Download Integration of populations and differentiation of species

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

The Selfish Gene wikipedia , lookup

Sexual selection wikipedia , lookup

Hologenome theory of evolution wikipedia , lookup

Hybrid (biology) wikipedia , lookup

Punctuated equilibrium wikipedia , lookup

Natural selection wikipedia , lookup

Inclusive fitness wikipedia , lookup

Organisms at high altitude wikipedia , lookup

Reproductive isolation wikipedia , lookup

Evidence of common descent wikipedia , lookup

Adaptation wikipedia , lookup

The eclipse of Darwinism wikipedia , lookup

Population genetics wikipedia , lookup

Introduction to evolution wikipedia , lookup

Transcript
NIH Public Access
Author Manuscript
New Phytol. Author manuscript; available in PMC 2008 December 12.
NIH-PA Author Manuscript
Published in final edited form as:
New Phytol. 2004 January ; 161(1): 59–69. doi:10.1046/j.1469-8137.2003.00933.x.
Integration of populations and differentiation of species
Loren H. Rieseberg, Sheri A. Church, and Carrie L. Morjan
Department of Biology, Indiana University, Bloomington, IN 47405, USA
Summary
NIH-PA Author Manuscript
The framework for modern studies of speciation was established as part of the Neo-Darwinian
synthesis of the early twentieth century. Here we evaluate this framework in the light of recent
empirical and theoretical studies. Evidence from experimental studies of selection, quantitative
genetic studies of species’ differences, and the molecular evolution of ‘isolation’ genes, all agree
that directional selection is the primary cause of speciation, as initially proposed by Darwin.
Likewise, as suggested by Dobzhansky and Mayr, gene flow does hold species together, but probably
more by facilitating the spread of beneficial mutants and associated hitchhiking events than by
homogenizing neutral loci. Reproductive barriers are important as well in that they preserve
adaptations, but as has been stressed by botanists for close to a century, they rarely protect the entire
genome from gene flow in recently diverged species. Contrary to early views, it is now clear that
speciation can occur in the presence of gene flow. However, recent theory does support the longheld view that population structure and small population size may increase speciation rates, but only
under special conditions and not because of the increased efficacy of drift as suggested by earlier
authors. Rather, low levels of migration among small populations facilitates the rapid accumulation
of beneficial mutations that indirectly cause hybrid incompatibilities.
Keywords
gene flow; introgression; population size; population subdivision; reproductive isolation; selection;
selective sweep; speciation
Introduction
NIH-PA Author Manuscript
The study of speciation was formally initiated by the publication of Darwin’s famous
monograph, On the Origin of Species by Natural Selection, which posited that species
differences are caused by natural selection (Darwin, 1859). However, Darwin did not fully
explain how species differed from locally adapted populations or how conspecific populations
were able to evolve as a unit. These issues were clarified by the biological species concept
(Dobzhansky, 1937; Mayr, 1942), which emphasized the importance of gene flow for holding
species together and reproductive barriers for keeping them apart. Mayr (1942, 1954) also
argued forcefully that geographic isolation of populations was critical to species formation and
that speciation was stimulated by small population size. However, note that Mayr’s (1954)
paper on founder effect speciation was preceded by Lewis (1953), who argued that speciation
in Clarkia was facilitated by founder events and the fixation of chromosomal rearrangements
in peripheral populations.
Although these ideas provided a framework for speciation studies over the past 60 yr, they
have not been universally accepted. Indeed, there is inherent conflict between Darwin’s focus
Author for correspondence: Loren H. Rieseberg, Tel: +1 (812) 855 7614, Fax: +1 (812) 855 6705, Email: [email protected].
Rieseberg et al.
Page 2
NIH-PA Author Manuscript
on selection, which operates most efficiently in large populations, and Mayr’s emphasis on the
creative role of small population size. Likewise, some authors have argued that there is too
little gene flow to hold species together (Ehrlich & Raven, 1969), and others have shown that
species boundaries often are semipermeable to introgression (Rieseberg & Wendel, 1993;
Arnold, 1997). It also has become increasingly apparent that geographic isolation is not
required for speciation (McNeilly & Antonovics, 1968), and that the effects of population size
and structure on speciation rates are complex (Rice & Hostert, 1993; Orr & Orr, 1996; Church
& Taylor, 2002).
In this paper, we briefly review evidence from recent theoretical and empirical studies that bear
on these problems. We focus mostly on aspects that remain controversial or where we
personally have made contributions toward their resolution. When possible, we have tried to
emphasize botanical contributions.
Selection and speciation
NIH-PA Author Manuscript
Many kinds of evidence have been used to assess the role of selection in the evolution of species
differences and reproductive isolation. These include: analyses of patterns of selection in
contemporary hybrid zones or experimental hybrids; the application of Orr’s (1998)
quantitative trait locus (QTL) sign test to species differences, which allows the history of
selection on a given trait to be determined; analyses of the molecular evolution of genes that
contribute to reproductive isolation; experimental population genetic studies; and comparative
analyses. Because of space constraints, we focus on the first three approaches. Note, however,
that detailed reviews of results from experimental population genetic studies and comparative
analyses have been published elsewhere (Rice & Hostert, 1993; Barraclough & Nee, 2001;
Schilthuizen, 2001). Also, it should be evident that different methods may apply best to
different questions or traits. For example, neither phenotypic selection analysis nor the QTL
sign test are useful for studying the role of selection in the origin of postzygotic barriers such
as hybrid sterility or inviability.
Contemporary patterns of selection
The simplest and most direct way to test for a role for selection in speciation is to ask whether
species differences are currently being maintained by selection. If so, this would imply that the
differences arose by selection as well. Such inferences are not robust, however, because of
temporal and spatial variation in the strength and direction of selection (Schemske &
Bierzychudek, 2001).
NIH-PA Author Manuscript
Most selection studies at the species level employ experimental or natural hybrids that have
been placed in the habitat of one or both parental species. A recent survey of the literature
yielded 47 such studies, of which 31 involved plants and 16 involved animals (Lexer et al.,
2003). Although evidence of selection for or against the hybrids was detected in all but one of
the studies, only eight experiments reported the strength of selection on individual phenotypic
traits. Fortunately, all eight were from plants. The eight studies generated 149 estimates of
directional selection gradients, of which 56% were significant, with a mean selection gradient
of 0.12 ± 0.01. By contrast, only 25% of trait polymorphisms within populations were found
to be under significant selection in a review of selection in natural populations (Kingsolver et
al., 2001). These results imply that the majority of trait differences between species (at least
for the taxa studied here) are indeed maintained by selection and that selection is quite strong.
Note, however, that results from selection studies are sensitive to sample size (Kingsolver et
al., 2001). With larger samples, the proportion of traits under significant selection is likely to
increase, although estimates of the strength of selection may decrease.
New Phytol. Author manuscript; available in PMC 2008 December 12.
Rieseberg et al.
Page 3
