You are on page 1of 9

REVIEWS

THE GENETIC THEORY OF


ADAPTATION: A BRIEF HISTORY
H. Allen Orr
Abstract | Theoretical studies of adaptation have exploded over the past decade. This work has
been inspired by recent, surprising findings in the experimental study of adaptation. For example,
morphological evolution sometimes involves a modest number of genetic changes, with some
individual changes having a large effect on the phenotype or fitness. Here I survey the history of
adaptation theory, focusing on the rise and fall of various views over the past century and the
reasons for the slow development of a mature theory of adaptation. I also discuss the challenges
that face contemporary theories of adaptation.

STANDING GENETIC VARIATION


Adaptation is not natural selection. As Ronald A. Fisher1 These questions have two things in common. First,
Allelic variation that is currently emphasized in 1930, adaptation is characterized by the they are important; indeed they are among the simplest
segregating within a population; movement of a population towards a phenotype that and most obvious questions that can be asked about the
as opposed to alleles that appear best fits the present environment. The result is an often changes underlying Darwinian evolution. Second, they
by new mutation events.
astonishingly precise match between an organism and are not answered by traditional evolutionary theory. In
FITNESS the world in which it lives. But as Fisher also empha- fact, the deeper problem is that they are not even asked
A quantity that is proportional sized, the steady increase in the frequency of an allele by traditional evolutionary theory.
to the mean number of viable, under selection need not invariably result in “the adap- Despite this near theoretical void, experimental evo-
fertile progeny produced by a tive modification of specific forms”. Competition lutionary geneticists have begun to address several of the
genotype.
among ‘selfish genes’, for instance, can change a popula- above questions. These studies — and the sometimes
tion’s sex ratio but we have no reason to expect any surprising answers that they have provided — have
improved fit between an organism and its environment. reinvigorated attempts to elaborate a mathematical
To an evolutionary geneticist, there is another, sim- theory of adaptation. Here I survey these attempts. My
pler way of distinguishing between selection and adap- approach is historical, considering the rise and fall of
tation: we know a lot about the former but little about various views on the genetic basis of adaptation, the rea-
the latter. Many important questions about the genetic sons a mature theory has been slow to develop and the
basis of adaptation remain unanswered. Do most prospects and problems facing current theory.
adaptations involve new mutations or STANDING GENETIC As we will see, recent models of adaptation seem to
VARIATION? Do most adaptations involve single genes of successfully explain certain qualitative patterns that
large phenotypic effect (‘major’ genes)? If so, can we say characterize morphological evolution in animals and
anything about the expected effect of this major gene? plants, as well as patterns that characterize fitness
Can we describe the distribution of phenotypic effects increase in microbes. Although this success is encourag-
among the mutations that are substituted during a typi- ing — and long overdue — future work must deter-
cal bout of adaptation? How does FITNESS change as a mine whether present theory can explain the genetic
Department of Biology, population approaches an optimum? For example, do data quantitatively.
University of Rochester, populations evolve quickly at first and then more
Rochester, New York 14627, slowly? Because random mutations are more likely to be Micromutationism: its rise and fall
USA.
e-mail: aorr@mail. deleterious in complex organisms than in simple organ- The theory. The earliest view of the genetic basis of adap-
rochester.edu isms, do complex organisms adapt more slowly than tation was pre-Mendelian. This view, which emphasized
doi:10.1038/nrg1523 simple ones? the extreme gradualness of phenotypic evolution, began

NATURE REVIEWS | GENETICS VOLUME 6 | FEBRUARY 2005 | 1 1 9


© 2005 Nature Publishing Group
REVIEWS

Box 1 | Experimental studies of adaptive evolution genetics7–11. William Bateson, for example — the
brashest and most articulate of the Mendelians —
Genetic analyses of phenotypic differences between species or populations often argued that the popularity of micromutationism merely
reveal major quantitative trait loci (QTL). A recent burst of work, for instance, has reflected the light demands it placed on naturalists.“By
focused on the genetic basis of armour-plate reduction and pelvic reduction in suggesting that the steps through which an adaptive
postglacial-lake forms of the threespine stickleback, Gasterosteus aculeatus.
mechanism arises are indefinite and insensible, all fur-
Although marine forms of this species are heavily armoured, lake populations
ther trouble is spared. While it could be said that
(which are recently derived from the marine form) are not. Instead, lake forms
have repeatedly and independently evolved reduced armour-plate and pelvic species arise by an insensible and imperceptible process
structures, leaving little doubt that these morphological changes are adaptive. of variation, there was clearly no use in tiring ourselves
QTL and developmental genetic studies indicate that this adaptive evolution by trying to perceive that process. This labor-saving
sometimes involves the same genes97–99. A strong candidate gene, Pitx1, has been counsel found great favor”12.
identified that has a major effect on pelvic reduction; the Pitx1 protein sequence is Despite such opposition — and despite the larger
identical between marine and lake forms, indicating that this case of morphological victory of Mendelism — the micromutationist view
change involved regulatory evolution99. won out among evolutionists by about 1930. To a con-
Similarly, Sucena and Stern100 showed that a qualitative difference in larval morphology siderable extent, this victory reflected the efforts of
(the presence or absence of a lawn of fine hairs) between two species of Drosophila Fisher, a founding father of population genetics and a
(Drosophila simulans and Drosophila sechellia) is due to a single gene. Through tireless champion of Darwinian gradualism. Fisher suc-
DEFICIENCY MAPPING and COMPLEMENTATION TESTS, they showed that the relevant gene is
cessfully fused micromutationism with Mendelism13,
ovo/shaven-baby, which resides on the X chromosome.
producing a mathematical framework known as the
Evidence for major genes is not limited to animals. Work in the 1990s showed,
infinitesimal model. Response to selection, he argued,
for example, that the evolution of maize from its wild ancestor teosinte involved
the substitution of a QTL that had large effects on morphology (for example, tb1, could be analysed through a kind of statistical mechan-
which affects lateral-branching pattern101,102 and tga1, which affects kernel ics: instead of following the effects of individual genes
architecture101,103). Although the evolution of maize involved human intervention, on a selected phenotype, one could calculate their aggre-
QTL studies of two wild species of Mimulus yielded qualitatively similar results. gate effects under the assumption that a character is
Mimulus lewisii is primarily bee-pollinated, whereas Mimulus cardinalis is underlaid by an infinite number of genes, each unlinked
primarily pollinated by hummingbirds; not surprisingly, the flowers of the two to all others, each having no EPISTATIC INTERACTIONS with
species differ markedly. In a QTL analysis of a suite of floral characters that the others, and each having an infinitesimally small
distinguish these species, Bradshaw et al.104,105 found that one to six QTL underlie effect on the character3,14. Although modern evolutionists
each of the 12 floral traits analysed. For 9 of these 12 traits, at least one QTL treat the infinitesimal model as little more than a math-
explains 25% or more of the species difference. Indeed, a single QTL at the ematical convenience15 (the model has many attractive
yup locus qualitatively affects carotenoid concentration between the species. properties, including constant ADDITIVE GENETIC VARIANCE
Recent work further shows that replacement of chromosomal material at the yup
under selection), there is some evidence that Fisher
region from one species with that from the other has a large effect on pollinator
visitation106.
went further and considered the infinitesimal view a
Despite these findings, there can be no doubt that many QTL of small effect also reasonable approximation to biological reality16 (see
contribute to adaptation. As experimental sample sizes and therefore statistical power especially Fisher’s role in the debate over the genetic
increase, more of these small-effect QTL will surely be found. basis of mimicry11).

