PY52CH14-Matute

ARI

V I E W

A

15:9

Review in Advance first posted online on May 30, 2014. (Changes may still occur before final publication online and in print.)

N

I N

C E

S

R

E

19 May 2014

D V A

Annu. Rev. Phytopathol. 2014.52. Downloaded from www.annualreviews.org by College of William & Mary on 06/13/14. For personal use only.

Speciation in Fungal and Oomycete Plant Pathogens Silvia Restrepo,1 Javier F. Tabima,2 Maria F. Mideros,1 2 Niklaus J. Grunwald, and Daniel R. Matute3,4 ¨ 1

Departamento de Ciencias Biologicas, Universidad de Los Andes, Bogot´a, Colombia ´

2

Horticultural Crops Research Laboratory, United States Department of Agriculture, Agricultural Research Service, Corvallis, Oregon 97333

3

Department of Human Genetics, University of Chicago, Chicago, Illinois 60637

4

Biology Department, University of North Carolina, Chapel Hill, North Carolina 27599p; email: [email protected]

Annu. Rev. Phytopathol. 2014. 52:14.1–14.28

Keywords

The Annual Review of Phytopathology is online at phyto.annualreviews.org

speciation, reproductive isolation, population genetics, hybrids, genomics

This article’s doi: 10.1146/annurev-phyto-102313-050056

Abstract

c 2014 by Annual Reviews. Copyright  All rights reserved

The process of speciation, by definition, involves evolution of one or more reproductive isolating mechanisms that split a single species into two that can no longer interbreed. Determination of which processes are responsible for speciation is important yet challenging. Several studies have proposed that speciation in pathogens is heavily influenced by host-pathogen dynamics and that traits that mediate such interactions (e.g., host mobility, reproductive mode of the pathogen, complexity of the life cycle, and host specificity) must lead to reproductive isolation and ultimately affect speciation rates. In this review, we summarize the main evolutionary processes that lead to speciation of fungal and oomycete plant pathogens and provide an outline of how speciation can be studied rigorously, including novel genetic/genomic developments.

14.1

Changes may still occur before final publication online and in print

PY52CH14-Matute

ARI

19 May 2014

15:9

INTRODUCTION

Annu. Rev. Phytopathol. 2014.52. Downloaded from www.annualreviews.org by College of William & Mary on 06/13/14. For personal use only.

The major evolutionary forces that lead to the emergence of new plant-pathogen species remain largely unknown; however, based on our understanding of speciation in general and from case studies of particular fungi or oomycetes a clearer picture is beginning to emerge. Understanding the demographic and ecological factors that lead to the evolution of plant-pathogen species can greatly advance our knowledge of how they adapt to agricultural systems and provide new insights for plant disease management. In addition to these potential applications, fungi and oomycetes are ideal models for understanding the biological features that lead to new species in general. Fungi and oomycetes comprise a large number of species (Oomycota: ∼800 species; Fungi: ∼70,000 species) (90), of which many can cause disease in a wide variety of plant hosts. These factors make them appealing models for evolutionary biologists because pathogen biology provides an extraordinary window into how species interactions shape divergence patterns in different organisms and allows for clear hypotheses formulation. For example, given the dependence of plant pathogens on their host species, it is expected that some of the major mechanisms underlying plant-pathogen species emergence include host-range expansion and host jumps (21, 70, 181). Following human migrations, host-range expansion might have been largely the result of migration or movement of pathogens or hosts into new geographic regions (18, 22, 56). Nonetheless, many of these hypotheses remain largely untested. A unified conceptual framework to properly define the biological features that give rise to new species, allow persistence over time, and characterize the divergence processes in plant pathogens is needed. Our aim here is twofold: (a) We want to summarize the current evidence for how new plant-pathogen species emerge, and (b) we want to outline a rigorous framework for characterizing the divergence, hybridization, and speciation observed in these organisms.

SPECIATION IN PLANT PATHOGENS: THE ROLE OF REPRODUCTIVE ISOLATION Evolutionary biologists have long studied the evolutionary processes that give rise to new genotypes, lineages, and eventually species, as well as the traits that make species unique (33, 147). Species are the largest and most inclusive reproductive community of cross-fertilizing individuals that share a common gene pool (45). This definition, known as the biological species concept, was coined to delimitate species boundaries in a hypothesis-testing framework. At its core, it has a simple premise: If two individuals show biological features that reduce or eliminate the possibility of gene flow between them, then they belong to different species. Importantly, speciation is a continuous process in which lineages diverge gradually to form separate species. On the surface, this definition does not seem to apply to clonal lineages because the potential for recombination and gene flow is not thought to exist. Nonetheless, alternative approaches have been proposed to detect the signature of reproductive isolation in these organisms (61, 124). Because gene flow is a homogenizing force that reduces genetic and phenotypic differences between groups of individuals, reproductive isolation must be a key feature of the process of speciation (and/or of species maintenance), even among asexual species. The appeal of this definition is twofold. First, it links species boundaries with the processes that lead to the evolution of traits that preclude gene flow between groups of individuals. Second, it provides a conceptual framework for recognizing species boundaries that allows for rigorous hypothesis testing: If gene flow between demes is reduced or nonexistent, then there must be reproductive barriers between the groups. Studying the processes involved in speciation thus boils down to understanding the barriers to genetic exchange and the mechanisms that shape these barriers. 14.2

Restrepo et al.

Changes may still occur before final publication online and in print

Annu. Rev. Phytopathol. 2014.52. Downloaded from www.annualreviews.org by College of William & Mary on 06/13/14. For personal use only.

PY52CH14-Matute

ARI

19 May 2014

15:9

These mechanisms can be classified into three main groups depending on where they occur in the reproductive cycle: premating, postmating-prezygotic, and postzygotic isolation (33, 46). Table 1 lists the reproductive isolating mechanisms (RIMs) that have been previously described in fungal and oomycete plant pathogens. Some major trends emerge from this literature review. Premating isolating barriers are by far the most often reported RIMs to date in plant pathogens (Table 1). Ecological isolation, especially by host specialization between potentially recombining species is reported as the most widespread form of reproductive isolation (177, 188) (Table 1). This observation might constitute an ascertainment bias because plant-pathogen research usually starts by investigating host range. However, ecological isolation is not the only type of reproductive isolation reported for plant pathogens (Table 1). Interspecific crosses between species of Phytophthora induce fewer sexual structures than intraspecific crosses, which indicates the existence of premating isolating mechanisms that hinder the production of hybrid progeny (170, 171). Postzygotic isolation, which reduces fitness of recombining individuals, has been less studied than ecologically based reproductive isolation in plant pathogens. It is clear that hybridization occurs in different groups of fungi and oomycetes (165). Extrinsic postzygotic isolation by reductions of hybrid fitness due to ecological factors seems to be common in plant pathogens. Hybrids between highly specialized pathogens that might not fare well in any of the parental niches consequently show low fitness in parental niches (see examples in section “Ecological Divergence” below) (70). However, some interspecific hybrids can have profound economic implications for agriculture, have a higher fitness than either parental (hybrid vigor), and show changes or expansions of host ranges (165). It remains unclear how frequently hybridization leads to range expansion or host jumps (64) as well as how frequently interspecific hybrids show decreased ecological fitness and are outperformed by the parental species. A second type of postzygotic isolating mechanism relates to the viability and fertility of the resulting hybrids and has been termed intrinsic reproductive isolation. This type of RIMs has been heavily studied in other taxa (137, 185, 192), but reports of their existence in plant pathogens are scarce. Examples of hybrid sterility have been documented for species of the genera Mycrobotryum (44, 113), Ascochyta (105), and Cochliobolus (131, 132), as well as in Gibberella zeae (17, 104), among others. The most extreme phenotype, hybrid inviability, has also been documented for species of the genera Ceratocystis (50, 88) and Phytophthora (41, 54, 129). It is crucial to understand how pervasive postzygotic mechanisms are given the substantial importance of host specialization in plant pathogens. An additional RIM in fungi and oomycetes is somatic incompatibility (SI) (117, 194). Mycelia from fungi and oomycetes can fuse vegetatively and give rise to hybrid individuals. SI is a system by which filamentous fungi can distinguish their own mycelia from those of other individuals and from individuals of different species. Even though SI occurs in several taxa, it has received little attention in the field of speciation research (118, 194). SI is of special importance for the generation of new genetic variation (reviewed in 139), especially in organisms for which sexual cycles do not exist or have yet not been observed. In somatic fusion between organisms of different species, hybrid individuals could be produced that might allow for gene flow between parental species. Similar to barriers to sexual reproduction, there can be barriers to the fusion of somatic tissues, especially when individuals belong to divergent parental species; such barriers are found between different populations and species of cereal rusts (Puccinia spp.) (117). Somatic barriers to gene flow can also act at different steps in the somatic fusion process (failure of anastomoses, genetic, and cytoplasmatic incompatibilities) (117, 194). Even though SI is unlikely to lead to speciation, it might have a role in keeping species apart. SI is doubtlessly important for stopping gene flow between individuals, populations, and ultimately species, so its role in plant pathogens needs to be studied further. www.annualreviews.org • Speciation in Fungal and Oomycete Plant Pathogens

Changes may still occur before final publication online and in print

RIMs: reproductive isolating mechanisms Hybrid vigor: higher fitness observed in some hybrids arising from interspecific crosses; also called heterosis Intrinsic reproductive isolation: RIMs that are not dependent on the environment (unlike extrinsic reproductive isolation). Hybrid inviability and sterility are examples of intrinsic RIMs SI: somatic incompatibility

14.3

PY52CH14-Matute

ARI

19 May 2014

15:9

Table 1 Reproductive isolating mechanisms observed in fungi and oomycete plant pathogens. Pairs of taxonomically recognized species are shown in two columns. Species complexes are shown in a single columna

Annu. Rev. Phytopathol. 2014.52. Downloaded from www.annualreviews.org by College of William & Mary on 06/13/14. For personal use only.