Historical patterns of selection: the QTL sign test
NIH-PA Author Manuscript
As mentioned earlier, the fact that species differences are currently maintained by selection
doesn’t necessarily prove that the differences were caused by selection. Thus, methods that
can estimate the history of selection on a trait or gene are needed to fully understand the
speciation process. One such method is based on the direction of effects of QTLs that contribute
to phenotypic differences (Orr, 1998). If a trait has a continuous history of directional selection,
then QTL effects should be in the same direction within a line. By contrast, if a trait has diverged
under neutrality, QTLs with opposing effects should be common. Orr (1998) has formalized
these predictions with the creation of a QTL sign test that compares observed numbers of ±
QTLs in a given line with those expected under neutral conditions.
A recent application of Orr’s (1998) sign test to the QTL literature in both wild and
domesticated plants (Rieseberg et al., 2002) revealed that, as predicted by theory,
domestication traits had a much lower proportion of opposing QTLs (0.06 ± 0.03, least square
mean ± standard error) than did other kinds of traits that were segregating in the same wild
domesticated mapping populations (0.18 ± 0.03). Analyses of crosses involving taxonomically
diverse wild populations also revealed widespread selection; all categories of traits had QTL
proportions that deviated significantly from neutral expectations (Rieseberg et al., 2002). Thus,
directional selection appears to be a major contributor to phenotypic differentiation in
essentially all kinds of traits and organisms.
NIH-PA Author Manuscript
There are caveats associated with this conclusion, however. In particular, QTL magnitudes
were not included in this analysis, and Orr (1998) has pointed out that major QTLs fixed during
an initial bout of natural selection may overshoot the phenotypic optimum, and minor QTLs
in the opposite direction may evolve later to bring the trait back to the phenotypic optimum.
Although the signature of directional selection is widespread, there was variation among trait
categories, suggesting that certain kinds of traits were exposed to stronger and more consistent
selection than other trait categories. For the purposes of this review, the most interesting result
was that intraspecific traits had almost twice the proportion of opposing QTLs as interspecific
traits (0.25 vs 0.14; P = 0.02), suggesting that species differences are more likely to be a product
of divergent selection than intraspecific differences.
NIH-PA Author Manuscript
We have extended this analysis further in the present paper by asking whether traits that
contribute directly to assortative mating (e.g. flowering time; flower color, pollination
syndrome, etc.) are more or less strongly selected than ordinary species differences. This was
accomplished by adding the category ‘assortative mating trait vs other’ to the ANOVA model
described by Rieseberg et al. (2002). As expected, the proportion of opposing QTLs (0.12 ±
0.15, least square mean ± tandard error) for assortative mating traits was significantly lower
than that predicted under a model of neutral divergence. Indeed, this is the lowest proportion
of opposing QTLs observed for any trait category in comparisons involving wild populations.
Nonetheless, QTL proportions did not differ significantly between assortative mating traits and
ordinary species differences. Thus, although most assortative mating traits appear to have
diverged as a consequence of divergent natural selection, there is no evidence that they have
been more strongly or consistently selected than other kinds of species differences.
Historical patterns of selection: molecular evolution of isolation genes
Another method for assessing historical patterns of selection during the speciation process is
to examine the molecular evolution of genes that contribute to reproductive isolation. This
approach tests for selection by asking whether there is an excess of nonsynonymous
substitutions (dN) relative to neutral expectations (i.e. dN/dS > 1), significant variation in
New Phytol. Author manuscript; available in PMC 2008 December 12.
Rieseberg et al.
Page 4
substitution rates among codons within a gene (Nielsen & Yang, 1998), or reduced variability
in the gene and flanking sequences caused by a recent selective sweep (Wang et al., 1999).
NIH-PA Author Manuscript
NIH-PA Author Manuscript
Only a handful of genes have been identified that are known to contribute to reproductive
isolation, and most of these are from animals. These ‘isolation genes’ can be divided into two
groups, based on whether they contribute to pre or postzygotic isolation. Those involved in
prezygotic isolation include: period, a clock gene which modifies song rhythm in
Drosophila (Wheeler et al., 1991) and timing of mating behavior in both Drosophila (Tauber
et al., 2003) and melon fly (Miyatake et al., 2002); bindin, a gamete recognition protein in sea
urchins that mediates species-specific attachment to an egg-surface receptor during fertilization
(Metz & Palumbi, 1996); lysin, a sperm protein that species-specifically creates a hole in the
egg envelope during abalone fertilization (Lee et al., 1995); VERL, the egg vitelline envelope
receptor for lysin (Galindo et al., 2003); desat2, which is responsible for a cuticular
hydrocarbon pheromone polymorphism that contributes to behavioral isolation between
geographic races of Drosophila melanogaster (Takahashi et al., 2001); and S-RNase-based
self-incompatibility (SI) that causes unilateral interspecific incompatibility in Nicotiana
(Hancock et al., 2003). Other plant genes that are likely to contribute to reproductive isolation,
but for which definitive proof is lacking, include rapidly evolving pollen coat proteins that may
mediate species-specific pollen recognition (Mayfield et al., 2001), flowering time genes such
as FRIGIDA (Johanson et al., 2000) and Hd1 (Yano et al., 2000), and flower color genes such
as the anthocyanin2 (an2) locus that is the main determinant of floral color differences between
Petunia integrifolia and P. axillaris (Quattrocchio et al., 1999).
The six genes proven to contribute to prezygotic isolation all show the signature of positive
selection. For bindin, lysin, VERL, and SI, the evidence of selection is in the form of an excess
of nonsynonymous substitutions across all or part of the studied gene. The rapid rate of protein
evolution exhibited by these genes is likely a result of continuous coevolution of the gamete
recognition system, driven by sexual selection or sexual conflict for the animal reproductive
proteins (bindin, lysin, and VERL), and frequency dependent selection for SI. By contrast,
period and desat2 are not unusually variable and evidence for positive selection comes from
evidence of selective sweeps (i.e. reduced variation) associated with the gene or selected
mutation (Ford & Aquadro, 1996; Wang & Hey, 1996; Takahashi et al., 2001). Note that in
all cases, selection is not for reproductive isolation. Rather, isolation appears to have evolved
as a byproduct of positive selection for some other function.
NIH-PA Author Manuscript
We are aware of four genes that have been shown to contribute to postzygotic isolation,
including Odysseus (Ods), which induces hybrid male sterility in crosses between Drosophila
mauritiana and D. simulans (Ting et al., 1998); Hybrid male rescue (Hmr), which causes
lethality and female sterility in hybrids among D. melanogaster and its sibling species (Barbash
et al., 2003); Nup96, a nuclear pore protein that causes hybrid lethality in crosses between D.
melanogaster and D. simulans (Presgraves et al., 2003); and xmrk-2, a growth factor receptor
that is overexpressed in hybrids between Xiphophorus (platyfish) species, causing tumor
development (Froschauer et al., 2001). Three of these (Ods, Hmr and Nup96) evolve rapidly
and display an excess of replacement substitutions, indicative of positive Darwinian selection.
Unlike the gamete recognition proteins described above, little is known about the selective