The data. Early empiricists, including Theodosius


Dobzhansky 17, Hermann J. Muller 18–20, Kenneth
with Charles Darwin himself, who argued that “natural Mather21,22 and Julian Huxley23,24, claimed that there
selection can act only by taking advantage of slight suc- was considerable empirical support for micromutation-
cessive variations; she can never take a leap, but must ism. For example, Huxley argued that “selective advan-
DEFICIENCY MAPPING advance by the shortest and slowest steps”2. Although tages so small as to be undetectable in any one generation,
A type of genetic mapping that Darwin had no clear understanding of the nature of are capable … of producing all the observed phenomena
uses chromosomal deletions to
inheritance (or more exactly, had an incorrect under- of biological evolution…. Evolutionary change is almost
‘uncover’ recessive alleles that
affect a trait. standing of it 3), he concluded that the heritable basis of always gradual”23. Despite such confident statements,
adaptive evolution was extremely fine-grained. After all, the case studies to which these early workers pointed
COMPLEMENTATION TESTS precise adaptation is possible only if organisms can were uniformly (and in retrospect, appallingly)
The use of genetically defined come to fit their environments by many minute adjust- weak11,25. Rigorous data bearing on micromutation-
knockout mutations to identify
loci that affect a trait.
ments (see also REF. 4). This ‘micromutational’ view of ism did not appear until remarkably recently, in the
adaptation proved extraordinarily influential, laying the 1980s. Although this delay partly reflected certain tech-
BIOMETRIC foundation for the British BIOMETRIC school of evolution nical difficulties — such as the development of dense
An approach to the study of led by Karl Pearson and Walter Weldon. The biometri- linkage maps — part of the problem was inherent in
phenotypes that emphasizes
cians used newly invented statistical tools, such as regres- the micromutational view itself. Although all scientific
quantitative measurements
(such as of body size) and sion, to analyse the inheritance of CONTINUOUS CHARACTERS, views are simplifications of nature, the micromuta-
statistical analysis. as well as the evolutionary response to selection5,6. (See tional view had the unfortunate effect of excluding
William B. Provine’s7 classic history of this period, entire classes of questions from empirical study. There
CONTINUOUS CHARACTER including the formation of the influential Royal Society is little reason, after all, to ask non-trivial questions
A trait (such as body size) that
varies smoothly (continuously)
Committee on Biometrics.) about the genes that underlie adaptation if one assumes
in magnitude; as opposed to This micromutational view of adaptation was vig- that there are thousands of them, each with small and
discrete characters. orously challenged by the rising Mendelian school of interchangeable effects on the phenotype.

120 | FEBRUARY 2005 | VOLUME 6 www.nature.com/reviews/genetics


© 2005 Nature Publishing Group
REVIEWS

In the 1980s, two new experimental approaches and theory. Evolutionary geneticists find themselves in a
were developed that finally allowed the collection of rig- decidedly awkward position. On the one hand, we can
orous data on the genetics of adaptation — QUANTITATIVE point to a rich and formidable body of mathematical
TRAIT LOCUS (QTL) analysis and MICROBIAL EXPERIMENTAL theory on phenotypic evolution, built largely on an
EVOLUTION. In QTL analysis, the genetic basis of pheno- infinitesimal foundation. On the other hand, we can
typic differences between populations or species can be point to a large and growing body of data on the genetic
analysed using a large suite of mapped molecular mark- basis of adaptation. The problem, of course, is that the
ers. In microbial evolution work, microbes are intro- formidable theory says little or nothing about the formi-
duced into a new environment and their adaptation to dable data. To ask what might seem an obvious ques-
this environment is allowed; genetic and molecular tools tion, where is the theory that tells us what we should
then allow the identification of some or all of the genetic find in QTL or microbial evolution experiments?
changes that underlie this adaptation. The results of More precisely, recent empirical findings force us to
both approaches were surprising: evolution often confront two questions. Can we construct a theory of
involved genetic changes of relatively large effect and, at adaptive evolution that speaks in the same terms as the
least in some cases, the total number of changes seemed data; that is, in terms of individual mutations that have
to be modest. As this empirical literature has been well individual effects? And if so, can this theory account for
reviewed26–30, I devote little space to it here. However, the observed phenomena?
BOX 1 describes several classical studies, including Several attempts to construct such a theory have
those that analyse the evolution of reduced body been made throughout the history of evolutionary
armour or pelvic structure in lake stickleback, the loss genetics. Although my survey of these attempts is far
of larval trichomes (fine ‘hairs’) in Drosophila species, from exhaustive, I consider each of the two main classes
and the evolution of new morphologies in maize and of adaptation model — those that are phenotype-based
the monkeyflower Mimulus spp. Microbial studies fur- and those that are DNA-sequence based.
ther revealed that genetic changes occurring early in
adaptation often have larger fitness effects than those Phenotypic evolution
that occur later31, and that parallel adaptive evolution Given his role as father of the infinitesimal model, it is
is surprisingly common32–35. surprising to learn that Fisher also presented the first
These experimental findings posed — and continue model of adaptation that allowed individual mutations to
to pose — a considerable challenge to evolutionary have different-sized phenotypic effects. In The Genetical
geneticists: to bridge the gap between adaptation data Theory of Natural Selection, Fisher offered his so-called
EPISTATIC INTERACTION
Any non-additive interaction
between two or more mutations
at different loci, such that their
combined effect on a phenotype Box 2 | Fisher’s geometric model of adaptation
deviates from the sum of their
Ronald A. Fisher argued that his geometric model 0.5
individual effects.
captures the statistical essence of adaptation — the
ADDITIVE GENETIC VARIANCE fact that “one thing [the organism] is made to 0.4
The part of the total genetic conform to another [the environment] in a large
variation that is due to the main number of different respects”1. Fisher represented
(or additive) effects of alleles on each character of an organism as an axis in a Cartesian 0.3
a phenotype; as opposed to the
Pa(x)

coordinate system. The optimal combination of trait


dominance and epistatic
variances. The additive variance values is represented by the origin of this coordinate 0.2
determines the degree of system (see FIG. 1). Because of a recent environmental
resemblance between relatives change, the population no longer resides at the 0.1
and therefore the response to optimum. As in actual Darwinian evolution, Fisher’s
selection. model has three features. First, that populations must
adapt by using mutations that are random with 0
QUANTITATIVE TRAIT LOCUS 0 0.5 1 1.5 2 2.5 3
(QTL). A mapped chromosomal respect to the needs of organisms (that is, that are
x
region that has a detectable effect random in phenotypic direction, pointing away from
on a phenotypic difference the optimum at least as often as towards). Second, that mutations have different phenotypic ‘sizes’ (that is, some mutations
between two populations or are vectors of large magnitude and others are vectors of small magnitude). Third, that populations must adapt in the face of
species. A QTL does not pleiotropy (that is, a mutation affects many characters and, although improving one character, might worsen many others).
necessarily correspond to a
Fisher showed that the probability, Pa (x), that a random mutation of a given phenotypic size, r, is favourable is
single gene, but can reflect
several linked genes. 1 – Φ(x), where Φ is the cumulative distribution function of a standard normal random variable and x is a standardized
mutational size, x = r √n /(2z), where n is the number of characters and z is the distance to the optimum. This
MICROBIAL EXPERIMENTAL probability, which is plotted in the figure, falls rapidly with mutational size. This famous plot from Fisher’s The Genetical
EVOLUTION Theory of Natural Selection1 shows that infinitesimally small mutations enjoy a 50% chance of being favourable but that
An experimental approach that this probability falls rapidly for progressively larger mutations. Although Fisher presented no derivation of his famous
involves the ‘real time’
probability, it has since been re-derived by Leigh107, and Hartl and Taubes44.
adaptation of microbes
(typically bacteria, phage or Essentially all studies of adaptation in Fisher’s geometric model assume that adaptation involves the appearance and
yeast) to defined laboratory substitution of new mutations (not standing genetic variation) and that the optimum does not move during the bout of
conditions. adaptation that is studied (but see REF. 108). Results might well differ under other scenarios92.

