• Aucun résultat trouvé

Strong phylogeographic costructure between the anther-smut fungus and its white campion host

N/A
N/A
Protected

Academic year: 2021

Partager "Strong phylogeographic costructure between the anther-smut fungus and its white campion host"

Copied!
27
0
0

Texte intégral

(1)

HAL Id: hal-01365319

https://hal.archives-ouvertes.fr/hal-01365319

Submitted on 16 Nov 2016

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci-entific research documents, whether they are pub-lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Distributed under a Creative Commons Attribution - NonCommercial| 4.0 International License

Strong phylogeographic costructure between the

anther-smut fungus and its white campion host

Alice Feurtey, Pierre Gladieux, Michael E. Hood, Alodie Snirc, Amandine

Cornille, Lisa Rosenthal, Tatiana Giraud

To cite this version:

Alice Feurtey, Pierre Gladieux, Michael E. Hood, Alodie Snirc, Amandine Cornille, et al.. Strong phylogeographic costructure between the anther-smut fungus and its white campion host. New Phy-tologist, Wiley, 2016, 3 (212), pp.668-679. �10.1111/nph.14125�. �hal-01365319�

(2)

1

S

TRONG PHYLOGEOGRAPHIC COSTRUCTURE BETWEEN THE ANTHER

-

SMUT

1

FUNGUS AND ITS WHITE CAMPION

H

OST

2 3 4

Alice Feurtey

1

, Pierre Gladieux

1,2

, Michael E Hood

3

, Alodie Snirc

1

, Amandine Cornille

1

, Lisa

5

Rosenthal

3

and Tatiana Giraud

1 6

1

Ecologie Systématique Evolution, Univ. Sud, CNRS, AgroParisTech, Université

Paris-7

Saclay, 91400 Orsay, France

8

2

INRA, UMR BGPI, 34398 Montpellier, France

9

3

Department of Biology, Amherst College, Amherst, Massachusetts 01002, USA

10

11

Corresponding author: Tatiana Giraud tatiana.giraud@u-psud.fr

12

Laboratoire Ecologie, Systématique et Evolution, Bâtiment 360, Université de Paris-Sud,

13

91405 Orsay cedex France

14

phone: +33 1 69 15 56 69 fax: +33 1 69 15 46 97

15

16

Running title: Strong host and pathogen costructure

17

Word count: 5916 without the references or figure legends

18

Number of tables: 0

19

Number of figures (3 in colour) : 5

20

Number of supporting information: 6 supplementary figures, 6 supplementary Tables and 1

21

supplementary Note

22

23 24

(3)

2

Summary

25

 While congruence between host and pathogen phylogenies has been extensively 26

investigated, the congruence between host and pathogen genetic structures at the within-27

species level has received little attention. 28

 Using an unprecedented and comprehensive collection of associated plant-pathogen 29

samples, we investigated the degree of congruence between the genetic structures across 30

Europe of two evolutionary and ecological model organisms, the anther-smut pathogen 31

Microbotryum lychnidis-dioicae and its host plant Silene latifolia. 32

 We demonstrated a significant and particularly strong level of host-pathogen costructure, 33

with three main genetic clusters displaying highly similar spatial ranges in Western, Eastern 34

Europe, and Italy, respectively. Correcting for the geographical component of genetic 35

variation, significant correlations were still found between the genetic distances of anther-36

smut and host populations. Inoculation experiments suggested plant local adaptation, at the 37

cluster level, for resistance to pathogens. 38

 These findings indicate that the pathogen has remained isolated in the same fragmented 39

southern refugia as its host plant during the last glaciation, and that little long-distance 40

dispersal has occurred since the recolonization of Europe in neither the plant nor the 41

pathogen, despite their known ability to travel across continents. This, together with the 42

inoculation results, suggests that coevolutionary and competitive processes may be drivers 43

of host-pathogen costructure. 44

45

Key words: coevolution, congruent population structure, cross-inoculations, local adaptation,

46

Microbotryum violaceum, partial Mantel test

47

48 49 50 51

(4)

3

Introduction

52

Understanding and management of fungal diseases of plants is aided by appreciating the processes 53

that generate and maintain spatial patterns in pathogen diversity (Williams, 2010). Predicting the 54

impacts of coevolving systems, and more generally our ability to manage biological interactions 55

embedded in a complex network of agricultural and natural landscapes, requires knowledge about 56

how spatial heterogeneity, selection, and environmental variables contribute to differentiation of 57

populations (Barrett et al., 2008). The degree of genetic variation in specialized pathogens is 58

particularly influenced by their host’s distribution, abundance, and genetic makeup (Wilson et al., 59

2005). This leads to the expectation that the population structure of pathogens would reflect those 60

of their hosts (Criscione et al., 2005). Under a very similar rationale, host and pathogen phylogenies 61

have long been expected to be congruent (Hafner & Page, 1995), while it has come as a recent 62

surprise that host shifts are actually the main diversification processes in the vast majority of host-63

pathogens associations (de Vienne et al., 2013). Unlike the between-species level, however, few 64

studies have addressed the degrees of congruence between host and pathogen genetic structures 65

within-species, although such studies may also reveal unexpected findings, as did the between-66

species phylogenetic congruence studies (de Vienne et al., 2013). In addition, within-species 67

coevolution and diversification on different host correspond to different evolutionary mechanisms 68

that can thus lead to different patterns (de Vienne et al., 2013). 69

The few published studies on host-pathogen costructure have been conducted mostly on parasites of 70

animals (Dybdahl & Lively, 1996; McCOY et al., 2005; Miura et al., 2006; Nieberding et al., 2008; Rich 71

et al., 2008; Amico & Nickrent, 2009; Nakao et al., 2009), with a few, notable exceptions in plant 72

parasites (Michalakis et al., 1993; Whiteman & Parker, 2005; Wirth et al., 2005; Tsai & Manos, 2010; 73

Magalhaes et al., 2011). The genetic structure of the pathogens has, in these cases, often been found 74

congruent with those of their host. Due to shorter generation times and the tendency toward selfing 75

mating systems or partial clonality, pathogen populations have nevertheless often been more 76

subdivided than their hosts (Nieberding & Olivieri, 2007; Barrett et al., 2008). This may help in 77

reconstructing the host population history (Nieberding & Olivieri, 2007; Barrett et al., 2008), provided 78

that pathogens have followed the same history as their hosts. Indeed, if pathogens show a stronger 79

genetic structure than their hosts’, analyses of pathogen population subdivision will be more 80

informative about the joint past demographic histories of the two partners than studies conducted 81

using host data alone. In fungal plant pathogens, however, studies of population structure have 82

focused on agents of crop diseases, where the host population is genetically homogeneous and 83

artificially distributed, and the pathogens are dispersed widely, in part by humans transport, and are 84

often recently introduced exotic species (Enjalbert et al., 2005 ; Barres et al., 2008; Fournier & 85

(5)

4 Giraud, 2008; Gladieux et al., 2008; Saleh et al., 2014).

86

Here, we report a study on the co-phylogeography of the castrating fungal pathogen Microbotryum 87

lychnidis-dioicae (previously named M. violaceum or Ustilago violacea) and its host plant Silene 88

latifolia. Both are model organisms for a variety of topics in ecology and evolution, and as such 89

numerous studies are available on their life-history, population biology and natural history 90

(Bernasconi et al., 2009; Gladieux et al., 2015). This system is ideal for studying the co-structure 91

between pathogens and their host plants, given that the fungus is an obligate pathogen that is not 92

capable of persisting in the environment (Hood et al., 2010) and is highly host specific in nature 93

(Vercken et al., 2010; Gladieux et al., 2011), the same vectors disperse the fungus and the host pollen 94