forces causing the high rates of protein evolution at these loci, although Ting et al. (1998)
speculate that sexual selection may drive rapid sequence evolution in Ods because it is
expressed in Drosophila testis. As far as we are aware, no evidence for positive selection has
been reported for xmrk-2, and variation (small chromosomal rearrangements) in regulatory
rather than coding sequence appears to be responsible for the hybrid inviability associated with
this gene (Froschauer et al., 2001).
New Phytol. Author manuscript; available in PMC 2008 December 12.
Rieseberg et al.
Page 5
NIH-PA Author Manuscript
In sum, 9 of the 10 genes currently known to contribute to reproductive isolation appear to
have diverged as a consequence of Darwinian natural selection, a result that agrees nicely with
the prominent role for selection implied by the experimental selection studies and the QTL
sign test. There is a major caveat, however. The identified ‘isolation’ genes are biased toward
animals and toward reproductive proteins, which may diverge more rapidly than other kinds
of characters because of sexual selection or sexual conflict (Wyckoff et al., 2000). Although
reproductive proteins in plants also appear to evolve rapidly (Mayfield et al., 2001), other kinds
of traits that contribute to reproductive isolation (e.g. differences in morphological, life history,
and physiological adaptations to different habitats) seem less likely to be involved in the
coevolutionary ‘chases’ that lead to rapid protein divergence. These ‘ordinary’ species
differences probably are caused by natural selection as well, but it may be that the selective
sweeps will sometimes be too ancient to detect them.
NIH-PA Author Manuscript
A prominent role for natural selection does not necessarily mean that stochastic forces are
unimportant. One means by which genetic drift might contribute, for example, is through the
fixation of neutral or underdominant chromosomal rearrangements (White, 1978). Unless
strongly underdominant (and thus unlikely to be fixed in the first place), chromosomal
rearrangements will act mostly to reduce effective gene flow rates in regions close to
chromosomal breakpoints (Rieseberg, 2001). As a consequence, selected differences are
predicted to accumulate most quickly on rearranged chromosomes (Navarro & Barton,
2003a), a prediction which has been confirmed for divergence among species of Drosophila
(Noor et al., 2001) and between humans and chimpanzees (Navarro & Barton, 2003b).
Rearrangements may be particularly important during the early stages of speciation when
effective gene flow rates may otherwise be high enough to prevent differentiation at most loci.
Of course, it may be that many rearrangements become established as a consequence of
selection favoring a particular combination of alleles (Charlesworth & Charlesworth, 1980) or
position effects, rather than drift. Unfortunately, little is known about the evolutionary forces
underlying the establishment of chromosomal rearrangements, although it does appear that
transposable elements often contribute to their origin (Gray, 2000).
NIH-PA Author Manuscript
Stochastic forces may also play a significant role in the divergence of duplicate genes following
polyploidy or segmental duplication, and models have been developed for the evolution of
hybrid incompatibilities as a result of reciprocal silencing of duplicated genes (Werth &
Windham, 1991) or divergent resolution of regulatory sequences (Lynch & Force, 2000).
Although both phenomena are well-documented in the literature, neither has yet been shown
to cause hybrid incompatibilities. Plant genomes tend to be more redundant than those of
animals, however, and may therefore provide a more fertile substrate for evolutionary changes
that involve duplicate genes.
Gene flow and speciation
Although battered and scarred from constant criticism over the past 70 yr, the biological species
concept remains the most widely employed species concept in evolutionary biology, and its
focus on gene flow and reproductive barriers continues to provide the motivation for most
empirical and theoretical studies of speciation (Schilthuizen, 2001). It is not possible to list the
many criticisms of the biological species concept in the space provided here, let alone respond
to them. A frequent objection by phylogeneticists is that the ability to interbreed is
symplesiomorphic. However, this criticism only makes sense if the phenotypic clusters
recognized by naturalists represent monophyletic lineages, and this is demonstrably not the
case. Not only are species often of recurrent origin (Levin, 2001), but newly derived sister
species follow a common time course of change subsequent to speciation of polyphyly →
paraphyly → monophyly, with 4 N generations required to achieve reciprocal monophyly for
New Phytol. Author manuscript; available in PMC 2008 December 12.
Rieseberg et al.
Page 6
NIH-PA Author Manuscript
organellar genes and even longer times for nuclear genes (Avise, 2000). More serious criticisms
relate to the observation that there appears to be too little gene flow among populations of some
species to hold them together (Ehrlich & Raven, 1969) and too much gene flow among other
species to keep them apart. Among Eukaryotes, these problems are perhaps most pronounced
in plants, resulting in the widespread rejection of the biological species concept (and species
reality) by botanists (Mishler, 1999). Below, we show that the problem of too little gene flow
is ameliorated by recognition that even very low levels of gene flow are sufficient for the spread
of advantageous alleles and that the problem of too much flow is mitigated by recognition that
whole genome isolation is not required for species divergence.
Too little gene flow
Students of speciation have primarily focused on the conservative role of gene flow, in which
high levels of gene flow (Nem > 4, where Nem is the effective number of migrants per
generation) serve to homogenize populations at neutral loci (Hartl & Clark, 1997). It was
recognized more than three decades ago, however, that levels of gene flow in many species
are not nearly this high (Ehrlich & Raven, 1969). Indeed, for many plant and animal species,
estimates of Nem fall well below one (Fig. 1), the level of gene flow required to prevent
divergence at neutral loci (Wright, 1931).
NIH-PA Author Manuscript
Consideration of the creative role of gene flow as a mechanism for the spread of advantageous
alleles offers a potential solution to this problem (Rieseberg & Burke, 2001). Only very low
levels of gene flow are required for the spread of advantageous alleles and fixation times are
much less than for their neutral counterparts (Slatkin, 1976). Thus, it is conceivable that
species’ populations could remain connected through repeated selective sweeps of favored
mutations and associated hitchhiking events or ‘genetic draft’ (Gillespie, 2001).
Is this scenario likely? In low gene flow species, population subdivision greatly reduces the
rate of allelic spread, particularly for weakly selected or neutral alleles (Slatkin, 1976;
Whitlock, 2003). Thus, one concern is whether a favored allele will spread to fixation before
it goes extinct. A second concern is whether selective sweeps are frequent enough to produce
cohesion. If they are rare or restricted to a handful of loci, the level of connectedness might
not be sufficient to account for the apparent cohesiveness observed for many species in nature.
NIH-PA Author Manuscript
We are aware of only two studies that have modeled the effects of population subdivision on
the probability of and/or time to fixation of beneficial alleles (Slatkin, 1976; Whitlock, 2003).
Slatkin (1976) employed a one-dimensional stepping stone model to estimate the time it would
take for a favored allele to spread across the range of a species that comprised 20 demes. He
showed that the rate of spread of a beneficial mutation depended strongly on the strength of
selection and the long-distance migration rate. We have extended Slatkin’s model by computer
simulation to include a second dimension and a leptokurtic dispersal function (Fig. 2). Our
results confirm that moderately selected alleles will spread rapidly across a species range