NATURE REVIEWS | GENETICS VOLUME 6 | FEBRUARY 2005 | 1 2 1


© 2005 Nature Publishing Group
REVIEWS

geometric model as an explicit attempt to capture the


“statistical requirements of the situation” of adaptation1.
In this model, an organism is represented as a set of
phenotypic characters, each measured on a Cartesian
axis and each having an optimal value in the present
environment (BOX 2; FIG. 1). Because of a recent environ-
mental change, the population has fallen off the opti-
mum; the problem for adaptation is to return to the
optimum. The essence of Darwinian evolution is that
populations must attempt this return by producing
mutations that are random with respect to the organ-
ism’s need, that is those that have random direction in
phenotypic space. Crucially, some of these mutations
might be larger than others.
Fisher used his geometric model to ask a simple ques-
tion — what is the probability that a random mutation of
a given phenotypic size will be beneficial? He showed Figure. 1 | Adaptation in Fisher’s geometric model. A bout
that, although infinitesimally small mutations enjoy a of adaptation in Ronald A. Fisher’s geometric model is shown.
For simplicity, the organism that is considered comprises only
50% chance of being beneficial, this probability falls to
three characters. The population begins on the surface of the
“exceedingly small values”with increasing mutational size sphere and, by substituting beneficial mutations (red vectors),
(see figure in BOX 1). Fisher therefore concluded that very evolves towards the phenotypic optimum at the centre of the
small mutations are the genetic basis of adaptation — a sphere. The mutations that are substituted become smaller on
conclusion that was extraordinarily influential and that average as the population nears the optimum. Modified, with
was cited by nearly all the founders of the modern syn- permission, from REF. 41  (2002) Macmillan Magazines Ltd.
thesis (reviewed in REF. 25). Interestingly, although Fisher
published his calculations only in 1930, brief comments
in previous publications36 reveal that he had arrived at his for subsequent substitutions, fall off by an almost con-
micromutational conclusion far earlier and for essentially stant proportion; that is, as an approximate geometric
the same reasons as he emphasized in 1930. sequence. Adaptation is therefore characterized by a pat-
Ironically, although Fisher offered the first sensible tern of diminishing returns — larger-effect mutations are
model of adaptation, the sole question he asked of it typically substituted early on and smaller-effect ones
suppressed all further interest in the model. His answer, later38–41. These results appear to be reasonably robust to
after all, suggested that micromutationism is plausible assumptions about the precise shape of the fitness func-
and that one could, therefore, study adaptation through tion and the distribution of mutational sizes provided to
infinitesimally based quantitative genetics. To make it natural selection38,39.
doubly ironic, Fisher erred here and his conclusion Although Fisher’s model has recently been used to
(although not his calculation) was flawed. Unfortunately, study the evolution of sex 42, the evolution of develop-
his error was only detected half a century later, by Motoo ment43, COMPENSATORY EVOLUTION44, the distribution of
Kimura37. Kimura pointed out that to contribute to adap- phenotypes under mutation–selection-drift balance45,
46
tation, mutations must do more than be beneficial — they MUTATION LOAD in finite populations and the effects of
47
must escape accidental loss when rare — and mutations species hybridization , some of the most surprising
of larger effect are more likely to escape such loss. Taking findings concern the so-called cost of complexity.
both factors into account, Kimura concluded that muta- Analysis of Fisher’s model shows that complex species
tions of intermediate size are the most likely to contribute (those having many characters) typically show slower
to adaptation, a conclusion that did curiously little to increases in fitness during adaptation than do simple
COMPENSATORY EVOLUTION
curb enthusiasm for micromutationism. species48. Part of the reason is that the distance travelled
Evolution in which a second In the late 1990s, however, it became clear that to the optimum by a beneficial mutation is smaller in a
substitution compensates for the Kimura’s conclusion was also not what it first seemed. complex than a simple species (this distance decreases
deleterious effects of an earlier Although Kimura derived the distribution of sizes among with the square root of the number of characters)48.
substitution.
mutations that are used at a particular step in adaptation Recent work by Welch and Waxman49 indicates that this
MUTATION LOAD this is not the same as the distribution of mutations that cost of complexity might be a general feature of adapta-
The decrease in population are substituted throughout an entire bout of adaptation, a tion; indeed, this cost may be little affected by the degree
fitness below its ideal value bout that might involve many steps (FIG. 1). Using a com- of organismal MODULARITY.
owing to recurrent deleterious
bination of analytical theory and computer simulation, it
mutation.
has been shown38,39 that the size distribution of mutations Sequence evolution
MODULARITY that are substituted over entire bouts of adaptation is Maynard Smith and sequence spaces. Although Fisher’s
The idea that organisms are nearly exponential. Adaptation in Fisher’s model there- model takes into account the Mendelian nature of muta-
broken developmentally into fore involves a few mutations of relatively large pheno- tion, it does not reflect the molecular basis of inheritance;
roughly independent modules,
such that mutations affecting
typic effect and many of relatively small effect. This work that is, the fact that DNA is a linear sequence of
traits in one module do not further showed that the mean sizes of mutations substi- nucleotides that, among other things, encodes a linear
affect traits in other modules. tuted at the first versus the second substitution, and so on sequence of amino acids. In the early 1960s, however,