Species 1

Type of isolating mechanism

Method and conclusions

Reference

Phytophthora infestans

Phytophthora mirabilis

Premating

In vitro interspecific crosses failed to produce hybrids

(79)

Phytophthora rubi

Phytophthora fragariae

Premating

Attempts to infect each other’s host have been unsuccessful

(120)

Phytophthora capsici

Phytophthora tropicalis

Postmating, prezygotic

In vitro interspecific crosses revealed the existence of hybrid sterility

(47)

Phytophthora megakarya

Phytophthora nicotianae

Postmating, postzygotic

In vitro interspecific crosses produced fewer progeny than conspecific crosses and inviable oospores

(14)

P. capcisi

Phytophthora nicotianae

Postzygotic

In vitro interspecific crosses produced mostly inviable oospores

(14)

P. capcisi

Phytophthora palmivora

Postzygotic

In vitro interspecific crosses produced mostly inviable oospores

(14)

Phytophthora vignae

Phytophthora sojae

Postmating postzygotic

Interspecific crosses in vitro and infection assays showed that F1 hybrids have reduced aggressiveness.

(121)

Phytophthora infestans

Phytophthora palmivora

Premating

Absence of natural hybrids in areas of secondary contact. Possible reproductive isolation

(163)

Phytophthora parasitica

Phytophthora cinnamomi

Premating

In vitro interspecific crosses produced mostly inviable oospores

(15)

P. parasitica

Phytophthora megakarya

Premating

In vitro interspecific crosses produced mostly inviable oospores

(55)

P. sojae

Phytophthora medicaginis

Premating

Putative species were isolated in different substrates, indicating habitat isolation

(87)

Phytophthora porri

Phytophthora brassicae

Premating

Putative species were isolated in different substrates, indicating habitat isolation

(40)

P. cinnamomi

Phytophthora cambivora

Premating

In vitro interspecific crosses produced mostly inviable oospores

(14)

Microbotryum violaceum complex

Postmating, postzygotic

In vitro interspecific crosses revealed the existence of hybrid sterility

(44)

M. violaceae complex

Premating and postmating, postzygotic

Interspecific crosses in vitro and infection assays indicated that F1 hybrids have lower fitness than parental species

(113)

Prezygotic

Putative species were isolated in different substrates, indicating habitat isolation

(188)

Ophiostoma piceae complex

Premating

Putative species were isolated in different substrates, indicating habitat isolation

(142)

Cylindrocladium floridanum

Premating

Interspecific crosses failed to produce hybrids

(106)

Melampsora medusae f. sp. deltoidae

Species 2

M. medusae f. sp. tremuloidae

Cylindrocladium canadence

(Continued )

14.4

Restrepo et al.

Changes may still occur before final publication online and in print

PY52CH14-Matute

ARI

Table 1

19 May 2014

15:9

(Continued ) Type of isolating

Annu. Rev. Phytopathol. 2014.52. Downloaded from www.annualreviews.org by College of William & Mary on 06/13/14. For personal use only.

Species 1

Method and conclusions

Reference

C. floridanum

Cylindrocladium pacificum

Premating

Interspecific crosses failed to produce hybrids

(106)

C. floridanum

Cylindrocladium pseudospataphylli

Premating

Interspecific crosses failed to produce hybrids

(106)

Ascochyta rabiei complex (Ascochyta fabae, Ascochyta lentis, Ascochyta pisi, A. rabiei, and Ascochyta viciae-villosae)

Premating and postmating

Host specificity assays revealed habitat isolation; inoculations with F1 hybrid spores suggested reduced fitness of the hybrids in the habitat of both parentals

(93)

Didymella rabiei demes from two different chickpeas

Postmating, postzygotic

Thermal fitness differences indicated possible habitat isolation; inoculations with F1 hybrid spores suggested reduced fitness of the hybrids in the habitat of both parentals

(62)

Microbotryum lychnidis-dioicae

Postmating, postzygotic

Interspecific crosses in vitro revealed the existence of hybrid sterility

(44)

Melampsora lini species complex

Premating

In vitro interspecific crosses in vitro detected vegetative incompatibilities

(59)

Fusarium oxysporum

Postmating

In vitro interspecific crosses in vitro detected vegetative incompatibilities

(32)

Gibberella fujikuroi (Fusarium moniliforme)

Premating

In vitro interspecific crosses in vitro detected vegetative incompatibilities

(149)

P. cinnamomi

Postmating, postzygotic

In vitro interspecific crosses in vitro detected vegetative incompatibilities

(171)

Phytophthora drechsleri, P. cinnamomi, P. parasitica, P. palmivora

Premating

Interspecific crosses in vitro detected hybrid inviability and sterility

(170)

A. fabae/A. lentis

Postzygotic

In vitro interspecific crosses revealed hybrid sterility

(93, 105)

Gibberella zeae (Fusarium graminearum) species complex

Postzygotic

In vitro interspecific crosses revealed hybrid sterility

(17, 104)

Fusarium acuminatum, Fusarium avenaceum, Fusarium culmorum, Fusarium crookwellense, F. oxysporum

Postzygotic

In vitro interspecific crosses revealed hybrid sterility and inviability

Phytophthora alni

Postmating, postzygotic

Lower fertility in interspecific crosses; hybrids showed reduced growth rate

(43)

Ceratocystis fimbriata, Ceratocystis cacaofunesta, Ceratocystis platani

Postmating, postzygotic

In vitro interspecific crosses revealed hybrid inviability

(50)

Ceratocystis coerulescens

Premating

Putative species were isolated in different substrates, indicating habitat isolation

(89)

a

Species 2

Microbotryum silenes-dioicae

mechanism

A total of 34 species pairs have been studied and six RIMs have been reported between them (36 papers).

www.annualreviews.org • Speciation in Fungal and Oomycete Plant Pathogens

Changes may still occur before final publication online and in print

14.5

ARI

19 May 2014

Annu. Rev. Phytopathol. 2014.52. Downloaded from www.annualreviews.org by College of William & Mary on 06/13/14. For personal use only.

PY52CH14-Matute

15:9

a

Locus 1

c

Locus 3

b

Locus 2

d

Locus 4

e

Species tree

Dxy

π1

π2

Figure 1 Recognizing species in plant pathogens. Gene genealogies can be used to reconstruct the phylogenetic history of sister species. Four hypothetical loci (a–d ) were used to build the genealogies shown in each panel. The size of each circle is proportional to the number of individuals that show a particular haplotype (black dots represent missing haplotypes and are a proxy of genetic distance between observed haplotypes). The goal of building gene genealogies is to identify potential speciation events and discern whether gene flow is still ongoing. In this example, the four loci depict the evolutionary history of two hypothetical species (Sp1, red; Sp2, yellow). The merged data set (e) can be used to calculate the intraspecific polymorphism (π) and the absolute divergence (i.e., Dxy). These quantities can be compared to determine how much divergence has accumulated between the two species.

SPECIES RECOGNITION IN PLANT PATHOGENS Because reproductive isolation is often not easily assessed, fungal and oomycete species have traditionally been defined with morphological and phenotypic traits. Contemporary species description integrates these classical approaches with sequence-based phylogenetic approaches that attempt to find evidence of reproductive isolation (8, 26, 34, 84, 98, 133). The premise of such approaches is that different gene genealogies of two groups that have ceased interchanging genes should reflect concordant topologies among loci. The most common practice involves the use of gene genealogies to detect long branches, which in turn can be interpreted as discontinuities in population structure that might signal lack of gene exchange (depicted in Figure 1). Such studies have revealed that the number of species in fungi and oomycetes tends to be underestimated by a factor of two (96, 110, 183, 184). In an attempt to unify molecular taxonomy, some general guidelines have been proposed as to what levels of divergence are required to designate species 14.6

Restrepo et al.

Changes may still occur before final publication online and in print

Annu. Rev. Phytopathol. 2014.52. Downloaded from www.annualreviews.org by College of William & Mary on 06/13/14. For personal use only.

PY52CH14-Matute

ARI

19 May 2014

15:9

in fungi (42, 43), although how these correlate with reproductive isolation is unknown. Accordingly, species description should rely on multiple, unlinked loci from neutrally evolving genes with strong phylogenetic signals (5, 24, 141, 162). Figure 1 shows an example of the use of gene genealogies to reconstruct the divergence history of a pair of cryptic sibling species. Phylogenetic species recognition practices and the application of other species concepts are extensively reviewed elsewhere (25, 154). Identifying species is a challenging task for asexual populations. The majority of species concepts have been developed for individuals that reproduce sexually and recombine with individuals from the same gene pool. Thus, a rigorous framework is required to study species formation in those groups of organisms that do not reproduce sexually or for which no sexual stage is known. Population genetics and evolutionary theory for asexual taxa provide some guidance and several quantitative extensions of the gene genealogies approach have been proposed. Two of the most comprehensive proposals are the K/θ method (or 4X rule) and the application of branching rates in phylogenetic trees. The K/θ method (12) uses DNA polymorphism and basic coalescent theory to identify clusters of organisms that are diverged enough to be considered separate species. The average number of differences between individuals of two species (after correcting for multiple hits) equals, on average, 8Ne μ, where Ne is the effective population size and μ is the mutation rate (7). This quantity is known as K. Intraspecific polymorphism, θ, is roughly equivalent to 2Ne μ. The K/θ metric reflects divergence scaled by the effective population size and is thought to reveal different species if greater than 4, hence the reference to the 4X rule. This approach identifies putative species only on the basis of reciprocal monophyly using rooted trees and ascribes a metric to delimitate populations from true species (11, 156). An alternative framework for considering whether a clade has diversified into discrete genetic clusters is to consider branching models in phylogenetic trees (145). This approach is more complex than the K/θ method but allows for a global test of the relative rates of divergence within a clade. Under a null model, the entire sample derives from a single asexual population, i.e., without divergence into independently evolving or ecologically distinct species. Under this null hypothesis, the pattern of branching is expected to conform to a standard coalescent model for a single population. However, if speciation has taken place, the branching pattern of a phylogenetic tree deviates from the coalescent null hypothesis and has longer branches in the deeper regions of the phylogenies. These long branches are thus a diagnostic that reveals species boundaries. A large body of theoretical work is available to specify the likelihood of a given pattern of branching under a particular model (107), ranging from a neutral coalescent in a population of constant size to populations that increased through time or experienced different forms of selection (157). To our knowledge, this approach has not been applied to plant pathogens but, along with the K/θ method, has the potential for application to the identification of species boundaries and determination of rates of speciation.