despite low levels of migration, although the time to fixation is somewhat reduced relative to
that of Slatkin’s calculations, presumably as a result of the use of the leptokurtic dispersal
function.
Whitlock (2003) provides estimates of both the time to and probability of fixation for beneficial
alleles in island, stepping stone, and extinction-recolonization models. Unfortunately, only a
subset of parameter space is illustrated by his figures. Nonetheless, his results appear to be
fully compatible with Slatkin’s work. As expected, the strength of selection has a huge impact
on both the probability of and time to fixation. Weakly selected alleles spread at a glacial pace
and almost always go extinct, whereas strongly beneficial alleles spread much faster and have
a much higher probability of fixation.
New Phytol. Author manuscript; available in PMC 2008 December 12.
Rieseberg et al.
Page 7
NIH-PA Author Manuscript
In sum, the effects of population subdivision are to greatly increase fixation times relative to
panmictic populations with a slight positive effect on fixation probabilities. More importantly,
however, population subdivision magnifies the differences in time to fixation for strongly and
weakly selected alleles. Thus, weakly selected alleles that spread to fixation in panmictic
populations are less likely to do so in subdivided populations, possibly biasing fixed
interspecific differences toward major genes.
NIH-PA Author Manuscript
So are selection coefficients for mutations underlying species differences large enough to
validate a model in which species are held together by the spread of beneficial alleles?
Rieseberg & Burke (2001) generated crude estimates of what selection coefficients might be
by calculating the average percentage variance explained by QTLs that contribute to species
differences in plants, multiplying by selection differentials for phenotypic traits in wild plant
populations and then halving to account for diploidy. However, their calculations were based
on a very small number of studies. We have updated this analysis with the inclusion of data
from Kingsolver et al. (2001), who report 993 linear selection gradients and 753 selection
differentials for phenotypic traits in 62 studies of natural populations; Lexer et al. (2003), who
provide 149 estimates of directional selection gradients and 27 selection differentials from 8
studies of experimental hybrids in natural populations; and C. L. Morjan & L. H. Rieseberg
(unpublished), in which 133 selection gradients and 96 selection differentials from
experimentally manipulated or disturbed populations were compiled from 26 studies. The
expanded analysis corroborates the general conclusions of Rieseberg & Burke (2001).
Selection coefficients associated with moderate to large QTLs are likely to be large enough to
ensure rapid spread and fixation, whereas very minor QTLs (PVE < 0.01) may be trapped in
local populations and are less likely to contribute to fixed differences between species (Fig.
3). Of course, if most phenotypic differentiation were to occur in a local population, followed
by range expansion, then even small QTLs might contribute to species differences in such a
scenario.
Another issue alluded to earlier concerns the frequency of selective sweeps. If they are frequent
and involve genes scattered across the genome, it is plausible that, in concert with hitchhiking
effects, the whole genome could remain connected. On the other hand, if cohesion is limited
to a small number of major genes, then we could have a situation in which a species is evolving
collectively at a handful of genes while simultaneously diverging at other loci through drift or
local adaptation (Rieseberg & Burke, 2001).
NIH-PA Author Manuscript
Comparative sequencing studies are beginning to shed light on this question. For example,
Smith & Eyre-Walker (2002) have recently shown that 45% of all amino-acid substitutions
between Drosophila species appear to have been fixed by selection. We do not yet have
comparable data from plants, although Barrier et al. (2003) suggest that c. 5% of genes
differentiating Arabidopsis thaliana and A. lyrata have diverged as a consequence of positive
selection. Comparative EST sequencing of wild sunflowers (S. Church et al., unpublished)
suggest a similar percentage of sunflower genes are under selection. These values, if reliable,
are consistent with the maintenance of genome-wide species cohesion through repeated
selective sweeps and genetic hitchhiking. Of course this doesn’t rule out the possibility that
some loci are simultaneously diverging through local selection. Indeed, fairly weak selection
can overcome the effects of migration at a given locus, so this is expected.
Too much gene flow
The fact that many otherwise ‘good’ species continue to exchange genetic material long after
they have embarked on independent and irreversible evolutionary trajectories has been known
and accepted by botanists for close to a century (Ostenfeld, 1927; Grant, 1981). Thus, there
seems little reason to exhaustively review the literature on the permeability of species barriers
other than to comment that the evidence for introgression is becoming increasingly widespread
New Phytol. Author manuscript; available in PMC 2008 December 12.
Rieseberg et al.
Page 8
NIH-PA Author Manuscript
and robust (Grant, 1981; Rieseberg & Wendel, 1993; Arnold, 1997; Wendel & Doyle, 1998)
and now extends beyond plants to animal groups such as Drosophila in which introgression
was once thought to be rare or absent (Wang et al., 1997).
Coincident with more rigorous evidence of introgression has been the development of theory,
which indicates that divergence in the presence of gene flow is possible as long as the strength
of selection at a given locus exceeds the migration rate (Maynard Smith, 1966). This is not an
unlikely condition, resulting in widespread acceptance of models for sympatric and parapatric
speciation. The same basic theory applies to zones of secondary contact (i.e. hybrid zones)
between species that have diverged in allopatry, but with one main difference. In hybrid zones
there often are genome–wide associations or linkage disequilibrium between the traits and
genes characteristic of a given taxon (Barton & Hewitt, 1985). That is, multilocus parental
genotypes are over-represented relative to recombinant genotypes. Although some linkage
disequilibrium among selected traits is likely in zones of primary contact, genome wide
disequilibrium is unlikely until the very latest stages of sympatric and parapatric speciation.
NIH-PA Author Manuscript
The primary consequence of linkage disequilibrium is to retard the movement of genes across
larger chromosomal segments, or across the entire genome if many genes contribute to
reproductive isolation. However, in most hybrid zones, the number of genes contributing to
reproductive isolation is not sufficient to create genome-wide reproductive isolation (Barton
& Hewitt, 1985; Rieseberg et al., 1999). As a consequence neutral or universally favorable
alleles move easily across the species barrier, unless they are tightly linked to negatively
selected alleles. This contrasts with the restricted movement of alleles that contribute to
reproductive isolation and genes or markers tightly linked to them, whose movement across
the zone will decline in proportion to the selection:recombination ratio (Barton, 1979). That
is, strongly selected alleles will have a greater impact on linked loci than will weakly selected
alleles, and tightly linked loci will be more affected than unlinked or loosely linked loci. For
example, Ting et al. (2000) has shown that the Drosophila sterility locus, Odysseus, protects
just 2 kb of the genome from introgression. These findings have led to a renewed interest in
the role of chromosomal rearrangements as reproductive barriers, because rearrangements
reduce effective recombination rates, thereby extending the effects of linked isolation genes
(Rieseberg, 2001).
In sum, both empirical and theoretical data indicate that reproductive barriers are likely to be
