate change research is clown fish (A. ocellaris). The purpose of this research is to determine the effect of temperature increases on the growth and survival of clown fish. The experiment was carried out in a laboratory with a completely randomized design. The treatments were 28°C, 29°C, 30°C, and 31°C to observe clown fish growth in length and weight, relative growth rate, increased life, and color change over an 8-week period. The treatment temperature resulted in the best growth in length, weight, specific growth rate, and survival. The treatment temperature of 29OC (P0.05) resulted in the best growth in length, weight, specific growth rate, and survival, with values of 1.81 0.09 cm, 1.06 0.15g, 0.78 0. 09%, and 100%, respectively. At 31°C, growth and life were 1.02 0.04cm, 0.18 0.16 g, 0.30 0.04%, and 66.67 5.44%, respectively. At 31 degrees Celsius, the color of Badut fish changed from red to green to blue (P, 0.05). Temperature can cause badut fish mortality and adversely affect their growth. I. PENDAHULUAN Habitat ikan badut tersebar luas di ekosistem terumbu karang tropis dan subtropis. Ikan badut termasuk ke dalam famili pomacentridae terdiri dari 31 spesies dengan dua genus yaitu Ampiprion dan Premnas (Colleye dan Iwata, 2016). Salah satu spesies yang sangat menarik dan memiliki pangsa pasar internasional yang cukup tinggi sebagai ikan hias adalah ikn badut (A. ocellaris) karena keindahan warnanya, ukruran yang kecil, aktivitas pergerakan dan kemudahan pemeliharaan di akuarium. Ikan ini memiliki variasi warna yang unik dengan kombinasi warna merah dan putih, pergerakan yang lincah dan sangat menghibur sehingga dikenal juga dengan nama ikan badut. Ikan badut hidup berasosiasi dengan abadutn yang tersebar di terumbu karang indo pasifik. Namun, karena kerusakan habitat dan penangkapan ikan yang berlebihan, populasi ikan badut mengalami penurunan populasi yang sangat drastis. Selain akibat penangkapan berlebihan, populasi ikan badut juga dihadapkan pada kondisi alam yang telah mengalami dampak negatif akibat perubahan iklim. Berdasarkan proyeksi iklim yang dilakukan oleh IPCC (2014) menyatakan bahwa suhu global akan mengalami peningkatan 4 o C pada akhir abad ini. Perubahan ini akan memberi dampak pada organisme dan
Article
Full-text available
Nocturnal geckos are active with body temperatures (Tb) that are low and variable relative to those of diurnal lizards. If the physiology of geckos is evolutionarily adapted to these low and variable Tb's, then the physiology of geckos should function best at relatively low and variable temperatures. In fact, optimal temperatures and performance breadths for sprinting of several geckos (Coleonyx brevis, C. variegatus, Hemidactylus frenatus, H. turcicus, Lepidodactylus lugubris) do not differ substantially from those of diurnal lizards from other families. As a result geckos normally forage at night at Tb's that should be suboptimal for sprinting. Potential evolutionary explanations (eg evolutionary inertia of thermal physiology, possible selection pressures favouring high optimal temperatures) for the similarity of the thermal dependence of sprinting of geckos and diurnal lizards are evaluated. -from Authors
Article
Full-text available
The tegus increase in body mass after hatching until early autumn, when the energy intake becomes gradually reduced. Resting rates of oxygen consumption in winter drop to 20% of the values in the active season ( V̇ O2=0.0636 ml g-1 h-1) and are nearly temperature insensitive over the range of 17-25°C (Q10=1.55). During dormancy, plasma glucose levels are 60% lower than those in active animals, while total protein, total lipids and β-hydroxybutyrate are elevated by 24%, 43% and 113%, respectively. In addition, a significant depletion of liver carbohydrate (50%) and of fat deposited in the visceral fat bodies (24%) and in the tail (25%) and a slight loss of skeletal muscle protein (14%) were measured halfway through the inactive period. Otherwise, glycogen content is increased 4-fold in the brain and 2.3-fold in the heart of dormant lizards, declining by the onset of arousal. During early arousal, the young tegus are still anorexic, although V̇ O2 is significantly greater than winter rates. The fat deposits analysed are further reduced (62% and 45%, respectively) and there is a large decrease in tail muscle protein (50%) together with a significant increase in glycogen (2-3-fold) and an increase in plasma glucose (40%), which suggests a role for gluconeogenesis as a supplementary energy source in arousing animals. No change is detectable in citrate synthase activity, but β-hydroxyacyl CoA dehydrogenase activities are strongly affected by season, reaching a 3-fold and 5-fold increase in the liver tissue of winter and arousing animals, respectively, and becoming reduced by half in skeletal muscle and heart of winter animals compared with late fall or spring active individuals. From hatching to late autumn, the increase of the fat body mass relatively to body mass is disproportionate ( b =1.44), and the mass exponent changes significantly to close to 1.0 during the fasting period. The concomitant shift in the V̇ O2 mass exponent in early autumn ( b =0.75) to values significantly greater than 1.0 in late autumn and during winter dormancy indicates an allometric effect on the degree of metabolic depression related to the size of the fat stores and suggests greater energy conservation in the smaller young.