Effective population size (Ne ): the size of an idealized population that would have the same population genetic properties of the population under study

HOW TO STUDY SPECIATION IN PLANT PATHOGENS? The completion of speciation is signaled by the existence of distinct RIMs. Thus, understanding how species originate necessarily focuses on understanding what biological features prevent gene flow. Although research in plant pathology is usually not characterized as speciation research, a series of approaches has aimed at dissecting the biological basis of reproductive isolation between plant-pathogen species. We divide these studies into two categories and refer to the studies that have identified RIMs between species as the classical studies. The second category, genomic studies, involves genome-scale studies that identify genomic features that are involved in reproductive isolation or in interspecies differences. There is no reason, however, for these categories to be www.annualreviews.org • Speciation in Fungal and Oomycete Plant Pathogens

Changes may still occur before final publication online and in print

14.7

PY52CH14-Matute

ARI

19 May 2014

15:9

distinct, and we are optimistic that the coming years will see an increase in studies that combine both of these approaches.

Classical Studies

Annu. Rev. Phytopathol. 2014.52. Downloaded from www.annualreviews.org by College of William & Mary on 06/13/14. For personal use only.

Because one of the most notable traits of plant pathogens is their habitat specificity, most of the studies dealing with reproductive isolation have addressed the possibility of isolation by ecological factors, such as reciprocal fitness in each other’s niches (common garden or transplant experiments) in two species that share a most recent common ancestor. In multiple cases (31), these experiments have demonstrated the existence of extreme specialization to plant hosts. In these instances, the main source of reproductive isolation is likely to be the separation of nascent species by these ecological factors. For example, species of the genus Ascochyta (teleomorph Didymella) (93, 105) show strong ecological isolation. Inoculations with ascospores from Ascochyta fabae caused disease in fava bean but not in lentil; inoculations of ascospores from Ascochyta lentis caused disease in lentil but not in fava beans, indicating strong host specificity and ecological isolation. The crosses between the two species produced fewer ascospores per ascus than pure species crosses, indicating the existence of a postmating nonecological RIM. Moreover, inoculation with hybrid ascospores (A. fabae × A. lentis) did not induce disease in either host. This suggests a fitness depression in the hybrids and the presence of extrinsic postzygotic isolation, and demonstrates an important role of environmental factors in keeping these species apart (105). Abiotic factors can also play a role in keeping species apart. Botrytis pseudocinerea, a recently identified species, and Botrytis cinerea show adaptation to different temperatures [as measured by at least one thermal fitness component (number of sclerotia in minimal media at different temperatures), which might explain the relative abundance of B. pseudocinerea in the springtime. This temperature adaptation might also cause temporal isolation, but this has not been tested. A strong influence of temperature has been reported for ecotypes of the species complex (189). This is also the case for Phytophthora drechsleri, where northern isolates thrive in cooler temperatures and southern isolates thrive in higher temperatures (172). It is still unclear whether these ecotypes are different species and whether there are any other RIMs between them. Some studies have analyzed the relative fitness of hybrids after interspecific crosses under laboratory conditions, and these studies have been successful in the characterization of postzygotic RIMs that keep species of plant pathogens apart at different phylogenetic scales (Table 1). These mechanisms are key in the persistence of species because they prevent nascent species from admixing (148, 151). Ceratocystis fimbriata sensu lato is a large complex of species that causes wild-type diseases of many economically important plants. Species within the group show strong specialization to cacao (C. fimbriata f. sp. cacaofunesta), sweet potato (C. fimbriata sensu stricto), and sycamore (C. fimbriata f. sp. platani ). Interspecific crosses between the different lineages show complex variation in reproductive isolation with high levels of phenotypic polymorphism. The results from these crosses can be categorized into three main types. The first type of cross produced no progeny (no perithecia were formed), indicating the existence of a prezygotic RIMs (or a very early postzygotic RIM). A second type produced perithecia with few ascospores, many of which were devoid of cytoplasm and showed deformed morphology. The colonies recovered from these ascospores were generally abnormal and showed restricted growth, indicating hybrid inviability at the ascospore level but also reduced fitness at the mycelial stage. The final type showed complete compatibility, indicating no intrinsic RIM between species (51). Similar variation in hybrid viability and fitness have been observed in conifer-specialized Ceratocystis (88), indicating that different Ceratocystis species show different types and degrees of reproductive isolation. It will be interesting to see whether the magnitude of reproductive isolation correlates with genetic divergence between species. 14.8

Restrepo et al.

Changes may still occur before final publication online and in print

PY52CH14-Matute

ARI

19 May 2014

15:9

These intrinsic mechanisms of reproductive isolation have also been characterized in oomycetes. In crosses involving divergent species of Phytophthora, only a very small proportion ( 1). This is a powerful approach but carries three major caveats. First, the identification of variants involved in habitat specialization is not equivalent with the identification of variants involved in speciation (e.g., if there is polymorphism and locally adapted populations within a population, dN/dS ratios might be significant, indicating positive selection without reproductive isolation). The second caveat is that dN/dS uses the number of synonymous substitutions as a proxy of neutral evolution and does not take into account codon bias (27, 144), thus underestimating the percentage of the genome under selection. Finally, and on a related note, dN/dS is a particularly conservative approach (111, 115, 197), and other approaches, such as modified McDonald-Kreitman’s tests, might be more informative (67, 100, 193).

Genetic Differentiation, Fst Fst (inbreeding coefficients as a proxy of population structure) is by far the most popular method to identify regions of the genome that are highly differentiated between populations or species. Calculating Fst values [or any of its analogs -ST , F ST , GST , θ, RST , (130, 195)] can be done using multiple approaches (97, 190) that are sensitive to particular demographic conditions (108). The premise of this approach is to compare variance components of phenotypic traits and try to determine whether there are molecular variants that are as differentiated as the phenotypes are. These comparisons can reveal what alleles are involved in divergent phenotypes (reviewed in 191). An expansion of this approach has been the study of populations that have had the chance to homogenize through gene flow (i.e., secondary contact of species in hybrid zones) in order to identify the genetic architecture of alleles involved in reproductive isolation. The identification of genomic islands of speciation (i.e., genomic clusters of sites with high Fst values) has at least three major caveats (28, 48, 134, 150). First, the calculation of Fst values is highly dependent on the quantification of genetic divergence within each of the species; if one of the species has a level of genetic polymorphisms that is much lower than the other, the average Fst value across the genome will be inflated. This might generate false positives. Second, regions with low genetic www.annualreviews.org • Speciation in Fungal and Oomycete Plant Pathogens

Changes may still occur before final publication online and in print

14.17

PY52CH14-Matute

ARI

19 May 2014

Annu. Rev. Phytopathol. 2014.52. Downloaded from www.annualreviews.org by College of William & Mary on 06/13/14. For personal use only.

Linkage disequilibrium: nonrandom association of alleles at two or more loci that descend from a single ancestral chromosome. Linkage disequilibrium does not require physical linkage

15:9

diversity driven by positive selection will show high Fst values regardless of whether the genomic region is involved in reproductive isolation (134). Fst calculations between species are also heavily affected by retention of ancestral polymorphism, and deviations from the neutral allele frequency spectrum within species can lead to higher Fst values in regions of low recombination, even in the complete absence of interspecies gene flow (91, 92, 102, 103, 123, 161). Finally, when Fst values are calculated genome-wide, it is absolutely necessary to generate a background distribution in order to correct for false positives and the differential influences of recombination and selection. These violations might lead to the spurious detection of Fst outliers (3, 134).

Genetic Differentiation and Allele Frequency Differences An alternative quantification of divergence is the calculation of allele frequency differences across the whole genome (13). Even though this approach still requires the generation of an underlying background distribution to identify outliers, it does not suffer from the issue of differential levels of polymorphism within each species. Similarly, it does not depend on the influence of recombination and selection in different locations of the genome. An additional advantage of this approach is that it can be used in pooled-DNA sequences (109). Thus far, this approach has not been used in plant pathogens.

Candidate Gene Approaches Unlike other methods, this approach relies on previous knowledge generated in other organisms (reviewed in 125). This technique has been successful in identifying point alleles involved in a phenotype but does not reveal the whole genetic architecture of an adaptive trait (i.e., there is a negative ascertainment bias to discover alleles with small effect). An additional caveat is that for the identification of alleles involved in reproductive isolation, especially for phenotypes that cause intrinsic hybrid breakdown (i.e., sterility, inviability), there are no candidate genes to follow up, as hybrid breakdown can occur at many different biological levels.