semipermeable to gene flow in recently diverged species or taxa in which hybrid
incompatibilities evolve slowly. Because the term ‘reproductive isolation’ implies wholegenome isolation (Wu, 2001), evidence of widespread introgression in both plants and animals
does appear to conflict with a strict interpretation of the biological species concept.
NIH-PA Author Manuscript
How should evolutionary biologists respond to this conflict? One possibility would be to apply
the biological concept very strictly, with absolute reproductive isolation required for species
status. This approach is logically attractive, but stumbles in its application to real organisms,
primarily because biological barriers to successful reproduction evolve at hugely different rates
in different organismal groups (Rieseberg, 2001), and such barriers are not a requirement for
divergent evolution. Strict implementation would lead to such absurdities as according similar
taxonomic rank to cryptic species of Drosophila, genera of plants and birds, and phyla of
bacteria.
Another possibility would be to employ alternative species concepts that do not rely on
reproductive isolation. A variety of evolutionary forces are assumed to contribute to species
cohesion (e.g. common descent, stabilizing and parallel selection, genetic constraints, gene
flow/reproductive isolation), and species concepts have been devised that emphasize one or
all these mechanisms (Templeton, 1989). However, gene flow/reproductive isolation is unique
New Phytol. Author manuscript; available in PMC 2008 December 12.
Rieseberg et al.
Page 9
NIH-PA Author Manuscript
among these cohesive mechanisms in that it acts primarily at the population and species level.
Selection, by contrast, acts most strongly on individuals, and common descent and genetic
constraint contribute equally to cohesion at all taxonomic levels. Thus, there is considerable
justification for focusing on the only distinguishing property of species (gene flow/reproductive
isolation), while at the same time recognizing its imperfections as a species diagnostic.
If complete reproductive isolation is not a requirement for species status, then how much is
required? One answer to this question relates to the key difference between locally adapted
populations/geographic races and species: Divergence among locally adapted populations and
even geographic races is almost always ephemeral, whereas evolutionary divergence among
species is preserved by reproductive isolation (Futuyma, 1989). Thus, the amount of
reproductive isolation required for species status is simply that necessary to preserve the
evolutionary changes associated with population systems following changes in the location of
suitable habitat (e.g. following climatic shifts) or in the proximity of other population systems
or species (i.e. changes in levels of gene flow).
NIH-PA Author Manuscript
This may not seem helpful, but because reproductive barriers themselves are sensitive to shifts
in selection or gene flow, the permanence of species may be equated with the permanence of
the reproductive barriers that protect them. Premating barriers may be erased by shifts in
patterns of selection, and two-locus hybrid incompatibilities are sensitive to changes in gene
flow. By contrast, chromosomal barriers and complex Dobzhansky–Muller (Dobzhansky,
1937; Muller, 1942) incompatibilities (epistatic incompatibilities that appear in hybrids as a
result of interaction between alleles that did not lower fitness when they were individually
substituted within the diverging lineages) tend to be robust to either kind of challenge and attest
to the likely longevity or permanence of the population system associated with them. Thus,
while speciation often is initiated by the development of premating barriers, the evolution of
complex postzygotic barriers is required to set these differences in ‘concrete’. In this context,
note that plant species seemingly isolated by prezygotic barriers alone often show considerable
transmission ratio distortion in hybrids (Whitkus, 1998; Fishman et al., 2001), implying that
postzygotic barriers are present as well.
Population structure and speciation
NIH-PA Author Manuscript
As alluded to in the introduction, early models of speciation emphasized the importance of
geographic isolation, because gene flow was thought to prevent geographically proximal
populations from diverging through drift and selection (Mayr, 1942). However, studies over
the past 40 yr have shown time and again that speciation may occur despite gene flow between
populations (McNeilly & Antonovics, 1968; Caisse & Antonovics, 1978; Church & Taylor,
2002). These studies have inspired the development of spatially explicit models to explore
more complex and realistic predictions of the effects of selection, drift, and gene flow on the
likelihood, rate, and tempo of speciation.
One of the earliest studies to incorporate spatially explicit models of population structure and
gene flow showed that speciation could occur between connected populations, regardless of
gene flow (Caisse & Antonovics, 1978). The authors used a computer simulation of 10 linearly
arranged populations that shared genes through stepping stone or exponential migration
patterns to investigate the evolution of reproductive isolation along a cline. Their model
confirmed the earlier study of Maynard Smith (1966), demonstrating that divergence will occur
when the strength of selection at a given locus (s) is greater than the migration rate (m).
Furthermore, if the gene under selection also caused assortative mating (or was linked to a gene
causing assortative mating) the development of reproductive isolation was straightforward
(Caisse & Antonovics, 1978). Reproductive isolation also occurred in the absence of linkage
or pleiotropy, but selection against hybrids had to be strong. These early models, and others
New Phytol. Author manuscript; available in PMC 2008 December 12.
Rieseberg et al.
Page 10
NIH-PA Author Manuscript
like them (Karlin & McGregor, 1972), have shown repeatedly that reproductive barriers may
arise and be maintained between hybridizing populations provided that selection against
hybrids is sufficiently strong relative to migration (Dieckmann & Doebeli, 1999; Kondrashov
& Kondrashov, 1999; Doebeli & Dieckmann, 2003).
NIH-PA Author Manuscript
More recently, speciation models have investigated the role of spatial structure and gene flow
on the evolution of Dobzhansky–Muller type incompatibilities, which may be the most
common cause of postzygotic reproductive isolation. In these models, all populations are
initially identical (genotype AAbb). Mutations occur independently across populations, with
one population becoming AABB and a second population becoming aabb. There is no reduced
fitness within either population as a result of these mutations; however, the hybrid genotype,
AaBb, has drastically reduced fitness or is inviable. These populations are now reproductively
isolated as a consequence of random mutation and genetic drift or natural selection acting to
fix alternate mutations. However, reproductive isolation between the populations was not the
target of selection. Once the barrier is established though, selection against hybrids maintains
the newly diverged species. Empirical studies have confirmed that hybrid sterility or inviability
can result from such multi–locus interactions (Orr, 1997). These incompatibilities can
accumulate rapidly in geographically isolated populations, in a ‘snowball effect’, which is the
square of the time since the separation of the populations (Orr, 1995). A similar effect occurs
in spatially structured populations, even with low levels of migration, where the number of
pairwise incompatibilities between populations can eventually increase quadratically
(Kondrashov, 2003).
If the accumulation of Dobzhansky–Muller incompatibilities represents a plausible scenario
for speciation, what are the consequences of population subdivision on the patterns and timing
of speciation events? Recent theoretical models have focused on this issue, with conflicting
results. Early students of speciation postulated that speciation was most likely to occur when