Article
Argues that the traditional view of latitudinal and seasonal acclimation of metabolic rate to temperature has no useful biological meaning, ignoring the heterogeneous nature of respiratory demand and imposing upon the system an apparent homogeneity that does not in fact exist. The author discusses temperature compensation, which is the maintenance of physiological rate in the face of a change in temperature, then outlines the process involved in respiratory demand in aquatic animals. Temperature is shown to have a major role in influencing seasonal and latitudinal patterns in metabolic rate, but this influence is subtle and difficult to distinguish from other more powerful influences. We should reformulate the energy budget to make explicit the heterogeneous nature of respiratory demand and avoid setting production and respiration as competing sinks. -P.J.Jarvis
Article
The aims of this paper are to compare the thermal ecology of four species of varanid lizards that occupy a range of habitats and climatic regions, and to assess the efficacy of methods for evaluating the extent to which ectothermic animals exploit their thermal environments. Hertz et al. (1993) have proposed several indices of thermoregulation, and these are evaluated with respect to our data from varanid lizards. The thermoregulatory characteristics of three tropical monitor lizards (Varanus panoptes, V. gouldii, and the semiaquatic V. mertensi), and the temperate-zone V. rosenbergi were studied throughout the year. Radiotelemetry was used to measure the body temperatures (Tb's) of free-ranging animals, and microclimatic data were collected to determine the range of possible Tb's that an animal could achieve. Operative temperatures (Tb's) were estimated by biophysical models for each set of animal characteristics and microclimatic conditions. The Tb's selected by animals in a laboratory thermal gradient were used to determine the set-point range of Tb's that the animals voluntarily select. Plots that superimpose Tb's, Te's, and the set-point range across the day are extremely useful for describing the thermoregulatory characteristics of ectotherms. These plots can be used to determine the extent to which the animals exploit their thermal environment: we define an index of thermal exploitation (Ex) as the time in which Tb's are within the set-point range, divided by the time available for the animal to have its Tb's within the set-point range. Only V. mertensi was active throughout the year. In general, during seasons of inactivity, the Tb's of inactive species fell outside the set-point range, but during periods of activity all species selected Tb's within their set-point range. The temperate-zone species (V. rosenbergi) thermoregulates very carefully during periods when environmental conditions allow the animals to attain the set-point range, and V. gouldii also thermoregulates carefully in the wet season. V. mertensi selects Tb's that are significantly lower than the other species both in the field and in the laboratory, and thermoregulatory indices of this species were intermediate relative to the other species. The amount of time spent in locomotion each day was not correlated with the indices of thermoregulation: the most active species, V. panoptes, was, with respect to several indices, the least careful thermoregulator. The type of question that is being addressed, with respect to the interactions between an animal's thermal environment and its thermoregulatory behavior, determines the appropriateness of the various indices of thermoregulation. The Ex index describes the thermoregulatory characteristics of ecotherms in a heterogeneous thermal environment, and in such an environment a large amount of information can easily be interpreted graphically. This index is less useful in a thermally homogeneous environment.
Article
Thermal constraint on energy assimilation is an important source of life history variation in geographically widespread ectotherms such as the eastern fence lizard (Sceloporus undulatus). Fence lizards in southern populations grow faster and produce more offspring per year than do those in northern populations. Biophysical models indicate that this difference in production is the result of thermal constraints on energy assimilation, but they do not exclude intraspecific variation in behavior or physiology. I quantified both thermoregulatory behavior and the thermal sensitivity of metabolizable energy intake (MEI) in lizards from New Jersey (NJ) and South Carolina (SC) populations of Sceloporus undulatus. In the laboratory, I conducted feeding trials to estimate MEI at body temperatures experienced by field-active lizards (20⚬, 30⚬, 33⚬, and 36⚬C). I also measured preferred body temperature (Tp) of lizards in a thermal gradient. In the field, I estimated the accuracy of thermoregulation by lizards. Both NJ and SC lizards exhibited a maximal MEI at their Tp (33⚬C), but lizards from SC had a significantly higher MEI at this temperature than lizards from NJ. Although lizards in both populations thermoregulated within 2⚬C of Tp, lizards in SC could maintain Tp for a longer duration on a daily and annual basis. Therefore, lizards in SC could assimilate more energy because they had a higher maximal MEI during activity and were active for longer durations than lizards in NJ. Geographic variation in the life history of S. undulatus may be caused by differentiation of physiology between populations, as well as by differences in the thermal environments of populations.