Admixture Mapping For those pathogens that do not recombine in the laboratory but are thought to do so in nature, it is possible to sample areas in which two species with phenotypic differences overlap and hybridize (i.e., hybrid zones) and use linkage disequilibrium to map the genetic basis of traits. The premise of this approach is to use local ancestry estimates from each of the parental species and do association studies with particular phenotypes. Currently, there are several models that can be applied to identify admixture patterns in plant pathogens [HAPMIX (146), frappe (182), and ADMIXTURE (2) among others]. This approach has more power than common QTL approaches but has not been used in plant pathogens.

SPECIATION AND THE EMERGENCE OF NEW DISEASES Understanding the appearance of new pathogenic species can inform control methods and prevention plans (1, 181). Population genetics inferences and the study of evolutionary processes in fungi and oomycetes can reveal how pathogens emerge to cause new diseases. Emergence proceeds by one of three processes. The simplest scenario does not involve the evolution of new species and occurs when a pathogen colonizes a new geographical area in which it finds the host species that it infected in its ancestral location. Mycosphaerella fijiensis is an example of a recent worldwide 14.18

Restrepo et al.

Changes may still occur before final publication online and in print

Annu. Rev. Phytopathol. 2014.52. Downloaded from www.annualreviews.org by College of William & Mary on 06/13/14. For personal use only.

PY52CH14-Matute

ARI

19 May 2014

15:9

epidemic that affects banana plantations. Molecular genealogies indicate that the range expansion of this pathogen has occurred through bottlenecks (i.e., severe population reductions) and that the spread of the disease might be caused by movement of plant material infected with ascospores (86, 155). A similar pattern is observed with the sudden oak death pathogen Phytophthora ramorum (82). This pathogen consists of several distinct genetic lineages that appear to have diverged a long time ago, possibly before modern agriculture (80). A second process of disease emergence occurs when a pathogen colonizes a new geographic area and develops the capability of infecting new hosts. This process, mediated by adaptation (via host shifts or expansion of the host range), enables the pathogen to thrive in hosts different from its ancestral plants. If the new adaptation leads to reproductive isolation from the parental species, then speciation is correlated to the emergence of the new pathogen and of the new pathogenic syndrome. Rhynchosporium, the causal agent of scald disease, seems to be composed of three cryptic pathogen species that have evolved by ecological divergence and host specialization to barley, rye, and Agropyron (198). Gene genealogies also suggest that Colletotrichum kahawae emerged as a pathogen that specializes in green coffee berries (173). The recent emergence of these pathogens is correlated with host shifts, suggesting that agriculture has played an important role in the emergence of diseases. A final possibility is that emergent pathogens and diseases originate after the hybridization between two species; the hybrid then shows a new suite of traits not observed in the parents (a type of inheritance known as transgressive segregation), allowing it to infect new hosts. Hybridization and admixture after bringing closely related pathogens into contact could favor cross-host-species disease transmission (71, 72). The demographic inference of gene exchange between plant-pathogen species in an explicitly spatial context can lead to a deeper understanding of the range constraints of plant-pathogen species.

CONCLUSIONS Evolutionary biology provides a powerful framework for understanding key aspects of the natural history of pathogens (114, 116, 196). The study of evolutionary processes in plant pathogens is a nascent field that promises to deliver new model systems and a comprehensive view of mechanisms involved in speciation and adaptation. On a more practical scale, the development of multilocus databases will allow for the identification of pathogenic lineages in a phylogeographic framework for additional comparative studies to understand the population dynamics of plant pathogens. One of the ultimate goals of plant pathology is to be able to predict the emergence of new pathogens, and this can be done using an evolutionary perspective. The combination of classical approaches coupled with high-resolution genomic studies will reveal demographic and genetic characteristics of how speciation occurs in oomycetes and fungal plant pathogens. The addition of new technological developments will bolster classical approaches and questions. Studying the evolutionary and ecological dynamics of plant pathogens can lead to a better understanding of how these organisms affect agricultural processes over time. The study of recently diverged pathogenic species can also address key questions in evolutionary theory. What is the level of gene flow required to override divergence by selection? What is the role of hybridization in the generation of more virulent pathogens? Do hybrids show new traits or do they always show reduced fitness compared with their parents? Does hybridization (and admixture) lead to the collapse of species boundaries? These questions, among others, are at the very core of evolutionary biology research but also have direct implications in the day-to-day management of plant diseases, as they essentially resolve the origins and mechanisms of emergence of new pathogens. www.annualreviews.org • Speciation in Fungal and Oomycete Plant Pathogens

Changes may still occur before final publication online and in print

14.19

PY52CH14-Matute

ARI

19 May 2014

15:9

SUMMARY POINTS 1. Plant pathogens are an ideal system to study the evolution of reproductive isolation and how new species originate. The study of evolutionary processes in plant pathogens gives rise to new model systems and a comprehensive view of speciation and adaptation. 2. Although plant pathology research is usually not characterized as speciation-centered research, a series of studies have successfully identified the biological basis of reproductive isolation between plant-pathogen species. 3. The combination of classical approaches coupled with high-resolution genomic studies reveals demographic and genetic characteristics of how speciation proceeds in oomycetes and fungal plant pathogens. Annu. Rev. Phytopathol. 2014.52. Downloaded from www.annualreviews.org by College of William & Mary on 06/13/14. For personal use only.

4. The study of ecological and geographical drivers has important implications for plant pathologists and evolutionary biologists alike, and will lead to a better understanding of the events that lead to the emergence of new pathogens and new diseases. 5. The relative prevalence of speciation with versus without gene flow is still unknown for all taxa. Although speciation with gene flow might be possible in plant pathogens, it has not yet been proven. Geographical isolation can at least be assumed to play a role in most cases of speciation among these organisms.

FUTURE ISSUES 1. What levels of divergence are required to define species in fungi? Studies that connect molecular taxonomy with reproductive isolation are sorely needed. 2. The application of quantitative approaches to recognize species will reveal the existence of new plant-pathogen species. The application of these methods will allow for a global test of speciation and extinction rates within clades of plant pathogens. 3. Pathogen host range appears to be constrained by the pathogen effector repertoires. The interplay of effectors and plant immunity genes might shape the evolution of new plant-pathogen species. This hypothesis remains untested. 4. It remains unclear how frequently hybridization leads to host-range expansion. The formal study of the consequences of hybridization in plant pathogens is an open research avenue.

DISCLOSURE STATEMENT The authors are not aware of any affiliations, memberships, funding, or financial holdings that might be perceived as affecting the objectivity of this review.

ACKNOWLEDGMENTS We would like to thank K.L. Gordon, M.F. Przeworski, J.A. Coyne, P. Gladieux, J.W. Taylor, B.J. Knaus, and C.D. Cadena for scientific discussions and comments at every stage of the manuscript. 14.20

Restrepo et al.

Changes may still occur before final publication online and in print

PY52CH14-Matute

ARI

19 May 2014

15:9

S.R. and M.F.M. are funded by a Basic Sciences project from Facultad de Ciencias and Vicerrector´ıa de Investigaciones (Universidad de los Andes). D.R.M is funded by a Chicago Fellowship.

Annu. Rev. Phytopathol. 2014.52. Downloaded from www.annualreviews.org by College of William & Mary on 06/13/14. For personal use only.

LITERATURE CITED 1. Ahmed S, de Labrouhe DT, Delmotte F. 2012. Emerging virulence arising from hybridisation facilitated by multiple introductions of the sunflower downy mildew pathogen Plasmopara halstedii. Fungal Gen. Biol. 49:847–55 2. Alexander DH, Novembre J, Lange K. 2009. Fast model-based estimation of ancestry in unrelated individuals. Genome Res. 19:1655–64 3. Andrew RL, Bernatchez L, Bonin A, Buerkle CA, Carstens BC, et al. 2013. A road map for molecular ecology. Mol. Ecol. 22:2605–26 4. Arnold ML. 1997. Natural Hybridization and Evolution. Oxford: Oxford Univ. Press 5. Avise JC, Wollenberg K. 1997. Phylogenetics and the origin of species. Proc. Natl. Acad. Sci. USA 94:7748–55 6. Bahri B, Kaltz O, Leconte M, de Vallavieille-Pope C, Enjalbert J. 2009. Tracking costs of virulence in natural populations of the wheat pathogen, Puccinia striiformis f. sp. tritici. BMC Evol. Biol. 9:26 7. Barraclough TG, Birky CW, Burt A. 2003. Diversification in sexual and asexual organisms. Evolution 57:2166–72 8. Baum DA, Donoghue MJ. 1995. Choosing among alternative “phylogenetic” species concepts. Syst. Bot. 20:560–73 9. Becquet C, Przeworski M. 2007. A new approach to estimate parameters of speciation models with application to apes. Genome Res. 17:1505–19 10. Bierne N, Welch J, Loire E, Bonhomme F, David P. 2011. The coupling hypothesis: why genome scans may fail to map local adaptation genes. Mol. Ecol. 20:2044–72 11. Birky CW Jr. 2013. Species detection and identification in sexual organisms using population genetic theory and DNA sequences. PLoS ONE 8:e52544 12. Birky CW Jr, Barraclough TG. 2009. Asexual Speciation. In Lost Sex. The Evolutionary Biology of Parthenogenesis, ed. I Schon, ¨ K Martens, P Van Dijk, pp. 201–16. Dordrecht, Neth.: Springer 13. Biswas S, Akey JM. 2006. Genomic insights into positive selection. Trends Genet. 22:437–46 14. Boccas BR. 1981. Interspecific crosses between closely related heterothallic Phytophthora species. Phytopathology 71:60–65 15. Boccas BR, Zentmyer GA. 1976. Genetical studies with interspecific crosses between Phytophthora cinnamomi and Phytophthora parasitica. Phytopathology 66:77–84 16. Bonants PJ, Hagenaar-de Weerdt M, Man In ‘t Veld WA, Baayen RP. 2000. Molecular characterization of natural hybrids of Phytophthora nicotianae and P. cactorum. Phytopathology 90:867–74 17. Bowden RL, Leslie JF. 1999. Sexual recombination in Gibberella zeae. Phytopathology 89:182–88 18. Brasier CM. 2008. The biosecurity threat to the UK and global environment from international trade in plants. Plant Pathol. 57:792–808 19. Brasier CM, Cooke DE, Duncan JM. 1999. Origin of a new Phytophthora pathogen through interspecific hybridization. Proc. Natl. Acad. Sci. USA 96:5878–83 20. Brasier CM, Kirk SA, Pipe N, Buck KW. 1998. Rare hybrids in natural populations of the Dutch elm disease pathogens Ophiostoma ulmi and O. novo-ulmi. Mycol. Res. 102:45–57 21. Broders KD, Boraks A, Sanchez AM, Boland GJ. 2012. Population structure of the butternut canker fungus, Ophiognomonia clavigignenti-juglandacearum, in North American forests. Ecol. Evol. 2:2114–27 22. Brown JK, Hovmoller MS. 2002. Aerial dispersal of pathogens on the global and continental scales and its impact on plant disease. Science 297:537–41 23. Browning M, Rowley LV, Zeng P, Chandlee JM, Jackson N. 1999. Morphological, pathogenic and genetic comparisons of Colletotrichum graminicola isolates from Poaceae. Plant Dis. 83:286–92 24. Buerkle CA, Gompert Z, Parchman TL. 2011. The n = 1 constraint in population genomics. Mol. Ecol. 20:1575–81 www.annualreviews.org • Speciation in Fungal and Oomycete Plant Pathogens

