ArticlePDF Available

Adaptive radiation: the interaction of ecological opportunity, adaptation, and speciation

Authors:

Figures

Content may be subject to copyright.
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated
in any form without express written permission from the publisher.
Chapter 15
Adaptive Radiation: The Interaction of
Ecological Opportunity, Adaptation,
and Speciation
Jonathan B. Losos and D. Luke Mahler
Darwin may have been the first to describe adaptive radiation when, con-
templating the variety of finches that now bear his name, he remarked:
“Seeing this gradation and diversity of structure in one small, intimately
related group of birds, one might really fancy that from an original paucity
of birds in this archipelago, one species has been taken and modified for
different ends” (Darwin 1845: 380). Since Darwin’s time, naturalists and
evolutionary biologists have been fascinated by the extraordinary diversity
of ecology, morphology, behavior, and species richness of some clades, but
interest in adaptive radiation has surged in recent years, largely as a result
of three developments.
First, evolutionary ecologists have focused on the mechanisms that pro-
duce adaptive radiation. Careful study of ecological divergence within and
among populations has yielded a wealth of information about how natu-
ral selection leads to evolutionary diversification (Schluter 2003; Nosil and
Crespi 2006; Grant and Grant 2008a) ( Figure 15.1). Schluter’s (2000) seminal
work, The Ecology of Adaptive Radiation was a watershed in the field, build-
ing upon and extending in important ways Simpson’s (1953) The Major
Features of Evolution published a half-century earlier.
Second, the explosion of molecular phylogenetics in the last two decades
(see Hillis, Chapter 16) has revealed the diversification histories of count-
less clades and has provided the raw material for a renaissance of adaptive
radiation studies. Molecular research has offered surprising discoveries
about the history and magnitude of many adaptive radiations, such as the
vangids of Madagascar (Yamagishi et al. 2001) (Figure 15.2), the corvoids
of Australia (Sibley and Ahlquist 1990; Barker et al. 2004), the cichlids of
Lake Victoria (Meyer et al. 1990; Seehausen 2006), the lobeliads of Hawaii
(Givnish et al. 2009) and a plethora of others. In each of these cases, the great
ecological and morphological diversity of a group had been thought to be
the result of independent colonization events from multiple, differently
adapted ancestral lineages. Instead, new molecular phylogenies revealed
382  Chapter 15 Losos • Mahler
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated in
any form without express written permission from the publisher.
Adaptive Radiation: The Interaction of Ecological Opportunity, Adaptation, and Speciation  383
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated
in any form without express written permission from the publisher.
that the great diversity in these groups is the result of in situ evolution, that
is, adaptive radiation (see references previously cited).
Not only have many previously unknown adaptive radiations been dis-
covered with molecular data, but also time-calibrated phylogenetic trees
of extant taxa allow classic hypotheses of adaptive radiation to be tested in
new ways (Glor 2010). Previously, the primary way to estimate the tempo
of macroevolution was to measure it directly using high-quality paleon-
tological data (see Wagner, Chapter 17), which are available for relatively
Darwin Bell
Sinauer Associates
Morales Studio
Figure 15.01 Date 04-19-10
75 9 11 13 15 17
10
5
15
20
25
(B)
Growth rate in open water (mg/d)
Growth rate in littoral zone (mg/d)
Benthic
Hybrid
Limnetic
–1.5 –1.0 –0.5 0.0 0.5 1.0 1.5
3.8
3.9
4.0
3.7
(C)
Morphological index
Logarithm of growth rate (mm per 90 days)
(A)
Figure 15.1  Natural Selection and Adaptation in British Columbian
Threespine Sticklebacks (Gasterosteus aculeatus) (A) A number of lakes in
British Columbia contain two species of sticklebacks, one that is slender-bodied,
with a small mouth, eats zooplankton, and occupies open water (limnetic,
above); the other is larger, deeper-bodied, has a large mouth, and eats inverte-
brates near the lake bottom (benthic, below). (B) Each species has higher growth
rates in its own habitat, and hybrids are inferior in both habitats. (C) Natural
selection (using growth rate as a proxy) on a highly variable hybrid population
favors more benthic-like individuals in the presence of the limnetic species (red
dots and solid line), but not in their absence (green dots and dashed line). Such
studies have shed light on the role of ecology and natural selection in driving
divergence during adaptive radiation. (A, from Rundle and Schluter 2004, photos
© Ernie Cooper; B, adapted from Schluter 1995; C, adapted from Schluter 1994.)
Figure15.2  Malagasy Vangids Molecular phylogenetic study indicates that the
vangids represent a monophyletic group, rather than being members of several dif-
ferent families with closest relatives elsewhere. (From Yamagishi et al. 2001.)
384  Chapter 15 Losos • Mahler
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated in
any form without express written permission from the publisher.
Adaptive Radiation: The Interaction of Ecological Opportunity, Adaptation, and Speciation  385
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated
in any form without express written permission from the publisher.
few lineages. However, nucleotide sequence data now provide an alter-
native means of reconstructing phylogenetic relationships and the timing
of lineage divergence events. With time-calibrated phylogenies, one may
ask questions such as whether the occurrence of adaptive radiation is cor-
related with historical events (e.g., mass extinctions, changes in climate)
or whether the pace of diversification decreases through time, as often is
expected of an adaptive radiation (Box 15.1).
Third, experimental studies of microbial evolution have added a new
dimension to the study of adaptive radiation (see Dykhuizen, Commentary
2). Such studies bring the benefits of experimental control, large sample
sizes and replication, and the ability to not only track lineages through the
diversification process, but also to freeze ancestral taxa and subsequently
resurrect them to interact with their descendants (Lenski and Travisano
1994). Although thus far primarily based on studies of short-lived asexually
reproducing organisms diversifying under simplified ecological conditions
(a situation that is changing), microbial evolution studies have permitted
experimental tests of many of the basic hypotheses of adaptive radiation,
allowing microevolutionary processes to be directly and experimentally
connected to macroevolutionary outcomes and often confirming predic-
tions of the adaptive radiation model (Kassen 2009) (Figure 15.3).
With time-calibrated phylogenies, researchers
can investigate whether lineage diversification
patterns match patterns expected from the
ecological process of adaptive radiation. These
approaches focus on how the pace of lineage
or phenotypic diversification alters over time or
with changing ecological conditions, such as eco-
logical opportunity. The most common approach
to testing for the signature of adaptive radiation
is to construct models in which parameters de-
scribe changes in the tempo of diversification as
a result of ecological conditions. Alternative eco-
logical models may be compared to each other,
to non-ecological models, or to a null model in
which diversification proceeds at a constant rate.
Tests for the signature of adaptive radiation
have a rich pedigree in quantitative paleontol-
ogy, in which numerous studies have tracked the
rise and fall of diversity and disparity in relation
to mass extinction events, the evolution of key
innovations, and colonization of new regions
(Simpson 1953; Sepkoski 1978; Foote 1997,
1999; reviewed in Erwin 2007; see Foote, Chap-
ter 18) (Figure 1A,B).
Box 15.1
TESTING FOR CHARACTERISTIC PATTERNS OF ADAPTIVE RADIATION
Darwin Bell
Sinauer Associates
Morales Studio
Figure 15.03 Date 04-19-10
It is very difficult to
see what is happening
in the tubes.
(A) Heterogeneous environment
(B) Homogeneous environment
Figure 15.3  Experiments on the
Evolutionary Diversification of the Bacterium
Pseudomonas fluorescens Resources were
distributed either heterogeneously (solu-
tion with red and green strata in A) or
homogeneously (uniform green solution in
B) distributed. After several days, bacteria
in the heterogeneous environments repeat-
edly diverged into the same three morpho-
types, which interact negatively and differ
in resource use (the yellow shading on the
right depicts typical resource use for each
morphotype; the different forms tend to
thrive as a surface film, in solution, or along
the substrate, respectively). By contrast, only
one morphotype occurred in the homog-
enous treatment. (Adapted from Rainey and
Travisano 1998.)
Figure 1  The “Early Burst” Pattern of Adap-
tive Radiation In the fossil record, an early
burst may be detected by tracking the number
of species and the total disparity of a radiation
over geological time. In trilobites, species rich-
ness (A) and disparity (B) both peaked early in
the history of the clade. Phylogenetic compara-
tive methods may be used to detect an early
burst from data on extant taxa. Greater Antil-
lean Anolis lizards exhibit both a rapid early
accumulation of lineages, depicted as a concave
lineages-through-time plot in (C), as well as an
early burst in phenotypic evolution, evident as
larger independent contrasts early in the radia-
tion (D). (A,B adapted from Foote 1993; C,D
adapted from Mahler et al. 2010.)
Darwin Bell
Sinauer Associates
Morales Studio
Figure Box 15.01 Date 04-19-10
0 0.000
497560 434 371 308 245
258
515
773
1031
1288
Geologic time (Ma)
Number of species
(A)
497560 434 371 308 245
0.014
0.028
0.043
0.057
0.071
1
0
2
3
4
5
6
7
8
9
10
–100 –80 –60 –40 –20 0
Relative time before present
Limb length contrasts
(absolute value)
(D)
Geologic time (Ma)
Disparity
(B)
–100 –80 –60 –40 –20 0
2
1
5
10
20
50
100
Relative time before present
Log (lineages)
(C)
(continued)
386  Chapter 15 Losos • Mahler
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated in
any form without express written permission from the publisher.
Adaptive Radiation: The Interaction of Ecological Opportunity, Adaptation, and Speciation  387
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated
in any form without express written permission from the publisher.
In this chapter, we review what is known about the mechanisms that
drive adaptive radiation and, more importantly, highlight those areas re-
quiring further research. Along the way, we will discuss what constitutes
an adaptive radiation and how one can be identified.
Evolutionary Radiation: What Are the Types, How Are
They Recognized, and Are They Special?
Adaptive radiations draw the attention of scientists and non-scientists alike
because their grandeur seems to imply that something special is respon-
sible: these groups are extraordinary and require explanation, invocation
of some special attribute, either intrinsic or external, that can explain why
these particular clades have diversified to such an extreme extent (Box
15.2). But, adaptive radiation describes only one part of the spectrum of
evolutionary radiations. Although distinguishing among types of evolu-
tionary radiation may involve making arbitrary distinctions (Olson and
Arroyo-Santos 2009), doing so provides a useful framework for further
study of the key features of radiations.
The most commonly used phylogenetic test
evaluates the early burst model, in which early
diversification is explosive as lineages rapidly
adapt to new ecological roles but slows as op-
portunities disappear due to ecological satura-
tion (Nee et al. 1992; Rabosky and Lovette 2008;
Phillimore and Price 2009; reviewed in Glor
2010). The majority of tests of the “early burst”
model have investigated patterns of lineage
diversification, but similar models have recently
been developed for phenotypic traits as well
(Freckleton and Harvey 2006; Agrawal et al. 2009;
Harmon et al. 2010, in press; Mahler et al. 2010,
in revision) (Figure 1C,D). Support for this model
varies—the common finding of temporally de-
clining lineage diversification has led some to
proclaim a strong role for ecological opportunity
in regulating cladogenesis, although the sample
of clades studied may be biased (McPeek 2008;
Phillimore and Price 2008; Ricklefs 2009). Among
phenotypic studies of the early burst model
(which, to date, are far fewer than studies focus-
ing on species richness), some studies report
declining rates of evolution with diminishing
ecological opportunity (Freckleton and Harvey
2006; Agrawal et al. 2009; Mahler et al. 2010, in
revision), whereas Harmon et al. (2010) find only
limited support for the early burst model in a
wide survey of animal radiations.
A key feature of the model-fitting approach is
the flexibility provided by the variety of models
that may be compared when evaluating adap-
tive radiation hypotheses. In addition to the ear-
ly burst model, a diversity of alternative models
are available for fitting data, including radiation
under flexible ecological limits, radiation with
high lineage turnover, and radiation in stages,
among others (Price 1997; Harvey and Rambaut
2000; McPeek 2008; Benton 2009; Gavrilets and
Vose 2009; Rabosky 2009). The ability to identify
patterns of adaptive radiation using phylogenet-
ic methods is likely to improve in coming years
as new and more refined models are developed.
Box 15.1  Continued
We suggest that the term adaptive radiation
should be reserved for those clades exhibiting
exceptional ecological and phenotypic disparity
(see Figure 15.4). The rationale for this argument
is that it is the unusually great degree of dispar-
ity in these clades that requires explanation—by
identifying such clades, researchers can focus on
them to understand what has triggered their ex-
traordinary evolutionary diversification. Implicit
in this approach is the need to develop statistical
methods to separate those clades that constitute
adaptive radiations from those that do not.
This approach can be criticized on two
counts. First, it creates an arbitrary dichotomy
in what is most likely a continuous distribution
(Olson and Arroyo-Santos 2009). That is, the de-
gree of adaptive disparity of clades is surely con-
tinuously distributed. How can one draw a line
and say that all clades with a greater amount of
disparity than the threshold constitute adaptive
radiations and that all with even a slightly lesser
amount, are not?
An alternative approach to the threshold-
based approach would be to quantify disparity
for a sample of clades and investigate whether
the degree of disparity is statistically related to
a factor, such as degree of ecological opportu-
nity (presuming it could be quantified), that has
been hypothesized to drive adaptive radiation.
Increasingly, tools for investigating the relation-
ship between such factors and patterns of eco-
logical diversification are being developed and
employed (Olson and Arroyo-Santos 2009; Glor
2010).
The second criticism of this definition of
adaptive radiation is based on the view that
adaptive radiation is a process as well as an out-
come (just as one might argue that adaptation is
both a process and an end-result). This argument
suggests that the same processes are involved in
adaptive diversification whether the result is an
enormous adaptive radiation, such as African Rift
Lake cichlids, or a small clade of species slightly
morphologically differentiated to adapt to minor
differences in habitat use. Because both cases
are the result of the same process of cladogen-
esis plus adaptive divergence driven by natural
selection, this view would suggest that all such
clades, no matter how disparate, should be con-
sidered adaptive radiations.
Such a view would render the term adap-
tive radiation meaningless. The vast majority
of clades are composed of species that exhibit
at least a small degree of phenotypic disparity
that has arisen as a result of adaptive divergence
driven by natural selection. Assuming this is true,
almost all clades would be adaptive radiations,
and the term would have little utility in identify-
ing clades of special interest. We feel that this
not only neuters the term “adaptive radiation,
but also departs from its use throughout the his-
tory of evolutionary biology.
One way out of this problem might be to
restrict the term adaptive radiation to clades in
which adaptive diversity arose in a burst of evo-
lution early in a clade’s history, with subsequent
deceleration in the rate of evolution. Indeed,
many definitions of adaptive radiation include
the proviso that radiation must occur quickly
(Givnish 1997). This view, however, also has prob-
lems. Either the definition must include clades
that achieve only modest disparity as long as
they accumulated it early, or it must rely on an
arbitrary disparity threshold.
From our perspective and that of many others
(Givnish 1997), the important aspect of adaptive
radiation is the disparity produced by the clade
(Foote 1997; Erwin 2007), rather than the pace at
which it accumulates. In some models of adap-
tive radiation (Harvey and Rambaut 2000), an
adaptive radiation may unfold at a rather steady
pace; indeed, ecological opportunity may ap-
Box 15.2
ARE ADAPTIVE RADIATIONS EXCEPTIONAL?
(continued)
388  Chapter 15 Losos • Mahler
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated in
any form without express written permission from the publisher.
