44
1 1 2 Dynamics and innovations within oomycete genomes: 3 insights into biology, pathology, and evolution 4 5 6 Howard S. Judelson 7 Department of Plant Pathology and Microbiology 8 University of California 9 Riverside, CA 92521 USA 10 11 Address correspondence to: [email protected], Tel: 951-827-4199, Fax: 951- 12 827-4294 13 14 Running title: Dynamics and innovations within oomycete genomes 15 16 17 18 Copyright © 2012, American Society for Microbiology. All Rights Reserved. Eukaryotic Cell doi:10.1128/EC.00155-12 EC Accepts, published online ahead of print on 24 August 2012 on August 21, 2018 by guest http://ec.asm.org/ Downloaded from

Dynamics and innovations within oomycete genomes: …ec.asm.org/content/early/2012/08/20/EC.00155-12.full.pdf · 51 sister genus Pythium where most members are saprophytic and

  • Upload
    ngodiep

  • View
    217

  • Download
    0

Embed Size (px)

Citation preview

1

1

2

Dynamics and innovations within oomycete genomes: 3

insights into biology, pathology, and evolution 4

5

6

Howard S. Judelson 7

Department of Plant Pathology and Microbiology 8

University of California 9

Riverside, CA 92521 USA 10

11

Address correspondence to: [email protected], Tel: 951-827-4199, Fax: 951-12

827-4294 13

14

Running title: Dynamics and innovations within oomycete genomes 15

16

17

18

Copyright © 2012, American Society for Microbiology. All Rights Reserved.Eukaryotic Cell doi:10.1128/EC.00155-12 EC Accepts, published online ahead of print on 24 August 2012

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

2

ABSTRACT 19

The eukaryotic microbes known as oomycetes are common inhabitants of terrestrial 20

and aquatic environments, and include saprophytes and pathogens. Lifestyles of the 21

pathogens extend from biotrophy to necrotrophy, obligate to facultative pathogenesis, 22

and narrow to broad host ranges on plants or animals. Sequencing of several 23

pathogens has revealed striking variation in genome size and content, a plastic set of 24

genes related to pathogenesis, and adaptations associated with obligate biotrophy. 25

Features of genome evolution include repeat-driven expansions, deletions, gene 26

fusions, and horizontal gene transfer, in a landscape organized into gene-dense and 27

gene-sparse sectors and influenced by transposable elements. Gene expression 28

profiles are also highly dynamic throughout oomycete life cycles, with transcriptional 29

polymorphisms as well as differences in protein sequence contributing to variation. The 30

genome projects have set the foundation for functional studies and should spur the 31

sequencing of additional species, including more diverse pathogens and nonpathogens. 32

33

INTRODUCTION 34

Oomycetes are known best for their plant pathogens, but also include saprophytes 35

and colonizers of insects, vertebrates, and microbes. Oomycetes outwardly resemble 36

fungi since both exhibit hyphal growth and heterotrophic absorptive nutrition, and reside 37

in similar ecological niches. Oomycetes were misclassified as Mycota until the last part 38

of the 20th century, but are now accepted to have a distinct evolutionary history and 39

belong to the Kingdom Stramenipila, which also includes brown algae and diatoms (7). 40

One theory joins stramenopiles with other protists into the Chromalveolata 41

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

3

superkingdom, linked by an ancestral endosymbiosis of a photosynthetic alga (44). 42

Of the hundreds of known oomycetes, those affecting agriculture and natural 43

ecosystems are the most studied. Most fall into two large orders, the Peronosporales 44

and Saprolegniales (Fig. 1). Most notorious is the potato and tomato late blight agent 45

Phytophthora infestans, a peronosporalean that triggered the Irish Famine in the 1840s 46

and still has major impact worldwide (22). Phytophthora contains over 100 species, all 47

plant pathogens (48). Many are relatively recent discoveries, such as the Sudden Oak 48

Death agent P. ramorum that has severely damaged woodlands in North America and 49