Changes may still occur before final publication online and in print

14.21

ARI

19 May 2014

15:9

25. Cai L, Giraud T, Zhang N, Begerow D, Cai G, Shivas RG. 2011. The evolution of species concepts and species recognition criteria in plant pathogenic fungi. Fungal Divers. 50:121–33 26. Cardenas M, Tabima J, Fry WE, Grunwald NJ, Bernal A, Restrepo S. 2012. Defining species boundaries ¨ in the genus Phytophthora: the case of Phytophthora andina: a response to “Phytophthora andina sp nov., a newly identified heterothallic pathogen of solanaceous hosts in the andean highlands” (Oliva et al. 2010). Plant Pathol. 61:215–20 27. Chamary JV, Hurst LD. 2005. Evidence for selection on synonymous mutations affecting stability of mRNA secondary structure in mammals. Genome Biol. 6:R75 28. Charlesworth B, Nordborg M, Charlesworth D. 1997. The effects of local selection, balanced polymorphism and background selection on equilibrium patterns of genetic diversity in subdivided populations. Genet. Res. 70:155–74 29. Chuma I, Isobe C, Hotta Y, Ibaragi K, Futamata N, et al. 2011. Multiple translocation of the AVR-Pita effector gene among chromosomes of the rice blast fungus Magnaporthe oryzae and related species. PLoS Pathog. 7:e1002147 30. Clay K, Kover PX. 1996. The Red Queen Hypothesis and plant/pathogen interactions. Annu. Rev. Phytopathol. 34:29–50 31. Condon BJ, Leng Y, Wu D, Bushley KE, Ohm RA, et al. 2013. Comparative genome structure, secondary metabolite, and effector coding capacity across Cochliobolus pathogens. PLoS Genet. 9:e1003233 32. Correll JC. 1991. The relationship between formae speciales, races, and vegetative compatibility groups in Fusarium oxysporum. Phytopathology 81:1061–64 33. Coyne JA, Orr HA. 2004. Speciation. Sunderland, MA: Sinauer Assoc. 34. Cracraft J. 1989. Speciation and its ontology: the empirical consequences of alternative species concepts for understanding patterns and processes of differentiation. In Speciation and its Consequences, ed. D Otte, JA Endler, pp. 28–59. Sunderland, MA: Sinauer Assoc. 35. Crandall ED, Sbrocco EJ, Deboer TS, Barber PH, Carpenter KE. 2012. Expansion dating: calibrating molecular clocks in marine species from expansions onto the Sunda Shelf following the last glacial maximum. Mol. Biol. Evol. 29:707–19 36. Croll D, McDonald BA. 2012. The accessory genome as a cradle for adaptive evolution in pathogens. PLoS Pathog. 8:e1002608 37. Crous PW, Groenewald JZ, Pongpanich K, Himaman W, Arzanlou M, Wingfield MJ. 2004. Cryptic speciation and host specificity among Mycosphaerella spp. occurring on Australian Acacia species grown as exotics in the tropics. Stud. Mycol. 50:457–69 38. Cui R, Schumer M, Kruesi K, Walter R, Andolfatto P, Rosenthal GG. 2013. Phylogenomics reveals extensive reticulate evolution in Xiphophorus fishes. Evolution 67:2166–79 39. Cuomo CA, Guldener U, Xu JR, Trail F, Turgeon BG, et al. 2007. The Fusarium graminearum genome reveals a link between localized polymorphism and pathogen specialization. Science 317:1400–2 40. de Cock AW, Ilieva E, L´evesque CA. 2002. Gene flow analysis of Phytophthora porri reveals a new species: Phytophthora brassicae sp. nov. Eur. J. Plant Pathol. 108:51–62 41. Delcan J, Brasier CM. 2001. Oospore viability and variation in zoospore and hyphal tip derivatives of the hybrid alder Phytophthoras. Forest Pathol. 31:65–83 42. Dettman JR, Jacobson DJ, Taylor JW. 2003. A multilocus genealogical approach to phylogenetic species recognition in the model eukaryote Neurospora. Evolution 57:2703–20 43. de Vienne DM, Giraud T, Martin OC. 2007. A congruence index for testing topological similarity between trees. Bioinformatics 23:3119–24 44. de Vienne DM, Refregier G, Hood ME, Guigue A, Devier B, et al. 2009. Hybrid sterility and inviability in the parasitic fungal species complex Microbotryum. J. Evol. Biol. 22:683–98 45. Dobzhansky T. 1950. Heredity, environment, and evolution. Science 111:161–66 46. Dobzhansky TG. 1937. Genetics and the Origin of Species. New York: Columbia Univ. Press 47. Donahoo RS, Lamour KH. 2008. Interspecific hybridization and apomixis between Phytophthora capsici and Phytophthora tropicalis. Mycologia 100:911–20 48. Edelaar P, Bjorklund M. 2011. If FST does not measure neutral genetic differentiation, then comparing it with QST is misleading. Or is it? Mol. Ecol. 20:1805–12

Annu. Rev. Phytopathol. 2014.52. Downloaded from www.annualreviews.org by College of William & Mary on 06/13/14. For personal use only.

PY52CH14-Matute

14.22

Restrepo et al.

Changes may still occur before final publication online and in print

Annu. Rev. Phytopathol. 2014.52. Downloaded from www.annualreviews.org by College of William & Mary on 06/13/14. For personal use only.

PY52CH14-Matute

ARI

19 May 2014

15:9

49. Ellwood SR, Syme RA, Moffat CS, Oliver RP. 2012. Evolution of three Pyrenophora cereal pathogens: recent divergence, speciation and evolution of non-coding DNA. Fungal Gen. Biol. 49:825–29 50. Engelbrecht CJ, Harrington TC. 2005. Intersterility, morphology and taxonomy of Ceratocystis fimbriata on sweet potato, cacao and sycamore. Mycologia 97:57–69 51. Engelbrecht CJ, Harrington TC, Steimel J, Capretti P. 2004. Genetic variation in eastern North American and putatively introduced populations of Ceratocystis fimbriata f. platani. Mol. Ecol. 13:2995–3005 ´ 52. Ersek T, English JT, Schoelz JE. 1995. Creation of species hybrids of Phytophthora with modified host ranges by zoospore fusion. Phytopathology 85:1343–47 ´ 53. Ersek T, Man in ‘t Veld WA. 2013. Phytophthora species hybrids: a novel threat to crops and natural ecosystems. In Phytophthora: A Global Perspective, ed. K Lamour, pp 37–47. Wallingford, UK: CABI ´ 54. Ersek T, Ribeiro OK. 2010. Mini review article: an annotated list of new Phytophthora species described post 1996. Acta Phytopathol. Entomol. Hung. 45:251–66 55. Erselius LJ, Shaw DS. 1982. Protein and enzyme differences between Phytophthora palmivora and P. megakarya. Evidence for self-fertilization in pairings of the two species. Trans. Br. Mycol. Soc. 78:227–38 56. Fisher MC, Henk DA, Briggs CJ, Brownstein JS, Madoff LC, et al. 2012. Emerging fungal threats to animal, plant and ecosystem health. Nature 484:186–94 57. Flier WG, Grunwald NJ, Kroon LP, Sturbaum AK, van den Bosch TB, et al. 2003. The population struc¨ ture of Phytophthora infestans from the Toluca Valley of central Mexico suggests genetic differentiation between populations from cultivated potato and wild Solanum spp. Phytopathology 93:382–90 58. Flier WG, Grunwald NJ, Kroon LPNM, van den Bosch TBM, Garay-Serrano E, et al. 2002. Phytophthora ¨ ipomoeae, a new homothallic species causing late blight on Ipomoeae longipedunculata in the Toluca Valley of central Mexico. Mycol. Res. 106:848–56 59. Flor HH. 1964. Genetics of somatic variation for pathogenicity in Melampsora lini. Phytopathology 54:823– 26 60. Fournier E, Giraud T. 2007. Sympatric genetic differentiation of a generalist pathogenic fungus, Botrytis cinerea, on two different host plants, grapevine and bramble. J. Evol. Biol. 21:122–32 61. Frankham R, Ballou JD, Dudash MR, Eldridge MDB, Fenster CB, et al. 2012. Implications of different species concepts for conserving biodiversity. Biol. Conserv. 153:25–31 62. Frenkel O, Peever TL, Chilvers MI, Ozkilinc H, Can C, et al. 2010. Ecological genetic divergence of the fungal pathogen Didymella rabiei on sympatric wild and domesticated Cicer spp. (chickpea). Appl. Environ. Microbiol. 76:30–39 63. Friesen TL, Faris JD. 2012. Characterization of plant-fungal interactions involving necrotrophic effector-producing plant pathogens. Methods Mol. Biol. 835:191–207 64. Friesen TL, Stukenbrock EH, Liu Z, Meinhardt S, Ling H, et al. 2006. Emergence of a new disease as a result of interspecific virulence gene transfer. Nat. Genet. 38:953–56 65. Galagan JE, Henn MR, Ma LJ, Cuomo CA, Birren B. 2005. Genomics of the fungal kingdom: insights into eukaryotic biology. Genome Res. 15:1620–31 66. Gale LR, Bryant JD, Calvo S, Giese H, Katan T, et al. 2005. Chromosome complement of the fungal plant pathogen Fusarium graminearum based on genetic and physical mapping and cytological observations. Genetics 171:985–1001 67. Garrigan D, Hedrick PW. 2003. Perspective: detecting adaptive molecular polymorphism: lessons from the MHC. Evolution 57:1707–22 68. Gayral P, Melo-Ferreira J, Gl´emin S, Bierne N, Carneiro M, et al. 2013. Reference-free population genomics from next-generation transcriptome data and the vertebrate-invertebrate gap. PLoS Genet. 9:e1003457 69. Giraud T. 2006. Speciation: selection against migrant pathogens: the immigrant inviability barrier in pathogens. Heredity 97:316–18 70. Giraud T, Gladieux P, Gavrilets S. 2010. Linking the emergence of fungal plant diseases with ecological speciation. Trends Ecol. Evol. 25:387–95 71. Gladieux P, Devier B, Aguileta G, Cruaud C, Giraud T. 2013. Purifying selection after episodes of recurrent adaptive diversification in fungal pathogens. Infect. Genet. Evol. 17:123–31 www.annualreviews.org • Speciation in Fungal and Oomycete Plant Pathogens