Adaptive Radiation: The Interaction of Ecological Opportunity, Adaptation, and Speciation  389
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated
in any form without express written permission from the publisher.
Evolutionary radiation results in the production of two components of di-
versity—species richness and phenotypic diversity (often termed “disparity”
to avoid confusion with “species diversity”). Adaptive radiation is a type of
evolutionary radiation, emphasizing the extent of phenotypic differentiation
among members of a clade as species adapt to use different ecological re-
sources; we henceforth refer to this as “adaptive disparity.” Although a wide
variety of definitions of adaptive radiation have been proposed (Givnish
1997), Futuyma’s (1998) definition seems to capture the sense of most of
these: “evolutionary divergence of members of a single phylogenetic lineage
into a variety of different adaptive forms.” More specifically, we propose
that the term adaptive radiation should refer to those clades that exhibit an
exceptional extent of adaptive disparity. Definitional issues and methods for
identifying adaptive radiations are reviewed in Box 15.3.1
Much of the literature on evolutionary radiations equates species rich-
ness with adaptive radiation. Indeed, many of the world’s most remark-
able and celebrated radiations are rich in both species and adaptive form,
including the African Rift Lake cichlids, Hawaiian Drosophila, Caribbean
Anolis lizards and, at a higher level, beetles and angiosperms. Such lineages
undoubtedly represent adaptive radiations and suggest that species prolif-
eration and ecological radiation occur hand in hand. Although often true,
this need not be the case.
Some clades are exceptionally diverse phenotypically and constitute
adaptive radiations, despite having unexceptional species diversity (Fig-
ure 15.4). For example, the lizard clades Pygopodidae and Cordylidae both
contain great ecological and morphological disparity despite being spe-
cies poor (Webb and Shine 1994; Branch 1998). Groups such as these are
commonly neglected in studies of adaptive radiation, but deserve more
1 Integral to the concept of adaptive radiation is the concept of adaptation
itself. Evolutionary divergence can occur for reasons other than adaptive
differentiation. Thus, investigation of the adaptive basis of trait differentiation
is essential in any study of adaptive radiation. Arnold (1994), Larson and
Losos (1996), and McPeek, Commentary 3 (in this volume) provide reviews of
adaptation and how it can be studied. For the purposes of discussion here, we
will assume that phenotypic differences among species are adaptively based.
attention, particularly in studies of the role of speciation in adaptive radia-
tion (see subsequent discussion).
Conversely, other clades contain a great number of species, but little eco-
logical or phenotypic disparity. For example, the eastern North American
slimy salamanders (Plethodon) are a species-rich clade that diversified rap-
idly early in their evolutionary history. However, these salamander species
are distributed almost entirely allopatrically and are ecologically similar
(Kozak et al. 2006). Such clades have been termed non-adaptive radiations
(Gittenberger 1991; Kozak et al. 2006; Rundell and Price 2009). Great species
richness without substantial phenotypic disparity could arise if a clade was
predisposed to speciate in ways that do not involve adaptive divergence,
as might result from allopatric isolation in similar environments or from
the operation of sexual selection occurring in divergent ways in different
populations.
pear suddenly in some cases (e.g., colonization
of an unoccupied island) but more gradually in
others (e.g., co-radiation with another clade). We
prefer to reserve the question of timing of di-
versification as a hypothesis to be tested among
adaptive radiations, rather than a criterion for
deciding whether a clade constitutes an adap-
tive radiation or not. At the least, if one takes the
more restrictive definition that includes timing,
then another term is needed for those clades
that produce exceptional disparity but in a non-
explosive way, perhaps such as Simpson’s (1953)
mostly forgotten “progressive occupation of
adaptive zones.
Box 15.2  Continued
Species richness
Frequency
Ecological
disparity
Cordylus
lizards
Welwitschia
Plethodon
salamanders
African
cichlids
Figure 15.4  The Axes of Evolutionary Radiation Clades can be diverse in
both species richness and ecological variety, termed disparity. Clades that have
exceptional ecological disparity are adaptive radiations, whether they have great
(e.g., African lake cichlids) or little species richness (e.g., cordylid lizards). Non-
adaptive radiations are those that are exceptional in species richness, but not in
ecological disparity, such as plethodontid salamanders. Some clades are excep-
tionally non-disparate and species-poor, such as Welwitschia.
390  Chapter 15 Losos • Mahler
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated in
any form without express written permission from the publisher.
Adaptive Radiation: The Interaction of Ecological Opportunity, Adaptation, and Speciation  391
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated
in any form without express written permission from the publisher.
If exceptional diversity is a key feature of adaptive
radiation, statistical methods are needed that can
test whether a radiation is exceptional. Such a
test must demonstrate that a clade contains more
disparity than would be expected under a neutral
model of diversification, or that it is exceptionally
disparate compared to other radiations.
In principle, such questions could be ad-
dressed either by examining the fossil record or
by using phylogenetic methods. Both of these
methods have advantages and disadvantages.
Paleontological methods directly track changes
in diversity over time. By contrast, with phylo-
genetic methods using extant taxa, macroevo-
lutionary changes must be inferred. As such, re-
sults arising from these methodologies are only
as sound as their assumptions, and inadequate
models for reconstructing evolutionary history
may lead to both Type I and Type II errors in test-
ing adaptive radiation hypotheses (Revell et al.
2005; see Foote, Chapter 18). Paleontological
studies incorporate extinct taxa, and although
phylogenetic methods exist for estimating the
influence of extinction (Nee et al. 1994; Kubo
and Iwasa 1995), they often suffer from low
statistical power (Bokma 2009; Quental and Mar-
shall 2009; Rabosky 2009). In contrast, the phy-
logenetic approach has several distinct advan-
tages. In particular, phylogenetic studies permit
investigation of long-term evolutionary patterns
in taxa that are also well-studied ecologically—
inferences about the ecology of fossil taxa can
be unreliable, particularly when the taxa do
not have comparable extant counterparts, and
morphological homoplasy may also undermine
accurate estimation of the relationships of fossil
taxa in paleontological studies. Moreover, the
quality of the fossil record varies greatly among
taxa; phylogenetic methods can be used even
for taxa with little or no fossil record. Of course,
the most reliable insights into the large-scale
pattern of adaptive radiation will be those that
are well supported by both paleontological and
phylogenetic investigations.
Statistical approaches to the identification
of exceptional clades have concentrated almost
entirely on patterns of lineage diversification
rather than phenotypic diversification. The first
lineage diversification models, in which specia-
tion occurred as a uniform stochastic process,
produced surprising results (Raup et al. 1973;
Slowinski and Guyer 1989). The resulting phy-
logenies tended to be topologically unbalanced,
suggesting that large differences in clade species
richness could occur by chance alone (Guyer and
Slowinski 1993; Barraclough and Nee 2001; Nee
2001, 2006). Nonetheless, many clades in the Tree
of Life are exceptionally species-rich (or species-
poor), even compared to the highly variable
neutral expectation. For instance, at a very broad
scale, Alfaro et al. (2009) identified nine such
clades among all vertebrates.
Of course, while such lineage diversification
models are useful for testing whether species
radiations are exceptional, they do not neces-
sarily identify adaptive radiations, which are
distinguished by ecological and morphological
disparity. Paleontologists have long used mea-
sures of disparity to trace patterns of phenotypic
evolution through time, and changes in disparity
in fossil lineages have been pivotal in testing
and refining adaptive radiation theory in recent
decades (Foote 1997; Roy and Foote 1997; Ciam-
paglio et al. 2001; Erwin 2001, 2007).
To date, few studies of extant taxa have at-
tempted to identify those clades that are excep-
tionally disparate using phylogenetic methods
(Losos and Miles 2002). Direct comparisons of
disparity among taxa are problematic, because
disparity is a function of the rate of evolution,
the age of a clade, and the phylogenetic topol-
ogy of the clade (O’Meara et al. 2006) (Figure 1).
One approach is to estimate the rate of pheno-
typic evolution in a phylogenetic context and to
compare rates among clades (Collar et al. 2005;
Harmon et al. 2008; Pinto et al. 2008).
Box 15.3
METHODS FOR IDENTIFYING EXCEPTIONAL CLADES
Box 15.3  Continued
Figure 1  The Effect of Phylogenetic Struc-
ture and Rate of Evolution on Expected
Disparity Under a Brownian motion model
of evolution, disparity among extant taxa is a
function of the rate of phenotypic evolution
as well as of the structure of the phylogenetic
tree relating those taxa. In (A), a phenotypic
trait evolved at a constant rate on phyloge-
netic trees that were identical in branching
order, but different in the relative timing of
divergence. In the first phylogeny, lineage
diversification was concentrated early, while
in the second phylogeny, lineage divergences
were concentrated late. Although the rate
of trait evolution was the same for each
phylogeny, greater disparity resulted when
divergences were concentrated earlier, be-
cause traits evolved independently for longer
periods of time. In (B), a trait evolved on
identical phylogenies, but the rate of evolu-
tion (σ2) differed by a factor of two between
the two phylogenies. In this case, when the
rate of evolution was higher, trait disparity
was greater. In both (A) and (B), disparity
differences are depicted graphically, using
minimum convex polygons for the evolved
trait values, and data in the upper panels are
colored to match the phylogenies and rates
under which they were generated.
Darwin Bell
Sinauer Associates
Morales Studio
Figure Box 15.03 Date 04-19-10
–5–6 –4 –3 –2 –1 0123
–2
–3
–1
0
1
2
3
4
5
6
Trait 1
Trait 2
(A)
–5 –2.5 0 2.5 5–7.5 7.5
–2
–1
0
1
2
3
–3
Trait 1
σ2 = 1.0σ2 = 1.0 σ2 = 0.5σ2 = 1.0
Trait 2
(B)
392  Chapter 15 Losos • Mahler
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated in
any form without express written permission from the publisher.
Adaptive Radiation: The Interaction of Ecological Opportunity, Adaptation, and Speciation  393
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated
in any form without express written permission from the publisher.
Ecological Opportunity and Adaptive Radiation
The idea that ecological opportunity is a necessary prerequisite for adap-
tive radiation dates to Simpson (1953) who argued that an ancestral species
must find itself in a setting in which “the [adaptive] zone must be occupied
by organisms for some reason competitively inferior to the entering group
or must be empty” (Simpson 1953: 207). More recently, Schluter (2000)
suggested that ecological opportunity is “loosely defined as a wealth of
evolutionarily accessible resources little used by competing taxa” (Schluter
2000: 69).
An ancestral species might find itself in the presence of ecological op-
portunity for a number of reasons:
1. colonization of isolated areas with a depauperate biota, such as
islands, lakes, or mountaintops;
2. arrival or evolution of a new type of resource;
3. occurrence in a post-mass extinction environment, again with a dep-
auperate biota;
4. evolution of a feature that provides the species with access to avail-
able resources that were previously unattainable; such a feature is
referred to as a key innovation2.
Many of the most famous examples of adaptive radiation occur on is-
lands, including such iconic radiations as Darwin’s finches, Hawaiian
honeycreepers and silverswords, and Anolis lizards. Lakes, the terrestrial
counterparts of islands, host additional renowned radiations, such as the
African Rift Lake cichlids and the many radiations in Lake Baikal. One
notable feature of many island radiations is that in the absence of many
types of organisms normally found in mainland settings, members of
the radiation have adapted in a wide variety of different ways, using re-
sources that on the mainland are utilized by taxa not present on the island
(Carlquist 1974; Leigh et al. 2007; Losos and Ricklefs 2009). The absence
of many types of predators on islands may also play a role, by allowing
organisms to use resources and habitats previously unavailable because
of predation threat (Carlquist 1974; Schluter 1988; Benkman 1991). As a
result, the ecological and phenotypic disparity of island radiations is of-
ten much greater than their mainland relatives (Carlquist 1974; Schluter
2 A key innovation is defined as a trait that allows a species to interact with the
environment in a fundamentally different way (Miller 1949; Liem 1974). For
a critical discussion of the concept of key innovations, see Cracraft (1990) and
Donoghue (2005). In recent years, the term has also been used to refer to a trait
that increases the rate of species diversification (Heard and Hauser 1995; Hodges
and Arnold 1995; Sanderson and Donoghue 1996; Ree 2005). This definition,
however, introduces a different concept; conflating the two under the same name
is confusing (Hunter 1998). A new term is needed for the latter phenomenon.
2000; Lovette et al. 2002; Figure 15.5). This oft-repeated phenomenon clearly
indicates the role of ecological opportunity in spurring adaptive radiation.
Species may also experience ecological opportunity without moving to
a new area if new resources appear, by immigration or evolution, where a
Darwin Bell
Sinauer Associates
Morales Studio
Figure 15.05 Date 04-19-10
Cardueline finches
Darwin’s finches
Hawaiian honeycreepers
Other passerines
–2 0 2 4
–2
0
2
4
Morphological trait axis I (bill width and depth)
Morphological trait axis II
(bill length)
(B)
(A)
Nectarivores
Generalists
Foragers
among leaves
Bark pickers
Seed and
fruit eaters
Figure 15.5  Adaptive Radiation of Island
Birds (A) Hawaiian honeycreepers exhibit stun-
ning morphological disparity in bill traits which
corresponds to diversity in resource use. (B) The
disparity of honeycreepers surpasses that of their
mainland relatives, the cardueline finches and
rivals that of all passerine birds. Another island
radiation, Darwin’s finches, also exhibits substan-
tial disparity. In both island radiations, species
occupy a diverse range of ecological niches uti-
lized by species in many different families on the
mainland. (A, reproduced with permission from
Pratt 2005; B, adapted from Lovette et al. 2002
and Losos and Ricklefs 2009.)
394  Chapter 15 Losos • Mahler
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated in
any form without express written permission from the publisher.
Adaptive Radiation: The Interaction of Ecological Opportunity, Adaptation, and Speciation  395
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated
in any form without express written permission from the publisher.
clade already occurs. For example, the radiation of horses is often attributed
to the spread of grasslands in the Miocene (MacFadden 1992). Similarly, the
plant genus Espeletia has radiated extensively in paramo habitats at high el-
evations in the northern Andes (Monasterio and Sarmiento 1991; Rauscher
2002; Hooghiemstra et al. 2006). The Andes are young, and presumably, the
ancestral Espeletia rode the rising mountain chain, adapting to new high-
elevation habitats as they appeared. Hughes and Eastwood (2006) infer a
similar history for Andean Lupinus.
Ecological opportunity also appears in the aftermath of mass extinc-
tions (Erwin 2001, 2007). Surviving taxa often radiate rapidly and relatively
quickly attain similar disparity to that seen before the extinction event. In
some cases, clades radiate into parts of morphological space that previously
were unoccupied by members of that clade (Foote 1999), whereas in other
cases a surviving subclade expands its disparity to encompass parts of mor-
phological space previously occupied by other subclades that perished in
the mass extinction (Foote 1996; see Foote, Chapter 18; Ciampaglio 2002;
McGowan 2004; Friedman 2010). The high rates of phenotypic evolution
exhibited after mass extinctions by these clades strongly support the role
of ecological opportunity in spurring radiation.
The evolution of a key innovation can allow access to previously un-
attainable resources, providing the stimulus for adaptive radiation. For
example, the evolution of wings in bats allowed them to prey upon a wide
range of flying insects probably unavailable to their earthbound ancestors.