species were subdivided into small populations (Wright, 1931; Lewis, 1953; Mayr, 1954)
because of the increased efficiency of genetic drift. However, Orr & Orr (1996) showed that,
if selection acts to fix alternate alleles in different populations, the time to speciation increases
as population size decreases. This results from the scattering of new mutations across many
small populations in a subdivided species. As a consequence, more mutations must arise in a
subdivided species before any two populations accumulate a sufficient number of mutations
to be incompatible.
NIH-PA Author Manuscript
However, more recent models show that the relationship between population size and
speciation rate breaks down when migration is introduced (Gavrilets et al., 2000b; Church &
Taylor, 2002; Gavrilets & Gibson, 2002). One such model considers a population subdivided
into demes of equal size and arranged spatially in a circular stepping stone model (Church &
Taylor, 2002). Initially, all demes were identical genetically and diverged through the
accumulation of beneficial mutations. Low levels of migration between neighboring demes
was allowed, with all compatible, positively selected genes entering and going to fixation in a
deme. The results of this model agreed with Orr & Orr (1996) in that small deme size inhibited
speciation in a strictly allopatric model. However, this result was reversed when low levels of
migration were introduced; speciation rates accelerate with nonzero rates of migration (Fig.
4). Simply put, migration increased the rate at which a given deme accumulated mutations,
with subsets of demes within the population evolving in concert.
Other models of parapatric speciation have elucidated geographic patterns of speciation in
spatially structured populations (Gavrilets et al., 2000b). It is thought that, in nature, most
migration occurs between neighboring populations, resulting in geographic differentiation of
populations (Endler, 1977). The extreme cases of this are found in so-called ‘ring species’,
which exchange migrants only with their neighbors, allowing high levels of genetic
New Phytol. Author manuscript; available in PMC 2008 December 12.
Rieseberg et al.
Page 11
NIH-PA Author Manuscript
differentiation among geographically isolated populations (Wake, 1997). In these situations,
new species may form as a result of extensive geographic differentiation, despite low levels of
gene flow. Theoretical studies suggest that speciation in such circumstances can occur as
rapidly as a few hundred to a few thousand generations in spite of the exchange of several
migrants per generation between neighboring populations (Gavrilets et al., 2000b).
Building on parapatric models, recent studies have also incorporated metapopulation
dynamics, the phenomenon of local extinction and colonization of populations that is common
among natural populations (Hanski & Gilpin, 1997). In these models, divergence among
populations is erased by extinction and recolonization rather than migration (Gavrilets et al.,
2000a). It follows that increased rates of population turnover (extinction and colonization)
decreases the rate of speciation (Gavrilets et al., 2000a). More generally, the model develops
an excellent framework for modeling speciation in a metapopulation, but more realistic spatial
structures and migration models need to be incorporated to make this model more biologically
relevant.
NIH-PA Author Manuscript
Analyses of Dobzhansky–Muller incompatibilities in clinal populations generate similar
results (Bengtsson, 1985; Barton & Bengtsson, 1986; Gavrilets, 1997). Although persistent
migration can break down barriers to gene exchange between populations (Barton &
Bengtsson, 1986), speciation can result from the accumulation of Dobzhansky–Muller
incompatibilities if migration rates are low and selection against intermediates is strong
(Gavrilets, 1997). Moreover, genotypes of intermediate populations in a cline tend to be quite
different from those that would be formed by hybridization between the more geographically
distant populations at either end of the cline (Gavrilets, 1997).
Empirical studies of spatially structured populations or species often reveal patterns of
ecological and genetic divergence that are strongly correlated with geographic structure. These
include parallel adaptation to the same range of habitats (Turesson, 1922, 1925, 1930; Clausen
et al., 1940, 1948), as well as parallel geographic patterns in the distribution of alleles among
populations and species (Soltis et al., 1997; Avise, 2000; Barraclough & Vogler, 2000; Church
et al., 2003). The theoretical models discussed above indicate that this geographic variation
could in fact contribute to speciation. They also provide a sounder basis for interpreting
geographic patterns seen in nature. The primary caveat associated with many of these studies
relates to their focus on Dobzhansky–Muller interactions, which, for reasons described earlier,
seem unlikely to initiate speciation. More complex models that include various kinds of
isolating mechanisms and empirically validated assumptions regarding mutation rate,
migration rate, and population size and structure will be needed before confident conclusions
can be made regarding the role of geographic structure in speciation.
NIH-PA Author Manuscript
Acknowledgements
The authors’ research in these areas is funded by the U.S. National Institutes of Health (GM059065 to L.H. Rieseberg),
and the U.S. National Science Foundation (Postdoctoral Research Grants in Bioinformatics to S.A. Church and C.L.
Morjan). We thank Troy Wood for constructive comments on the manuscript.
References
Arnold, ML. Natural hybridization and evolution. New York, USA: Oxford University Press; 1997.
Avise, JC. Phylogeography. Cambridge, MA, USA: Harvard University Press; 2000.
Barbash DA, Siino DF, Tarone AM, Roote J. A rapidly evolving MYB-related protein causes species
isolation in Drosophila. Proceedings of the National Academy of Sciences, USA 2003;100:5302–
5307.
Barraclough TG, Nee S. Phylogenetics and speciation. Trends in Ecology and Evolution 2001;16:391–
399. [PubMed: 11403872]
New Phytol. Author manuscript; available in PMC 2008 December 12.
Rieseberg et al.
Page 12
NIH-PA Author Manuscript
NIH-PA Author Manuscript
NIH-PA Author Manuscript
Barraclough TG, Vogler AP. Detecting the geographic pattern of speciation from species-level
phylogenies. American Naturalist 2000;155:419–434.
Barrier M, Bustamante CD, Yu JY, Purugganan MD. Selection on rapidly evolving proteins in the
Arabidopsis genome. Genetics 2003;163:723–733. [PubMed: 12618409]
Barton NH. Gene flow past a cline. Heredity 1979;43:333–339.
Barton NH, Bengtsson BO. The barrier to genetic exchange between hybridizing populations. Heredity
1986;56:357–376. [PubMed: 3804765]
Barton NH, Hewitt GM. Analysis of hybrid zones. Annual Review of Ecology and Systematics
1985;16:113–148.
Bengtsson, BO. The flow of genes through a genetic barrier. In: Greenwood, JJ.; Harvey, PH.; Slatkin,
M., editors. Evolution essays in honor of John Maynard Smith. Cambridge, UK: Cambridge
University Press; 1985. p. 31-42.
Caisse M, Antonovics J. Evolution in closely adjacent plant populations. IX. Evolution of reproductive
isolation in clinal populations. Heredity 1978;40:371–384.
Charlesworth D, Charlesworth B. Sex differences in fitness and selection for centric fusions between sex
chromosomes and autosomes. Genetical Research 1980;35:205–214. [PubMed: 6930353]
Church SA, Kraus J, Mitchell JC, Church DR, Taylor DR. Evidence for multiple Pleistocene refugia in
the postglacial expansion of the eastern tiger salamander, Ambystoma tigrinum tigrinum. Evolution
2003;57:372–383. [PubMed: 12683533]
Church SA, Taylor DR. The evolution of reproductive isolation in spatially structured populations.
Evolution 2002;56:1859–1862. [PubMed: 12389731]
Clausen, JD.; Keck, DD.; Hiesey, WM. Experimental studies of the nature of species. I. Effect of varied
environments on western North American plants. Stanford, CA, USA: Carnegie Institute of
Washington Publishers; 1940. p. 520
Clausen, JD.; Keck, DD.; Hiesey, WM. Experimental studies of the nature of species. III. Environmental