122 | FEBRUARY 2005 | VOLUME 6 www.nature.com/reviews/genetics


© 2005 Nature Publishing Group
REVIEWS

theorists began to develop models of adaptation that Their approach emphasized the idea that different
were sequence-based.As with so many innovative ideas in sequence spaces might feature different numbers of
late twentieth-century evolutionary genetics, the key local optima. This is most easily seen by picturing a ‘fit-
insight was that of John Maynard Smith50. Maynard ness landscape’. Fitness, or adaptive landscapes were
Smith50,51 emphasized what would become a dominant introduced by Sewall Wright70,71 in his studies of the
theme in the theory of adaptation — real adaptation SHIFTING BALANCE THEORY of evolution (for an excellent
occurs in a sequence space that, unlike the phenotypic review of landscape types and theory, see REF. 72).
space considered above, is discrete. He also emphasized Wright’s landscapes typically plotted the fitnesses of dif-
that this discreteness imposes certain constraints on ferent combinations of genotypes across multiple loci
adaptation. For example, the number of possible sequ- (Wright spoke of a “field of possible gene combina-
ences at a gene is limited — given a gene that is L base- tions”). The fitness landscapes considered by NK theo-
pairs long, only 4L different sequences are possible. The rists were similar, except that each ‘locus’ was typically
constraints on adaptation are, however, far more severe taken to represent a particular site in a DNA sequence.
than this number implies. Because the per site mutation To see this, imagine all possible DNA sequences at a
rate is low (~10–9 per base pair), Maynard Smith argued gene as points on a grid (FIG. 2); sequences that differ
that multiple (for example, double or triple) mutations slightly from each other reside near each other on this
are “probably too rare to be important in evolution”51. grid, whereas sequences that differ more substantially
Instead, natural selection is constrained to surveying reside farther apart; the fitness of each sequence is then
sequences that differ from the wild type at a single site. plotted as its height above this grid. The resulting pic-
There are only 3L such one-mutational step sequences; a ture of hills and valleys represents a fitness landscape.
finite number indeed. The key point is that fitness landscapes can differ in
Maynard Smith also emphasized that adaptation their ruggedness. The smoothest possible landscape fea-
involves what came to be called ‘adaptive walks’ through tures a single optimum and all adaptation necessarily
sequence space: “if evolution by natural selection is to involves walks up this single peak; the most rugged
occur, functional proteins [or DNA sequences] must landscape possible features many local optima and adap-
form a continuous network which can be traversed by tation involves walks up the nearest optimum. The NK
unit mutational steps without passing through non- model allows one to ‘tune’ the ruggedness of fitness land-
functional intermediates”51. In particular, if a wild-type scapes between these extremes by varying two mathe-
sequence can mutate to one or more fitter sequences, matical parameters (N and K), allowing the analysis of
natural selection will ultimately substitute one of these adaptation for many families of landscapes.
beneficial alleles; this new wild-type sequence will in
turn produce its own suite of one-mutational step
sequences and, if any is fitter than the current wild-type, Global optimum
selection will again substitute one of these beneficial
alleles. This adaptive walk ends only when the popula-
tion arrives at a wild-type sequence that is fitter than all
of its 3L one-mutational-step neighbours. At that point,
the population has arrived at a local optimum.
Although Maynard Smith’s work appeared early in
the molecular revolution, his ideas on adaptive walks
Local optimum
were almost entirely ignored for two decades. The cause
of this neglect seems clear: the rise of the neutral theory of
molecular evolution. Throughout the 1960s and 1970s, W
evolutionary geneticists grew increasingly convinced
that much, if not most, molecular evolution reflects
the substitution of neutral52,53 or nearly neutral37,54–56
mutations, not beneficial ones. Throughout much of
this period, the study of adaptation itself grew intellec-
tually suspect57 and the theoretical study of molecular
adaptation essentially ceased.
Figure 2 | A fitness landscape. The underlying grid represents
a DNA sequence space. Although such a space cannot actually
SHIFTING BALANCE THEORY Kauffman and NK models. In the 1980s, theoretical be represented in two dimensions (and instead must be
A largely verbal theory of study of adaptation at the sequence level finally depicted as a high dimensional hypercube), we can loosely
evolution which maintains that
resumed. One of the best known of these new efforts imagine that alleles that are similar in sequence sit near each
the interaction between natural
selection, genetic drift and involved so-called NK models. This work, launched by other on the grid shown, whereas alleles that are different in
migration is more important Kauffman and Levin58, ultimately grew into a large and sequence sit farther apart. Allelic fitnesses (W) are plotted above
the grid. Adaptation in a large population will involve adaptive
that the action of any single sophisticated body of mathematical and computational
force. Sewall Wright argued that walks up the nearest fitness peak. Because natural selection will
literature59–69. Arguing that evolutionary geneticists not allow a population to descend into an adaptive valley,
this theory helped to explain
how species could effectively
possess “essentially no theory of adaptation”, Kauffman adaptation drives a population to a local optimum, which might
search for the global, and not and colleagues58 set out to discover the laws, if any, that or might not correspond to the global optimum. Modified, with
merely local, optimum. describe adaptation through sequence space. permission, from REF. 111  (2004) Sinauer Associates Inc.

NATURE REVIEWS | GENETICS VOLUME 6 | FEBRUARY 2005 | 1 2 3


© 2005 Nature Publishing Group
REVIEWS

Gillespie and the mutational landscape. How many


sequences at a gene should be highly fit and how
many lethal? Just what does the distribution of fitnesses
look like on a fitness landscape? Unfortunately, we have
almost no data relating to these questions. Given this
situation, Ohta54 and Kimura55 suggested that we con-

f(W)
sider fitnesses as randomly drawn from a probability
distribution. The question is, of course, which distribu-
tion? In the early 1980s, John Gillespie75,76 surprisingly
argued that it might not matter much. The reason
unexpectedly follows from the point noted above —
that adaptation typically begins from a high fitness
W sequence.
Figure 3 | The distribution of fitnesses at a locus. An Gillespie’s insight was that, although we do not
arbitrary distribution of fitnesses among alleles at some gene is know the detailed distribution of fitnesses at any gene,
shown (W, fitness value; f(W), frequency of a certain fitness we do know two things: that the wild-type represents a
value). Although the precise shape of this distribution will, in draw from the right tail of the fitness distribution; and
reality, almost always be unknown, John Gillespie75–77 that beneficial mutations represent even more extreme
emphasized that, in most cases, the wild-type allele will be
draws from this tail (see FIG. 3). And this, Gillespie75–77
highly fit; that is, drawn from the right-hand tail of the
distribution (black arrow). Consequently, beneficial mutations argued, means that we can import extreme value
represent even more extreme draws (green arrows). This theory (EVT) into the study of adaptation. EVT is a
implies that extreme value theory (EVT) can be used to study body of probability theory that is concerned with the
adaptation (see also BOX 3). properties of draws from the tails of distributions78,79.
Remarkably, EVT shows that these draws have certain
properties that are asymptotically independent of the
NK theorists derived several approximate results that (usually unknown) details of the distribution, a robust-
characterize such landscapes. These included the proba- ness that is reminiscent of the central limit theorem79.
bility that a sequence is a local optimum, the number of Although EVT is important in modern mathematical
local optima, the proportion of local optima that can be finance and risk analysis (extreme swings in security
reached from a given sequence and the average length of prices and insurance claims can have devastating effects
adaptive walks on landscapes of varying ruggedness on portfolios and insurance firms, respectively80), it
(reviewed in REFS 60,69). NK and related models (such as had, until Gillespie’s work, little role in evolutionary
the Block model73) were also applied to the study of genetics. But Gillespie saw that, as most adaptation
58–60,64
AFFINITY MATURATION during immune response , occurs in the right tail of fitness distributions, EVT
genetic regulatory networks58,60 and RNA folding67. might tell us a good deal about the adaptation of
Although NK models successfully explained certain DNA sequences (BOX 3).
features of affinity maturation60 and received consider- Gillespie75–77 used EVT to study adaptation over what
able attention in computer science and in physics (in he called the “mutational landscape”. He was primarily
which problems in SPIN GLASSES are similar74), it seems fair concerned with evolution under strong selection–weak
to say that these models ultimately had a lesser impact on mutation (SSWM) conditions. His weak mutation
evolutionary genetics where they began. In retrospect, assumption was essentially equivalent to that of Maynard
two features of many NK studies were unbiological. The Smith: mutation rates per site are sufficiently low for us to
first involves the ‘move rule’ used to decide which of sev- ignore double mutants; his strong selection assumption
eral beneficial mutant sequences a population will sub- meant roughly that mutations are either definitely bene-
stitute at the next step in adaptation. The NK literature ficial or deleterious (neutral mutations are not allowed).
focused on two such rules: random, in which a beneficial Under SSWM assumptions, a population is essentially
sequence is randomly chosen from those avail- fixed for a single wild-type sequence at any moment in
able58,61,63–67 and gradient, in which the fittest of available time and this wild-type sequence recurrently mutates to
beneficial sequences is always chosen58,63,67,68. The prob- 3L neighbouring sequences. Because the wild-type
lem is that natural selection uses neither of these rules sequence enjoys high fitness, it is assumed that few, if
(the correct move rule is described in the next section). any, of these 3L mutant sequences are beneficial. Each
AFFINITY MATURATION Second, the NK literature typically considered adapta- beneficial allele is likely to be lost accidentally each time it
An increase in the affinity of an tion from a random starting sequence or even from the appears but, because mutation is recurrent, one will
antibody for an antigen which is
worst possible sequence58,62–64,66,67,69. Real adaptation, eventually be substituted; this completes one step in
seen as an immune response
improves. however, often begins from a wild-type sequence of high an adaptive walk. The new wild-type sequence now
fitness. The wild-type allele, after all, produces a func- produces it own set of 3L mutants and the process is
SPIN GLASSES tional protein and was, until a recent environmental repeated. Gillespie assumed that mutant fitnesses are
Magnetic objects that are change, the fittest sequence that was locally available. As drawn from the same distribution throughout a
disordered and in which
adjacent dipoles can either
we will see (and as Kauffman later acknowledged60), the (brief) adaptive walk; adaptation is therefore charac-
‘point’ in the same direction or simple fact that the wild-type allele is highly fit turns out terized by the movement of a population out along a
in opposite directions. to be of considerable importance. tail of fitnesses.