(i.e., the pollinators) (Thrall et al., 1995), Silene and Microbotryum species have identical generation 95

times (i.e., one per year) (Thrall & Jarosz, 1994), and there has been no attempt to modify plant 96

distribution or genotypes or to control the disease because its effects are without economic 97

consequences (Bernasconi et al., 2009). A stronger phylogeographic structure is expected in the 98

fungal pathogen than its host plant because M. lychnidis-dioicae is highly selfing (Hood & Antonovics, 99

2000; Giraud et al., 2005; Giraud et al., 2008) while S. latifolia is dioecious and shows inbreeding 100

depression (Teixeira et al., 2009), because the fungus only disperses via pollinators, while gene flow in 101

the plant also occurs via its seeds in addition to pollen, and because pollinators discriminate against 102

diseased plants (Shykoff & Bucheli, 1995), likely reducing spore transport compared to pollen flow. 103

Note that M. lychnidis-dioicae, unlike most fungi, cannot proliferate by asexual reproduction beyond 104

a few cells after meiosis, and a sex event is required for infecting a new plant. However, the mating 105

system is highly automictic, i.e., with mostly diploid selfing by mating among products of the same 106

meiosis (Oudemans et al., 1998; Hood & Antonovics, 2000; Thomas et al., 2003; Badouin et al., 2015). 107

A stronger population genetic structure has in fact been found in M. lychnidis-dioicae than in S. 108

latifolia, although costructure has only been investigated at a very small spatial scale so far (Delmotte 109

et al., 1999). Local adaptation in this system has similarly been investigated at a very small scale to 110

date, with local plant genotypes being more resistant to their local anther-smut pathogen than to 111

pathogen from other nearby populations (Kaltz et al., 1999). Differences were found in terms of the 112

percentage of plants that became diseased after experimental inoculations, while no gene-for-gene 113

interactions have been detected in this system. 114

Previous studies on the population genetics of M. lychnidis-dioicae provided one of the most clear-115

cut examples of phylogeographic structure in pathogens (Vercken et al., 2010; Gladieux et al., 2011; 116

Fontaine et al., 2013; Gladieux et al., 2013). Clustering analyses revealed the existence of three main 117

genetically distinct groups within M. lychnidis-dioicae, with distributions strikingly similar to the 118

(6)

5 major intraspecific lineages identified in Europe for many temperate plant and animal species

119

(Hewitt, 1999). This within-species structure is believed to reflect recolonization from southern 120

refugia after glaciation. The Eastern and Western clusters were each further split into two groups, 121

consistent with colonization from distinct refugia located in the Balkans and further East in Eurasia, 122

and survival of the pathogen in distinct regional refugia in Western Europe. 123

Phylogeographic analyses of chloroplast genetic data in the host plant, S. latifolia, also revealed 124

patterns of population structure consistent with post-glacial expansion from distinct Mediterranean 125

refugia, although less clear-cut than for the anther-smut fungus, with large geographical overlaps of 126

maternal lineages (Taylor & Keller, 2007). However, thorough comparisons between host and 127

pathogen structures were prohibited by the facts that the plant and pathogen samples were not 128

collected at the same localities, and that a selective sweep in S. latifolia chloroplast DNA may have 129

obscured historical patterns (Muir & Filatov, 2007). Other studies using microsatellite markers in the 130

host plant S. latifolia were done, but at smaller geographical scales (Delmotte et al., 1999; Barluenga 131

et al., 2011; Magalhaes et al., 2011), or with less dense sampling (Keller et al., 2012). A clear 132

differentiation was nevertheless found between an Eastern and a Western European clusters (Keller 133

et al., 2012). 134

Thanks to a collection of associated plant/pathogen samples (fungal spores and plant material from 135

the same disease plants) with an unprecedented sampling of the plant in terms of both density and 136

geographical scale in Europe, we investigated here the costructure of M. lychnidis-dioicae and S. 137

latifolia using nuclear microsatellite markers for both the plant and the fungus and we used 138

inoculation experiments to test for the existence of local adaptation at the cluster level. More 139

specifically, we addressed the following questions: 1) Can we detect footprints of three glacial refugia 140

in the plant population structure as occurs in its fungal pathogen? 2) How congruent are the 141

population structures of the host and the pathogen at the European scale? 3) Are pairwise genetic 142

distances between pathogen individuals and host individuals significantly correlated? 4) In this case, 143

does the correlation still hold when removing the effect of the isolation by geographic distance, i.e., 144

are the host and pathogen population genetic structures more congruent than expected if they were 145

just similarly affected by the same drift-migration equilibrium after similar post-glacial recolonization 146

histories? 5) Is there experimental evidence of local adaptation in the host-pathogen association that 147

may contribute to congruent spatial structure at the cluster level? 148

149 150

(7)

6

Material and methods

151

Sampling, genotyping and species identification 152

We used the same diseased S. latifolia plants from which the anther-smut samples had been 153

previously genotyped (Vercken et al., 2010). Because Microbotryum spores were collected in flower 154

buds of S. latifolia, samples included both materials from the fungus and from the parasitized plants. 155

DNA was extracted using the Nucleospin Plant II kit following manufacturer instructions (Macherey 156

Nagel R). We genotyped the plants using 16 microsatellite markers distributed across the genome 157

(Moccia et al., 2009). Microsatellite markers were amplified by multiplex PCR, with the Multiplex PCR 158

Kit (QIAGEN) following manufacturer instructions (details of the multiplex PCR programs used in 159

Tables S1 and S2). Genotyping was performed on an ABI PRISM X3730XL at the GENTYANE 160

Genotyping Platform (INRA, Clermont, France). Alleles were scored with GENEMAPPER 4.0 (Applied 161

Biosystems). 162

For the anther-smut samples, we used a previously published dataset of genotypes produced with a 163

set of 11 microsatellite markers (Vercken et al., 2010). The Microbotryum dataset was rendered 164

haploid by randomly removing one allele in the rare cases of heterozygosity (Vercken et al., 2010). 165

Indeed, individuals are almost completely homozygous because of high selfing rates (Hood & 166

Antonovics, 2000; Giraud, 2004; Giraud et al., 2005; Giraud et al., 2008), and this may bias analyses 167

of population subdivision using Bayesian clustering methods assuming random mating. Apparent 168

heterozygosity may in principle result from co-infection of the plant by multiple genotypes 169

(Tollenaere et al., 2012), but previous studies have shown that a given S. latifolia bud typically hosts a 170

single M. lychnidis-dioica genotype, even in cases of multiple infections at the plant scale (Lopez-171

Villavicencio et al., 2007). For all the analyses we kept only the individuals (both for the fungus and 172

the plant) successfully genotyped at eight loci or more. 173

Our plant dataset contained individuals that were identified by the samplers as S. latifolia (n=592 174

individuals), S. dioica (n=74) and putative hybrids between these species (n=50). The anther-smut 175

fungi dataset initially contained samples that were identified by the samplers as collected on S. 176

latifolia (n=669), on S. dioica (371) and on hybrids between the two plant species (n=30). We first ran 177

analyses to ensure that our samples all belonged to the species studied, S. latifolia for the plant and 178

M. lychnidis-dioicae for the fungus. Indeed, the sister plant species S. dioica and S. latifolia have 179

largely overlapping distributions and S. dioica is also affected by anther-smut disease, caused by M. 180

silenes-dioicae, the sister species of M. lychnidis-dioicae. Both the plants and the anther-smut fungi 181

pairs hybridize in nature and plant and fungal hybrids are difficult to detect morphologically 182

(Karrenberg & Favre, 2008; Vercken et al., 2010; Gladieux et al., 2011). We therefore used Bayesian 183

(8)

7 multilocus genotype assignments to detect potential hybrids in plants and anther-smut fungi (Falush 184

et al., 2003). We identified 578 and 679 genotypes as ‘pure’ representative of S. latifolia and M. 185

lychnidis-dioicae, respectively (Supplementary Notes S1; Fig. S1). Only these ‘pure’ individuals were 186

retained for subsequent analyses, from which we could obtain both the genotype of the plant and of 187

its associated anther-smut pathogen for 391 samples. 188

189

Population structure 190

We investigated the population structure within the M. lychnidis-dioicae and S. latifolia species. First, 191

we used the model-based Bayesian clustering methods implemented in TESS v2.3 (Guillot, 2009). 192