Europe (29). No Phytophthora is known to persist as a saprophyte, in contrast to its 50

sister genus Pythium where most members are saprophytic and opportunistic 51

phytopathogens. Mycoparasites and an animal pathogen also populate Pythium. 52

Saprolegnian genera, like those in the Peronosporales, also include saprophytes and 53

plant pathogens, but there is a greater tendency to an aquatic lifestyle. Most species 54

within Achlya and Saprolegnia, for example, are usually innocuous inhabitants of 55

freshwater ponds and streams, where they grow on dead vegetable matter. They 56

however can also parasitize a range of aquatic animals, particularly when immune 57

systems are stressed. Saprolegnia parasitica in particular has become a growing 58

problem in commercial aquaculture, where high fish density and imperfect water quality 59

increase stress and favor disease (88). 60

Much diversity in the nature of pathogenesis is displayed by different oomycetes. 61

Species such as P. infestans and the soybean pest Phytophthora sojae colonize only a 62

limited number of hosts. Other Phytophthora spp. are more cosmopolitan, such as P. 63

ramorum which infects over 60 trees and shrubs, and P. cinnamomi which has over 64

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

4

1000 hosts spanning both herbaceous and woody plants (19). Broad host range is also 65

typical of most species of Pythium, and Saprolegnia which infects both fish and 66

crustaceans. In contrast, with a few exceptions each downy mildew and white rust 67

colonizes only a single type of host. The white rust Albugo laibachii and the downy 68

mildew Hyaloperonospora arabidopsidis only infect Arabidopsis thaliana, for example. 69

Besides diversity in host range, pathogenic interactions involving oomycetes are also 70

varied, ranging from biotrophy to necrotrophy. Pythium is largely necrotrophic, 71

extracting resources from dying cells. At the other extreme, white rusts and downy 72

mildews feed on living cells and do not directly kill their hosts; one Albugo is even 73

reported as an asymptomatic endophyte (64). Phytophthora are hemibiotrophs, which 74

are biotrophic initially but shift to necrotrophy at the end of the disease cycle. 75

Genome sequences are now available for several oomycetes, and valuable for 76

resolving long-standing issues about their biology and evolution. Questions that can be 77

addressed include: What accounts for the diversity of oomycete lifestyles? How do the 78

genes that oomycetes use to infect hosts compare to those of pathogens in other taxa? 79

To what extent has horizontal gene transfer affected the pathogenic, metabolic, or 80

regulatory characteristics of oomycetes? Do oomycete genomes have features that 81

help oomycetes evolve to survive environmental changes, or defeat efforts to control 82

them? 83

84

Genome size and topography. Oomycete genomes are estimated to range in size 85

from a low of about 37 Mb, as in the cases of Pythium sylvaticum and A. laibachii, to a 86

high of 280 Mb, as in some Phytophthora spp. Repetitive DNA, mostly in the form of 87

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

5

transposable elements (TEs), is responsible for the bulk of variation (45, 55, 65). 88

Assemblies annotated with gene models are now available publicly for eight plant 89

pathogens, including five Phytophthora spp. (P. capsici, P. cinnamomi, P. infestans, P. 90

parasitica, P. ramorum, P. sojae), Pythium ultimum, H. arabidopsidis, and A. laibachii 91

(6, 31, 45, 50, 51, 86). Data are also released for one animal pathogen, S. parasitica, 92

but not for an exclusively saprophytic species. Predicted gene contents range from 93

about 13 to 26 thousand (Fig. 1). A striking feature is the wide variation in repetitive 94

DNA content, which ranges from 7% for Py. ultimum to 74% for P. infestans. Some 95

correlation exists between gene number and genome size in P. infestans, P. ramorum, 96

and P. sojae but the trend loosens when considering other species. The recently 97

released sequence of P. cinnamomi, for example, includes 26,131 predicted genes 98

within a 78 Mb genome, compared to 16,988 for the 95 Mb genome of P. sojae. Initial 99

gene counts are often overestimates, however it is intriguing to consider that the large 100

number of genes in P. cinnamomi relates to its ability to infect more than a thousand 101