124 | FEBRUARY 2005 | VOLUME 6 www.nature.com/reviews/genetics


© 2005 Nature Publishing Group
REVIEWS

MOLECULAR CLOCK Gillespie75–77 described several results that character- two to five. Molecular evolution at a gene therefore
The empirical finding that a ize adaptation over the mutational landscape. Perhaps occurs in small bursts of substitutions. Put differently,
particular type of protein or most importantly, he calculated the ‘move rule’ that is molecular evolution by natural selection leaves a signa-
DNA sequence evolves at a
used by positive natural selection. He showed that, when ture that differs from that left by neutrality — whereas
nearly constant rate through
time. several beneficial mutations are available to a wild-type neutral evolution yields a simple MOLECULAR CLOCK (with
sequence, the probability that a population jumps to a substitutions that occur as a POISSON PROCESS), natural
POISSON PROCESS beneficial allele j at the next step in evolution is, under selection does not (the molecular clock is instead
A simple statistical process in SSWM conditions, sj /Σs, where s is the selective advan- ‘overdispersed’). Gillespie77,85,86 argued that adaptation
which there is a small and
constant probability of change
tage of an allele, and the summation involves all avail- explained the existing molecular evolutionary data better
during each short interval of able beneficial alleles. Explained in words, this formula than neutrality did (see also REF. 87).
time. means that the chance that a particular beneficial allele Much of the recent work on adaptation theory has
is the next substituted is proportional to the selective focused on Gillespie’s mutational landscape model.
advantage of that allele. Beneficial mutations of greater Several results have been described over the past several
effect are therefore more likely, although not guaran- years. Perhaps most surprising, two studies88,89 showed
teed, to be the next substituted. This move rule differs that beneficial mutations should have exponentially dis-
from those used in NK studies (despite REF. 81). tributed fitness effects that are independent of many bio-
Most of Gillespie’s work focused on entire adaptive logical details. Most recent analyses of the mutational
walks. He noted two key facts. First, because the fitness landscape model, however, have focused on the unit
of wild-type sequences increases during adaptation, event in adaptation — the substitution of a beneficial
beneficial mutations become increasingly difficult to mutation. It has been shown90 that, if the current wild-
come by. The theory of records82,83 shows that the type sequence is the ith fittest allele (that is, i – 1 benefi-
cumulative number of alleles that break the previous cial mutations are available), the population will jump to
‘fitness record’ (that is, that are beneficial) only a beneficial allele that has mean fitness rank (i + 2)/4 at
increases logarithmically during adaptive walks84. the next substitution. Adaptation is, therefore, character-
Second, Gillespie77 showed that the mean number of ized by surprisingly large jumps in fitness rank. If thir-
steps taken during adaptive walks is small — typically teen beneficial mutations are available to a wild-type
sequence, populations will typically jump to the fourth
best allele at the next step in adaptation. Adaptation
therefore leap-frogs many moderately beneficial muta-
Box 3 | Adaptation and extreme value theory
tions, arriving at a strongly beneficial one. Recent work
Classical extreme value theory (EVT) emerged from early work by the biologist Ronald A. also shows that parallel evolution at the DNA sequence
Fisher and the probabilists Richard von Mises and Boris V. Gnedenko, among others level is more common under natural selection than
(reviewed in REFS 78,79,109). EVT is concerned with the asymptotic properties of the largest under neutrality. In fact, the probability of parallel evolu-
draws from a probability distribution, as the number of draws becomes large. The most tion approximately doubles under positive selection91.
fundamental, and best known result from EVT concerns the distribution of maxima. If Computer simulations of entire adaptive walks further
many values from a given ‘parent’ distribution are drawn, the largest value saved, and show that at least half the total gain in fitness that occurs
then this process is repeated many times, the distribution of these maxima approaches during adaptation is due to a single substitution90. Once
the so-called extreme value distribution. There are in fact three extreme value
again, therefore, adaptation features relatively large
distributions — one for ‘ordinary’ parent distributions (the Gumbel type), another for
jumps. Indeed, adaptation seems to be characterized by a
many parent distributions that are truncated on the right (the Weibull type) and a last
‘Pareto principle’, in which the majority of an effect
for parent distributions that lack all or higher moments (the Fréchet type). The Gumbel
(increased fitness) is due to a minority of causes (one
type includes most ordinary distributions — for example, the normal, lognormal, gamma,
exponential,Weibull and logistic — and was the focus of classical EVT. The Gumbel substitution). Finally, recent theory shows that, condi-
extreme value distribution is often referred to as ‘the’ extreme value distribution. The tional on the substitutions occurring, the mean selection
biologically important point is that extreme draws from a surprisingly wide array of coefficient, s, among mutations that are fixed at subse-
parent distributions all converge to the same extreme value distribution. EVT is generally quent steps in adaptation decreases as an approximate
characterized by such robustness to distributional details, allowing conclusions to be made geometric sequence — again revealing a pattern of
about adaptation that depend only weakly on the (generally unknown) distribution of diminishing returns — and that the overall distribu-
fitnesses at a gene. tion of s among mutations that are substituted during
Gillespie75–77 assumed that the distribution of fitnesses (see FIG. 3) belongs to the adaptation is roughly exponential90.
Gumbel type and subsequent work on the mutational landscape has followed suit. Recent These results are obviously reminiscent of those
work provides some support for this assumption. It can be shown, for example, that the from Fisher’s geometric model — a similarity that is
distribution of mutational effects in Fisher’s geometric model is of the Gumbel type surprising given the fundamental differences between
(A.O., unpublished data). EVT describes not only the distribution of largest values from a the older phenotypic and newer DNA sequence-based
parent distribution, but the distribution of the second, third, and so on, largest values79. models92. This congruence represents one of the most
EVT also describes the asymptotic distributions of the ‘extreme spacings’ between the tantalizing results to emerge from recent adaptation
largest and next-largest (and so on) draws — given a Gumbel-type parent distribution, theory, indicating (although certainly not proving) that
these spacings are independent exponential random variables with averages that behave adaptation to a fixed optimum by new mutations may
in a certain simple way110. This result has a fundamental role in recent theory of be characterized by certain robust patterns. Future work
adaptation, as fitness differences among beneficial mutations represent extreme spacings.
could of course show that different biological scenarios
Recent work has begun to explore the consequences of non-Gumbel fitness distributions
(for example, adaptation from the standing genetic
on adaptation over the mutational landscape (A.O., unpublished data).
variation) are characterized by different patterns.