Analyses were performed anew on the genetic marker dataset created for the fungus in our previous 193

study (Vercken et al., 2010) in order to have the same software tools applied on the plant and fungal 194

datasets. We only kept one representative of each genotype in the dataset for clustering analyses 195

and assigned the repeated genotypes to the same cluster as the corresponding representative 196

genotypes. We used the conditional auto-regressive (CAR) Gaussian model of admixture with 197

200,000 MCMC iterations after a burn-in of 100,000 iterations. We used the CAR Gaussian model of 198

admixture with linear trend surface, setting the spatial interaction parameter (ρ) to 0.6. These 199

parameters (ρ and trend) affect the weight given to spatial distance when clustering genotypes. TESS 200

outputs were processed with CLUMPP v1.1.2 (Jakobsson & Rosenberg, 2007), to identify distinct 201

clustering solutions in the replicated runs for each K value. We considered that an individual 202

belonged to a given genetic cluster if its membership coefficient for this cluster was superior to 0.8. 203

As described above, repeated genotypes were not included in the analyses that established cluster 204

assignment per population, but they were added back for subsequent analyses. 205

We also performed population structure analyses not relying on any population genetic model and 206

therefore not affected by deviations from Hardy-Weinberg assumptions, such as those caused by 207

selfing. Principal component analyses (PCA) were performed using the ade4 R package and 208

discriminant analyses of principal component (DAPC) using the adegenet R package with the R 209

software (find.clusters function). 210

For both the host and the parasite, we tested the existence of isolation by distance (IBD) patterns by 211

assessing the significance of the correlations between matrices of genetic and geographic distances, 212

using a Mantel test (100,000 resamples for the null distribution) as implemented in the ncf R package 213

(Paradis et al., 2004). Genetic distances between individuals were calculated using Nei’s Da estimator 214

as implemented in POPULATION v1.2.32 (Langella, 2002). Geographic distances were computed with 215

TESS (with the option Compute Great Circle Distances). A Mantel test was also performed to evaluate 216

(9)

8 the significance of the correlation between the genetic distances between pathogen individual pairs 217

and plant individual pairs. 218

To investigate the congruence of genetic structure between the host and the pathogen, we 219

performed partial Mantel tests (100,000 resamples for the null distribution) using the ncf R package 220

to assess the significance of the correlation of pairwise genetic distances between pathogen 221

individuals and plant individuals, as above, but removing the effect of the correlation between 222

genetic and geographic distances. 223

The genetic variation for both plants and pathogens was quantified using the allelic richness, Ar, 224

calculated with ADZE1.0 (Szpiech et al., 2008). Expected heterozygosity (HE), observed heterozygosity 225

(HO) and FIS were computed using GENEPOP 4.3 (Rousset, 2008). The statistical analyses on allelic 226

richness and HE (correlations and Wilcoxon rank sign tests) were performed with JMP v9 (SAS 227

Institute) using the value of the statistics per marker.

228 229

Experimental inoculations 230

To generate host material for the inoculation experiment, seeds of S. latifolia were produced under

231

greenhouse conditions from plant originating from field-collected seeds. Each location was

232

represented by a pooled set of seeds from at least three crosses between male and female parental

233

plants. Because sufficient parental material was not available from all regions (e.g. the Southern

234

cluster), we could not perform a full design inoculation experiment. For each available seed-fungus

235

population pair, inoculations were performed using a local pathogen strain and a strain chosen at

236

random among available strains from a different cluster (Table S3). Inoculum suspensions were

237

prepared from field-collected teliospores of M. lychnidis-dioicae, suspended in water at a

238

concentration of 500 spores/mL. Teliospores germination was confirmed to be >90% by plating on

239

potato dextrose agar and examination after 24 hr incubation.

240

Inoculation procedures followed prior studies (Hood & Antonovics, 2000). Briefly, seeds were

241

germinated in Petri dishes on 0.8% agar medium containing Murashige and Skoog (MS) basal salt.

242

After ten days of growth the apical meristems were exposed by expansion of the cotyledons, and 4

243

uL of inoculum suspension was applied. Inoculated plants were maintained at 15°C for three days,

244

and then transplanted to soil and maintained under greenhouse conditions until flowering. Infection

245

status was assessed by the presence of M. lychnidis-dioicae spores in the flowers.

246

For each source of seeds, experimental treatments included inoculation with M. lychnidis-dioicae

247

collected from the same location as the source of seeds or with M. lychnidis-dioicae from one of the

(10)

9

other identified clusters (Table S3). Two hundred seeds per location were used, with one hundred

249

seedling each randomly assigned to one of the two treatments (local versus inter-cluster inoculum

250

type). Positions of the plants in the Petri dishes and the greenhouse were randomized. A logistic

251

regression was used to test for differences in disease rates between treatments, performed with JMP

252 v9 (SAS Institute). 253 254

Results

255

Population structure and isolation by distance in the host plant Silene latifolia 256

The population structure of the host plant S. latifolia (n=578) was analyzed using the TESS Bayesian 257

clustering method. The Deviance Information Criterion (DIC), a statistical measure of the model 258

prediction abilities for TESS, sharply dropped at K=3 and decreased much more slowly with further 259

increase in K (Fig. S2). This indicates that the 3-clusters structure is the strongest genetic structure. In 260

fact, the barplots only showed three clusters even at higher K values (Fig. S3). Also consistent with 261

this finding, principal component analysis (PCA) and discriminant analyses of principal components 262

(DAPC) clearly separated the three same genetic clusters (Figs. 1a and S4), and no further clear and 263

geographically relevant cluster could be identified at higher K using DAPC or looking at further axes in 264

the PCA. A first cluster was found in France, the UK, Belgium and Germany (hereafter called the 265

“Western” cluster, n=319), the second cluster in Italy (hereafter called the “Southern” cluster, n=91) 266

and the third one in Hungary, Ukraine, Russia and the Czech Republic (hereafter called the “Eastern” 267

cluster, n=132) (Figs. 1b and 1c). We found 36 admixed individuals that could not be assigned to any 268

cluster, i.e. genotypes whose highest membership coefficient was < 0.8. Half of these admixed 269

individuals came from locations in contact areas among clusters. 270

The correlation between the matrices of genetic and geographic distances between pairs of S. 271

latifolia individuals was significant, at the European scale (r=0.3351, p<0.0001), as well as within each 272

cluster (Mantel tests; Eastern: r=0.1486, p<0.0001; Southern: r=0.2091, p<0.0001; Western r=0.1151, 273

p<0.0001), indicating IBD patterns at both spatial scales. 274

Allelic richness (Ar), expected heterozygosity (HE), observed heterozygosity (HO) and inbreeding index 275

FIS were computed for the tree genetic clusters (Table S4). The Eastern cluster appeared the most 276

genetically diverse but no significant differences were detected between clusters. FIS values were 277

relatively high for a dioecious plant (FIS >0.16). 278

(11)

10 Population structure and isolation by distance in the fungal pathogen Microbotryum lychnidis-280

dioicae

281

Genetic structure patterns for M. lychnidis-dioicae (n=679) was also investigated using TESS, and 282

gave results very close to the clustering results found in the host plant. Three clusters were 283

differentiated at K=3: one in Western (n=308), one in Southern (mainly Italy, n=160) and one in 284

Eastern Europe (n=178) (Figs. 2a and 2b). Only 33 individuals could not be assigned to any cluster, 285

half of which coming from contact areas between two clusters. The Deviance Information Criterion 286

(DIC) decreased steadily with increasing number of K, with no sharp drop (Fig. S2b). The PCA however 287

showed that the structure at K=3 was the strongest, with three distinct clusters corresponding to 288

those obtained with the TESS software and a separation of the Western cluster by the first axis, then 289

a subsequent split into Southern and Eastern clusters along the second axis (Fig. 2c). 290