Changes may still occur before final publication online and in print

14.23

ARI

19 May 2014

15:9

72. Gladieux P, Vercken E, Fontaine MC, Hood ME, Jonot O, et al. 2011. Maintenance of fungal pathogen species that are specialized to different hosts: allopatric divergence and introgression through secondary contact. Mol. Biol. Evol. 28:459–71 73. Gladieux P, Zhang XG, Roldan-Ruiz I, Caffier V, Leroy T, et al. 2010. Evolution of the population structure of Venturia inaequalis, the apple scab fungus, associated with the domestication of its host. Mol. Ecol. 19:658–74 74. Goker M, Voglmayr H, Riethmuller A, Weiβ M, Oberwinkler F. 2003. Taxonomic aspects of Per¨ ¨ onosporaceae inferred from Bayesian molecular phylogenetics. Can. J. Bot. 81:672–83 75. Gonthier P, Garbelotto M. 2011. Amplified fragment length polymorphism and sequence analyses reveal massive gene introgression from the European fungal pathogen Heterobasidion annosum into its introduced congener H. irregulare. Mol. Ecol. 20:2756–70 76. Gonthier P, Nicolotti G, Linzer R, Guglielmo F, Garbelotto M. 2007. Invasion of European pine stands by a North American forest pathogen and its hybridization with a native interfertile taxon. Mol. Ecol. 16:1389–400 77. Gonzalez-Vera AD, Bernardes-de-Assis J, Zala M, McDonald BA, Correa-Victoria F, et al. 2010. Divergence between sympatric rice- and maize-infecting populations of Rhizoctonia solani AG-1 IA from Latin America. Phytopathology 100:172–82 78. Goodwin SB, Fry WE. 1994. Genetic analyses of interspecific hybrids between Phytophthora infestans and Phytophthora mirabilis. Exp. Mycol. 18:20–32 79. Goodwin SB, Legard DE, Smart CD, Levy M, Fry WE. 1999. Gene flow analysis of molecular markers confirms that Phytophthora mirabilis and P. infestans are separate species. Mycologia 91:796–810 80. Goss EM, Carbone I, Grunwald NJ. 2009. Ancient isolation and independent evolution of the three ¨ clonal lineages of the exotic sudden oak death pathogen Phytophthora ramorum. Mol. Ecol. 18:1161–74 81. Grunwald NJ, Flier WG. 2005. The biology of Phytophthora infestans at its center of origin. Annu. Rev. ¨ Phytopathol. 43:171–90 82. Grunwald NJ, Garbelotto M, Goss EM, Heungens K, Prospero S. 2012. Emergence of the sudden oak ¨ death pathogen Phytophthora ramorum. Trends Microbiol. 20:131–38 83. Grunwald NJ, Goss EM. 2011. Evolution and population genetics of exotic and re-emerging pathogens: ¨ novel tools and approaches. Annu. Rev. Phytopathol. 49:249–67 84. Grunwald NJ, Werres S, Goss EM, Taylor CR, Fieland VJ. 2012. Phytophthora obscura sp nov., a new ¨ species of the novel Phytophthora subclade 8D. Plant Pathol. 61:610–22 85. Guerin F, Gladieux P, Le Cam B. 2007. Origin and colonization history of newly virulent strains of the phytopathogenic fungus Venturia inaequalis. Fungal Gen. Biol. 44:284–92 86. Halkett F, Coste D, Platero GG, Zapater MF, Abadie C, Carlier J. 2010. Genetic discontinuities and disequilibria in recently established populations of the plant pathogenic fungus Mycosphaerella fijiensis. Mol. Ecol. 19:3909–23 87. Hansen EM, Maxwell DP. 1991. Species of the Phytophthora megasperma complex. Mycologia 83:376–81 88. Harrington TC, McNew DL. 1998. Partial interfertility among the Ceratocystis species on conifers. Fungal Genet. Biol. 25:44–53 89. Harrington TC, Steimel JP, Wingfield MJ, Kile GA. 1996. Isozyme variation and species delimitation in the Ceratocystis coerulescens complex. Mycologia 88:104–13 90. Hawksworth DL. 1997. Fungi and biodiversity: international incentives. Microbiologia 13:221–26 91. Hedrick PW. 2005. A standardized genetic differentiation measure. Evolution 59:1633–38 92. Heller R, Siegismund HR. 2009. Relationship between three measures of genetic differentiation GST , DEST and G ST : How wrong have we been? Mol. Ecol. 18:2080–83; discuss. 8–91 93. Hernandez-Bello MA, Chilvers MI, Akamatsu H, Peever TL. 2006. Host specificity of Ascochyta spp. infecting legumes of the Viciae and Cicerae tribes and pathogenicity of an interspecific hybrid. Phytopathology 96:1148–56 94. Hey J, Nielsen R. 2004. Multilocus methods for estimating population sizes, migration rates and divergence time, with applications to the divergence of Drosophila pseudoobscura and D. persimilis. Genetics 167:747–60 95. Hey J, Nielsen R. 2007. Integration within the Felsenstein equation for improved Markov chain Monte Carlo methods in population genetics. Proc. Natl. Acad. Sci. USA 104:2785–90

Annu. Rev. Phytopathol. 2014.52. Downloaded from www.annualreviews.org by College of William & Mary on 06/13/14. For personal use only.

PY52CH14-Matute

14.24

Restrepo et al.

Changes may still occur before final publication online and in print

Annu. Rev. Phytopathol. 2014.52. Downloaded from www.annualreviews.org by College of William & Mary on 06/13/14. For personal use only.