NATURE REVIEWS | GENETICS VOLUME 6 | FEBRUARY 2005 | 1 2 5


© 2005 Nature Publishing Group
REVIEWS

Limitations and conclusions the data. There are in fact two problems. The first is
The history of evolutionary genetics indicates that two that current theory is limited in several ways — all the
main factors slowed the development of a mature theory models that have been mentioned rest on important
of adaptation. The micromutational view of quantitative assumptions and idealizations. Fisher himself noted
genetics and the neutral theory of population genetics. such limitations in correspondence about his geo-
We cannot, after all, construct a meaningful theory of metric model96,92. And the mutational landscape
adaptation if we assume away the existence of muta- model assumes that adaptation occurs at a single gene
tions that have different-sized phenotypic effects or if we (or small genome), that fitness distributions show cer-
assume that substitutions at the DNA-sequence level tain tail behaviour, and that mutant fitnesses are taken
have no effect on fitness. The fact that we possess little from the same distribution throughout adaptation.
theory of adaptation went largely unnoticed until the Although they are reasonable starting points for theory,
1980s, when QTL and microbial evolution approaches none of these assumptions is necessarily correct and
yielded abundant data that revealed a finite (and often changing any might well change our predictions
modest) number of substitutions that have definite (and (however, see BOX 3).
often different) effects on the phenotype or fitness. The second problem concerns testability. Although
In this article, I have emphasized that the appearance adaptation models have become more concrete and per-
of such data forced evolutionary geneticists to confront haps more realistic, it is not clear that they have become
two questions. First, can we construct a theory of adapta- more testable. The difficulty is practical, not principled.
tion that speaks in the same terms as the data? And, if so, Whereas current theory does make testable predictions,
can this theory actually account for the empirical phe- the effort required to perform these tests is often enor-
nomena observed? The answer to the first question is mous (particularly as the theory is probabilistic, making
clearly yes. Evolutionists can and have built models of predictions over many realizations of adaptation). Given,
adaptation that speak in terms of individual mutations for example, the inevitable and often severe limits on
that have individual effects (whether phenotypic or fit- replication in microbial evolution work, we can usually
ness). The answer to the second question is less clear. do no more than test qualitative predictions.
Present theory appears to adequately explain certain Despite such limitations, experiments to test the more
qualitative patterns that characterize genetic data on tractable predictions of present theory (for example, those
adaptation. Four such patterns are: that there are more that focus on a single step in adaptation, where greater
beneficial mutations of small than large effect89,93,94; that replication is possible) are important. The main reason is
QTL studies reveal more substitutions of small than large that the two problems noted above are related; only by
effect26,28,95; that microbial studies show that early substi- testing present predictions can we determine which, if
tutions have larger fitness effects than later ones (that is, any, of the assumptions that underlie current theory are
there is pattern of diminishing returns)31; and that parallel inappropriate. Only then will theorists know which
evolution is common at the DNA sequence level32–35. assumptions to change, deriving new — and perhaps
However, it is unclear whether current theory accom- different — predictions and ultimately allowing the
plishes much more than qualitative agreement with elaboration of a more realistic theory of adaptation.

1. Fisher, R. A. The Genetical Theory of Natural Selection 14. Bulmer, M. G. The Mathematical Theory of Quantitative 29. Orr, H. A. The genetics of species differences. Trends Ecol.
(Oxford Univ. Press, Oxford, 1930). Genetics (Oxford Univ. Press, Oxford, 1980). Evol. 16, 343–350 (2001).
One of the founding documents of modern 15. Barton, N. H. & Turelli, M. Evolutionary quantitative genetics: 30. Mackay, T. F. C. Quantitative trait loci in Drosophila. Nature
evolutionary biology. It includes Fisher’s classical how little do we know? Annu. Rev. Genetics 23, 337–370 Rev. Genet. 2, 11–20 (2001).
discussions of his geometric model of adaptation. (1989). 31. Holder, K. & Bull, J. Profiles of adaptation in two similar
2. Darwin, C. R. The Origin of Species (J. Murray, London, 1859). 16. Turner, J. R. G. in The Probabilistic Revolution Vol. 2 viruses. Genetics 159, 1393–1404 (2001).
3. Mayr, E. The Growth of Biological Thought (Harvard Univ. (eds. Kruger, L., Gigerenzer, G. & Morgan, M. S.) 313–354 32. Bull, J. J. et al. Exceptional convergent evolution in a virus.
Press, Cambridge, Massachusetts, 1982). (The MIT Press, Cambridge, Massachusetts, 1987). Genetics 147, 1497–1507 (1997).
4. Maynard Smith, J. (ed.) Evolution Now: a Century After 17. Dobzhansky, T. Genetics and the Origin of Species 33. Wichman, H. A., Badgett, M. R., Scott, L. A., Boulianne, C. M.
Darwin (W. H. Freeman and Co., San Francisco, 1982). (Columbia Univ. Press, New York, 1937). & Bull, J. J. Different trajectories of parallel evolution during
5. Pearson, K. Mathematical contributions to the theory of 18. Muller, H. J. Eugenics, Genetics and the Family: viral adaptation. Science 285, 422–424 (1999).
evolution: on the law of ancestral heredity. Proc. R. Soc. Proceedings of the Second International Congress of 34. Riehle, M. M., Bennett, A. F. & Long, A. D. Genetic
Lond. 62, 386–412 (1898). Eugenics Vol. 1 (ed. History, A. M.) (William and Wilkens, architecture of thermal adaptation in Escherichia coli. Proc.
6. Weldon, W. F. R. Attempt to measure the death-rate due to Baltimore, 1923). Natl Acad. Sci. USA 98, 525–530 (2001).
the selective destruction of Carcinus moenas with respect to 19. Muller, H. J. in The New Systematics (ed. Huxley, J. S.) 35. Anderson, J. et al. Mode of selection and experimental
a particular dimension. Proc. R. Soc. Lond. 58, 360–379 185–268 (Clarendon Press, Oxford, 1940). evolution of antifungal drug resistance in Saccharomyces
(1895). 20. Muller, H. J. The Darwinian and modern conceptions of cerevisiae. Genetics 163, 1287–1298 (2003).
7. Provine, W. B. The Origins of Theoretical Population natural selection. Proc. Am. Phil. Soc. 93, 459–470 (1949). 36. Fisher, R. A. Darwinian evolution of mutations. Eugen. Rev.
Genetics (Univ. Chicago Press, Chicago, 1971). 21. Mather, K. Polygenic inheritance and natural selection. Biol. 14, 31–34 (1922).
8. Morgan, T. H. Evolution and Adaptation (Macmillan, New Rev. 18, 32–64 (1943). 37. Kimura, M. The Neutral Theory of Molecular Evolution
York, 1903). 22. Mather, K. Biometrical Genetics (Dover, New York, 1949). (Cambridge Univ. Press, Cambridge, 1983).
9. Bateson, W. Mendel’s Principles of Genetics (Cambridge 23. Huxley, J. Evolution as a Process (Unwin Bros, Woking; 38. Orr, H. A. The population genetics of adaptation: the
Univ. Press, Cambridge, 1913). London, 1954). distribution of factors fixed during adaptive evolution.
10. Punnett, R. C. Mimicry in Butterflies (Cambridge Univ. Press, 24. Huxley, J. Evolution, the Modern Synthesis (George Allen & Evolution 52, 935–949 (1998).
Cambridge, 1915). Unwin, London, 1942). A theoretical study of the genetic basis of phenotypic
11. Turner, J. R. G. Fisher’s evolutionary faith and the challenge 25. Orr, H. A. & Coyne, J. A. The genetics of adaptation evolution using Fisher’s geometric model of
of mimicry. Oxford Surv. Evol. Biol. 2, 159–196 (1985). revisited. Am. Nat. 140, 725–742 (1992). adaptation. This study argues that previous claims
12. Bateson, W. in Darwin and Modern Science 26. Tanksley, S. D. Mapping polygenes. Annu. Rev. Genet. 27, about adaptation based on Fisher’s model were partly
(ed. Seward, A. C.) 85–101 (Cambridge Univ. Press, 205–233 (1993). mistaken.
Cambridge, 1909). 27. Falconer, D. S. & Mackay, T. F. C. Introduction to 39. Orr, H. A. The evolutionary genetics of adaptation: a
13. Fisher, R. A. The correlations between relatives on the Quantitative Genetics (Longman, Harlow, England, 1996). simulation study. Genet. Res. 74, 207–214 (1999).
supposition of Mendelian inheritance. Trans. R. Soc. Edinb. 28. Kearsey, M. J. & Farquhar, G. L. QTL analysis in plants; 40. Barton, N. The geometry of natural selection. Nature 395,
52, 399–433 (1918). where are we now? Heredity 80, 137–142 (1998). 751–752 (1998).