The correlations between the matrices of genetic and geographic distances between pairs of M. 291

lychnidis-dioicae were significant at the European scale as well as within clusters (Mantel tests; 292

European scale, r=0. 3937, p=0.0002; Eastern cluster, r=0. 3459, p=0.0002; Southern cluster, r=0. 293

3905, p=0.0002; Western cluster, r=0. 2347, p=0.0002). This indicated IBD patterns in M. lychnidis-294

dioicae, both within clusters and at the European scale. 295

The allelic richness and the expected heterozygosity in M. lychnidis-dioicae were much lower than in 296

the plant (Table S5), and no significant differences were detected between clusters. FIS values were 297

very high, as expected for this highly selfing fungus. 298

In agreement with the steady decline of the DIC value from TESS, we found a finer population 299

subdivision, with well-defined clusters at K values higher than 3. Seven geographically circumscribed 300

clusters were thus identified by both model-based and non model-based methods (Figs. S5 and S6). 301

302

Costructure between the host Silene latifolia and pathogen Microbotryum lychnidis-dioicae 303

At K=3, the host and pathogen population structures appeared highly similar (Fig. 3a and b). In 304

particular, the limits of the M. lychnidis-dioicae clusters appeared extremely congruent with those 305

found in S. latifolia. Of the 80 sampling localities where at least two joint samples of both pathogen 306

and host were available, only five were assigned to different clusters between the host and the 307

pathogen. 308

Microsatellite data were available for 391 associated pairs of both hosts and pathogens, i.e., the 309

genotypes from the fungus and its host plant individual. Directly associated host and pathogen 310

samples, collected on the same plant individual, were predominantly assigned to the same clusters 311

(12)

11 for the host and the pathogen (i.e., either Western, Eastern or Southern cluster). In only 8 (2%) host 312

and pathogen pairs, the plant and fungal individuals were assigned to different clusters. In three host 313

and pathogen pairs, the plant and fungal individuals were both admixed (i.e., 8-9% of the admixed 314

plant or fungal individuals), all of these being located at the suture zones between the West and 315

South clusters. 316

The correlations between genetic distance matrices for the host and the pathogen were significant at 317

the European scale as well as within clusters (Mantel tests; European scale, r=0.3505, p<0.0001, Fig. 318

4; Eastern, r=0.1389, p<0.0001; Southern, r=0.1518, p=0.0003; Western, r=0.0837, p=0.0081). 319

A partial Mantel test was then performed for testing the congruence between the genetic structures 320

of the host and the pathogen while controlling for the effect of isolation by geographic distance, 321

using the two matrices of genetic distances (host and pathogen) and a matrix of geographic distance 322

between pairs of samples. The partial Mantel test at the European scale was again significant 323

(r=0.2277; p<0.0001), indicating that the genetically closer two S. latifolia individuals were, the 324

genetically closer their anther-smut pathogens were, independently of the shared pattern of 325

isolation by distance. With a threshold of 0.8 for assignation of individuals to a genetic cluster, the 326

partial Mantel test was significant for the Eastern cluster (r=0.0840, p<0.0001) and close to 327

significance for the two other clusters (Western: r=0.0450; p=0.0903; Southern: r=0.0710, p=0.0538). 328

When lowering the threshold of assignation to a cluster to 0.5 (thus increasing the number of sample 329

pairs and adding admixed individuals of plants and pathogens), all the partial Mantel tests were 330

significant (Europe: r=0.2277; p<0.0001; Western: r=0.0665; p=0.0229; Eastern: r=0.0889, p<0.0001; 331 Southern: r=0.0987, p=0.0089). 332 333 Inoculation experiments 334

In order to test the existence of local adaptation in the plant and/or the pathogen at the cluster level, 335

seedlings from one location in the western cluster and three locations in the eastern cluster were 336

inoculated with spores coming either from their local population (hereafter, “local treatment”) or 337

from a population belonging to a different cluster, randomly assigned (hereafter, “inter-cluster 338

treatment”). The probability of a plant to be diseased varied significantly with the treatment (i.e., 339

local vs. inter-cluster, Fig. 5); while the seedling origin was not significant as a main effect, the 340

interaction between the seedling origin and the treatment was significant (Table S6). The disease 341

rates were however never lower in inoculations by pathogens from a different cluster than by local 342

pathogens (Fig. 5). This altogether means that the plant better resists to their local anther-smut fungi 343

(13)

12 than to one from other clusters, while not significantly so in all populations. The results thus indicate 344

that local adaptation by the host plants often occurs. 345

346

Discussion

347

Footprints of glacial refugia in the population structure of the host Silene latifolia 348

A sampling of an unprecedented density and spatial scale for S. latifolia enabled the investigation of 349

genetic structure across Europe for this important model in ecology and evolution (Bernasconi et al., 350

2009). The strongest pattern of microsatellite variation corresponded to three genetic clusters: a 351

Western cluster, a Southern cluster in Italy, and an Eastern cluster. The pattern obtained here using 352

microsatellite markers appears much more clear-cut and revealed further geographical structure 353

(i.e., the Italian clusters) than the one previously obtained with chloroplast markers (Taylor & Keller, 354

2007), which showed large geographic overlaps between lineages. This discrepancy reinforces the 355

idea of a selective sweep having occurred throughout several clusters in the chloroplast DNA 356

(Ironside & Filatov, 2005), while the nuclear gene pools from different southern glacial refugia 357

remained differentiated after recolonization of northern Europe. Population structure of S. latifolia 358

has previously been investigated using microsatellite markers across Europe, but only the East/West 359

structure had been highlighted (Keller et al., 2012). This previous study nevertheless suggested that a 360

finer genetic subdivision was present. In fact, re-analysing this independent dataset, that includes 361

different markers and samples from ours, also reveal the Italian cluster at K>2 (not shown). 362

The biogeographic pattern we found in S. latifolia is typical of the genetic structure originating from 363

contracted range in the south of Europe during the last glacial maximum followed by recolonization 364

northward after climate warming at the beginning of the Holocene. The structure thus indicates that, 365

during the last glaciation, S. latifolia likely survived in three main glacial refugia (the Iberian, the 366

Italian and the Balkan refugia), and then recolonized Northern Europe without much large-scale 367

eastern or western dispersal that would have admixed the different clusters. A similar pattern has 368

been found in numerous animal and plant species, these three regions being the main European 369

lands where climatic conditions stayed mild enough for temperate species during the glacial phases 370

(Hewitt, 1999; Hewitt, 2004). 371

The identification of a new differentiated gene pool in S. latifolia will be useful for future studies on 372

the different topics for which this plant is a model, from the study of sex chromosomes and mating 373

systems, to speciation and biotic interactions (Marais et al., 2008; Bernasconi et al., 2009; Blavet et 374

al., 2011; Muyle et al., 2012). The knowledge of diversity and population subdivision is essential for 375

(14)

13 such evolutionary and ecological studies; for example differentiated clusters may display different 376

degrees of reproductive isolation with closely related species or different degrees of degeneration or 377

suppressed recombination on sex chromosomes. In addition, the invasion of S. latifolia in the US has 378

been studied with only knowledge of two large genetic clusters in Europe (keller & Taylor, 2008; 379

Keller et al., 2009; Keller & Taylor, 2010; Keller et al., 2012), and it will be highly informative to 380

include the newly identified Italian cluster in future studies investigating invasion dynamics and the 381

evolutionary and ecological processes underlying invasions. 382

383

Costructure between the host Silene latifolia and pathogen Microbotryum lychnidis-dioicae 384