responses of climatic races of Achillea. Stanford, CA, USA: Carnegie Institute of Washington
Publishers; 1948. p. 581
Darwin, D. On the origin of species. London, UK: John Murray; 1859.
Dieckmann U, Doebeli M. On the origin of species by sympatric speciation. Nature 1999;400:354–357.
[PubMed: 10432112]
Dobzhansky, TH. Genetics and the origin of species. New York, USA: Columbia University Press; 1937.
Doebeli M, Dieckmann U. Speciation along environmental gradients. Nature 2003;421:259–264.
[PubMed: 12529641]
Ehrlich PR, Raven PH. Differentiation of populations. Science 1969;165:1228–1232. [PubMed:
5803535]
Endler, JA. Geographic variation, speciation, and clines. Princeton, NJ, USA: Princeton University Press;
1977.
Fishman L, Kelly AJ, Morgan E, Willis JH. A genetic map in the Mimulus guttatus species complex
reveals transmission ratio distortion due to heterospecific interactions. Genetics 2001;159:1701–
1716. [PubMed: 11779808]
Ford MJ, Aquadro CF. Selection on X-linked genes during speciation in the Drosophila athabasca
complex. Genetics 1996;144:689–703. [PubMed: 8889530]
Froschauer A, Korting C, Bernhardt W, Nanda I, Schmid M, Schartl M, Volff JN. Genomic plasticity
and melanoma formation in the fish Xiphophorus. Marine Biotechnology 2001;3:S72–S80. [PubMed:
14961302]
Futuyma, DJ. Macroevolutionary consequences of speciation: Inferences from phytophagous insects. In:
Otte, D.; Endler, JA., editors. Speciation and its Consequences. Sunderland, MA USA: Sinauer
Associates; 1989. p. 557-578.
Galindo E, Vacquier VD, Swanson WJ. Positive selection in the egg receptor for abalone sperm lysine.
Proceedings of the National Academy of Sciences, USA 2003;100:4639–4643.
Gavrilets S. Hybrid zones with Dobzhansky-type epistatic selection. Evolution 1997;51:1027–1035.
Gavrilets S, Acton R, Gravner J. Dynamics of speciation and diversification in a metapopulation.
Evolution 2000a;54:1493–1501. [PubMed: 11108578]
New Phytol. Author manuscript; available in PMC 2008 December 12.
Rieseberg et al.
Page 13
NIH-PA Author Manuscript
NIH-PA Author Manuscript
NIH-PA Author Manuscript
Gavrilets S, Gibson N. Fixation probabilities in a spatially heterogeous environment. Population Ecology
2002;44:51–58.
Gavrilets S, Li H, Vose MD. Patterns of parapatric speciation. Evolution 2000b;54:1126–1134. [PubMed:
11005282]
Gillespie JH. Is the population size of a species relevant to its evolution? Evolution 2001;55:2161–2169.
[PubMed: 11794777]
Grant, V. Plant speciation. New York, USA: Columbia University Press; 1981.
Gray YHM. It takes two transposons to tango – transposable-mediated chromosomal rearrangements.
Trends in Genetics 2000;16:461–468. [PubMed: 11050333]
Hancock CN, Kondo K, Beecher B, McClure B. The S-locus and unilateral incompatibility. Philosophical
Transactions of The Royal Society of London Series B-Biological Sciences 2003;358:1133–1140.
Hanski, I.; Gilpin, ME., editors. Metapopulation biology: ecology, genetics, and evolution. San Diego,
CA, USA: Academic Press; 1997.
Hartl, DL.; Clark, AG. Principles of population genetics. underland, MA, USA: Sinauer Associates; 1997.
Johanson U, West J, Lister C, Michaels S, Amasino R, Dean C. Molecular analysis of FRIGIDA, a major
determinant of natural variation in Arabidopsis flowering time. Science 2000;290:344–347.
[PubMed: 11030654]
Karlin S, McGregor J. Application of method of small parameters in multiniche population genetics
models. Theoretical Population Biology 1972;3:180–209.
Kingsolver JG, Hoekstra HE, Hoekstra JM, Berrigan D, Vignieri SN, Hill CE, Hoang A, Gibert P, Beerli
P. The strength of phenotypic selection in natural populations. American Naturalist 2001;157:245–
261.
Kondrashov AS. Accumulation of Dobzhansky–Muller incompatibilities within a spatially structured
population. Evolution 2003;57:151–153. [PubMed: 12643575]
Kondrashov AS, Kondrashov FA. Interactions among quantitative traits in the course of sympatric
speciation. Nature 1999;400:351–354. [PubMed: 10432111]
Lee YH, Ota T, Vacquier VD. Positive selection is a general phenomenon in the evolution of abalone
sperm lysin. Molecular Biology and Evolution 1995;12:231–238. [PubMed: 7700151]
Levin DA. The recurrent origin of plant races and species. Systematic Botany 2001;26:197–204.
Lewis H. The mechanism of evolution in the genus Clarkia. Evolution 1953;7:1–20.
Lexer C, Randell RA, Rieseberg LH. Experimental hybridization as a tool for studying selection in the
wild. Ecology 2003;84:1688–1699.
Lynch M, Force AG. The origin of interspecific genomic incompatibility via gene duplication. American
Naturalist 2000;156:590–605.
Mayfield JA, Fiebig A, Johnstone SE, Preuss D. Gene families from the Arabidopsis thaliana pollen coat
proteome. Science 2001;292:2482–2485. [PubMed: 11431566]
Maynard Smith J. Sympatric speciation. American Naturalist 1966;100:637–650.
Mayr, E. Systematics and the origin of species. New York, USA: Columbia University Press; 1942.
Mayr, E. Change of genetic environment and evolution. In: Huxley, J., editor. Evolution as a process.
London, UK: Allen & Unwin; 1954. p. 156-180.
McNeilly T, Antonovics J. Evolution in closely adjacent plant populations. IV. Barriers to gene flow.
Heredity 1968;23:205–218.
Metz EC, Palumbi SR. Positive selection and sequence rearrangements generate extensive polymorphism
in the gamete recognition protein bindin. Molecular Biology and Evolution 1996;13:397–406.
[PubMed: 8587504]
Mishler, BD. Getting rid of species. In: Wilson, RA., editor. Species: new interdisciplinary essays.
Cambridge, USA: Harvard University Press; 1999. p. 307-316.
Miyatake T, Matsumoto A, Matsuyama T, Ueda HR, Toyosato T, Tanimura T. The period gene and
allochronic reproductive isolation in Bactrocera cucurbitae. Proceedings of The Royal Society of
London Series B-Biological Sciences 2002;269:2467–2472.
Muller HJ. Isolating mechanisms, evolution, and temperature. Biological Symposia 1942;6:71–125.
New Phytol. Author manuscript; available in PMC 2008 December 12.
Rieseberg et al.
Page 14
NIH-PA Author Manuscript
NIH-PA Author Manuscript
NIH-PA Author Manuscript
Navarro A, Barton NH. Accumulating postzygotic isolation genes in parapatry: a new twist on
chromosomal speciation. Evolution 2003a;57:447–459. [PubMed: 12703935]
Navarro A, Barton NH. Chromosomal speciation and molecular divergence – Accelerated evolution in
rearranged chromosomes. Science 2003b;300:321–324. [PubMed: 12690198]
Nielsen R, Yang ZH. Likelihood models for detecting positively selected amino acid sites and applications
to the HIV-1 envelope gene. Genetics 1998;148:929–936. [PubMed: 9539414]
Noor MAF, Grams KL, Bertucci LA, Reiland J. Chromosomal inversions and the reproductive isolation
of species. Proceedings of the National Academy of Sciences, USA 2001;98:12084–12088.
Orr HA. The population genetics of speciation: the evolution of hybrid incompatibilities. Genetics
1995;139:1805–1813. [PubMed: 7789779]
Orr HA. Haldane’s rule. Annual Review of Ecology and Systematics 1997;28:195–218.
Orr HA. Testing natural selection vs. genetic drift in phenotypic evolution using quantitative trait locus
data. Genetics 1998;149:2099–2104. [PubMed: 9691061]
Orr HA, Orr LH. Waiting for speciation – the effect of population subdivision on the time to speciation.
Evolution 1996;50:1742–1749.
Ostenfeld CH. The present state of knowledge on hybrids between species of flowering plants. Journal
of the Royal Horticultural Society 1927;53:31–44.
Presgraves DC, Balagopalan L, Abmayr SM, Orr HA. Adaptive evolution drives divergence of a hybrid
inviability gene between two species of Drosophila. Nature 2003;423:715–719. [PubMed:
12802326]
Quattrocchio F, Wing J, van der Woude K, Souer E, de Vetten N, Mol J, Koes R. Molecular analysis of
the anthocyanin2 gene of petunia and its role in the evolution of flower color. Plant Cell
1999;11:1433–1444. [PubMed: 10449578]
Rice WR, Hostert EE. Laboratory experiments on speciation – what have we learned in 40 years.
Evolution 1993;47:1637–1653.
Rieseberg LH. Chromosomal rearrangements and speciation. Trends in Ecology and Evolution
2001;16:351–358. [PubMed: 11403867]
Rieseberg LH, Burke JM. The biological reality of species: gene flow, selection, and collective evolution.