126 | FEBRUARY 2005 | VOLUME 6 www.nature.com/reviews/genetics


© 2005 Nature Publishing Group
REVIEWS

41. Barton, N. H. & Keightley, P. D. Understanding quantitative 67. Fontana, W. et al. RNA folding and combinatory landscapes. 96. Fisher, R. A. The Genetical Theory of Natural Selection: a
genetic variation. Nature Rev. Genet. 3, 11–21 (2002). Phys. Rev. E 47, 2083–2099 (1993). Complete Variorum Edition (Oxford Univ. Press, Oxford,
42. Peck, J. R., Barreau, G. & Heath, S. C. Imperfect genes, 68. Stadler, P. F. & Happel, R. Random field models for fitness 2000).
Fisherian mutation and the evolution of sex. Genetics 145, landscapes. J. Math. Biol. 38, 435–478 (1999). 97. Colosimo, P. F. et al. The genetic architecture of parallel
1171–1199 (1997). 69. Macken, C. A. & Stadler, P. F. in 1993 Lectures in Complex armor plate reduction in threespine sticklebacks. PLoS Biol.
43. Rice, S. A geometric model for the evolution of Systems (eds Nadel, L. & Stein, D. L.) 43–86 (Addison- 2, e109 (2004).
development. J. Theor. Biol. 143, 319–342 (1990). Wesley, Reading, Massachusetts, 1995). 98. Cresko, W. A. et al. Parallel genetic basis for repeated evolution
44. Hartl, D. & Taubes, C. H. Compensatory nearly neutral 70. Wright, S. Evolution in Mendelian populations. Genetics 16, of armor loss in Alaskan threespine stickleback populations.
mutations: selection without adaptation. J. Theor. Biol. 182, 97–159 (1931). Proc. Natl Acad. Sci. USA 101, 6050–6055 (2004).
303–309 (1996). 71. Wright, S. The roles of mutation, inbreeding, crossbreeding, 99. Shapiro, M. D. et al. Genetic and developmental basis of
45. Hartl, D. L. & Taubes, C. H. Towards a theory of evolutionary and selection in evolution. Proc. Sixth Intl Cong. Genet. 1, evolutionary pelvic reduction in threespine sticklebacks.
adaptation. Genetica 102/103, 525–533 (1998). 356–366 (1932). Nature 428, 717–723 (2004).
46. Poon, A. & Otto, S. P. Compensating for our load of 72. Gavrilets, S. Fitness Landscapes and the Origin of Species Experimental analysis that identified a candidate
mutations: freezing the meltdown of small populations. (Princeton Univ. Press, Princeton, 2004). gene, Pitx1, that has an important role in the
Evolution 54, 1467–1479 (2000). 73. Perelson, A. S. & Macken, C. A. Protein evolution on partially morphological adaptation of marine sticklebacks to
47. Barton, N. H. The role of hybridization in evolution. Mol. Ecol. correlated landscapes. Proc. Natl Acad. Sci. USA 92, lake environments.
10, 551–568 (2001). 9657–9661 (1995). 100. Sucena, E. & Stern, D. L. Divergence of larval morphology
48. Orr, H. A. Adaptation and the cost of complexity. Evolution 74. Anderson, P. W. in Emerging Synthesis in Science: between Drosophila sechellia and its sibling species caused
54, 13–20 (2000). Proceedings of the Founding Workshop of the Santa Fe by cis-regulatory evolution of ovo/shaven-baby. Proc. Natl
49. Welch, J. J. & Waxman, D. Modularity and the cost of Institute (ed. Pines, D.) (Santa Fe Inst., Santa Fe, 1985). Acad. Sci. USA 97, 4530–4534 (2000).
complexity. Evolution 57, 1723–1734 (2003). 75. Gillespie, J. H. A simple stochastic gene substitution model. 101. White, S. & Doebley, J. Of genes and genomes and the
50. Maynard Smith, J. The Scientist Speculates: an Anthology Theor. Popul. Biol. 23, 202–215 (1983). origin of maize. Trends Genet. 14, 327–332 (1998).
of Partly-Baked Ideas (ed. Good, I. J.) 252–256 (Basic 76. Gillespie, J. H. Molecular evolution over the mutational 102. Wang, R.-L., Stec, A., Hey, J., Lukens, L. & Doebley, J. The
Books, New York, 1962). landscape. Evolution 38, 1116–1129 (1984). limits of selection during maize domestication. Nature 398,
This introduced the idea of adaptation through a This is Gillespie’s most extensive discussion of the 236–239 (1999).
‘sequence space’. Although Maynard Smith use of extreme value theory in the study of molecular 103. Doebley, J. The genetics of maize evolution. Annu. Rev.
considered protein spaces, his ideas had a key role in evolution. Genet. 38, 37–59 (2004).
later thinking about adaptative walks through DNA 77. Gillespie, J. H. The Causes of Molecular Evolution (Oxford A comprehensive and up-to-date review of Doebley’s
sequence spaces. Univ. Press, Oxford, 1991). classical studies of the genetic basis of maize
51. Maynard Smith, J. Natural selection and the concept of a 78. Gumbel, E. J. Statistics of Extremes (Columbia Univ. Press, domestication.
protein space. Nature 225, 563–564 (1970). New York, 1958). 104. Bradshaw, H. D., Wilbert, S. M., Otto, K. G. &
52. Kimura, M. Evolutionary rate at the molecular level. Nature 79. Leadbetter, M. R., Lindgren, G. & Rootzen, H. Extremes and Schemske, D. W. Genetic mapping of floral traits associated
217, 624–626 (1968). Related Properties of Random Sequences and Processes with reproductive isolation in monkeyflowers (Mimulus).
53. King, J. L. & Jukes, T. H. Non-Darwinian evolution: random (Springer, New York, 1983). Nature 376, 762–765 (1995).
fixation of selectively neutral mutations. Science 164, 80. Embrechts, P. (ed.) Extremes and Integrated Risk 105. Bradshaw, H. D., Otto, K. G., Frewen, B. E., McKay, J. K. &
788–798 (1969). Management (Risk Books, London, 2000). Schemske, D. W. Quantitative trait loci affecting differences