In M. lychnidis-dioicae, the strongest genetic structure divided the samples into three genetic 385

clusters, as previously reported using the same microsatellite dataset as well as using nuclear gene 386

sequences (Vercken et al., 2010; Gladieux et al., 2011; Gladieux et al., 2013). These three genetic 387

clusters closely matched those observed in S. latifolia, with the same Southern, Western and Eastern 388

clusters. This indicates that the anther-smut pathogen has remained during the last glaciation in all 389

the three refugia of its host, i.e., in the Iberian, Italian and Balkan Peninsula, leaving no major 390

disease-free area. 391

Our extensive sampling of host-pathogen pairs from the same individual plants allowed comparing 392

the population structures of the two interacting species at a very fine level across Europe. The 393

population structures were strikingly congruent between the pathogen and its host. It is very striking 394

that the population structures of both S. latifolia and its pathogen indicates the rarity of recent 395

effective long distance dispersal events, while both are known to be able to disperse across large 396

distances. For instance, both have invaded large ranges of North America within the last centuries 397

(Keller et al., 2012; Fontaine et al., 2013; Gladieux et al., 2015). In Europe, recolonization after 398

glaciation may also have occurred through long distance dispersal events (Bialozyt et al., 2006). Yet 399

the three main clusters identified in Europe in the plant and the pathogen were clearly 400

geographically circumscribed, suggesting that they remained isolated one from each other for 401

thousands of generations. 402

One of the factors possibly explaining these discrepancies is local adaptation to the host. A previous 403

study at smaller geographical scale showed that the plant was locally adapted to its anther-smut 404

fungus, i.e., that S. latifolia individuals were more resistant to their local M. lychnidis-dioicae strains 405

than to geographically more distant strains (Kaltz et al., 1999; Kaltz & Shykoff, 2002). The results 406

presented here suggest that this pattern of local adaptation also holds at the cluster level. Lower 407

resistance to foreign pathogens may cause a form of immigrant sterility by preventing S. latifolia 408

genotypes to establish in regions where local M. lychnidis-dioicae populations belong to a cluster that 409

(15)

14 is different from the cluster present in their region of origin, as the immigrant plant genotypes would 410

be more often sterilized by local pathogens. The pattern of local adaptation supports suggestions of 411

ongoing cycles of coevolution, with the plant running faster in the arms race thanks to its higher 412

levels of inter-population migration levels (Kaltz & Shykoff, 2002). 413

Local adaptation of S. latifolia to their anther-smut pathogens could in contrast favor long-distance 414

establishment of the fungus, as our results suggest that immunity against foreign pathogens is lower 415

that immunity against local ones. However, another factor may limit pathogen migration and 416

contribute to the relative isolation among geographically-circumscribed genetic clusters of the 417

anther-smut fungus. A competitive exclusion of distantly related genotypes by resident genotypes 418

has previously been shown to occur in M. lychnidis-dioicae (Hood, 2003; Lopez-Villavicencio et al., 419

2007; López-Villavicencio et al., 2011; Buono et al., 2014). This relatedness-dependent competitive 420

exclusion may prevent long-distance migrants to establish in genetically different clusters. Moreover, 421

for both the host and the pathogen, local adaptation to climatic conditions may also play a role in the 422

low effective dispersal across Europe. If the different genetic clusters are locally adapted to abiotic 423

conditions, migrants may be prevented to establish by being poorer competitors in abiotic conditions 424

to which they are ill-suited. 425

The significant correlations between genetic distance matrices for the host and the pathogen further 426

revealed that their population structures were congruent even at the within-cluster scale. Such 427

highly congruent population structures are also likely due, in part, to the anther-smut fungus being 428

an obligate pathogen, i.e., being unable to grow outside its host plant. In addition, the pathogen 429

spores are dispersed by the same insect vectors serving as pollinators for the plant (Jennersten, 430

1988), increasing the probability that the host plant and its anther-smut fungus followed the same 431

recolonization pathways at fine scales. It would be interesting in future studies to also assess the 432

genetic costructure between the plants the fungus and the pollinators. In fact, a previous work at a 433

smaller geographical scale found a correlation between among-population genetic distances of S. 434

latifolia and Hadena biscuris, a pollinator and seed-eater of the plant (Magalhaes et al., 2011). 435

The congruence of population structures between S. latifolia and its M. lychnidis-dioicae pathogen 436

appeared even stronger than what could be expected by geography and dispersal alone. Indeed, the 437

partial Mantel tests revealed that the correlations between genetic distances of pairs of individuals 438

between the host and the pathogen remained significant or close to significant when controlling for 439

isolation by distance effect. This suggests either that similar dispersal pathways have been taken by 440

the host and the pathogen, that were not linear with geographical distance, and/or that coevolution 441

has played a role in the congruence of the population structures. Indeed the anther-smut fungus may 442

(16)

15 more frequently disperse to host populations that carry similar host genotypes as its population-of-443

origin if it is fitter on these particular genotypes. As explained above, genotype-by-genotype effects 444

on disease probability have actually been reported in experiments in the S. latifolia-M. lychnidis-445

dioicae system, showing mostly local adaptation in the host more than in the pathogen (Kaltz et al., 446

1999; Kaltz & Shykoff, 2002). 447

448

Stronger genetic subdivision in the pathogen Microbotryum lychnidis-dioicae than in the host 449

Silene latifolia

450

The strongest genetic structure in M. lychnidis-dioicae was the three-cluster structure across Europe, 451

with the Eastern, Northern and Southern clusters. However, TESS detected further subdivisions in the 452

fungus. As previously suggested (Vercken et al., 2010), these subdivisions may be a clue to a more 453

complex history than the classic model of the Iberian, Italian and Balkan refugia. The split in the 454

Eastern cluster is consistent with recolonization from two distinct refugia, one in the Balkans and the 455

other further East. The lack of such subdivision in the host S. latifolia, despite more markers being 456

used than in the fungus, and despite the markers being more polymorphic, might suggests that the 457

host plant distribution was larger than that of its anther-smut pathogen during the last glaciation, 458

spanning across both the Balkans and further East, while the distribution of M. lychnidis-dioicae 459

would have been more fragmented. This would be consistent with contemporaneous observations 460

that large areas colonized by S. latifolia are disease-free (Giraud T, pers. obs.). 461

An alternative, although not exclusive, interpretation is that several life history traits of both host 462

and pathogen may contribute to a higher genetic subdivision in the fungus than in the plant, even if 463

their distribution ranges have remained quite similar in the southern refugia. First, S. latifolia gene 464

flow is mediated both by dispersal of its seed and pollination, while M. lychnidis-dioicae gene flow 465

occurs only via spore transport by pollinators as seeds do not carry the fungus (Baker, 1947). In 466

addition, the pollinators discriminate against the infected flowers (Shykoff & Bucheli, 1995) which 467

may further reduce the extent of pathogen gene flow compared to its host. Finally, the mating 468

systems are important life history traits that may explain lower levels of gene flow among 469

populations in the fungus than in the plant. Although FIS values obtained here and previously suggest 470

that some inbreeding occurs in this dioecious plant (Delmotte et al., 1999; Magalhaes et al., 2011; 471

Keller et al., 2012) despite dioecy and inbreeding depression (Teixeira et al., 2009), the fungus has a 472

much more closed mating system, being highly selfing (Delmotte et al., 1999; Hood & Antonovics, 473

2000; Giraud, 2004; Giraud et al., 2008). Therefore, it may well be that the stronger subdivision in the 474

anther-smut fungus reflects fragmented ranges in southern refugia and/or more northern refugia, 475

(17)

16 but that the higher level of gene flow in the plant has erased their footprints by homogenizing allele 476 frequencies. 477 478 Conclusion 479

In fungal plant pathogens, studies of genetic structures at continental scales have been almost 480

exclusively performed for crop diseases, where the host population is often genetically 481

homogeneous across the landscape and pathogen dispersal is impacted heavily by anthropogenic 482

introductions and regional transport (Enjalbert et al., 2005 ; Barres et al., 2008; Fournier & Giraud, 483