Many other examples of key innovations have been proposed, although
such hypotheses are usually difficult to test because they represent unique
historical events. One way around this difficulty is to look for putative key
innovations that have evolved multiple times to see if they have repeat-
edly led to adaptive radiation (Mitter et al. 1988; de Queiroz 2002); several
examples, such as toe pads in lizards (Russell 1979; Larson and Losos 1996),
phytophagy in insects (Mitter et al. 1988; Farrell 1998), wings in vertebrates,
and pharyngeal jaws in fish (Stiassny and Jensen 1987; Mabuchi et al. 2007),
seem to pass this test.3 However, key innovations do not necessarily lead
to adaptive radiation; the evolution of such a trait may allow the species to
interact with the environment in new ways without leading to substantial
evolutionary diversification (Fürsich and Jablonski 1984; Levinton 1988;
de Queiroz 2002).4 For example, aardvarks have evolved a suite of skeletal
3 All of these examples are cases in which taxa with the putative key innovation
exhibit substantially greater disparity than their sister taxa, although this
observation has not been statistically tested. Instead, quantitative tests in these
papers have compared species richness, rather than disparity.
4 Similarly, it is possible that the evolution of the same key innovation could lead
to radiation in some clades and not others, depending on the context in which it
evolves. Consequently, failure to find a statistical relationship between evolution
of a putative key innovation and adaptation radiation does not indicate the
modifications permitting a termitophagous existence, but this specializa-
tion has not led to evolutionary diversification (Hunter 1998; see also Baum
and Larson 1991 on Aneides salamanders).
Although not usually discussed in these terms, the evolution of mu-
tualistic interactions may also function in a manner analogous to that of
a key innovation by providing access to resources that were previously
inaccessible (see Lane, Commentary 4). For example, the great diversity of
herbivorous insects and vertebrates would not be possible if it were not for
their mutualistic microbial gut inhabitants that allow them to digest cellu-
lose (Janson et al. 2008). Similarly, the great variety among modern corals
is likely in part the result of a mutualism between scleractinian corals and
their endosymbiotic zooxanthellae, which provides photosynthetic energy
in return for protection and nutrition (Stanley 1981).
The concept of ecological opportunity is straightforward and intuitive.
Given the number of examples in which ecological opportunity (achieved
in the four ways just enumerated) has led to adaptive radiation, there can
be little doubt that it is an important trigger to adaptive radiation. None-
theless, we may wonder whether ecological opportunity is a prerequisite
for adaptive radiation.
Can Adaptive Radiation Occur in the Absence of Pre-Existing
Ecological Opportunity?
Adaptive radiation could occur in the absence of pre-existing ecological op-
portunity either by members of a clade wresting resources away from other
taxa that had been using them or by creating their own opportunity.
COMPETITIVE REPLACEMENT OF ONE CLADE BY ANOTHER Much of the
older paleontological literature is peppered with proposals that one group
has radiated by outcompeting another, usurping its resources and forcing
it into evolutionary decline (e.g., the decline of mammal-like reptiles puta-
tively as a result of the purported superiority of dinosaurs in locomotion
or thermoregulation; Bakker 1968). However, very few of these cases hold
up to close scrutiny. Usually, when one group replaces another ecologi-
cally similar group, the explanation is that the first group went extinct, fol-
lowed by radiation of the second group (Rosenzweig and McCord 1991;
Benton 1996; Brusatte et al. 2008). In a few cases, the fossil record does
support the argument that one group has radiated at the expense of a clade
that previously utilized the same resources. Perhaps, the best example is
the interaction between cheilostome and cyclostome bryozoans, in which
cheilostomes have outcompeted cyclostomes in local interactions over the
course of many millions of years, all the while diversifying adaptively,
while cyclostomes decreased in diversity (Sepkoski et al. 2000).
evolution of the trait did not lead to adaptive radiation in any of the clades in
which it occurs (de Queiroz 2002).
396  Chapter 15 Losos • Mahler
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated in
any form without express written permission from the publisher.
Adaptive Radiation: The Interaction of Ecological Opportunity, Adaptation, and Speciation  397
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated
in any form without express written permission from the publisher.
This phenomenon of one group preventing evolutionary radiation of an-
other has been termed niche incumbency and is seen in experimental micro-
bial systems as well as in the fossil record (Brockhurst et al. 2007). It parallels
the ecological phenomenon of the priority effect, which occurs when one
species can prevent another from becoming established in a community by
dint of prior occupancy (MacArthur 1972; Chase 2007). Moreover, the obser-
vation that the introduced invasive species almost never causes the global
extinction of native species by outcompeting them5 lends support to the idea
that a radiating clade is unlikely to drive another clade to extinction through
competitive interactions (Simberloff 1981; Davis 2003; Sax et al. 2007).
CAN ADAPTIVE RADIATIONS CREATE THEIR OWN OPPORTUNITY? It would
seem, then, that ecological opportunity is usually necessary for adaptive
radiation and that in most well-studied cases, such opportunity exists prior
to adaptive radiation. However, a rarely considered alternative possibility
is that lineages create their own opportunity as they radiate. Such self-
propagation of an evolving radiation could occur in two ways.
First, as a clade radiates, the increased number of co-occurring clade
members may create opportunities for exploitation by other species. The
standard view among evolutionary biologists is that ecological opportunity
decreases through the course of a radiation as niches are filled. However,
an alternative view is that the more species that occur in a community,
the more opportunity there is for other species to take advantage of them
through predation, parasitism, mutualism, or other processes (Whittaker
1977; Tokeshi 1999; Erwin 2008). Most communities are composed of spe-
cies from many different clades, but in cases (usually on islands) in which
members of a single radiation are extremely diverse ecologically, the pos-
sibility exists for an evolutionary component to this view: as the clade radi-
ates, it may create additional opportunities, spurring further radiation, thus
creating further opportunities, and so on.
Implicit in this hypothesis is the view that interactions driving adaptive
radiation occur not only as competition for resources, but also between
species on different trophic levels. Indeed, adaptive radiations are known
in which some members of the radiation prey on other members. Among
African Rift Lake cichlids, for example, some species prey on others by eat-
ing their young, plucking their eyeballs, or rasping scales off their body.
Social parasites in hymenopterans are often closely related and are some-
times the sister taxa of the species they parasitize, which is referred to as
“Emery’s rule” (Bourke and Franks 1991; Savolainen and Vepsalainen
2003). These observations certainly indicate the possibility that adaptive
radiations create additional opportunity as they unfold. Although Schluter
5 The same is not true of introduced predators, which are responsible for scores of
species extinctions (Davis 2003; Sax et al. 2007).
(2000) suggested this idea a decade ago,6 no empirical work has addressed
this question; however, theoretical literature on the evolution of food webs
is beginning to develop (Ingram et al. 2009).
Radiations may also be self-propagating when two clades co-radiate.
Just as a clade may continually create its own ecological opportunity as it
radiates, two clades may reciprocally generate opportunity for each other
as they coevolve. Such coevolutionary radiations could take many forms.
For example, the radiation of one group, perhaps driven by interspecific
competition or predation, may create new ecological opportunities for a
second clade that utilizes members of the first as a resource. In turn, radia-
tion of the first group may continue if a species evolves some adaptation
that frees it from attack by the second group, allowing it to utilize resources
that were previously inaccessible. This last scenario is the basis of Ehrlich
and Raven’s (1964) famous escape and radiation theory of plant–herbi-
vore coevolution, and plays an important part in Vermeij’s (1987) theory
of “evolution and escalation” between predators and prey. In other cases,
co-radiations may simply result if species in one clade are each specialized
to use a single member of a second clade; in this case, adaptive radiation in
the host clade may be mirrored by radiation in their specialists. The recent
divergence of the parasitic wasp, Diachasma alloeum, in response to diver-
gence of its host, the apple maggot fly, Rhagoletis pomonella, illustrates how
such matched divergence may unfold (Forbes et al. 2009). A similar process
may have contributed to the extraordinary parallel radiations of figs and
fig wasps as well as of Glochidion trees and Epicephala moths (Herre et al.
1996; Weiblen and Bush 2002; Kato et al. 2003).
As with the role of mutualism in adaptive radiation, the processes driv-
ing coevolutionary radiations are the same as in other radiations, but the
synergistic interactions among co-radiating clades are distinctive. Such co-
evolutionary radiations may be particularly important in plant–herbivore
systems (Farrell and Mitter 1994, 1998; Roderick and Percy 2008; Winkler
and Mitter 2008; see Berenbaum and Schuler, Chapter 11).
Adaptive radiations may create their own opportunity in a second way:
members of a radiating clade may alter their environment, creating ecologi-
cal opportunities that did not previously exist. The concept of “ecosystem
engineering” refers to the role that organisms play in altering their physical
environment (Jones et al. 1994, 1997); ecosystem engineers, such as corals
or rainforest trees, change the physical environment in ways that allow the
existence of new species, potentially even leading to the evolution of eco-
logical types that otherwise would not exist (Erwin 2008). Although several
studies have documented that ecosystem engineering in one clade may
trigger adaptive radiation in another clade (e.g., corals and tetraodontiform
6 And Wilson in 1992 proposed: “what I like to call the test of a complete adaptive
radiation: the existence of a species specialized to feed on other members of its
own group, other products of the same adaptive radiation” (Wilson 1992: 118).
398  Chapter 15 Losos • Mahler
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated in
any form without express written permission from the publisher.
Adaptive Radiation: The Interaction of Ecological Opportunity, Adaptation, and Speciation  399
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated
in any form without express written permission from the publisher.
fishes; Alfaro et al. 2007), few examples from nature document members
of an adaptive radiation altering the environment in such a way to create
ecological opportunities that have then been utilized by other members
of their own radiation. One possible example is the evolution of lobeliad
plants in Hawaii. Among the first plant clades to arrive and radiate in
this archipelago, lobeliads diversified to fill ecological roles ranging from
canopy trees to shrubs, epiphytes, and vines—one-eighth of the Hawaiian
flora in total (Givnish et al. 2009). Given their early arrival and the extent
of their ecological diversification, it is plausible that some lobeliads (e.g.,
shade-dependent shrubs) radiated in microhabitats created by other mem-
bers of the radiation (e.g., woody trees), a hypothesis that could be tested
with further phylogenetic analyses.
Laboratory studies of microbial adaptive radiation also support a role for
ecosystem engineering driving diversification. Several studies have shown
that the waste product of an ancestral microbial species created a food
source subsequently used by a second type that evolved from its ancestor
(Kassen et al. 2009).
Does “Ecological Opportunity” Have More than Heuristic Value?
The conclusion of the preceding discussion is that ecological opportunity
is usually necessary for adaptive radiation, but whether we can predict a
priori if a clade will radiate is not clear for two reasons. First, ecological
opportunity, though usually necessary, may not be sufficient for radia-
tion. Second, ecological opportunity may be extremely difficult to define
objectively a priori and, as a result, the concept may have limited utility for
predicting when clades would be expected to radiate adaptively, as op-
posed to explaining retrospectively why they did so.
Many clades fail to radiate despite apparently abundant ecological op-
portunity. For example, the Galápagos and Hawaii are famous for their bird
radiations (finches and honeycreepers, respectively), but many other bird
lineages on these islands have failed to produce adaptive radiations, in-
cluding mockingbirds on the Galápagos (Arbogast et al. 2006) and thrushes
on Hawaii (Lovette et al. 2002). The same is true of many other island taxa.
Radiation may not occur in such circumstances for several reasons:
1. The perception of ecological opportunity may be mistaken.
2. Inability to access or utilize resources (e.g., from the lack of genetic
variation or phenotypic plasticity that could produce phenotypes
capable of taking advantage of novel available resources).
3. Lack of speciation: if, for some reason, speciation cannot occur (as dis-
cussed subsequently), then adaptive radiation cannot result.
4. Inability to diversify ecologically : even when speciation can occur,
if the resulting descendant species are unable to diverge phenotypi-
cally to specialize on different resources, then a clade may not be able
to radiate adaptively. Lack of such evolvability (Liem 1974; Vermeij
1974; Cheverud 1996; Wagner and Altenberg 1996; Gerhart and
Kirschner 1998; Rutherford and Lindquist 1998; see G. Wagner, Chap-
ter 8) may occur for a variety of reasons, often referred to as “evolu-
tionary constraints” (see Wray, Chapter 9).
The first hypothesis that one might pose about a clade that has failed to
radiate could invoke lack of ecological opportunity, but it is not clear how
one would test this idea.7 A way of doing so might involve estimating selec-
tion on an adaptive landscape (Fear and Price 1998; Schluter 2000; Arnold
et al. 2001). The existence of unutilized adaptive peaks might suggest that a
species had the opportunity to diversify and occupy those peaks. Of course,
the existence of multiple adaptive peaks on a landscape does not guarantee
that selection would push a clade to diversify to produce species occupy-
ing all of these peaks: speciation must occur, and the landscape itself will
change when other species are present. Research of this type has rarely been
conducted, the most thorough being studies on Darwin’s finches (Schluter
and Grant 1984; Schluter 2000; see also Case 1979). Development of these
sorts of ideas is needed to make ecological opportunity a fully operational
and predictive concept.
As a result, currently ecological opportunity is only recognizable after the
fact, as a plausible and often surely correct explanation for why a clade radi-
ated.8 As such, the concept may have great heuristic value in understanding
what causes adaptive radiation, but it may have little operational value in
the absence of a radiation to predict whether radiation could occur.
Adaptive Divergence, Speciation, and Adaptive Radiation
Adaptive radiation has two components: proliferation of species (specia-
tion) and divergence of species into different ecological niches (Losos 2009;
see Harrison, Chapter 13). Two important questions concern: (1) whether
any process other than natural selection could produce adaptive divergence
and (2) whether speciation and adaptation are causally connected.
Divergence of species to utilize different aspects of the environment could
occur in two ways, either with genetic drift playing a large role or by diver-
gent natural selection. Although drift by itself would not be expected to lead
7 In this regard, one might suggest that the concept of ecological opportunity
suffers from the same problems as the empty niche concept (Chase and Leibold
2003). Both empty niches and ecological opportunity are difficult to identify in
the absence of species that fill or take advantage of them. Moreover, one might
question whether resources are ever truly unutilized. How often, for example, is
some type of food resource not eaten by any organism? If nothing else, they are a
resource for decomposers, which is the reason for the wording of the definitions
of ecological opportunity presented earlier.
8 Ecological opportunity is also an explanation for why some species exhibit
exceptionally broad ecological and phenotypic diversity, which is arguably the
first step in adaptive radiation (Parent and Crespi 2009).
400  Chapter 15 Losos • Mahler
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated in
any form without express written permission from the publisher.
Adaptive Radiation: The Interaction of Ecological Opportunity, Adaptation, and Speciation  401
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated
in any form without express written permission from the publisher.
to adaptive change, in an adaptive landscape, it is possible that it could move
a population off one adaptive peak and into the domain of another (the basis
of Wright’s 1932 famous shifting balance theory). Alternatively, drift could
move a population along an adaptive ridge, from one high point in an adap-
tive landscape to another equally high point, assuming that the two points
are connected by a ridge of equally high fitness (Schluter 2000). While these
scenarios could ultimately result in the evolution of a suite of adaptively
differentiated species, they have received relatively little empirical support.
Rather, the standard and widely accepted view is that adaptive radiation
is driven by divergent natural selection, in which species diverge as they
adapt to use different parts of the environment. Most theories of adap-
tive radiation assume that a trade-off exists, such that enhanced adapta-
tion to use one part of the environment comes with a concomitant cost
of decreased adaptation to another part of the environment. Evidence for
such trade-offs is strongly implied by work on polymorphisms and local
adaptation (Schluter 2000), although the specific traits involved are often
unknown (see Agrawal et al., Chapter 10).