PY52CH14-Matute

ARI

19 May 2014

15:9

96. Hibbett DS, Taylor JW. 2013. Fungal systematics: Is a new age of enlightenment at hand? Nat. Rev. Microbiol. 11:129–33 97. Holsinger KE, Weir BS. 2009. Genetics in geographically structured populations: defining, estimating and interpreting FST . Nat. Rev. Genet. 10:639–50 98. Huai W-X, Tian G, Hansen EM, Zhao W-X, Goheen EM, et al. 2013. Identification of Phytophthora species baited and isolated from forest soil and streams in northwestern Yunnan Province, China. Forest Pathol. 43:87–103 99. Hudson RR. 1990. Gene genealogies and the coalescent process. In Oxford Surveys in Evolutionary Biology, ed. DJ Futuyma, J Antonovics, pp. 1–44. Oxford: Oxford Univ. Press 100. Hughes AL. 2007. Looking for Darwin in all the wrong places: the misguided quest for positive selection at the nucleotide sequence level. Heredity 99:364–73 101. Ioos R, Andrieux A, Marcais B, Frey P. 2006. Genetic characterization of the natural hybrid species Phytophthora alni as inferred from nuclear and mitochondrial DNA analyses. Fungal Gen. Biol. 43:511–29 102. Jost L. 2008. GST and its relatives do not measure differentiation. Mol. Ecol. 17:4015–26 103. Jost L. 2009. Reply: D versus G ST : response to Heller and Siegismund 2009 and Ryman and Leimar 2009. Mol. Ecol. 18:2088–91 104. Jurgenson JE, Bowden RL, Zeller KA, Leslie JF, Alexander NJ, Plattner RD. 2002. A genetic map of Gibberella zeae (Fusarium graminearum). Genetics 160:1451–60 105. Kaiser WJ, Wang B-C, Rogers JD. 1997. Ascochyta fabae and A. lentis: Host specificity, teleomorphs (Didymella), hybrid analysis, and taxonomic status. Plant Dis. 81:809–16 106. Kang JC, Crous PW, Schoch CL. 2001. Species concepts in the Cylindrocladium floridanum and Cy. spathiphylli complexes (Hypocreaceae) based on multi-allelic sequence data, sexual compatibility and morphology. Syst. Appl. Microbiol. 24:206–17 107. Kaplan N, Hudson RR, Iizuka M. 1991. The coalescent process in models with selection, recombination and geographic subdivision. Genet. Res. 57:83–91 108. Karl SA, Toonen RJ, Grant WS, Bowen BW. 2012. Common misconceptions in molecular ecology: echoes of the modern synthesis. Mol. Ecol. 21:4171–89 109. Kofler R, Pandey RV, Schlotterer C. 2011. PoPoolation2: identifying differentiation between populations using sequencing of pooled DNA samples (Pool-Seq). Bioinformatics 27:3435–36 110. Koufopanou V, Burt A, Szaro T, Taylor JW. 2001. Gene genealogies, cryptic species, and molecular evolution in the human pathogen Coccidioides immitis and relatives (Ascomycota, Onygenales). Mol. Biol. Evol. 18:1246–58 111. Kryazhimskiy S, Plotkin JB. 2008. The population genetics of dN/dS. PLoS Genet. 4:e1000304 112. Leach JE, Vera Cruz CM, Bai J, Leung H. 2001. Pathogen fitness penalty as a predictor of durability of disease resistance genes. Annu. Rev. Phytopathol. 39:187–224 113. Le Gac M, Hood ME, Giraud T. 2007. Evolution of reproductive isolation within a parasitic fungal species complex. Evolution 61:1781–87 114. Lemey P, Lott M, Martin DP, Moulton V. 2009. Identifying recombinants in human and primate immunodeficiency virus sequence alignments using quartet scanning. BMC Bioinform. 10:126 115. Lemey P, Minin VN, Bielejec F, Kosakovsky Pond SL, Suchard MA. 2012. A counting renaissance: combining stochastic mapping and empirical Bayes to quickly detect amino acid sites under positive selection. Bioinformatics 28:3248–56 116. Lemey P, Rambaut A, Welch JJ, Suchard MA. 2010. Phylogeography takes a relaxed random walk in continuous space and time. Mol. Biol. Evol. 27:1877–85 117. Little R, Manners JG. 1969. Somatic recombination in yellow rust of wheat (Puccinia striiformis). 2. Germ tube fusions, nuclear number and nuclear size. Trans. Br. Mycol. Soc. 53:259–67 118. Malik M, Vilgalys R. 1999. Somatic incompatibility in fungi. In Structure and Dynamics of Fungal Populations, ed. JJ Worrall, pp. 123–38. Dordrecht, Neth.: Kluwer Acad. Publ. 119. Mallet J. 2007. Hybrid speciation. Nature 446:279–83 120. Man in ‘t Veld WA. 2007. Gene flow analysis demonstrates that Phytophthora fragariae var. rubi constitutes a distinct species, Phytophthora rubi comb. nov. Mycologia 99:222–26 121. May KJ, Drent A, Irwin JAG. 2003. Interspecific hybrids between the homothallic Phytophthora sojae and Phytophthora vignae. Australas. Plant Pathol. 32:353–59 www.annualreviews.org • Speciation in Fungal and Oomycete Plant Pathogens

Changes may still occur before final publication online and in print

14.25

ARI

19 May 2014

15:9

122. McCoy KD. 2003. Sympatric speciation in parasites: What is sympatry? Trends Parasitol. 19:400–4 123. Meirmans PG, Hedrick PW. 2011. Assessing population structure: F(ST) and related measures. Mol. Ecol. Resour. 11:5–18 124. Mendelson TC, Shaw KL. 2012. The (mis)concept of species recognition. Trends Ecol. Evol. 27:421–27 125. Michelmore R. 2000. Genomic approaches to plant disease resistance. Curr. Opin. Plant Biol. 3:125–31 126. Mitchell MN, Ocamb CM, Grunwald NJ, Mancino LE, Gent DH. 2011. Genetic and pathogenic ¨ relatedness of Pseudoperonospora cubensis and P. humuli. Phytopathology 101:805–18 127. Montarry J, Hamelin FM, Glais I, Corbi R, Andrivon D. 2010. Fitness costs associated with unnecessary virulence factors and life history traits: evolutionary insights from the potato late blight pathogen Phytophthora infestans. BMC Evol. Biol. 10:283 128. Moon CD, Craven KD, Leuchtmann A, Clement SL, Schardl CL. 2004. Prevalence of interspecific hybrids amongst asexual fungal endophytes of grasses. Mol. Ecol. 13:1455–67 129. Nagel JH, Gryzenhout M, Slippers B, Wingfield MJ, Hardy GE, et al. 2013. Characterization of Phytophthora hybrids from ITS clade 6 associated with riparian ecosystems in South Africa and Australia. Fungal Biol. 117:329–47 130. Nei M. 1973. Analysis of gene diversity in subdivided populations. Proc. Natl. Acad. Sci. USA 70:3321–23 131. Nelson RR. 1959. Genetics of Cochliobolus heterostrophus. II. Genetic factors inhibiting ascospore formation. Mycologia 51:24–30 132. Nelson RR. 1964. Bridging interspecific incompatibility in the ascomycetous genus Cochliobolus. Evolution 18:700–4 133. Nixon KC, Wheeler QD. 1990. An amplification of the phylogenetic species concept. Cladistics 6:211–23 134. Noor MA, Bennett SM. 2009. Islands of speciation or mirages in the desert? Examining the role of restricted recombination in maintaining species. Heredity 103:439–44 135. O’Donnell K, Ward TJ, Aberra D, Kistler HC, Aoki T, et al. 2008. Multilocus genotyping and molecular phylogenetics resolve a novel head blight pathogen within the Fusarium graminearum species complex from Ethiopia. Fungal Gen. Biol. 45:1514–22 136. Ordonez ˜ ME, Hohl HR, Velasco JA, Ramon MP, Oyarzun PJ, et al. 2000. A novel population of Phytophthora, similar to P. infestans, attacks wild Solanum species in Ecuador. Phytopathology 90:197–202 137. Orr HA, Presgraves DC. 2000. Speciation by postzygotic isolation: forces, genes and molecules. BioEssays 22:1085–94 138. Ozkilinc H, Frenkel O, Abbo S, Eshed R, Sherman A, et al. 2010. A comparative study of Turkish and Israeli populations of Didymella rabiei, the ascochyta blight pathogen of chickpea. Plant Pathol. 59:492– 503 139. Park RF, Wellings CR. 2012. Somatic hybridization in the Uredinales. Annu. Rev. Phytopathol. 50:219–39 140. Parker IM, Gilbert GS. 2004. The evolutionary ecology of novel plant pathogen interactions. Annu. Rev. Ecol. Syst. 35:675–700 141. Petit RJ, Excoffier L. 2009. Gene flow and species delimitation. Trends Ecol. Evol. 24:386–93 142. Pipe ND, Buck KW, Brasier CM. 1995. Genomic fingerprinting supports the separation of Ophiostoma piceae into two species. Mycol. Res. 99:1182–86 143. Ploch S, Choi YJ, Rost C, Shin HD, Schilling E, Thines M. 2010. Evolution of diversity in Albugo is driven by high host specificity and multiple speciation events on closely related Brassicaceae. Mol. Phylogenet. Evol. 57:812–20 144. Plotkin JB, Kudla G. 2011. Synonymous but not the same: the causes and consequences of codon bias. Nat. Rev. Genet. 12:32–42 145. Pons JD, Barraclough TG, Gomez-Zurita J, Cardoso A, Duran DP, et al. 2006. Sequence-based species delimitation for the DNA taxonomy of undescribed insects. Syst. Biol. 55:1–15 146. Price AL, Tandon A, Patterson N, Barnes KC, Rafaels N, et al. 2009. Sensitive detection of chromosomal segments of distinct ancestry in admixed populations. PLoS Genet. 5:e1000519 147. Price PW. 2007. Parasite patterns. Evolution 61:2450–51 148. Price TD, Bouvier MM. 2002. The evolution of F1 postzygotic incompatibilities in birds. Evolution 56:2083–89 149. Puhalla JE, Spieth PT. 1985. A comparison of heterokaryosis and vegetative incompatibility among varities of Gibberella fujikuroi (Fusarium moniliforme). Exp. Mycol. 9:39–47

Annu. Rev. Phytopathol. 2014.52. Downloaded from www.annualreviews.org by College of William & Mary on 06/13/14. For personal use only.

PY52CH14-Matute

14.26

Restrepo et al.

Changes may still occur before final publication online and in print

Annu. Rev. Phytopathol. 2014.52. Downloaded from www.annualreviews.org by College of William & Mary on 06/13/14. For personal use only.