2008; Gladieux et al., 2008; Saleh et al., 2014). The pathogen population structures in these cases 484

were either very weak at a continental scale (Barres et al., 2008; Fournier & Giraud, 2008; Saleh et 485

al., 2014) or reflected temperature adaptation (Enjalbert et al., 2005 ; Mboup et al., 2012). The 486

present study therefore brings novel insights into the costructure and coevolutionary patterns of 487

plants and their pathogens, in a natural system, revealing strikingly strong congruent subdivisions. 488

This study is one of the first to be conducted at this spatial scale with such comprehensive and paired 489

sampling, associated with inoculation experiments, and thus provides a benchmark to develop future 490

studies. Studies on a variety of other systems are needed to assess whether congruence population 491

structure between host and pathogen is a general phenomenon. The knowledge of the costructure is 492

essential for understanding process of co-local adaptation between hosts and pathogens (Gandon et 493 al., 1996). 494 495

Acknowledgments

496

We thank the Plateforme de Génotypage GENTYANE INRA UMR1095 and William Fontanaud for 497

assistance with genotyping. We thank all those who helped with sampling seeds, plant material and 498

pathogens (Vercken et al., 2010; Gladieux et al., 2015). This work was supported by the European 499

Research Council (starting grant GenomeFun 309403 to TG), the Agence Nationale de la Recherche 500

(grants GANDALF ANR-12-ADAP-0009 and Emerfundis 07-BDIV-003), and the National Science 501

Foundation (grant DEB-1115765 to MEH). 502

503

Author contribution

504

T. G. and M.E.H. designed and supervised the study; A.F. performed the population genetics

505

analyses with the help of T.G., P.G. and A.C.; M.E.H. and L.R. performed the inoculation

(18)

17

experiments; A.F. and A.S. genotyped the plants; T.G. and M.E.H. contributed to the funding

507

of the study.

508 509 510

References

511

Amico GC, Nickrent DL. 2009. Population structure and phylogeography of the mistletoes Tristerix 512

corymbosus and T. aphyllus (Loranthaceae) using chloroplast DNA sequence variation. 513

American Journal of Botany 96(8): 1571-1580. 514

Badouin H, Hood M, Gouzy J, Aguileta G, Siguenza S, Perlin M, Cuomo C, Fairhead C, Branca A, 515

Giraud T. 2015. Chaos of rearrangements in in the mating-type chromosomes of the anther-516

smut fungus Microbotryum lychnidis-dioicae. Genetics 200: 1275-1284. 517

Baker HG. 1947. Infection of species of Melandrium by Ustilago violacea (Pers.) Fuckel and the 518

transmission of the resultant disease. Annals of Botany 11(43): 333-348. 519

Barluenga M, Austerlitz F, Elzinga J, Teixeira S, Goudet J, Bernasconi G. 2011. Fine-scale spatial 520

genetic structure and gene dispersal in Silene latifolia. Heredity 106: 13-24. 521

Barres B, Halkett F, Dutech C, Andrieux A, Pinon J, Frey P. 2008. Genetic structure of the poplar rust 522

fungus Melampsora larici-populina: Evidence for isolation by distance in Europe and recent 523

founder effects overseas. Infection Genetics and Evolution 8(5): 577-587. 524

Barrett LG, Thrall PH, Burdon JJ, Linde CC. 2008. Life history determines genetic structure and 525

evolutionary potential of host-parasite interactions. Trends in Ecology & Evolution 23(12): 526

678-685. 527

Bernasconi G, Antonovics J, Biere A, Charlesworth D, Delph LF, Filatov D, Giraud T, Hood ME, 528

Marais GAB, McCauley D, et al. 2009. Silene as a model system in ecology and evolution. 529

Heredity 103(1): 5-14. 530

Bialozyt R, Ziegenhagen B, Petit RJ. 2006. Contrasting effects of long distance seed dispersal on 531

genetic diversity during range expansion. Journal of Evolutionary Biology 19(1): 12-20. 532

Blavet N, Charif D, Oger-Desfeux C, Marais G, Widmer A. 2011. Comparative high-throughput 533

transcriptome sequencing and development of SiESTa, the Silene EST annotation database. 534

BMC Genomics 12(1): 376. 535

Buono L, López-Villavicencio M, Shykoff J, Snirc A, Giraud T. 2014. Influence of multiple infection 536

and relatedness on virulence: disease dynamics in an experimental plant population and its 537

castrating parasite. PLoS One 9: e98526. 538

Criscione CD, Poulin R, Blouin MS. 2005. Molecular ecology of parasites: elucidating ecological and 539

microevolutionary processes. Mol Ecol 14: 2247-2258. 540

de Vienne DM, Refregier G, Lopez-Villavicencio M, Tellier A, Hood ME, Giraud T. 2013. Cospeciation 541

vs host-shift speciation: methods for testing, evidence from natural associations and relation 542

to coevolution. New Phytologist 198(2): 347-385. 543

Delmotte F, Bucheli E, Shykoff JA. 1999. Host and parasite population structure in a natural plant-544

pathogen system. Heredity 82: 300-308. 545

Dybdahl MF, Lively CM. 1996. The geography of coevolution: Comparative population structures for 546

a snail and its trematode parasite. Evolution 50(6): 2264-2275. 547

Enjalbert J, Duan X, Leconte M, Hovmoller M, De Vallavieille-Pope C. 2005 Genetic evidence of local 548

adaptation of wheat yellow rust (Puccinia striiformis f. sp tritici) within France Mol Ecol 14 549

2065-2073 550

Falush D, Stephens M, Pritchard‡ JK. 2003. Inference of Population Structure Using Multilocus 551

Genotype Data: Linked Loci and Correlated Allele Frequencies. Genetics 164: 1567-1587 552

(19)

18 Fontaine MC, Gladieux P, Hood ME, Giraud T. 2013. History of the invasion of the anther smut 553

pathogen on Silene latifolia in North America. New Phytologist 198(3): 946-956. 554

Fournier E, Giraud T. 2008. Sympatric genetic differentiation of populations of a generalist 555

pathogenic fungus, Botrytis cinerea, on two different host plants, grapevine and bramble. J 556

Evol Biol 21: 122-132. 557

Gandon S, Capowiez Y, Dubois Y, Michalakis Y, Olivieri I. 1996. Local adaptation and gene-for-gene 558

coevolution in a metapopulation model. Proceedings of the Royal Society of London Series B-559

Biological Sciences 263(1373): 1003-1009. 560

Giraud T. 2004. Patterns of within population dispersion and mating of the fungus Microbotryum 561

violaceum parasitising the plant Silene latifolia. Heredity 93: 559-565. 562

Giraud T, Jonot O, Shykoff JA. 2005. Selfing propensity under choice conditions in a parasitic fungus, 563

Microbotryum violaceum, and parameters influencing infection success in artificial 564

inoculations Int. J. Plant Sci. 166: 649-657. 565

Giraud T, Yockteng R, Lopez-Villavicencio M, Refregier G, Hood ME. 2008. The mating system of the 566

anther smut fungus, Microbotryum violaceum: selfing under heterothallism. Euk. Cell 7: 765-567

775. 568

Gladieux P, Devier B, Aguileta G, Cruaud C, Giraud T. 2013. Purifying selection after episodes of 569

recurrent adaptive diversification in fungal pathogens. Infection Genetics and Evolution 17: 570

123-131. 571

Gladieux P, Feurtey A, Hood M, Snirc A, Clavel J, Dutech C, Roy M, Giraud T. 2015. The population 572

biology of fungal invasions. Mol Ecol. 24:1969-86. 573

Gladieux P, Vercken E, Fontaine M, Hood M, Jonot O, Couloux A, Giraud T. 2011. Maintenance of 574

fungal pathogen species that are specialized to different hosts: allopatric divergence and 575

introgression through secondary contact. Mol. Biol. Evol. 28: 459-471. 576

Gladieux P, Zhang X-G, Afoufa-Bastien D, Sanhueza R-MV, Sbaghi M, Le Cam B. 2008. On the origin 577

and spread of the scab disease of apple: out of Central Asia. Plos One 3:e1455. 578