With regard to the second question, speciation and adaptive divergence
could be related in a number of ways:
1. Allopatric speciation could occur with adaptive divergence.
2. Allopatric speciation could occur without adaptive divergence, fol-
lowed by adaptive divergence when species secondarily establish
sympatry.
3. Some degree of adaptive divergence and evolution of reproductive
isolation could occur in allopatry, followed by enhanced adaptive
divergence and completion of the speciation process in sympatry (if
not completed in allopatry).
4. Speciation could occur in sympatry accompanied by adaptive diver-
gence. With respect to adaptive radiation, adaptive divergence is usu-
ally invoked as an integral part of the speciation process, though in
theory, sympatric speciation could occur in non-adaptive ways (e.g.,
by polyploidy), followed by adaptive divergence, which would occur
as in the first scenario in this list.
Few workers have suggested that most or all of the adaptive divergence
during adaptive radiation evolves in allopatry (option 1). Rather, sympatric
divergence, either during or after speciation (options 2, 3, and 4), is the pri-
mary, though not exclusive, focus of theoretical and empirical discussion.
Sympatric speciation driven by disruptive selection is in a way the most
biogeographically parsimonious explanation for the occurrence of a clade of
co-occurring, ecologically differentiated species, because it does not require
the invocation of one or more rounds of range contraction and expansion to
permit allopatric speciation followed by current-day sympatry. Nonethe-
less, the possibility of sympatric speciation of this sort is highly controver-
sial, and probably the majority of workers consider the prerequisites for
it to occur to be very stringent and likely to be met by few organisms in
most settings (Coyne and Orr 2004; Gavrilets 2004; Bolnick and Fitzpatrick
2007; see Harrison, Chapter 13). In a few examples, the case for adaptive
radiation by sympatric speciation is strong. The most convincing is the
occurrence of two clades of ecologically differentiated cichlid fishes, each
one in different volcanic crater lakes in Cameroon (Schliewen et al. 1994).
The monophyly of the clades (comprised of 9 and 11 species) makes in situ
speciation far more plausible than the alternative of many colonization
events followed by extinction of related forms outside the lakes (which
would make the lake species monophyletic with respect to extant species).
The lakes are small and homogeneous, so it is hard to imagine an allopatric
phase in the speciation process, leading to the conclusion that sympatric
speciation likely occurred.
Many clades of insects, some extremely species-rich, are composed of
species adapted to specialize on different host plants (see Berenbaum and
Schuler, Chapter 11). In many cases, multiple—sometimes many—clade
members occur sympatrically. Some workers in this area consider the evo-
lution of sympatric host-specialist species to be most readily explicable
by sympatric speciation (Berlocher and Feder 2002; Drès and Mallet 2002;
Abrahamson and Blair 2008). However, controversy over the theoretical
likelihood of sympatric speciation also pertains to host race speciation
(Coyne and Orr 2004; Gavrilets 2004; Bolnick and Fitzpatrick 2007). Fu-
tuyma (2008) argued that the empirical evidence in support of this view is
not generally strong and suggests that host shifts in allopatry are a likely
alternative possibility. Although the literature on these insects usually is
not couched in terms of adaptive radiation (exceptions include Després
and Cherif 2004; Price 2008; Roderick and Percy 2008), many host-specialist
complexes surely are adaptive radiations, and if sympatric speciation is
a common mode of speciation in these groups, then they may represent
examples of adaptive radiation by sympatric speciation.
The extreme alternative to adaptive divergence during sympatric specia-
tion is speciation in allopatry without adaptive ecological divergence. Sub-
sequently, the new species become sympatric and as a result of divergent
natural selection, adapt to different ecological niches (i.e., ecological char-
acter displacement). Speciation without adaptive divergence could occur
if allopatric populations diverge and become reproductively isolated as a
result of genetic drift or if sexual selection pressures in the two popula-
tions lead to the evolution of different mating preferences (Gittenberger
1991; Price 2008). One difficulty with these scenarios is that, if resources
are limiting, ecologically undifferentiated species may compete when they
become sympatric and thus, may not be able to coexist long enough for evo-
lutionary divergence to occur (MacArthur and Levins 1967; Slatkin 1980;
Gomulkiewicz and Holt 1995).
An intermediate possibility emphasizes the role that different selec-
tive environments may play in causing allopatric populations to diverge.
Studies in both the laboratory and in nature clearly indicate that isolated
402  Chapter 15 Losos • Mahler
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated in
any form without express written permission from the publisher.
Adaptive Radiation: The Interaction of Ecological Opportunity, Adaptation, and Speciation  403
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated
in any form without express written permission from the publisher.
populations experiencing different selective pressures are more likely to
evolve reproductive isolation as an incidental result of adaptive differentia-
tion (Rice and Hostert 1993; Funk et al. 2006; Funk and Nosil 2008) (Figure
15.6). Consequently, speciation is more likely to occur, or at least be initiated,
by populations experiencing different selective pressures. Moreover, if such
populations become sympatric, the evolved differences in ecology increase
the possibility that the populations can coexist long enough for character
displacement to occur, leading to greatly enhanced ecological differentia-
tion that permits coexistence. This scenario corresponds to the archipelago
model of adaptive radiation (Grant and Grant 2008a; Price 2008).
Because species proliferation is required for adaptive radiation to occur,
taxa that are more likely to speciate may also be more likely to adaptively
radiate. Thus, factors that predispose taxa to speciate, such as a mating
system emphasizing female choice, may be linked to adaptive diversifica-
tion (Schluter 2000; Rundell and Price 2009). In the same vein, the evolu-
tion of a trait that leads to enhanced speciation rates may be the trigger
that promotes adaptive radiation in a group, if ecological opportunity is
already present.
Overall, the role of speciation in adaptive radiation is poorly under-
stood. In recent years, “ecological speciation”—the idea that divergent ad-
aptation to different environments leads to the speciation—has attracted
considerable attention (for insightful reviews, see Hendry 2009 and see
Harrison, Chapter 13). There is no doubt that adaptation to different en-
vironments by allopatric populations enhances the likelihood that those
populations will become reproductively isolated, and this model is a key
stage in the standard model for adaptive radiation on islands. Whether or
not ecological speciation in sympatry (i.e., sympatric speciation) commonly
occurs and leads to adaptive radiation is an open question. Similarly, the
role of nonadaptive modes of speciation, such as founder effect speciation,
in adaptive radiation is also controversial (Coyne and Orr 2004; Gavrilets
2004; Futuyma 2005; Price 2008). More generally, the extent to which adap-
tive radiation is limited by the production of new species is unclear and
thus the degree to which factors that promote speciation may be important
indirect promoters of adaptation radiation remains uncertain.
Adaptation to Different Aspects of the Environment in the Absence
of Interspecific Interactions
Regardless of how speciation occurs, what ecological mechanisms drive
adaptive divergence? Divergent natural selection is the underlying cause,
but the question is whether the driving force behind such selection is sim-
ply adaptation to different aspects of the environment or whether inter-
specific interactions change the selective landscape, producing divergent
selection that would not occur in the absence of the interacting species
(Schluter 2000).9
It is easy to envision how two allopatric populations experiencing differ-
ent environments would evolve different adaptations. Moreover, because the
environment and hence, the adaptive landscape is rarely identical in different
places, it is possible that two adaptive peaks that are separated by an adap-
tive valley in one location may be connected by an adaptive ridge in another.
Suppose a population on one island occupied the lower of two adaptive
peaks, which were separated by an adaptive valley, so that the higher peak
was not attainable. If members of that population colonized a second island
on which the peaks were connected by an adaptive ridge, then the popula-
tion would adapt to the higher peak. Subsequently, if members of the second
population colonized the first island, they would occur on the higher peak,
producing sympatry of the species on different peaks (Figure 15.7).
In theory, this sort of scenario, in which adaptive divergence occurs
entirely in allopatry, could lead to adaptive radiation, producing species
adapted to many different aspects of the environment. Most views of adap-
tive radiation do not take such an extreme view, though proponents of
the role of interspecific interactions envision allopatric differentiation as
the initial step in divergence, setting the stage for subsequent, much more
substantial divergence in sympatry that is driven by sympatric interactions
(i.e., character displacement).
An alternative view is that divergent selection in sympatry can split an
initially homogeneous population into distinct, reproductively isolated and
9 In this discussion, we consider the food an organism eats to be part of the
environment. By interspecific interactions, we refer to interactions among species
on the same trophic level or between the focal species and other species on
higher trophic levels.
Darwin Bell
Sinauer Associates
Morales Studio
Figure 15.06 Date 04-21-10
0.0 0.25 0.5 0.75
r = +0.43
1.0
Habitat divergence
Residual postmating isolation
Figure 15.6  The Relationship between Divergent Adaptation and
Reproductive Isolation The more dissimilar two species are in habitat use, the
greater the degree of reproductive isolation that has evolved between them.
Each point represents the degree of difference between two closely related spe-
cies or populations. (Adapted from Funk et al. 2006.)
404  Chapter 15 Losos • Mahler
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated in
any form without express written permission from the publisher.
Adaptive Radiation: The Interaction of Ecological Opportunity, Adaptation, and Speciation  405
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated
in any form without express written permission from the publisher.
adaptively differentiated populations. Some variants of this model envision
intraspecific resource competition or interspecific interactions as the driv-
ing force, but many prominent proposals, particularly those concerning
host–plant specialists (see following discussion), simply invoke adaptation
to different aspects of the environment, requiring a tradeoff such that dis-
ruptive selection produces a bimodal distribution of phenotypes. As just
discussed, the prerequisites for reproductive isolation to evolve in such a
scenario are very strict, and controversy still exists on how likely it is to
occur, even in host–plant specialized insects.
Adaptation to Different Aspects of the Environment
as a Result of Interspecific Interactions
Interspecific Competition
Dating back to the seminal work by Simpson (1953), which focused pri-
marily on vertebrates, a commonly held view is that adaptive radiation
primarily results from interspecific competition for resources, leading to
character displacement and adaptation to different resources (Schluter
2000; Grant and Grant 2008a). Repeated numerous times, this process leads
to sympatric coexistence of species adapted to a variety of different ecologi-
cal niches (i.e., an adaptive radiation).
Although character displacement was controversial in the 1970s and
1980s, a growing consensus exists that it is an important evolutionary phe-
nomenon (Schluter 2000; Dayan and Simberloff 2005; Pfennig and Pfennig
2009). Experimental studies have confirmed that interspecific competition
can lead to strong divergent selection (Schluter 1994; Schluter 2003), and the
process of character displacement has been directly observed in the field in
Darwin’s finches (Grant and Grant 2006) (Figure 15.8) as well as in labora-
tory studies (e.g., Barrett and Bell 2006; Tyerman et al. 2008; Kassen 2009).
It seems safe to conclude that interspecific competition-driven character
displacement is a common means by which adaptive radiation occurs.
The question is, then, whether interactions other than competition can
lead to adaptive radiation, and if so, whether they have been important in
driving adaptive radiations throughout the history of life.
Predation, Parasitism, and Herbivory
In ecological terms, predation, parasitism, and herbivory are similar pro-
cesses in that they refer to individuals of one species directly consuming
members of another species (hence, we refer to all as “predators” in the
following discussion). These processes could produce divergent natural se-
lection pressures when species adapt to different predators or when species
adapt in divergent ways to the same predator. Moreover, species initially
preyed upon by the same predator may diverge so as not to share the same
predator. Although predation may spur divergence and diversification, it
also can hinder it, and the extent to which predation plays an important
role in driving adaptive radiation remains poorly understood (Vamosi
2005; Langerhans 2006).
One can easily envision how species occurring in different places, with
different predation regimes, would face selection to evolve different pheno-
types. Several recent studies have demonstrated the importance of preda-
tion in driving such divergence. For example, both mosquitofish and dam-
selflies exhibit differences in their behavior, habitat use, and morphology
depending on which predators are present in the lakes in which they occur
(McPeek et al. 1996; Stoks et al. 2003; Langerhans et al. 2007).
In sympatry, predation-driven selection may be divergent for several
reasons. First, multiple ways may exist to avoid predation. For example,
Timema walking sticks that use different host plants have diverged to en-
hance crypsis on the differently colored plants; manipulative experimental
studies demonstrate that selection is divergent when predators are pres-
ent (Nosil and Crespi 2006) (Figure 15.9). Among garter snakes, disruptive
selection favors striped individuals that rapidly flee from predators and
Darwin Bell
Sinauer Associates
Morales Studio
Figure 15.07 Date 04-29-10
Mean
fitness
Island A
Mean
fitness
Island B
(A)
(B)
Colonization
Backcolonization
Trait 2
Trait 1
Figure 15.7Peak Shifts along an Adaptive Ridge (A) The ancestral species
occurs on the lower adaptive peak on Island A. Because an adaptive valley lies
between the two peaks, natural selection cannot drive the population to the sec-
ond peak. (B) However, on Island B, an adaptive ridge connects the ancestral posi-
tion to a higher peak, and when individuals immigrate to Island B from Island
A, the new population on Island B evolves a new morphology. Subsequently, the
population on Island B evolves reproductive isolation. When individuals from
Island B then recolonize Island A, they may evolve to the second peak on that
island, leading to the coexistence of two species—one on each adaptive peak.
(Adapted from Schluter 2000.)
406  Chapter 15 Losos • Mahler
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated in
any form without express written permission from the publisher.
Adaptive Radiation: The Interaction of Ecological Opportunity, Adaptation, and Speciation  407
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated
in any form without express written permission from the publisher.
blotched individuals that remain hidden (Brodie 1992). In
theory, species evolving different anti-predator adapta-
tions might not differ in any other way with regard to re-
source or habitat use, in which case ecological opportunity
might not be involved in adaptive divergence. However,
most documented examples include correlated shifts in other ecological and
behavioral aspects; the evolution of body armor, for example, has conse-
quences for locomotion, which in turn may affect where and how an animal
can forage (Bergstrom 2002; Losos et al. 2002). Consequently, in this pred-
ator-driven scenario for adaptive radiation, ecological opportunity would
still be required; multiple distinct habitats or resources to which different
prey species could adapt would be necessary, and these different niches
could not already be preempted by other species.
Selection may also favor sympatric species to diverge in habitat use so
as to avoid being preyed upon by the same predator. If prey species are
preyed upon by the same predator, then under some circumstances, in-
creased population size of one of the prey species would lead to increased
population size of the predator; thus, the population size of the other prey
species would decrease, as they are preyed upon by the greater number of
predators. The result is that a negative relationship would exist between the
population sizes of the prey species, just as would occur through interspe-
cific competition (Holt 1977). Assuming the predator species is not able to
function equally successfully in all parts of the environment, prey species
may diverge to use different resources or habitats, if they are available, and
thus no longer share predators. Subsequently, prey species would adapt to
the different habitats or resources they were utilizing, producing the same
outcome as competition-driven character displacement: an adaptive radia-
tion driven by “competition for enemy free space” (Jeffries and Lawton
1984) or “apparent competition” (Holt 1977).
In theory, adaptive radiation also could result from predation-driven di-
vergent selection by the means just outlined. However, we are aware of few
purported cases. The diversity of some tropical butterfly clades, involving
multiple mimicry complexes and a suite of other ecological and behavioral
differences, might be one example (Elias et al. 2008).