54. Ohta, T. Molecular Evolution and Polymorphism 81. Stadler, P. F. Biological Evolution and Statistical Physics in floral morphology between two species of Monkeyflower
(ed. Kimura, M.) 148–167 (Natl Inst. Genet., Mishima, 1977). (eds Lassig, M. & Valleriani, A.) 183–204 (Springer, Berlin, (Mimulus). Genetics 149, 367–382 (1998).
55. Kimura, M. Model of effectively neutral mutations in which 2002). An early and important QTL analysis that implicates
selective constraint is incorporated. Proc. Natl Acad. Sci. 82. Glick, N. Breaking records and breaking boards. Am. Math. the major genes in the natural adaptation of plant
USA 76, 3440–3444 (1979). Monthly 85, 2–26 (1978). species to their pollinators.
56. Ohta, T. The nearly neutral theory of molecular evolution. 83. Arnold, B. C., Balakrishnan, N. & Nagaraja, H. N. Records 106. Bradshaw, H. D. & Schemske, D. W. Allele substitution at a
Annu. Rev. Ecol. Syst. 23, 263–286 (1992). (John Wiley and Sons, New York, 1998). flower colour locus produces a pollinator shift in
57. Gould, S. J. & Lewontin, R. C. The spandrels of San Marco 84. Gillespie, J. H. Is the population size of a species relevant to monkeyflowers. Nature 426, 176–178 (2003).
and the Panglossian paradigm: a critique of the its evolution? Evolution 55, 2161–2169 (2003). 107. Leigh, E. G. Ronald Fisher and the development of
adaptationist program. Proc. R. Soc. Lond. B 205, 581–598 85. Gillespie, J. H. Natural selection and the molecular clock. evolutionary theory. II. Influences of new variation on
(1979). Mol. Biol. Evol. 3, 138–155 (1986). evolutionary process. Oxford Surv. Evol. Biol. 4, 212–264
58. Kauffman, S. & Levin, S. Towards a general theory of 86. Gillespie, J. H. Molecular evolution and the neutral allele (1987).
adaptive walks on rugged landscapes. J. Theor. Biol. 128, theory. Oxford Surv. Evol. Biol. 4, 10–37 (1989). 108. Otto, S. P. & Jones, C. D. Detecting the undetected:
11–45 (1987). 87. Gerrish, P. The rhythm of microbial adaptation. Nature 413, estimating the total number of loci underlying a trait in QTL
This paper introduced the idea of adaptation over 299–302 (2001). analyses. Genetics 156, 2093–2107 (2000).
fitness landscapes of varying ruggedness. Although it 88. Orr, H. A. The distribution of fitness effects among beneficial 109. David, H. A. Order Statistics (John Wiley and Sons, New
largely focused on ‘random’ landscapes, the paper mutations. Genetics 163, 1519–1526 (2003). York, 1970).
gave rise to a large literature on landscapes of 89. Rozen, D. E., de Visser, J. A. G. M. & Gerrish, P. J. Fitness 110. Weissman, I. Estimation of parameters and large quantiles
different ruggedness. effects of fixed benefical mutations in microbial populations. based on the kth largest observations. J. Am. Stat. Assoc.
59. Kauffman, S. A., Weinberger, E. D. & Perelson, A. S. in Curr. Biol. 12, 1040–1045 (2002). 73, 812–815 (1978).
Theoretical Immunology: Part One (ed. Perelson, A. S.) 90. Orr, H. A. The population genetics of adaptation: the 111. Coyne, J. A. & Orr, H. A. Speciation (Sinauer Associates
349–382 (Addison-Wesley, New York, 1988). adaptation of DNA sequences. Evolution 56, 1317–1330 Inc., Sunderland, Massachusetts, 2004).
60. Kauffman, S. A. The Origins of Order (Oxford Univ. Press, (2002).
New York, 1993). This paper brought the extreme value theory to bear Acknowledgements
61. Weinberger, E. A more rigorous derivation of some on several key questions in the study of DNA This work was supported by a grant from the US National Institutes
properties of uncorrelated fitness landscapes. J. Theor. Biol. sequence adaptation. of Health.
134, 125–129 (1988). 91. Orr, H. A. The probability of parallel adaptation. Evolution (in
62. Weinberger, E. Correlated and uncorrelated fitness the press). Competing interests statement
landscapes and how to tell the difference. Biol. Cybern. 63, 92. Orr, H. A. Theories of adaptation: what they do and don’t The author declares no competing financial interests.
325–336 (1990). say. Genetica (in the press).
63. Weinberger, E. D. Local properties of Kauffman’s N-k model: 93. Imhof, M. & Schlotterer, C. Fitness effects of advantageous
a tunably rugged energy landscape. Phys. Rev. A 44, mutations in evolving Escherichia coli populations. Proc. Online links
6399–6413 (1991). Natl Acad. Sci. USA 98, 1113–1117 (2001).
64. Macken, C. A. & Perelson, A. S. Protein evolution on rugged 94. Sanjuan, R., Moya, A. & Elena, S. F. The distribution of DATABASES
landscapes. Proc. Natl Acad. Sci. USA 86, 6191–6195 fitness effects caused by single-nucleotide substitutions in The following terms in this article are linked online to:
(1989). an RNA virus. Proc. Natl Acad. Sci. USA 101, 8396–8401 Entrez: http://www.ncbi.nih.gov/Entrez/
65. Macken, C. A., Hagan, P. S. & Perelson, A. S. Evolutionary (2004). ovo/shaven-baby
walks on rugged landscapes. SIAM J. Appl. Math. 51, 95. Liu, J. et al. Genetic analysis of a morphological FURTHER INFORMATION
799–827 (1991). shape difference in the male genitalia of Drosophila H. Allen Orr’s web page:
66. Flyvbjerg, H. & Lautrup, B. Evolution in a rugged fitness simulans and D. mauritiana. Genetics 142, 1129–1145 http://www.rochester.edu/College/BIO/faculty/Orr.html
landscape. Phys. Rev. A 46, 6714–6723 (1992). (1996). Access to this interactive links box is free online.

NATURE REVIEWS | GENETICS VOLUME 6 | FEBRUARY 2005 | 1 2 7


© 2005 Nature Publishing Group

You might also like