Guillot G. 2009. On the inference of spatial structure from population genetics data using the Tess 579

program. BIOINFORMATICS 25:1796-801. 580

Hafner MS, Page RDM. 1995. Molecular phylogenies and host-parasite cospeciation: gophers and 581

lice as a model system. Philos Trans R Soc London 349: 77-83. 582

Hewitt GM. 1999. Post-glacial re-colonization of European biota. Biological Journal of the Linnean 583

Society 68: 87-112. 584

Hewitt GM. 2004. Genetic consequences of climatic oscillations in the Quaternary. Philosophical 585

Transactions of the Royal Society of London Series B-Biological Sciences 359(1442): 183-195. 586

Hood M, Mena-Alí J, Gibson A, Oxelman B, Giraud T, Yockteng R, Arroyo M, Conti F, Pedersen A, 587

Gladieux P, et al. 2010. Distribution of the anther-smut pathogen Microbotryum on species 588

of the Caryophyllaceae assessed from natural history collections. New Phytol 187: 217-229. 589

Hood ME. 2003. Dynamics of multiple infection and within-host competition by the anther-smut 590

pathogen. The American Naturalist 162(1): 122-133. 591

Hood ME, Antonovics J. 2000. Intratetrad mating, heterozygosity, and the maintenance of 592

deleterious alleles in Microbotryum violaceum (=Ustilago violacea). Heredity 85: 231-241. 593

Ironside JE, Filatov DA. 2005. Extreme population structure and high interspecific divergence of the 594

Silene Y chromosome. Genetics 171(2): 705-713. 595

Jakobsson M, Rosenberg N. 2007. CLUMPP: a cluster matching and permutation program for dealing 596

with label switching and multimodality in analysis of population structure. Bioinformatics 23: 597

1801-1806. 598

Jennersten O. 1988. Insect dispersal of fungal disease: effects of Ustilago infection on pollinator 599

attraction in Viscaria vulgaris. Oikos 51: 163-170. 600

Kaltz O, Gandon S, Michalakis Y, Shykoff J. 1999. Local maladaptation of the plant pathogen 601

Microbotryum violaceum to its host Silene latifolia : Evidence from a cross-inoculation 602

experiment. Evolution 53: 395-407. 603

(20)

19 Kaltz O, Shykoff J. 2002. Within- and among-population variation in infectivity, latency and spore 604

production in a host-pathogen system. JEB 15: 850-860. 605

Karrenberg S, Favre A. 2008. Genetic and ecological differentiation in the hybridizing campions Silene 606

dioica and S. latifolia. Evolution 62(4): 763-773. 607

keller S, Taylor D. 2008. History, chance and adaptation during biological invasion: separating 608

stochastic phenotypic evolution from response to selection. Ecology Letters 11: 852-866. 609

Keller SR, Gilbert KJ, Fields PD, Taylor DR. 2012. Bayesian inference of a complex invasion history 610

revealed by nuclear and chloroplast genetic diversity in the colonizing plant, Silene latifolia. 611

Molecular Ecology 21(19): 4721-4734. 612

Keller SR, Sowell DR, Neiman M, Wolfe LM, Taylor DR. 2009. Adaptation and colonization history 613

affect the evolution of clines in two introduced species. New Phytologist 183(3):678-90. 614

Keller SR, Taylor DR. 2010. Genomic admixture increases fitness during a biological invasion. Journal 615

of Evolutionary Biology 23(8): 1720-1731. 616

Langella O. 2002. POPULATIONS 1.2.28. Population genetic software (individuals or populations 617

distances, phylogenetic trees). Available from 618

http://bioinformatics.org/~tryphon/populations/. 619

López-Villavicencio M, Courjol F, Gibson A, Hood M, Jonot O, Shykoff J, Giraud T. 2011. 620

Competition, cooperation among kin and virulence in multiple infections. Evolution 65:1357-621

1366. 622

Lopez-Villavicencio M, Jonot O, Coantic A, Hood M, Enjalbert J, Giraud T. 2007. Multiple infections 623

by the anther smut pathogen are frequent and involve related strains. PloS Pathogens 3: 624

e176. 625

Magalhaes IS, Gleiser G, Labouche AM, Bernasconi G. 2011. Comparative population genetic 626

structure in a plant-pollinator/seed predator system. Molecular Ecology 20(22): 4618-4630. 627

Marais GAB, Nicolas M, Bergero R, Chambrier P, Kejnovsky E, Moneger F, Hobza R, Widmer A, 628

Charlesworth D. 2008. Evidence for degeneration of the Y chromosome in the dioecious 629

plant Silene latifolia. Current Biology 18(7): 545-549. 630

Mboup M, Bahri B, Leconte M, De Vallavieille-Pope C, Kaltz O, Enjalbert J. 2012. Genetic structure 631

and local adaptation of European wheat yellow rust populations: the role of temperature-632

specific adaptation. Evolutionary Applications 5(4): 341-352. 633

McCOY KD, BOULINIER T, TIRARD C. 2005. Comparative hostparasite population structures: 634

disentangling prospecting and dispersal in the black-legged kittiwake Rissa tridactyla. Mol 635

Ecol 14: 2825-2838. 636

Michalakis Y, Sheppard AW, Noel V, Olivieri I. 1993. POPULATION-STRUCTURE OF A HERBIVOROUS 637

INSECT AND ITS HOST-PLANT ON A MICROGEOGRAPHIC SCALE. Evolution 47(5): 1611-1616. 638

Miura O, Torchin ME, Kuris AM, Hechinger RF, Chiba S. 2006. Introduced cryptic species of parasites 639

exhibit different invasion pathways. Proceedings of the National Academy of Sciences of the 640

United States of America 103(52): 19818-19823. 641

Moccia MD, Oger-Desfeux C, Marais GAB, Widmer A. 2009. A White Campion (Silene latifolia) floral 642

expressed sequence tag (EST) library: annotation, EST-SSR characterization, transferability, 643

and utility for comparative mapping. Bmc Genomics 10: 243. 644

Muir G, Filatov D. 2007. A Selective Sweep in the Chloroplast DNA of Dioecious Silene (Section 645

Elisanthe). Genetics 177(2): 1239-1247. 646

Muyle A, Zemp N, Deschamps C, Mousset S, Widmer A, Marais GAB. 2012. Rapid de novo evolution 647

of X chromosome dosage compensation in Silene latifolia, a plant with young sex 648

chromosomes. Plos Biology 10(4): e1001308. 649

Nakao M, Xiao N, Okamoto M, Yanagida T, Sako Y, Ito A. 2009. Geographic pattern of genetic 650

variation in the fox tapeworm Echinococcus multilocularis. Parasitology International 58(4): 651

384-389. 652

Nieberding CM, Durette-Desset MC, Vanderpoorten A, Casanova JC, Ribas A, Deffontaine V, Feliu C, 653

Morand S, Libois R, Michaux JR. 2008. Geography and host biogeography matter for 654

(21)

20 understanding the phylogeography of a parasite. Molecular Phylogenetics and Evolution 655

47(2): 538-554. 656

Nieberding CM, Olivieri I. 2007. Parasites: proxies for host genealogy and ecology? Trends in Ecology 657

& Evolution 22(3): 156-165. 658

Oudemans PV, Alexander HM, Antonovics J, Altizer S, Thrall PH, Rose L. 1998. The distribution of 659

mating-type bias in natural populations of the anther-smut Ustilago violacea on Silene alba in 660

Virginia. Mycologia 90: 372-381. 661

Paradis E, Claude J, Strimmer K. 2004. APE: Analyses of Phylogenetics and Evolution in R language. 662

Bioinformatics 20(2): 289-290. 663

Rich KA, Thompson JN, Fernandez CC. 2008. Diverse historical processes shape deep 664

phylogeographical divergence in the pollinating seed parasite Greya politella. Molecular 665