Processes Driving Radiation: Conclusions
The role of interspecific competition in driving evolutionary radiation is
well established and likely to be of paramount importance. Other ecologi-
cal processes may be important, either directly (e.g., predation, herbivory,
Darwin Bell
Sinauer Associates
Morales Studio
Figure 15.08 Date 04-19-10
(A)
1
2
3
4
(C)
7 8 9 10 11 12 13 14 15 16 17 18 19
10
0
20
30
40
Beak depth (mm)
Individuals
G. fortis G. magnirostris
(a)
(b)
1975 1980 1985 1990 1995 2000 2005
–1.0
–1.5
–0.5
0.0
0.5
1.0
Year
Beak size
(B)
Figure 15.8  Character Displacement in Darwin’s Finches (A) Daphne Major, in
the Galápagos Islands, harbors both the large ground finch, Geospiza magnirostris
(A2), and the medium ground finch, Geospiza fortis (A1, A3), the latter of which
exhibits substantial variability in beak shape. Only large-beaked birds can eat
large seeds, such as those from Tribulus cistoides (A4). (B) During the drought
of 1977, when only the medium ground finch occurred on the island, small
seeds were rapidly consumed and only large seeds remained. Selection strongly
favored large-beaked medium ground finches (as indicated by the calculation
of selection gradients, which are not shown here), and the population evolved
larger beak size. Another drought occurred in 2003 and 2004; however, in the
intervening years, the large ground finch had colonized Daphne Major. During
this drought, the large ground finches monopolized the larger seeds. Mortality
was very high in both species, and in the medium ground finch, smaller-beaked
birds that could eat the few remaining small seeds were favored, and the popula-
tion evolved smaller beak size, the opposite of what occurred in the absence of
the large-beaked ground finch. Beak size units represent scores on the first axis of
a principal components analysis of six bill measurements. (C) Thus, in the 2003–
2004 drought, selection favored smaller-beaked birds in the medium ground finch
and larger-beaked birds in the large ground finch, and the phenotypic distribu-
tions before the drought (indicated by blue bars) and afterwards (indicated by red
bars) can be compared; as a result, differences in beak size between the two spe-
cies were greater after the drought (arrow a) than before (arrow b), making this a
classic example of character displacement. (A, photos from Grant and Grant 2006;
B, adapted from Grant and Grant 2006; C, adapted from Grant and Grant 2008b.)
s
408  Chapter 15 Losos • Mahler
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated in
any form without express written permission from the publisher.
Adaptive Radiation: The Interaction of Ecological Opportunity, Adaptation, and Speciation  409
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated
in any form without express written permission from the publisher.
and parasitism) or indirectly (e.g., mutualism and coevolution), by creat-
ing the opportunity for other processes to operate. However, empirical
evidence demonstrating the role of other ecological processes is sparse. To
some extent, this paucity of case studies, may be a result of the historical
focus on competition as a driver of character displacement, but attention
among both theorists and ecologists to other mechanisms has occurred
for long enough now that one might expect more examples to have been
documented. The lack of such examples would seem to be indicative of
their limited importance relative to that of direct competition. Certainly,
the evolution of mutualisms and coevolution are events that can trigger
adaptive radiation. How frequently they occur is unclear.
Laboratory experiments on microbial evolution support these conclu-
sions (see Dykhuizen, Commentary 2). Kassen (2009) reviewed the rapidly
growing literature in this field and found overwhelming support for the role
of interspecific competition as the primary driver of adaptive radiation. The
addition of predators as an experimental treatment occasionally enhanced
adaptive radiation but only when the variety of resources was limited.
In situations with a wide variety of resources, the presence of predators
slowed, rather than enhanced, the rate of adaptive diversification (Meyer
and Kassen 2007; see also Benmayor et al. 2008). Other possible triggers
for adaptive radiation, such as ecosystem engineering or the evolution of
mutualisms, also rarely are important (Kassen 2009). The correspondence
between the results from experimental laboratory studies and the empirical
literature from nature suggests that interspecific competition is the primary
driver of adaptive radiation.
Conclusions and Future Directions
Many of what we consider the classic ideas about adaptive radiation are
well supported. In particular, ecological opportunity usually is the key
to adaptive radiation, and interspecific competition often is the driving
force behind it. Nonetheless, these are generalizations, and further work
is needed to understand the relative frequency and significance of alterna-
tives. In many cases, other possibilities, such as the role of predation in
driving adaptive radiation or the extent to which radiations create their
own ecological opportunity, have been little explored. Whether further
work will alter the bigger picture of our understanding of adaptive radia-
tion remains to be seen.
We also conclude that while ecological opportunity is typically necessary,
it is not sufficient for adaptive radiation, as many clades seemingly in the
presence of opportunity fail to radiate. Several factors likely contribute, but
at present, these are too poorly understood to predict with any certainty
whether adaptive radiation will occur in the presence of ecological opportu-
nity. Why, for example, did Darwin’s finches, but not mockingbirds, radiate
in the Galápagos? Experimental diversification studies have come to the fore-
front in filling this gap; the combination of replication, strict environmental
control, and fine-scale genetic characterization of lineages makes for a prom-
ising approach to dissect the nature of historical contingency in adaptive ra-
diation (Lenski and Travisano 1994; Fukami et al. 2007; Blount et al. 2008).
While we have offered a review of the current knowledge pertaining
to adaptive radiation, numerous basic questions about adaptive radiation
remain unanswered, and recent conceptual and technological advances
promise to lead adaptive radiation research into exciting new directions.
Toward this end, we suggest that new approaches that take advantage of
both conceptual and technological advances may be instrumental in mov-
ing forward understanding of adaptive radiation in years to come.
Darwin Bell
Sinauer Associates
Morales Studio
Figure 15.09 Date 04-21-10
0.0–4.0 –4.04.0
0.10
0.30
0.50
0.70
Brightness contrast
Fitness
(A) Ceanothus: predation present
0.0 4.0
Brightness contrast
(B) Adenostoma: predation present
0.0–4.0 4.0
0.10
0.30
0.50
0.70
Brightness contrast
Fitness
(C) Ceanothus: predation absent
0.0–4.0 4.0
Brightness contrast
(D) Adenostoma: predation absent
Figure 15.9  Selection for Background Matching in the Walking Stick, Timema
cristinae Experimental studies of selection reveal that in the presence of avian
predators, dull stripes, and bright bodies (“brightness contrast” is body bright-
ness minus stripe brightness) are favored in the Ceanothus ecotype (A) and
bright stripes and dull bodies in the Adenostoma ecotype (B). In the absence of
predators, selection does not differ between insects on the two bushes (C and D).
(Adapted from Nosil and Crespi 2006.)
410  Chapter 15 Losos • Mahler
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated in
any form without express written permission from the publisher.
Adaptive Radiation: The Interaction of Ecological Opportunity, Adaptation, and Speciation  411
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated
in any form without express written permission from the publisher.
Our review has focused, for the most part, on the role of external envi-
ronmental factors in shaping adaptive radiation. However, over the last
several decades, evolutionary biologists have debated the role that internal
constraints—manifested in limitations and directionality in the availability
of phenotypic variation upon which selection can operate—play in limiting
and directing evolutionary change (Gould 2002).
The role of such constraints on adaptive radiation is now being addressed
in two distinctive ways. On one hand, a number of researchers have sug-
gested that interspecific hybridization can provide enhanced variation that
may be critical in allowing extensive adaptive diversification. On the other
hand, the burgeoning field of genomics is now at last permitting evolution-
ary biologists to truly understand the relationship between genetic change
and phenotypic response.
Hybridization
Several authors have recently suggested that hybridization among closely
related species may play an important role in generating adaptive diversity
during the early stages of adaptive radiation (Rieseberg et al. 1999; See-
hausen 2004; Grant and Grant 2008a,b; Mallet 2009). Increasingly sophis-
ticated methods are being developed to distinguish past gene flow from
incomplete lineage sorting, permitting detection of historical hybridiza-
tion in a phylogenetic framework (Hey and Nielsen 2004; Hey et al. 2004;
Kubatko 2009), which may facilitate investigation of the contribution of
hybridization to adaptive diversification. Nonetheless, directly measuring
the effect of hybridization on divergence in the wild is often difficult be-
cause of the low frequency of hybridization events and the infeasibility of
tracking hybrid lineages over many generations, although Grant and Grant
(2008b, 2009) provide remarkable examples in Darwin’s finches. However,
future studies might assess the role of hybridization in adaptive radiation
by testing whether radiations that exhibit historical signatures of hybrid-
ization exhibit greater adaptive diversity than radiations in which there is
little evidence of hybridization.
Genomics
Thanks to the ever-decreasing cost of genome sequencing, the ability to
examine the genomes of multiple members of an adaptive radiation will
soon be readily available. With such information, researchers will be able
to examine the extent to which genetic architecture constrains and directs
the pathways by which adaptive diversification has occurred. Combined
with studies of natural selection in the field, we soon will have the ability
to fully integrate genetics and the study of natural selection to understand
how and why adaptive radiation has occurred. Experimental studies of this
sort have already been conducted (e.g., microbial laboratory experiments:
Bantinaki et al. 2007 and Spencer et al. 2007; sticklebacks: Barrett et al. 2008)
and no doubt will soon become commonplace. The next 10–20 years should
prove extremely enlightening as the integration of genomic and selection
studies, in a phylogenetic context, ushers in a golden age for the study of
adaptive radiation.
The Impact of Adaptive Radiation on Communities and Ecosystems
Finally, virtually all research on adaptive radiation has investigated the in-
fluence of ecological factors on adaptive diversification, but until recently,
few have asked how the process of adaptive radiation may alter the struc-
ture of communities and ecosystems. Adaptive radiation typically involves
a dramatic change in the ecological diversity of a lineage, usually within a
local setting. Diversification may influence community interactions, alter
food webs, and ultimately affect nutrient and energy flow in ecosystems,
and radiations have been implicated in major ecological changes in Earth’s
history (e.g., the rise and proliferation of autotrophs changing the atmo-
spheric oxygen concentration roughly 2 billion years ago). However, only
recently has research specifically sought to quantify the influence of evo-
lutionary diversification on community structure and ecosystem function
(Loueille and Loreau 2005, 2006; Harmon et al. 2009; Ingram et al. 2009).
Continued research in this direction may result in a richer understanding
of the complex and interactive relationship between ecological and evolu-
tionary diversity.
Acknowledgments
This chapter benefited from discussions with and critiques by: Adam Al-
gar, Mike Bell, Joel Cracraft, Doug Futuyma, Luke Harmon, Mark McPeek,
Trevor Price, Bob Ricklefs, John Thompson, and Peter Wainwright.
Literature Cited
Abrahamson, W. G. and C. P. Blair. 2008. Sequential radiation through host-race
formation: Herbivore diversity leads to diversity in natural enemies. In K. J.
Tilmon (ed.), Specialization, Speciation, and Radiation: The Evolutionary Biology of
Herbivorous Insects, pp. 188–202. University of California Press, Berkeley.
Adams, D. C., C. M. Berns, K. H. Kozak, and 1 other. 2009. Are rates of species
diversification correlated with rates of morphological evolution? Proc. Roy.
Soc. Lond. B 276: 2729–2738.
Agrawal, A. A., M. Fishbein, R. Halitschke, and 3 others. 2009. Evidence for adap-
tive radiation from a phylogenetic study of plant defenses. Proc. Natl. Acad. Sci.
USA 106: 18067–18072.
Alfaro, M. E., F. Santini, and C. D. Brock. 2007. Do reefs drive diversification in
marine teleosts? Evidence from the pufferfish and their allies (Order Tetrao-
dontiformes). Evolution 61: 2104–2126.
Alfaro, M. E., F. Santini, C. Brock, and 5 others. 2009. Nine exceptional radiations
plus high turnover explain species diversity in jawed vertebrates. Proc. Natl.
Acad. Sci. USA 106: 13410–13414.
412  Chapter 15 Losos • Mahler
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated in
any form without express written permission from the publisher.
Adaptive Radiation: The Interaction of Ecological Opportunity, Adaptation, and Speciation  413
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated
in any form without express written permission from the publisher.
Arbogast, B. S., S. V. Drovetski, R. L. Curry, and 6 others. 2006. The origin and
diversification of Galápagos mockingbirds. Evolution 60: 370–382.
Arnold, S. J., M. E. Pfrender, and A. G. Jones. 2001. The adaptive landscape as a
conceptual bridge between micro- and macroevolution. Genetica 112/113: 9–32.
Bakker, R. T. 1968. The superiority of dinosaurs. Discovery 3: 11–22.
Bantinaki, E., R. Kassen, C. G. Knight, and 3 others. 2007. Adaptive divergence in
experimental populations of Pseudomonas fluorescens. III. Mutational origins of
wrinkly spreader diversity. Genetics 176: 441–453.
Barker, K. F., A. Cibois, P. Schikler, and 2 others. 2004. Phylogeny and diversifica-
tion of the largest avian radiation. Proc. Natl. Acad. Sci. USA 101: 1040–1045.
Barraclough, T. G. and S. Nee. 2001. Phylogenetics and speciation. Trends Ecol.
Evol. 16: 391–399.
Barrett, R. D. H. and G. Bell. 2006. The dynamics of diversification in evolving
Pseudomonas populations. Evolution 60: 484–490.
Barrett, R. D. H., S. M. Rogers, and D. Schluter. 2008. Natural selection on a major
armor gene in threespine stickleback. Science 322: 255–257.
Baum, D. A. and A. Larson. 1991. Adaptation reviewed: A phylogenetic method-
ology for studying character macroevolution. Syst. Zool. 40: 1–18.
Benkman, C. W. 1991. Predation, seed size partitioning and the evolution of body
size in seed-eating finches. Evol. Ecol. 5: 118–127.
Benmayor, R., A. Buckling, M. B. Bonsall, and 2 others. 2008. The interactive
effects of parasites, disturbance, and productivity on experimental adaptive
radiations. Evolution 62: 467–477.
Benton, M. J. 1996. On the nonprevalence of competitive replacement in the evo-
lution of tetrapods. In D. Jablonski, D. H. Erwin, and J. Lipps (eds.), Evolution-
ary Paleobiology, pp. 185–210. University of Chicago Press, Chicago.
Benton, M. J. 2009. The red queen and the court jester: Species diversity and the
role of biotic and abiotic factors through time. Science 323: 728–732.
Bergstrom, C. A. 2002. Fast-start swimming performance and reduction in lateral
plate number in threespine stickleback. Can. J. Zool. 80: 207–213.
Berlocher, S. H. and J. L. Feder. 2002. Sympatric speciation in phytophagous
insects: Moving beyond controversy? Ann. Rev. Entom. 47: 773–815.
Blount, Z. D., C. Z. Borland, and R. E. Lenski. 2008. Historical contingency and
the evolution of a key innovation in an experimental population of Escherichia
coli. Proc. Natl. Acad. Sci. USA 105: 7899–7906.
Bokma, F. 2009. Problems detecting density-dependent diversification on phylog-
enies. Proc. Roy. Soc. Lond. B 276: 993–994.
Bolnick, D. I. and B. M. Fitzpatrick. 2007. Sympatric speciation: Models and
empirical evidence. Ann. Rev. Ecol. Evol. Syst. 38: 459–487.
Bourke, A. F. G. and N. R. Franks. 1991. Alternative adaptations, sympatric
speciation and the evolution of parasitic, inquiline ants. Biol. J. Linn. Soc. 43:
157–178.
Branch, W. R. 1998. Field Guide to Snakes and Other Reptiles of Southern Africa.
Ralph Curtis Books, Sanibel Island, FL.
Brockhurst, M. A., N. Colegrave, D. J. Hodgson, and 1 other. 2007. Niche occupa-
tion limits adaptive radiation in experimental microcosms. PLoS One 2: e193.