Taxon 2001;50:47–67.
Rieseberg, LH.; Wendel, J. Introgression and its consequences in plants. In: Harrison, R., editor. Hybrid
zones and the evolutionary process. New York, USA: Oxford University Press; 1993. p. 70-109.
Rieseberg LH, Whitton J, Gardner K. Hybrid zones and the genetic architecture of a barrier to gene flow
between two wild sunflower species. Genetics 1999;152:713–727. [PubMed: 10353912]
Rieseberg LH, Widmer A, Arntz AM, Burke JM. Directional selection is the primary cause of phenotypic
diversification. Proceedings of the National Academy of Sciences, USA 2002;99:12242–12245.
Schemske DW, Bierzychudek P. Evolution of flower color in the desert annual Linanthus parryae: Wright
revisited. Evolution 2001;55:1269–1282. [PubMed: 11525452]
Schilthuizen, M. Frogs, flies, and dandelions. Oxford, UK: Oxford University Press; 2001.
Slatkin, M. The rate of spread of an advantageous allele in a subdivided population. In: Karlin, S.; Nevo,
E., editors. Population genetics and ecology. New York, USA: Academic Press, Inc; 1976. p.
767-780.
Smith NGC, Eyre-Walker A. Adaptive protein evolution in Drosophila. Nature 2002;415:1022–1024.
[PubMed: 11875568]
Soltis DE, Gitzendanner MA, Strenge DD, Soltis PS. Chloroplast DNA intraspecific phylogeography of
plants from the Pacific Northwest of North America. Plant Systematics and Evolution 1997;206:353–
373.
Takahashi A, Tsaur SC, Coyne JA, Wu CI. The nucleotide changes governing cuticular hydrocarbon
variation and their evolution in Drosophila melanogaster. Proceedings of the National Academy of
Sciences, USA 2001;98:3920–3925.
Tauber E, Roe H, Costa R, Hennessy JM, Kyriacou CP. Temporal mating isolation driven by a behavioral
gene in Drosophila. Current Biology 2003;13:140–145. [PubMed: 12546788]
Templeton, AR. The meaning of species and speciation: a genetic perspective. In: Otte, D.; Endler, JA.,
editors. Speciation and its consequences. Sunderland, MA USA: Sinauer Associates; 1989. p. 3-27.
New Phytol. Author manuscript; available in PMC 2008 December 12.
Rieseberg et al.
Page 15
NIH-PA Author Manuscript
NIH-PA Author Manuscript
NIH-PA Author Manuscript
Ting CT, Tsaur SC, Wu CI. The phylogeny of closely related species as revealed by the genealogy of a
speciation gene Odysseus. Proceedings of the National Academy of Sciences, USA 2000;97:5313–
5316.
Ting CT, Tsaur SC, Wu ML, Wu CI. A rapidly evolving homeobox at the site of a hybrid sterility gene.
Science 1998;282:1501–1504. [PubMed: 9822383]
Turesson G. The genotypical response of the plant species to the habitat. Hereditas 1922;3:211–350.
Turesson G. The plant species in relation to habitat and climate. Hereditas 1925;6:147–236.
Turesson G. The selective effect of climate upon the plant species. Hereditas 1930;14:99–152.
Wake DB. Incipient species formation in salamanders of the Ensatina complex. Proceedings of the
National Academy of Sciences, USA 1997;94:7761–7767.
Wang RL, Hey J. The speciation history of Drosophila pseudoobscura and close relatives: Inferences
from DNA sequence variation at the period locus. Genetics 1996;144:1113–1126. [PubMed:
8913754]
Wang RL, Stec A, Hey J, Lukens L, Doebley J. The limits of selection during maize domestication. Nature
1999;398:236–239. [PubMed: 10094045]
Wang RL, Wakeley J, Hey J. Gene flow and natural selection in the origin of Drosophila
pseudoobscura and close relatives. Genetics 1997;147:1091–1106. [PubMed: 9383055]
Ward RD, Skibinski DO, Woodwark M. Protein heterozygosity, protein structure, and taxonomic
differentiation. Evolutionary Biology 1992;26:73–159.
Wendel, JF.; Doyle, JJ. Phylogenetic incongruence: window into genome history and molecular
evolution. In: Soltis, DE.; Soltis, PS.; Doyle, JJ., editors. Molecular systematics of plants II: DNA
sequencing. Boston, MA, USA: Kluwer Academic Publishers; 1998. p. 256-296.
Werth CR, Windham MD. A model for divergent, allopatric speciation of polyploid pteridophytes
resulting from silencing of duplicate-gene expression. American Naturalist 1991;137:515–526.
Wheeler DA, Kyriacou CP, Greenacre ML, Yu Q, Rutila JE, Rosbash M, Hall JC. Molecular transfer of
a species-specific behavior from Drosophila-simulans to. Drosophila-Melanogaster. Science
1991;251:1082–1085. [PubMed: 1900131]
White, MJD. Modes of speciation. San Francisco, CA, USA: W.H. Freeman Co; 1978.
Whitkus R. Genetics of adaptive radiation in Hawaiian and Cook Islands species of Tetramolopium
(Asteraceae). II. Genetic linkage map and its implications for interspecific breeding barriers. Genetics
1998;150:1209–1216. [PubMed: 9799272]
Whitlock MC. Fixation probability and time in subdivided populations. Genetics 2003;164:767–779.
[PubMed: 12807795]
Wright S. Evolution in Mendelian populations. Genetics 1931;16:97–159. [PubMed: 17246615]
Wu CI. The genic view of the process of speciation. Journal of Evolutionary Biology 2001;14:851–865.
Wyckoff GJ, Wang W, Wu CI. Rapid evolution of male reproductive genes in the descent of man. Nature
2000;403:304–309. [PubMed: 10659848]
Yano M, Katayose Y, Ashikari M, Yamanouchi U, Monna L, Fuse T, Baba T, Yamamoto K, Umehara
Y, Nagamura Y, Sasaki T. Hd1, a major photoperiod sensitivity quantitative trait locus in rice, is
closely related to the Arabidopsis flowering time gene CONSTANS. Plant Cell 2000;12:2473–2483.
[PubMed: 11148291]
New Phytol. Author manuscript; available in PMC 2008 December 12.
Rieseberg et al.
Page 16
NIH-PA Author Manuscript
Fig. 1.
NIH-PA Author Manuscript
Frequency histogram of migration rate (Nem) for total nuclear data from 290 plant studies.
Data are derived from 130 studies reporting Fst, Gst Nem, and Fst values published in Molecular
Ecology from May 1992 until December 2002, and Gst values from 160 allozyme studies
(Ward et al., 1992). For studies not reporting Nem values, Nem was calculated from Fst, Gst,
or analogous statistics for nuclear markers as Nem = 1 − Fst/4 * Fst (Wright, 1931). Data are
displayed on a logarithmic scale, binned for Nem = 0.01, 0.1, 0.25, 0.5, 1, and binned into 1unit intervals thereafter.
NIH-PA Author Manuscript
New Phytol. Author manuscript; available in PMC 2008 December 12.
Rieseberg et al.
Page 17
NIH-PA Author Manuscript
NIH-PA Author Manuscript
Fig. 2.
Fixation time (in generations) by selection coefficient and Nem for a subdivided population of
106 diploid individuals with nonoverlapping generations. Subpopulations, each consisting of
2500 individuals, are arranged in a 20 × 20 grid. Migration occurs among subpopulations
according to a leptokurtic dispersal function, with a mean dispersal distance of one
subpopulation.
NIH-PA Author Manuscript
New Phytol. Author manuscript; available in PMC 2008 December 12.
Rieseberg et al.
Page 18
NIH-PA Author Manuscript
NIH-PA Author Manuscript
NIH-PA Author Manuscript
Fig. 3.
Distributions for the estimated strength of selection (s) for leading quantitative trait locis
(QTLs) (top panel) and minor QTLs (bottom panel) underlying phenotypic traits in plants. s
was calculated by multiplying the average percent variance explained (PVE) for leading QTLs
for 50 traits (31.1%) by 604 selection gradients for phenotypic traits from a literature review
and halving for diploidy. The bottom panel was calculated by multiplying the 604 selection
gradients by expected PVE values for minor QTLs (1%) and halving for diploidy.
New Phytol. Author manuscript; available in PMC 2008 December 12.
Rieseberg et al.
Page 19
NIH-PA Author Manuscript
Fig. 4.
NIH-PA Author Manuscript
The effect of population fragmentation and migration on the time to speciation in a stepping
stone model with 5 or 20 subpopulations. Time to speciation was the number of mutational
events until a pair of populations accumulated an arbitrarily defined number of
incompatibilities required for speciation. Migration was the average number of migration
events occurring across the entire population. The simulation was stopped when speciation
was not completed before mutations occurred at all 250 loci therefore speciation times of 250
signify that the time to speciation was at least 250 mutational events.
NIH-PA Author Manuscript
New Phytol. Author manuscript; available in PMC 2008 December 12.