Ecology 17(10): 2430-2448. 666

Rousset F. 2008. Genepop'007: a complete reimplementation of the Genepop software for Windows 667

and Linux. Mol. Ecol. Resources 8: 103-106. 668

Saleh D, Milazzo J, Adreit H, Fournier E, Tharreau D. 2014. South-East Asia is the center of origin, 669

diversity and dispersion of the rice blast fungus, Magnaporthe oryzae. New Phytologist 670

201(4): 1440-1456. 671

Shykoff JA, Bucheli E. 1995. Pollinator visitation patterns, floral rewards and the probability of 672

transmission of Microbotryum violaceum, a veneral disease plant. Journal of Ecology 83: 189-673

198. 674

Szpiech Z, Jakobsson M, Rosenberg N. 2008. ADZE: a rarefaction approach for counting alleles 675

private to combinations of populations. Bioinformatics 24: 2498-2504. 676

Taylor DR, Keller SR. 2007. Historical range expansion determines the phylogenetic diversity 677

introduced during contemporary species invasion. Evolution 61(2): 334-345. 678

Teixeira S, Foerster K, Bernasconi G. 2009. Evidence for inbreeding depression and post-pollination 679

selection against inbreeding in the dioecious plant Silene latifolia. Heredity 102(2): 101-112. 680

Thomas A, Shykoff J, Jonot O, Giraud T. 2003. Mating-type ratio bias in populations of the 681

phytopathogenic fungus Microbotryum violaceum from several host species. Int. J. Plant Sci. 682

164: 641-647. 683

Thrall PH, Biere A, Uyenoyama MK. 1995. Frequency-dependant disease transmission and the 684

dynamics of the Silene-Ustilago host-pathogen system. Am. Nat. 145(1): 43-62. 685

Thrall PH, Jarosz AM. 1994. Host-pathogen dynamics in experimental populations of Silene alba and 686

Ustilago violacea. I. Ecological and genetic determinants of disease spread. Journal of Ecology 687

82: 549-559. 688

Tollenaere C, Susi H, Nokso-Koivisto J, Koskinen P, Tack A, Auvinen P, Paulin L, Frilander MJ, 689

Lehtonen R, Laine AL. 2012. SNP design from 454 sequencing of Podosphaera plantaginis 690

transcriptome reveals a genetically diverse pathogen metapopulation with high levels of 691

mixed-genotype infection. Plos One 7(12): e52492. 692

Tsai Y-HE, Manos PS. 2010. Host density drives the postglacial migration of the tree parasite, 693

Epifagus virginiana. Proceedings of the National Academy of Sciences of the United States of 694

America 107(39): 17035-17040. 695

Vercken E, Fontaine M, Gladieux P, Hood M, Jonot O, Giraud T. 2010. Glacial refugia in pathogens: 696

European genetic structure of anther smut pathogens on Silene latifolia and S. dioica. PloS 697

Pathogens 6: e1001229. 698

Whiteman NK, Parker PG. 2005. Using parasites to infer host population history: a new rationale for 699

parasite conservation. Animal Conservation 8: 175-181. 700

Williams PD. 2010. Darwinian interventions: taming pathogens through evolutionary ecology. Trends 701

in Parasitology 26(2): 83-92. 702

Wilson DJ, Falush D, McVean G. 2005. Germs, genomes and genealogies. Trends in Ecology & 703

Evolution 20(1): 39-45. 704

Wirth T, Meyer A, Achtman M. 2005. Deciphering host migrations and origins by means of their 705

microbes. Molecular Ecology 14(11): 3289-3306. 706

(22)

21

Supporting Information

707

Figure S1. Identification of species. 708

Figure S2. Statistics indicating the number of clusters K corresponding to the strongest genetic 709

structure. 710

Figure S3. Membership proportions per Silene latifolia individual (as vertical bars) from the TESS 711

results, from K=2 to K=6. 712

Figure S4. Discriminant analyses of principal component (DAPC) on the Silene latifolia dataset. 713

Figure S5. Membership proportions per Microbotryum lychnidis-dioicae individual (as vertical bars) 714

inferred with TESS from K=2 to K=6. 715

Figure S6. Discriminant analyses of principal component (DAPC) on the Microbotryum lychnidis-716

dioicae dataset. 717

Table S1. Details of the multiplex PCR used to amplify microsatellite markers in Silene latifolia. 718

Table S2. Details of the PCR program used to amplify microsatellite markers in Silene latifolia 719

Table S3. Cross-inoculation design, origin of the Microbotryum lychnidis-dioicae strains and Silene 720

latifolia seeds. 721

Table S4. Summary statistics for Silene latifolia for each of the three TESS genetic clusters. 722

Table S5. Summary statistics for Microbotryum lychnidis-dioicae for each of the three TESS genetic 723

clusters. 724

Table S6. Logistic regression testing local adaptation at the cluster level in the system Microbotryum 725

lychnidis-dioicae - Silene latifolia. 726

Notes S1. Identification of species 727

728 729

(23)

22

Figure Legends

730

Figure 1. Genetic structure of the host plant, Silene latifolia, representing the three clusters 731

obtained at K=4 and in the majority of the K=3 solutions. (a) Principal component analysis on the 732

microsatellite alleles. The dot color corresponds to the three clusters assigned to each individual by 733

the Bayesiand computations. The gray dots are the individuals not assigned to any cluster when using 734

a threshold of 0.8. (b) Barplots of membership probabilities per individual (as vertical bars) from the 735

TESS analysis. (c) Map of mean cluster membership probabilities per locality inferred with TESS. The 736

pie diameter reflects the number of individuals sampled in the corresponding locality: the smallest 737

circles correspond to N ≤ 3 and the largest circles to N ≥ 16. 738

739 740

(24)

23 741

Figure 2. Genetic structure of the anther-smut fungus, Microbotryum lychnidis-dioicae, at K=3. (a) 742

Map of mean cluster membership proportions per locality inferred with TESS at K=3. The pie 743

diameter reflects the number of individuals sampled in the corresponding locality: the smallest 744

circles correspond to N ≤ 3 and the largest circles to N ≥ 16. (b) Barplots of membership proportions 745

per individual (as vertical bars) from the TESS analysis for K=3. (c). Principal component analysis on 746

the microsatellite alleles. The dot color corresponds to the cluster assigned for each individual by the 747

Bayesian computations. The gray dots are the individuals not assigned to any cluster when using a 748

threshold of 0.8. 749

750 751

Figure

Figure Legends
Figure 2.  Genetic structure of the anther-smut fungus, Microbotryum lychnidis-dioicae, at K=3
Figure 3. Genetic structure of the host plant, Silene latifolia, and its pathogen, Microbotryum 753
Figure 4. Genetic distances (Nei’s Da) between plant (Silene latifolia) pairs plotted against genetic 763
+2

Références

Documents relatifs

Here we aimed to benefit from a very peculiar eco- logical situation (the large and long lasting selection pressure exerted by R7 cultivated poplars on the rust fungus

We investigated multiple infections in natural populations of the systemic fungal plant parasite Microbotryum violaceum, the anther smut of Caryophyllaceae, on its host,

Indeed, the more prey they capture, the more they provide the host plant with nutrients that then has greater growth (more leaves equals more housing available for

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

Serp-2 is, however, an essential MV virulence factor, and the ma- jority of rabbits infected with the Serp-2 knockout virus only developed moderate clinical signs of myxomatosis

Ortholog conservation (right panel) highlights genes conserved in all species (core, green), genes conserved in at least two species (conserved, blue), genes unique to

Indeed, the more prey they capture, the more they provide the host plant with nutrients that then has greater growth (more leaves equals more housing available for guest ants)

The five tools were run on three different datasets, chosen for their complementarity and the possibility to validate the reconstructed candidate 16S rRNA sequences: a simulated