Brodie, E. D. III. 1992. Correlational selection for color pattern and antipredator
behavior in the garter snake Thamnophis ordinoides. Evolution 46: 1284–1298.
Brusatte, S. L., M. J. Benton, M. Ruta, and 1 other. 2008. Superiority, competi-
tion, and opportunism in the evolutionary radiation of dinosaurs. Science 321:
1485–1488.
Carlquist, S. J. 1974. Island Biology. Columbia University Press, New York.
Case, T. J. 1979. Character displacement and coevolution in some Cnemidophorus
lizards. Forts. Zool. 25: 235–282.
Chase, J. M. 2007. Drought mediates the importance of stochastic community
assembly. Proc. Natl. Acad. Sci. USA 104: 17430–17434.
Chase, J. M. and M. A. Leibold. 2003. Ecological Niches: Linking Classical and Con-
temporary Approaches. University of Chicago Press, Chicago.
Cheverud, J. M. 1996. Developmental integration and the evolution of pleiotropy.
Am. Zool. 36: 44–50.
Ciampaglio, C. N. 2002. Determining the role that ecological and developmental
constraints play in controlling disparity: Examples from the crinoid and blas-
tozoan fossil record. Evol. Dev. 4: 170–188.
Ciampaglio, C. N., M. Kemp, and D. W. McShea. 2001. Detecting changes in mor-
phospace occupation patterns in the fossil record: Characterization and analy-
sis of measures of disparity. Paleobiology 27: 695–715.
Collar, D. C., T. J. Near, and P. C. Wainwright. 2005. Comparative analysis of
morphological diversity: Does disparity accumulate at the same rate in two
lineages of centrarchid fishes? Evolution 59: 1783–1794.
Coyne, J. A., and H. A. Orr. 2004. Speciation. Sinauer Associates, Sunderland, MA.
Cracraft, J. 1990. The origin of evolutionary novelties: Pattern and process at
different hierarchical levels. In M. Nitecki (ed.), Evolutionary Innovations, pp.
21–44. University of Chicago Press, Chicago.
Darwin, C. 1845. Journal of Researches into the Natural History and Geology of the
Countries Visited During the Voyage of H.M.S. Beagle Round the World, under the
Command of Capt. FitzRoy, R.N, 2nd ed. John Murray, London.
Davis, M. A. 2003. Biotic globalization: Does competition from introduced species
threaten biodiversity? BioScience 53: 481–489.
Dayan, T. and D. Simberloff. 2005. Ecological and community-wide character dis-
placement: The next generation. Ecol. Lett. 8: 875–894.
de Queiroz, A. 2002. Contingent predictability in evolution: Key traits and diver-
sification. Syst. Biol. 51: 917–929.
Després, L. and M. Cherif. 2004. The role of competition in adaptive radiation: A
field study on sequentially ovipositing host-specific seed predators. J. Anim.
Ecol. 73: 109–116.
Donoghue, M. J. 2005. Key innovations, convergence, and success: Macroevolu-
tionary lessons from plant phylogeny. Paleobiology 31(suppl.): 77–93.
Drès, M. and J. Mallet. 2002. Host races in plant-feeding insects and their impor-
tance in sympatric speciation. Phil. Trans. Roy. Soc. Lond. B 357: 471–492.
Ehrlich, P. R. and P. H. Raven. 1964. Butterflies and plants: A study in coevolu-
tion. Evolution 18: 586–608.
Elias, D. O., M. M. Kasumovic, D. Punzalan, and 2 others. 2008. Assessment
during aggressive contests between male jumping spiders. Anim. Behav. 76:
901–910.
Erwin, D. H. 2001. Lessons from the past: Biotic recoveries from mass extinctions.
Proc. Natl. Acad. Sci. USA 98: 5399–5403.
Erwin, D. H. 2007. Disparity: Morphological pattern and developmental context.
Palaeontology 50: 57–73.
Erwin, D. H. 2008. Macroevolution of ecosystem engineering, niche construction
and diversity. Trends Ecol. Evol. 23: 304–310.
Farrell, B. D. 1998. “Inordinate fondness” explained: Why are there so many
beetles? Science 281: 553–557.
414  Chapter 15 Losos • Mahler
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated in
any form without express written permission from the publisher.
Adaptive Radiation: The Interaction of Ecological Opportunity, Adaptation, and Speciation  415
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated
in any form without express written permission from the publisher.
Farrell, B. D. and C. Mitter. 1994. Adaptive radiation in insects and plants: Time
and opportunity. Am. Zool. 34: 57–69.
Farrell, B. D. and C. Mitter. 1998. The timing of insect/plant diversification: Might
Tetraopes (Coleoptera: Cerambycidae) and Asclepias (Asclepiadaceae) have co-
evolved? Biol. J. Linn. Soc. 63: 553–577.
Fear, K. K. and T. Price. 1998. The adaptive surface in ecology. Oikos 82: 440–448.
Foote, M. 1993. Discordance and concordance between morphological and taxo-
nomic diversity. Paleobiology 19: 185–204.
Foote, M. 1996. Ecological controls on the evolutionary recovery of post-Paleozoic
crinoids. Science 274: 1492–1495.
Foote, M. 1997. The evolution of morphological diversity. Ann. Rev. Ecol. Syst. 28:
129–152.
Foote, M. 1999. Morphological diversity in the evolutionary radiation of Paleozoic
and post-Paleozoic crinoids. Paleobiology 25: 1–115.
Forbes, A. A., T. H. Q. Powell, L. L. Stelinski, and 2 others. 2009. Sequential sym-
patric speciation across trophic levels. Science 323: 776–779.
Freckleton, R. P. and P. H. Harvey. 2006. Detecting non-Brownian trait evolution
in adaptive radiations. PLoS Biol. 4: 2104–2111.
Friedman, M. 2010. Explosive morphological diversification of spiny-finned
teleost fishes in the aftermath of the end-Cretaceous extinction. Proc. Roy. Soc.
Lond. B 277: 1675–1683.
Fukami, T., H. J. E. Beaumont, X. X. Zhang, and 1 other. 2007. Immigration his-
tory controls diversification in experimental adaptive radiation. Nature 446:
436–439.
Funk, D. J. and P. Nosil. 2008. Comparative analyses of ecological speciation. In
K. J. Tilmon (ed.), Specialization, Speciation, and Radiation: The Evolutionary Biol-
ogy of Herbivorous Insects, pp. 117–135. University of California Press, Berkeley.
Funk, D. J., P. Nosil, and W. J. Etges. 2006. Ecological divergence exhibits consis-
tently positive associations with reproductive isolation across disparate taxa.
Proc. Natl. Acad. Sci. USA 103: 3209–3213.
Fürsich, F. T. and D. Jablonski. 1984. Late Triassic naticid drillholes: Carnivorous
gastropods gain a major adaptation but fail to radiate. Science 224: 78–80.
Futuyma, D. J. 1998. Evolutionary Biology. Sinauer Associates, Sunderland, MA.
Futuyma, D. J. 2005. Progress on the origin of species. PLoS Biol. 3: 197–199.
Futuyma, D. J. 2008. Sympatric speciation: Norm or exception? In K. J. Tilmon
(ed.), Specialization, Speciation, and Radiation: The Evolutionary Biology of Herbivo-
rous Insects, pp. 136–148. University of California Press, Berkeley.
Gavrilets, S. 2004. Fitness Landscapes and the Origin of Species. Princeton University
Press, Princeton.
Gavrilets, S. and A. Vose. 2009. Dynamic patterns of adaptive radiation: Evolu-
tion of mating preferences. In R. K. Butlin, J. R. Bridle, and D. Schluter (eds.),
Speciation and Patterns of Diversity, pp. 102–126. Cambridge University Press,
Cambridge, UK.
Gerhart, M. and J. Kirschner. 1998. Cells, Embryos, and Evolution: Toward a Cellular
and Developmental Understanding of Phenotypic Variation and Evolutionary Adapt-
ability. Blackwell, Oxford.
Gittenberger, E. 1991. What about nonadaptive radiation? Biol. J. Linn. Soc. 43:
263–272.
Givnish, T. J. 1997. Adaptive radiation and molecular systematics: Issues and
approaches. In T. J. Givnish and K. J. Sytsma (eds.), Molecular Evolution and
Adaptive Radiation, pp. 1–54. Cambridge University Press, Cambridge, UK.
Givnish, T. J., K. C. Millam, A. R. Mast, and 7 others. 2009. Origin, adaptive radia-
tion and diversification of the Hawaiian lobeliads (Asterales: Campanulaceae).
Proc. Roy. Soc. Lond. B 276: 407–416.
Glor, R. E. 2010. Phylogenetic approaches to the study of adaptive radiation. Ann.
Rev. Ecol. Evol. Syst. in revision.
Gomulkiewicz, R. and R. D. Holt. 1995. When does evolution by natural selection
prevent extinction. Evolution 49: 201–207.
Gould, S. J. 2002. The Structure of Evolutionary Theory. Harvard University Press,
Cambridge, MA.
Grant, B. R. and P. R. Grant. 2008b. Fission and fusion of Darwin’s finch popula-
tions. Phil. Trans. Roy. Soc. Lond. B 363: 2821–2829.
Grant, P. R. and B. R. Grant. 2006. Evolution of character displacement in Dar-
win’s finches. Science 313: 224–226.
Grant, P. R. and B. R. Grant. 2008a. How and Why Species Multiply: The Radiation of
Darwin’s Finches. Princeton University Press, Princeton.
Grant, P. R. and B. R. Grant. 2009. The secondary contact phase of allopatric spe-
ciation in Darwin’s finches. Proc. Natl. Acad. Sci. USA 106: 20141–20148.
Guyer, C. and J. B. Slowinski. 1993. Adaptive radiation and the topology of large
phylogenies. Evolution 47: 253–263.
Harmon, L. J., J. B. Losos, T. J. Davies, and 16 others. 2010. Early bursts of body
size and shape evolution are rare in comparative data. Evolution, in press.
Harmon, L. J., B. Matthews, S. Des Roches, and 3 others. 2009. Evolutionary diver-
sification in stickleback affects ecosystem functioning. Nature 458: 1167–1170.
Harmon, L. J., J. Melville, A. Larson, and 1 other. 2008. The role of geography and
ecological opportunity in the diversification of day geckos (Phelsuma). Syst.
Biol. 57: 562–573.
Harvey, P. H. and A. Rambaut. 2000. Comparative analyses for adaptive radia-
tions. Phil. Trans. Roy. Soc. Lond. B 355: 1599–1605.
Heard, S. B. and D. L. Hauser. 1995. Key evolutionary innovations and their eco-
logical mechanisms. Hist. Biol. 10: 151–173.
Hendry, A. P. 2009. Ecological speciation! Or the lack thereof? Can. J. Fish. Aquat.
Sci. 66: 1383–1398.
Herre, E. A., C. A. Machado, E. Bermingham, and 5 others. 1996. Molecular phy-
logenies of figs and their pollinator wasps. J. Biogeog. 23: 521–530.
Hey, J. and R. Nielsen. 2004. Multilocus methods for estimating population sizes,
migration rates and divergence time, with applications to the divergence of
Drosophila pseudoobscura and D. persimilis. Genetics 167: 747–760.
Hey, J., Y. J. Won, A. Sivasundar, and 2 others. 2004. Using nuclear haplotypes
with microsatellites to study gene flow between recently separated cichlid spe-
cies. Mol. Ecol. 13: 909–919.
Hodges, S. A. and M. L. Arnold. 1995. Spurring plant diversification: Are floral
nectar spurs a key innovation? Proc. Roy. Soc. Lond. B 262: 343–348.
Holt, R. D. 1977. Predation, apparent competition and the structure of prey com-
munities. Theor. Pop. Biol. 12: 197–229.
Hooghiemstra, H., V. M. Wijninga, and A. M. Cleef. 2006. The paleobotanical
record of Colombia: Implications for biogeography and biodiversity. An. Mo.
Bot. Gard. 93: 297–324.
Hughes, C. and R. Eastwood. 2006. Island radiation on a continental scale: Excep-
tional rates of plant diversification after uplift of the Andes. Proc. Natl. Acad.
Sci. USA 103: 10334–10339.
Hunter, J. P. 1998. Key innovations and the ecology of macroevolution. Trends
Ecol. Evol. 13: 31–36.
416  Chapter 15 Losos • Mahler
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated in
any form without express written permission from the publisher.
Adaptive Radiation: The Interaction of Ecological Opportunity, Adaptation, and Speciation  417
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated
in any form without express written permission from the publisher.
Ingram, T., L. J. Harmon, and J. B. Shurin. 2009. Niche evolution, trophic struc-
ture, and species turnover in model food webs. Am. Nat. 174: 56–67.
Janson, E. M., J. O. Stireman III, M. S. Singer, and 1 other. 2008. Phytophagous
insect-microbe mutualisms and adaptive evolutionary diversification. Evolu-
tion 62: 997–1012.
Jeffries, M. J. and J. H. Lawton. 1984. Enemy free space and the structure of eco-
logical communities. Biol. J. Linn. Soc. 23: 269–286.
Jones, C. G., J. H. Lawton, and M. Shachak. 1994. Organisms as ecosystem engi-
neers. Oikos 69: 373–386.
Jones, C. G., J. H. Lawton, and M. Shachak. 1997. Positive and negative effects of
organisms as physical ecosystem engineers. Ecology 78: 1946–1957.
Kassen, R. 2009. Toward a general theory of adaptive radiation: Insights from
microbial experimental evolution. Yr. Evol. Biol. 1168: 3–22.
Kato, M., A. Takimura, and A. Kawakita. 2003. An obligate pollination mutual-
ism and reciprocal diversification in the tree genus Glochidion (Euphorbiaceae).
Proc. Natl. Acad. Sci. USA 100: 5264–5267.
Kozak, K. H., R. A. Blaine, and A. Larson. 2006. Gene lineages and eastern North
American palaeodrainage basins: Phylogeography and speciation in salaman-
ders of the Eurycea bislineata species complex. Mol. Ecol. 15: 191–207.
Kubatko, L. S. 2009. Identifying hybridization events in the presence of coales-
cence via model selection. Syst. Biol. 58: 478–488.
Kubo, T., and Y. Iwasa. 1995. Inferring the rates of branching and extinction from
molecular phylogenies. Evolution 49: 694–704.
Langerhans, R. B. 2006. Evolutionary consequences of predation: Avoidance,
escape, reproduction, and diversification. In A. M. T. Elewa (ed.), Predation in
Organisms: A Distinct Phenomenon, pp. 177–220. Springer-Verlag, Heidelberg.
Langerhans, R. B., M. E. Gifford, and E. O. Joseph. 2007. Ecological speciation in
Gambusia fishes. Evolution 61: 2056–2074.
Larson, A. and J. B. Losos. 1996. Phylogenetic systematics of adaptation. In M. R.
Rose and G. V. Lauder (eds.), Adaptation. pp. 187–220. Academic Press, San
Diego.
Leigh, E. G. Jr., A. Hladik, C. M. Hladik, and 1 other. 2007. The biogeography of
large islands, or how does the size of the ecological theater affect the evolu-
tionary play? Revue E’cole (Terre Vie) 62: 105–168.
Lenski, R. E. and M. Travisano. 1994. Dynamics of adaptation and diversification:
A 10,000-generation experiment with bacterial populations. Proc. Natl. Acad.
Sci. USA 91: 6808–6814.
Levinton, J. 1988. Genetics, Paleontology, and Macroevolution. Cambridge University
Press, Cambridge, UK.