PY52CH14-Matute

ARI

19 May 2014

15:9

150. Pujol B, Wilson AJ, Ross RI, Pannell JR. 2008. Are QST-FST comparisons for natural populations meaningful? Mol. Ecol. 17:4782–85 151. Rabosky DL, Matute DR. 2013. Macroevolutionary speciation rates are decoupled from the evolution of intrinsic reproductive isolation in Drosophila and birds. Proc. Natl. Acad. Sci. USA 110:15354–59 152. Raffaele S, Farrer RA, Cano LM, Studholme DJ, MacLean D, et al. 2010. Genome evolution following host jumps in the Irish potato famine pathogen lineage. Science 330:1540–43 153. Ree RH. 2005. Detecting the historical signature of key innovations using stochastic models of character evolution and cladogenesis. Evolution 59:257–65 154. Rintoul TL, Eggertson QA, L´evesque CA. 2011. Multigene phylogenetic analyses to delimit new species in fungal plant pathogens. Methods Mol. Biol. 835:549–69 155. Rivas GG, Zapater MF, Abadie C, Carlier J. 2004. Founder effects and stochastic dispersal at the continental scale of the fungal pathogen of bananas Mycosphaerella fijiensis. Mol. Ecol. 13:471–82 156. Rosenberg NA. 2003. The shapes of neutral gene genealogies in two species: probabilities of monophyly, paraphyly, and polyphyly in a coalescent model. Evolution 57:1465–77 157. Rosenberg NA, Nordborg M. 2002. Genealogical trees, coalescent theory and the analysis of genetic polymorphisms. Nat. Rev. Genet. 3:380–90 158. Rouxel M, Mestre P, Comont G, Lehman BL, Schilder A, Delmotte F. 2013. Phylogenetic and experimental evidence for host-specialized cryptic species in a biotrophic oomycete. New Phytol. 197:251–63 159. Roy BA. 2001. Patterns of association between crucifers and their flower-mimic pathogens: Host jumps are more common than coevolution or cospeciation. Evolution 55:41–53 160. Runge F, Thines M. 2012. Reevaluation of host specificity of the closely related species Pseudoperonospora humuli and P. cubensis. Plant Dis. 96:55–61 161. Ryman N, Leimar O. 2009. G(ST) is still a useful measure of genetic differentiation: a comment on Jost’s D. Mol. Ecol. 18:2084–87 162. Salichos L, Rokas A. 2013. Inferring ancient divergences requires genes with strong phylogenetic signals. Nature 497:327–31 163. Sansome ER, Brasier CM, Griffin MJ. 1975. Chromosome size differences in “Phytophthora palmivora” on cocoa in West Africa. Nature 255:704–5 164. Sarver BA, Ward TJ, Gale LR, Broz K, Kistler HC, et al. 2011. Novel Fusarium head blight pathogens from Nepal and Louisiana revealed by multilocus genealogical concordance. Fungal Gen. Biol. 48:1096– 107 165. Schardl CL, Craven KD. 2003. Interspecific hybridization in plant-associated fungi and oomycetes: a review. Mol. Ecol. 12:2861–73 166. Schluter D. 2009. Evidence for ecological speciation and its alternative. Science 323:737–41 167. Schmidt SM, Panstruga R. 2011. Pathogenomics of fungal plant parasites: What have we learnt about pathogenesis? Curr. Opin. Plant Biol. 14:392–99 168. Schroder S, Telle S, Nick P, Thines M. 2011. Cryptic diversity of Plasmopara viticola (Oomycota, Per¨ onosporaceae) in North America. Org. Divers. Evol. 11(1):3–7 169. Schulze-Lefert P, Panstruga R. 2011. A molecular evolutionary concept connecting nonhost resistance, pathogen host range, and pathogen speciation. Trends Plant Sci. 16:117–25 170. Shepherd CJ. 1978. Mating behaviour of Australian isolates of Phytophthora species. I. Inter- and intraspecific mating. Aust. J. Bot. 26:123–38 171. Shepherd CJ, Cunningham RB. 1978. Mating behaviour of Australian isolates of Phytophthora species. II. Phytophthora cinnamoni. Aust. J. Bot. 26:139–51 172. Shepherd CJ, Pratt BH, Taylor PA. 1974. Comparative morphology and behaviour of A1 and A2 isolates of Phytophthora cinnamomi. Aust. J. Bot. 22:461–70 173. Silva DN, Talhinhas P, Cai L, Manuel L, Gichuru EK, et al. 2012. Host-jump drives rapid and recent ecological speciation of the emergent fungal pathogen Colletotrichum kahawae. Mol. Ecol. 21:2655–70 174. Sousa V, Hey J. 2013. Understanding the origin of species with genome-scale data: modelling gene flow. Nat. Rev. Genet. 14:404–14 175. Staats M, van Baarlen P, van Kan JAL. 2005. Molecular phylogeny of the plant pathogenic genus Botrytis and the evolution of host specificity. Mol. Biol. Evol. 22:333–46 www.annualreviews.org • Speciation in Fungal and Oomycete Plant Pathogens

Changes may still occur before final publication online and in print

14.27

ARI

19 May 2014

15:9

176. Strasburg JL, Rieseberg LH. 2010. How robust are “isolation with migration” analyses to violations of the IM model? A simulation study. Mol. Biol. Evol. 27:297–310 177. Stukenbrock EH. 2013. Evolution, selection and isolation: a genomic view of speciation in fungal plant pathogens. New Phytol. 199:895–907 178. Stukenbrock EH, Banke S, McDonald BA. 2006. Global migration patterns in the fungal wheat pathogen Phaeosphaeria nodorum. Mol. Ecol. 15:2895–904 179. Stukenbrock EH, Christiansen FB, Hansen TT, Dutheil JY, Schierup MH. 2012. Fusion of two divergent fungal individuals led to the recent emergence of a unique widespread pathogen species. Proc. Natl. Acad. Sci. USA 109:10954–59 180. Stukenbrock EH, Jorgensen FG, Zala M, Hansen TT, McDonald BA, Schierup MH. 2010. Wholegenome and chromosome evolution associated with host adaptation and speciation of the wheat pathogen Mycosphaerella graminicola. PLoS Genet. 6:e1001189 181. Stukenbrock EH, McDonald BA. 2008. The origins of plant pathogens in agro-ecosystems. Annu. Rev. Phytopathol. 46:75–100 182. Tang H, Peng J, Wang P, Risch NJ. 2005. Estimation of individual admixture: analytical and study design considerations. Genet. Epidemiol. 28:289–301 183. Taylor JW, Jacobson DJ, Kroken S, Kasuga T, Geiser DM, et al. 2000. Phylogenetic species recognition and species concepts in fungi. Fungal Gen. Biol. 31:21–32 184. Taylor JW, Turner E, Townsend JP, Dettman JR, Jacobson D. 2006. Eukaryotic microbes, species recognition and the geographic limits of species: examples from the kingdom Fungi. Philos. Trans. R. Soc. B 361:1947–63 185. Turner E, Jacobson DJ, Taylor JW. 2010. Reinforced postmating reproductive isolation barriers in Neurospora, an ascomycete microfungus. J. Evol. Biol. 23:1642–56 186. Vaillancourt LJ, Hanau RM. 1992. Genetic and morphological comparisons of Glomerella (Colletotrichum) isolates from maize and from sorghum. Exp. Mycol. 16:219–29 187. van de Wouw AP, Howlett BJ. 2011. Fungal pathogenicity genes in the age of “omics.” Mol. Plant Pathol. 12:507–14 188. Vialle A, Feau N, Frey P, Bernier L, Hamelin RC. 2013. Phylogenetic species recognition reveals hostspecific lineages among poplar rust fungi. Mol. Phylogenet. Evol. 66:628–44 189. Walker AS, Gautier AL, Confais J, Martinho D, Viaud M, et al. 2011. Botrytis pseudocinerea, a new cryptic species causing gray mold in French vineyards in sympatry with Botrytis cinerea. Phytopathology 101:1433–45 190. Waples RS, Gaggiotti O. 2006. What is a population? An empirical evaluation of some genetic methods for identifying the number of gene pools and their degree of connectivity. Mol. Ecol. 15:1419–39 191. Whitlock MC, Guillaume F. 2009. Testing for spatially divergent selection: comparing QST to FST. Genetics 183:1055–63 192. Widmer A, Lexer C, Cozzolino S. 2009. Evolution of reproductive isolation in plants. Heredity 102:31–38 193. Wilson DJ, Hernandez RD, Andolfatto P, Przeworski M. 2011. A population genetics–phylogenetics approach to inferring natural selection in coding sequences. PLoS Genet. 7:e1002395 194. Worrall JJ. 1997. Somatic incompatibility in basidiomycetes. Mycologia 89:24–36 195. Wright S. 1969. Evolution and the Genetics of Populations, Vol. 2: The Theory of Gene Frequencies. Chicago: Univ. Chicago Press. 196. Xhaard C, Fabre B, Andrieux A, Gladieux P, Barres B, et al. 2011. The genetic structure of the plant pathogenic fungus Melampsora larici-populina on its wild host is extensively impacted by host domestication. Mol. Ecol. 20:2739–55 197. Yang Z, Bielawski JP. 2000. Statistical methods for detecting molecular adaptation. Trends Ecol. Evol. 15:496–503 198. Zaffarano PL, McDonald BA, Linde CC. 2008. Rapid speciation following recent host shifts in the plant pathogenic fungus Rhynchosporium. Evolution 62:1418–36 199. Zaffarano PL, McDonald BA, Linde CC. 2011. Two new species of Rhynchosporium. Mycologia 103:195– 202 200. Zhong S, Steffenson BJ, Martinez JP, Ciuffetti LM. 2002. A molecular genetic map and electrophoretic karyotype of the plant pathogenic fungus Cochliobolus sativus. Mol. Plant-Microbe Interact. 15:481–92

Annu. Rev. Phytopathol. 2014.52. Downloaded from www.annualreviews.org by College of William & Mary on 06/13/14. For personal use only.

PY52CH14-Matute

14.28

Restrepo et al.

Changes may still occur before final publication online and in print

Speciation in fungal and oomycete plant pathogens.

The process of speciation, by definition, involves evolution of one or more reproductive isolating mechanisms that split a single species into two tha...
660KB Sizes 3 Downloads 4 Views