Liem, K. F. 1974. Evolutionary strategies and morphological innovations: Cichlid
pharyngeal jaws. Syst. Zool. 22: 425–441.
Loeuille, N. and M. Loreau. 2005. Evolutionary emergence of size-structured food
webs. Proc. Natl. Acad. Sci. USA 102: 5761–5766.
Loeuille, N. and M. Loreau. 2006. Evolution of body size in food webs: Does the
energetic equivalence rule hold? Ecol. Lett. 9: 171–178.
Losos, J. B. 2009. Lizards in an Evolutionary Tree: Ecology and Adaptive Radiation of
Anoles. University of California Press, Berkeley.
Losos, J. B. and D. B. Miles. 2002. Testing the hypothesis that a clade has adap-
tively radiated: Iguanid lizard clades as a case study. Am. Nat. 160: 147–157.
Losos, J. B. and R. E. Ricklefs. 2009. Adaptation and diversification on islands.
Nature 457: 830–836.
Losos, J. B., P. L. N. Mouton, R. Bickel, and 2 others. 2002. The effect of body
armature on escape behaviour in cordylid lizards. Anim. Behav. 64: 313–321.
Lovette, I. J., E. Bermingham, and R. E. Ricklefs. 2002. Clade-specific morphologi-
cal diversification and adaptive radiation in Hawaiian songbirds. Proc. Roy.
Soc. Lond. B 269: 37–42.
Mabuchi, K., M. Miya, Y. Azuma, and 1 other. 2007. Independent evolution of the
specialized pharyngeal jaw apparatus in cichlid and labrid fishes. BMC Evol.
Biol. 7: 10.
MacArthur, R. H. 1972. Geographical Ecology: Patterns in the Distribution of Species.
Princeton University Press, Princeton.
MacArthur, R. and R. Levins. 1967. The limiting similarity convergence, and
divergence of coexisting species. Am. Nat. 101: 377–385.
MacFadden, B. J. 1992. Fossil Horses: Systematics, Paleobiology, and Evolution of the
Family Equidae. Cambridge University Press, Cambridge, UK.
Mahler, D. L., L. J. Revell, R. E. Glor, and 1 other. 2010. Ecological opportunity
and the rate of morphological evolution in the diversification of Greater Antil-
lean anoles. Evolution, in press.
Mallet, J. 2009. Rapid speciation, hybridization and adaptive radiation in the Heli-
conius melpomene group. In R. Butlin, J. Bridle, and D. Schluter (eds.), Speciation
and Patterns of Diversity, pp. 177–194. Cambridge University Press, Cambridge,
UK.
McGowan, A. J. 2004. Ammonoid taxonomic and morphologic recovery patterns
after the Permian-Triassic. Geology 32: 665–668.
McPeek, M. A. 2008. The ecological dynamics of clade diversification and com-
munity assembly. Am. Nat. 172: E270–E284.
McPeek, M. A., A. K. Schrot, and J. M. Brown. 1996. Adaptation to predators in a
new community: Swimming performance and predator avoidance in damsel-
flies. Ecology 77: 617–629.
Meyer, A., T. D. Kocher, P. Basasibwaki, and 1 other. 1990. Monophyletic origin
of Lake Victoria cichlid fishes suggested by mitochondrial-DNA sequences.
Nature 347: 550–553.
Meyer, J. R. and R. Kassen. 2007. The effects of competition and predation on
diversification in a model adaptive radiation. Nature 446: 432–435.
Miller, A. H. 1949. Some ecologic and morphologic considerations in the evolu-
tion of higher taxonomic categories. In E. Mayr and E. Schüz (eds.), Ornitholo-
gie als Biologische Wissenschaft, pp. 84–88. Carl Winter, Heidelberg.
Mitter, C., B. Farrell, and B. Wiegmann. 1988. The phylogenetic study of adap-
tive zones: Has phytophagy promoted insect diversification? Am. Nat. 132:
107–128.
Monasterio, M. and L. Sarmiento. 1991. Adaptive radiation of Espeletia in the cold
Andean tropics. Trends Ecol. Evol. 6: 387–391.
Nee, S. 2001. Inferring speciation rates from phylogenies. Evolution 55: 661–668.
Nee, S. 2006. Birth-death models in macroevolution. Ann. Rev. Ecol. Evol. Syst. 37:
1–17.
Nee, S., E. C. Holmes, R. M. May, and 1 other. 1994. Extinction rates can be esti-
mated from molecular phylogenies. Phil. Trans. Roy. Soc. Lond. B 344: 77–82.
Nee, S., A. Ø. Mooers, and P. H. Harvey. 1992. Tempo and mode of evolution
revealed from molecular phylogenies. Proc. Natl. Acad. Sci. USA 89: 8322–8326.
Nosil, P. and B. J. Crespi. 2006. Experimental evidence that predation promotes
divergence in adaptive radiation. Proc. Natl. Acad. Sci. USA 103: 9090–9095.
418  Chapter 15 Losos • Mahler
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated in
any form without express written permission from the publisher.
Adaptive Radiation: The Interaction of Ecological Opportunity, Adaptation, and Speciation  419
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated
in any form without express written permission from the publisher.
Olson, M. E. and A. Arroyo-Santos. 2009. Thinking in continua: Beyond the
“adaptive radiation” metaphor. BioEssays 31: 1337–1346.
O’Meara, B. C., C. Ane, M. J. Sanderson, and 1 other. 2006. Testing for different
rates of continuous trait evolution using likelihood. Evolution 60: 922–933.
Parent, C. E. and B. J. Crespi. 2009. Ecological opportunity in adaptive radiation
of Galápagos endemic land snails. Am. Nat. 174: 898–905.
Pfenning, K. S. and D. W. Pfennig. 2009. Character displacement: Ecological and
reproductive responses to a common evolutionary problem. Quart. Rev. Biol.
84: 253–276.
Phillimore, A. B. and T. D. Price. 2008. Density-dependent cladogenesis in birds.
PLoS Biol. 6: 483–489.
Pinto, G., D. L. Mahler, L. J. Harmon, and 1 other. 2008. Testing the island effect
in adaptive radiation: Rates and patterns of morphological diversification in
Caribbean and mainland Anolis lizards. Proc. Roy. Soc. Lond. B 275: 2749–2757.
Pratt, H. D. 2005. The Hawaiian Honeycreepers: Drepanididae. Oxford University
Press, Oxford.
Price, T. 1997. Correlated evolution and independent contrasts. Phil. Trans. Roy.
Soc. B 352: 519–529.
Price, T. D. 2008. Speciation in Birds. Roberts and Company, Greenwood Village,
CO.
Quental, T. B. and C. R. Marshall. 2009. Extinction during evolutionary radia-
tions: Reconciling the fossil record with molecular phylogenies. Evolution 63:
3158–3167.
Rabosky, D. L. 2009. Ecological limits on clade diversification in higher taxa. Am.
Nat. 173: 662–674.
Rabosky, D. L. and I. J. Lovette. 2008. Density-dependent diversification in North
American wood warblers. Proc. Roy. Soc. Lond. B 275: 2363–2371.
Rainey, P. B. and M. Travisano. 1998. Adaptive radiation in a heterogeneous envi-
ronment. Nature 394: 69–72.
Raup, D. M., S. J. Gould, T. J. M. Schopf, and 1 other. 1973. Stochastic models of
phylogeny and the evolution of diversity. J. Geol. 81: 525–542.
Rauscher, J. T. 2002. Molecular phylogenetics of the Espeletia complex (Asterace-
ae): Evidence from nrDNA ITS sequences on the closest relatives of an Andean
adaptive radiation. Am. J. Bot. 89: 1074–1084.
Ree, R. H. 2005. Detecting the historical signature of key innovations using sto-
chastic models of character evolution and cladogenesis. Evolution 59: 257–265.
Revell, L. J., L. J. Harmon, and R. E. Glor. 2005. Underparameterized model of
sequence evolution leads to bias in the estimation of diversification rates from
molecular phylogenies. Syst. Biol. 54: 973–983.
Rice, W. R. and E. E. Hostert. 1993. Laboratory experiments on speciation: What
have we learned in forty years? Evolution 47: 1637–1653.
Ricklefs, R. E. 2009. Aspect diversity in moths revisited. Am. Nat. 173: 411–416.
Rieseberg, L. H., M. A. Archer, and R. K. Wayne. 1999. Transgressive segregation,
adaptation, and speciation. Heredity 83: 363–372.
Roderick, G. K. and D. M. Percy. 2008. Insect-plant interactions, diversification,
and coevolution: Insights from remote oceanic islands. In K. J. Tilmon (ed.),
Specialization, Speciation, and Radiation: The Evolutionary Biology of Herbivorous
Insects, pp. 151–161. University of California Press, Berkeley.
Rosenzweig, M. L. and R. D. McCord. 1991. Incumbent replacement: Evidence for
long-term evolutionary progress. Paleobiology 17: 202–213.
Roy, K. and M. Foote. 1997. Morphological approaches to measuring biodiversity.
Trends Ecol. Evol. 12: 277–281.
Rundell, R. J. and T. D. Price. 2009. Adaptive radiation, nonadaptive radia-
tion, ecological speciation and nonecological speciation. Trends Ecol. Evol. 24:
394–399.
Rundle, H. D. and D. Schluter. 2004. Natural selection and ecological speciation
in sticklebacks. In U. Dieckmann, M. Doebeli, J. A. J. Metz, and D. Tautz (eds.),
Adaptive Speciation, pp. 192–209. Cambridge University Press, Cambridge, UK.
Russell, A. P. 1979. Parallelism and integrated design in the foot structure of gek-
konine and diplodactyline geckos. Copeia 1979: 1–21.
Rutherford, S. L. and S. Lindquist. 1998. Hsp90 as a capacitor for morphological
evolution. Nature 396: 336–342.
Sanderson, M. J. and M. J. Donoghue. 1996. Reconstructing shifts in diversifica-
tion rates on phylogenetic trees. Trends Ecol. Evol. 11: 15–20.
Savolainen, R. and K. Vepsalainen. 2003. Sympatric speciation through intraspe-
cific social parasitism. Proc. Natl. Acad. Sci. USA 100: 7169–7174.
Sax, D. F., J. J. Stachowicz, J. H. Brown, and 9 others. 2007. Ecological and evolu-
tionary insights from species invasions. Trends Ecol. Evol. 22: 465–471.
Schliewen, U. K., D. Tautz, and S. Pääbo. 1994. Sympatric speciation suggested by
monophyly of crater lake cichlids. Nature 368: 629–632.
Schluter, D. 1988. Character displacement and the adaptive divergence of finches
on islands and continents. Am. Nat. 131: 799–824.
Schluter, D. 1994. Experimental evidence that competition promotes divergence
in adaptive radiation. Science 266: 798–801.
Schluter, D. 1995. Adaptive radiation in sticklebacks: Trade-offs in feeding perfor-
mance and growth. Ecology 76: 82–90.
Schluter, D. 2000. The Ecology of Adaptive Radiation. Oxford University Press,
Oxford.
Schluter, D. 2003. Frequency dependent natural selection during character dis-
placement in sticklebacks. Evolution 57: 1142–1150.
Schluter, D. and P. R. Grant. 1984. Determinants of morphological patterns in
communities of Darwin’s finches. Am. Nat. 123: 175–196.
Seehausen, O. 2004. Hybridization and adaptive radiation. Trends Ecol. Evol. 19:
198–207.
Seehausen, O. 2006. African cichlid fish: A model system in adaptive radiation
research. Proc. Roy. Soc. Lond. B 273: 1987–1998.
Sepkoski, J. J. 1978. A kinetic model of Phanerozoic taxonomic diversity. I. Analy-
sis of marine orders. Paleobiology 4: 223–251.
Sepkoski, J. J., F. K. McKinney, and S. Lidgard. 2000. Competitive displacement
among post-Paleozoic cyclostome and cheilostome bryozoans. Paleobiology 26:
7–18.
Sibley, C. G. and J. E. Ahlquist. 1990. Phylogeny and Classification of Birds. Yale Uni-
versity Press, New Haven.
Simberloff, D. 1981. Community effects of introduced species. In T. H. Nitecki
(ed.), Biotic Rises in Ecological and Evolutionary Time, pp. 53–81. Academic Press,
New York.
Simpson, G. G. 1953. The Major Features of Evolution. Columbia University Press,
New York.
Slatkin, M. 1980. Ecological character displacement. Ecology 61: 163–177.
Slowinski, J. B. and C. Guyer. 1989. Testing null models in questions of evolution-
ary success. Syst. Zool. 38: 189–191.
420  Chapter 15 Losos • Mahler
© 2010 Sinauer Associates, Inc. This material cannot be copied, reproduced, manufactured or disseminated in
any form without express written permission from the publisher.
Spencer, C. C., M. Bertrand, M. Travisano, and 1 other. 2007. Adaptive diversi-
fication in genes that regulate resource use in Escherichia coli. PLoS Genet. 3:
0083–0088.
Stanley, G. D. 1981. Early history of scleractinian corals and its geological conse-
quences. Geology 9: 507–511.
Stiassny, M. L. J. and J. S. Jensen. 1987. Labroid intrarelationships revisited: Mor-
phological complexity, key innovations, and the study of comparative diver-
sity. Bull. Mus. Comp. Zool. 151: 269–319.
Stoks, R., M. A. McPeek, and J. L. Mitchell. 2003. Evolution of prey behavior in
response to changes in predation regime: Damselflies in fish and dragonfly
lakes. Evolution 57: 574–585.
Tokeshi, M. 1999. Species Coexistence: Ecological and Evolutionary Perspectives. Black-
well, Oxford.
Tyerman, J. G., M. Bertrand, C. C. Spencer, and 1 other. 2008. Experimental dem-
onstration of ecological character displacement. BMC Evol. Biol. 8: 34.
Vamosi, S. M. 2005. On the role of enemies in divergence and diversification of
prey: A review and synthesis. Can. J. Zool. 83: 894–910.
Vermeij, G. J. 1974. Adaptation, versatility, and evolution. Syst. Zool. 22: 466–477.
Vermeij, G. J. 1987. Evolution and Escalation: An Ecological History of Life. Princeton
University Press, Princeton.
Wagner, G. P. and L. Altenberg. 1996. Perspective: Complex adaptations and the
evolution of evolvability. Evolution 50: 967–976.
Webb, J. K. and R. Shine. 1994. Feeding habits and reproductive biology of Aus-
tralian pygopodid lizards of the genus Aprasia. Copeia 1994: 390–398.
Weiblen, G. D. and G. L. Bush. 2002. Speciation in fig pollinators and parasites.
Mol. Ecol. 11: 1573–1578.
Whittaker, R. H. 1977. Evolution of species diversity in land communities. Evol.
Biol. 10: 1–67.
Wilson, E. O. 1992. The Diversity of Life. Harvard University Press, Cambridge,
MA.
Winkler, I. S. and C. Mitter. 2008. The phylogenetic dimension of insect/plant
interactions: A summary of recent evidence. In K. J. Tilmon (ed.), Specializa-
tion, Speciation, and Radiation: The Evolutionary Biology of Herbivorous Insects, pp.
240–263. University of California Press, Berkeley.
Wright, S. 1932. The roles of mutation, inbreeding, crossbreeding, and selection in
evolution. Proc. 6th Intl. Cong. Genet. 1: 356–366.
Yamagishi, S., M. Honda, K. Eguchi, and 1 other. 2001. Extreme endemic radia-
tion of the Malagasy vangas (Aves: Passeriformes). J. Mol. Evol. 53: 39–46.
... Key innovations are defined as traits that allow organisms to interact with their environment in novel ways (Miller, 1949;Liem, 1973;Losos & Mahler, 2010) and thereby enable them to enter into a previously inaccessible ecological state (Miller et al., 2022). The acquisition of such novel traits can promote differential evolutionary success among lineages (Larouche et al., 2020) through the creation of new adaptive zones, access to previously unattainable resources, and increased fitness (Losos & Mahler, 2010;Fernández-Mazuecos et al., 2019). ...
... Key innovations are defined as traits that allow organisms to interact with their environment in novel ways (Miller, 1949;Liem, 1973;Losos & Mahler, 2010) and thereby enable them to enter into a previously inaccessible ecological state (Miller et al., 2022). The acquisition of such novel traits can promote differential evolutionary success among lineages (Larouche et al., 2020) through the creation of new adaptive zones, access to previously unattainable resources, and increased fitness (Losos & Mahler, 2010;Fernández-Mazuecos et al., 2019). The concept of a key innovation has been rather controversial in recent years with multiple studies incorporating increased lineage diversification as a core aspect of its definition. ...
... Individual genotypes may create different phenotypes when exposed to varying environmental conditions. This is termed phenotypic plasticity (Garland & Kelly, 2006;Losos & Mahler, 2010;Pigliucci, 2001;Pigliucci et al., 2006). Understanding the underlying reasons for phenotypic plasticity and the resulting morphological disparity is one of the key topics of evolutionary biological and paleobiological research (Dewitt & Scheiner, 2004;Fusco & Minelli, 2010;Gilbert & Epel, 2009;Greene, 1999;Jablonka & Lamb, 2005;Lewontin, 2000;Minelli & Fusco, 2008;Schlichting & Pigliucci, 1998;West-Eberhard, 2003, 2005. ...
Article
Full-text available
Understanding the underlying reasons for phenotypic plasticity and resulting morphological disparity is one of the key topics of evolutionary research. The phenotypic plasticity of extant and fossil melanopsids has been widely documented. Yet millennial‐resolution, well‐dated records from small aquatic habitats harboring endemics are scarce. The thermal spring‐fed Lake Pețea is an ice age refugia harboring a unique endemic warm‐water fauna. Subfossil melanopsids display incredible morphological variability from smooth to keeled, elongated to ribbed, shouldered forms. Numerous morphotypes have been considered as individual taxa with a fluent succession from the smooth elongated to the ribbed, shouldered types. This study presents an extensive morphometric analysis of subfossil melanopsids (ca. 3500 specimens) derived from stratified samples with an independent chronology. The aim was to separate morphotypes for investigations of temporal morphological disparity. Our results challenge the widely accepted hypothesis that proposes the evolution of shouldered, compressed, ribbed shells through a two‐step process from smooth elongated spindle‐shaped shells. Instead, it suggests that the subfossil shells belong to two distinct taxa present throughout the available stratigraphic data. The main components of shape variation, shape globularity, and shell coiling seem allometry‐related. Ribs, striation, and keels appear randomly. High‐spired spindle‐shaped forms were considered to represent specimens of Microcolpia daudebartii hazayi . Bulkier low‐spired and shouldered specimens represent phenotypes of Mi. parreyssii parreyssii . The collective and random distribution of morphotypes from the early stages of the lake's history also refutes the idea of a continuous transformation of the elongated forms into compressed, shouldered ones. Rather points to multiple events and environmental stimuli triggering development. Melanopsids appear in Late Glacial horizons, with Theodoxus prevostianus preferring temperatures above 23°C which may indicate the subordinate presence of hot water microhabitats in cooler waters.
... Simpson (1944Simpson ( , 1953 proposed a modification of this idea arguing that to better understand the role of divergent natural selection on survival and reproductive success in a particular environment we should account for phenotype differences in the adaptive landscape (Schluter 2000). Since then, the study of the adaptive landscapes has been focused on determining how morphological adaptation and ecological performance relate to ecological di-vergence (Grant and Grant 2002;Losos and Mahler 2010;Mahler et al. 2010;Brawand et al. 2014). Adaptive landscape topologies may represent the effects of stabilizing selection as a unimodal distribution in trait space, while disruptive selection should be related to multimodal distributions in trait space. ...
Article
Neotropical ecosystems are renowned for numerous examples of adaptive radiation in both plants and animals resulting in high levels of biodiversity and endemism. However, we still lack a comprehensive review of the abiotic and biotic factors that contribute to these adaptive radiations. To fill this gap, we delve into the geological history of the region, including the role of tectonic events such as the Andean uplift, the formation of the Isthmus of Panama, and the emergence of the Guiana and Brazilian Shields. We also explore the role of ecological op- portunities created by the emergence of new habitats, as well as the role of key innovations, such as novel feeding strategies or reproductive mechanisms. We discuss different examples of adaptive radiation, including classic ones like Darwin’s finches and Anolis lizards, and more recent ones like bromeliads and lupines. Finally, we propose new examples of adaptive radiations mediated by ecological interactions in their geological context. By doing so, we provide insights into the complex interplay of factors that contributed to the remarkable diversity of life in the Neotropics and highlight the importance of this region in understanding the origins of biodiversity.
... Puede ocurrir que exista evolución adaptativa en la que unas especies suceden o sustituyen a otras, sin que exista diversificación (evolución filética). En contraste, cuando existe diversificación de un ancestro hacia varias especies nuevas (cladogénsis), ésta puede ser el resultado de diferentes procesos, en algunos casos, ser resul-tado meramente del aislamiento geográfico o por otra parte, mediante la radiación adaptativa, entre otros Rundell y Price, 2009, Losos y Mahler, 2010. ...
... We purposefully refrain from characterizing this rapid diversification as an adaptive radiation. Adaptive radiations involve a pairing of species diversification and adaptive diversification (Schluter, 2000;Losos and Miles, 2002;Losos and Mahler, 2010), and the ecological diversity of the genus remains too poorly known to make this claim. Support for an adaptive radiation awaits data on phenotypic disparity and ecological divergence in this remarkably morphologically uniform clade (Ibáñez and Juste, 2019). ...
Article
African-Malagasy species of the bat genus Miniopterus are notable both for the dramatic increase in the number of newly recognized species over the last 15 years, as well as for the profusion of new taxa from Madagascar and the neighboring Comoros. Since 2007, seven new Malagasy Miniopterus species have been described compared to only two new species since 1936 from the Afrotropics. The conservative morphology of Miniopterus and limited geographic sampling in continental Africa have undoubtedly contributed to the deficit of continental species. In addition to uncertainty over species limits, phylogenetic relationships of Miniopterus remain mostly unresolved, particularly at deeper backbone nodes. Previous phylogenetic studies were based on limited taxon sampling and/ or limited genetic sampling involving no more than five loci. Here, we conduct the first phylogenomic study of the Afrotropical Miniopteridae by analyzing up to 3772 genome-wide ultraconserved elements (UCEs) from historic and modern samples of 70 individuals from 25 Miniopterus species/lineages. We analyze multiple datasets of varying degrees of completeness (70, 90, and 100 percent complete) using partitioned concatenated maximum likelihood and multispecies coalescent methods. Our well-supported, species-level phylogenies resolved most (6/8 or 7/8) backbone nodes and strongly support for the first time the monophyly of the Malagasy radiation. We inferred the crown age of African Miniopteridae in the late Miocene (10.4 Ma), while the main lineages of Miniopterus appear to have contemporaneously diversified in two sister radiations in the Afrotropics and Madagascar. Species-level divergence of 23 of 25 African + Malagasy Miniopterus were estimated to have 95 % HPDs that overlap with the late Miocene (5.3-10.4 Ma). We present ancestral range estimates that unambiguously support a continental African radiation that originated in the Zambezian and Somalian/Ethiopian biogeographic regions, but we cannot rule out back colonization of Africa from Madagascar. The phylogeny indicates genetic support for up to seven new species.
... In summary, it is fitting that the evolutionary affinities of the largest gecko-Gigarcanum delcourti-lie with a New Caledonian clade notable for its extensive morphological and ecological diversity. Rapid evolution of phenotypic disparity is a hallmark of adaptive radiations 46 . The rapid body size evolution in the New Caledonia clade outstrips that of both Australian and New Zealand diplodactylids ( 37,38 and this study), and we document a larger frequency of transitions in tested phenotypic traits than in other diplodactylids following the mid-Cenozoic colonization of New Caledonia. ...
Article
Full-text available
Hoplodactylus delcourti is a presumably extinct species of diplodactylid gecko known only from a single specimen of unknown provenance. It is by far the largest known gekkotan, approximately 50% longer than the next largest-known species. It has been considered a member of the New Zealand endemic genus Hoplodactylus based on external morphological features including shared toe pad structure. We obtained DNA from a bone sample of the only known specimen to generate high-throughput sequence data suitable for phylogenetic analysis of its evolutionary history. Complementary sequence data were obtained from a broad sample of diplodactylid geckos. Our results indicate that the species is not most closely related to extant Hoplodactylus or any other New Zealand gecko. Instead, it is a member of a clade whose living species are endemic to New Caledonia. Phylogenetic comparative analyses indicate that the New Caledonian diplodactylid clade has evolved significantly more disparate body sizes than either the Australian or New Zealand clades. Toe pad structure has changed repeatedly across diplodactylids, including multiple times in the New Caledonia clade, partially explaining the convergence in form between H. delcourti and New Zealand Hoplodactylus. Based on the phylogenetic results, we place H. delcourti in a new genus.
Article
En la última década, el campo del diseño de juegos ha experimentado diversas iteraciones con una amplia gama de áreas.
Preprint
Full-text available
The Canidae are an ecologically important group of dog-like carnivores that arose in North America and spread across the planet around 10 million years ago. The current distribution patterns of species, coupled with their phylogenetic structure, suggest that Canidae diversification may have occurred at varying rates across different biogeographic areas. However, such extant-only analyses undervalued the rich fossil history of the group because of a limitation in methods development. Current State-dependent Speciation and Extinction (SSE) models are (i) often parameter-rich which hinders reliable application to relatively small clades such as the Caninae (the only extant subclade of the Canidae consisting of 36 extant species); and (ii) often assume as possible states only the states that extant species present. Here we extend the SSE method SecSSE to apply to phylogenies with extinct species as well (111 Caninae species) and compare the results to those of analyses with the extant-species-only phylogeny. The results on the extant-species tree suggest that distinct diversification patterns are related to geographic areas, but the results on the complete tree do not support this conclusion. Furthermore, our extant-species analysis yielded an unrealistically low estimate of the extinction rate. These contrasting findings suggest that information from extinct species is different from information from extant species. A possible explanation for our results is that extinct species may have characteristics (causing their extinction), which may be different from the characteristics of extant species that caused them to be extant. Hence, we conclude that differences in biogeographic areas probably did not contribute much to the variation in diversification rates in Caninae.
Article
Full-text available
Geographical ranges and physiological tolerances of species are correlated, and it can be expected that widespread species encounter higher climatic variation across their distributions than restricted species. Widespread species should consequently be more tolerant to extreme or variable weather conditions, and may have the ability to better conserve their current geographical ranges under future climate change scenarios. We tested this hypothesis by studying the relationship between the climatic variation experienced by restricted and widespread Anolis lizards from different ecomorphs and regions of Cuba and the distributional shifts induced by climate change. We selected seven bioclimatic variables from WorldClim to characterize the realized climatic niche of 12 Cuban anoles, where the coefficients of variation of each variable were taken as a measure of climatic variation. We used niche modeling to predict changes in suitable habitats under future climatic scenarios. We found that species from Eastern Cuba occupy areas with the highest climatic variation, likely related to the topography of the region. Crown giant anoles experienced habitats with lower climatic variation in comparison with species from other ecomorphs, which together with their tree canopy habitat and large body size may represent a disadvantage to face changing climates. All species will experience a severe decrease in their habitat suitability, with the Western species being predicted to lose a higher proportion of suitable habitat. Combining niche modeling with physiological data would better predict the effects of climate change on Cuban lizards and might allow taking management actions for species and habitats to mitigate the possible negative impacts of this phenomenon.
Chapter
Full-text available
Bringing together the viewpoints of leading ecologists concerned with the processes that generate patterns of diversity, and evolutionary biologists who focus on mechanisms of speciation, this book opens up discussion in order to broaden understanding of how speciation affects patterns of biological diversity, especially the uneven distribution of diversity across time, space and taxa studied by macroecologists. The contributors discuss questions such as: Are species equivalent units, providing meaningful measures of diversity? To what extent do mechanisms of speciation affect the functional nature and distribution of species diversity? How can speciation rates be measured using molecular phylogenies or data from the fossil record? What are the factors that explain variation in rates? Written for graduate students and academic researchers, the book promotes a more complete understanding of the interaction between mechanisms and rates of speciation and these patterns in biological diversity.
Article
Molecular techniques provide ancestral phylogenies of extant taxa with estimated branching times. Here we studied the pattern of ancestral phylogeny of extant taxa produced by branching (or cladogenesis) and extinction of taxa, assuming branching processes with time-dependent rates. (1) If the branching rate b and extinction rate c are constant, the semilog plot of the number of ancestral lineages over time is not a straight line but is curvilinear, with increasing slope toward the end, implying that ancestral phylogeny shows apparent increase in the branching rate near the present. The estimate of b and c based on nonlinear fitting is examined by computer simulation. The estimate of branching rate can be usable for a large phylogeny if b is greater than c, but the estimate of extinction rate c is unreliable because of large bias and variance. (2) Gradual decrease in the slope of the semilog plot of the number of ancestral lineages over time, as was observed in a phylogeny of bird families based on DNA hybridization data, can be explained equally well by either the decreasing branching rate or the increasing extinction rate. Infinitely many pairs of branching and extinction rates as functions of time can produce the same ancestral phylogeny. (3) An explosive branching event in the past would appear as a quick increase in the number of ancestral lineages. In contrast, mass extinction occurring in a brief period, if not accompanied by an increase in branching rate, does not produce any rapid change in the number of ancestral lineages at the time. (4) The condition in which the number of ancestral lineages of extant species changes in parallel with the actual number of species in the past is derived.
Article
A simple equilibrial model for the growth and maintenance of Phanerozoic global marine taxonomie diversity can be constructed from considerations of the behavior of origination and extinction rates with respect to diversity. An initial postulate that total rate of diversification is proportional to number of taxa extant leads to an exponential model for early phases of diversification. This model appears to describe adequately the “explosive” diversification of known metazoan orders across the Precambrian-Cambrian Boundary, suggesting that no special event, other than the initial appearance of Metazoa, is necessary to explain this phenomenon. As numbers of taxa increase, the rate of diversification should become “diversity dependent.” Ecological factors should cause the per taxon rate of origination to decline and the per taxon rate of extinction to increase. If these relationships are modeled as simple linear functions, a logistic description of the behavior of taxonomie diversity through time results. This model appears remarkably consistent with the known pattern of Phanerozoic marine ordinal diversity as a whole. Analysis of observed rates of ordinal origination also indicates these are to a large extent diversity dependent; however, diversity dependence is not immediately evident in rates of ordinal extinction. Possible explanations for this pattern are derived from considerations of the size of higher taxa and from simulations of their diversification. These suggest that both the standing diversity and the pattern of origination of orders may adequately reflect the behavior of species diversity through time; however, correspondence between rates of ordinal and species extinction may deteriorate with progressive loss of information resulting from incomplete sampling of the fossil record.