plant species. 102

Genome-wide ortholog mapping suggests that a core set of about 8,000 to 9,500 103

genes is conserved between species (31, 75). The variable component thus represents 104

nearly half of some proteomes. Interspecific differences are largely attributable to family 105

expansion and/or gene loss, which to some extent relates to the novel partitioning of 106

oomycete genomes into gene-dense and gene-sparse regions. In P. infestans, for 107

example, about 80% of genes reside in clusters of tightly-spaced genes, with the rest 108

distant to each other (~2 or more kb). This is illustrated in Fig. 2B for a portion of the P. 109

infestans genome, which contains a gene-rich cluster of 24 genes with small intergenic 110

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

6

distances near a gene-sparse region with larger intergenic distances. Insight towards 111

the diversity of pathogenic lifestyles has come from studying the gene-sparse regions, 112

as these are enriched for loci encoding host-modulating "effectors" (31, 66). Many 113

genes in such regions have evolved faster than average in sequence and copy number, 114

possibly aided by the nearby repeated DNA that fosters illegitimate recombination and a 115

low gene density that minimizes lethal events. 116

Some clusters of repeated genes can nevertheless be found in gene-dense regions, 117

including some protein kinases and genes involved in sexual development (14, 34). P. 118

infestans encodes 354 proteins with classic eukaryotic protein kinase domains, with 119

other members of the genus having similar numbers. As with some other gene families, 120

unequal crossing-over appears to have contributed to its expansion as about 20% of 121

kinase loci reside in clusters of 2 to 13 genes. The same clusters tend to occur in other 122

members of the genus, indicating their presence in the last common ancestor of 123

Phytophthora, and a few are shared with H. arabidopsidis. The number of kinases in 124

different oomycetes varies dramatically, ranging from less than 100 in Py. ultimum to 125

about 500 in S. parasitica. 126

127

Transcriptional landscape. A typical oomycete life cycle involves vegetative 128

filamentous growth followed by sporulation, spore germination, and for pathogens the 129

formation of structures used to enter and feed from host tissues (37). Most species also 130

form both sexual and asexual spores. mRNA profiles exhibit dynamic changes 131

throughout these developmental transitions, with many genes expressed specifically in 132

spore or infection stages. In P. infestans, mRNA levels for nearly one-half of genes 133

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

7

change >2-fold during life-stage changes, with about 14% being expressed exclusively 134

in one stage or another (35). Studies in P. infestans, the lima bean pathogen 135

Phytophthora phaseoli, and the cucumber downy mildew agent Pseudoperonospora 136

cubensis also documented changes between growth on artificial media and a plant host 137

(31, 49). Although P. infestans and Ps. cubensis have different lifestyles 138

(hemibiotrophic and obligately biotrophic, respectively), most orthologs had similar 139

expression patterns suggesting that the non-orthologous genes may play roles in 140

determining the mode of pathogenesis (71). 141

Many differentially-expressed genes have functions that are logically required at only 142

some stages of the life cycle. Examples include structural components of asexual or 143

sexual spores, cell cycle proteins that may act to establish spore dormancy, and 144

pathogenicity factors. Others however have roles that are more challenging to discern, 145

such as new isoforms of metabolic enzymes. Many transcription factors also show 146

strong differential expression, especially those in the bZIP and Myb families which have 147

patterns suggestive of a transcription factor cascade (35, 71, 92). 148

Most oomycetes genes are tightly spaced, with median intergenic distances being 149

only 435 nt within the gene-dense regions of P. infestans. By comparison, median 150

intergenic regions in S. cerevisiae, Arabidopsis thaliana, and Homo sapiens are 0.45, 151

1.5 and 35 kb, respectively (13, 47, 94). It is remarkable that despite the close spacing 152

of genes in Phytophthora, neighboring genes typically lack similar patterns of 153

developmental regulation. This can be seen in Fig. 2A, where each color indicates up-154

regulation in a particular developmental stage with black representing constitutive 155

expression. Tandemly duplicated genes are frequently co-expressed, as illustrated in 156

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

8

Fig. 2B where the right border of the gene-rich interval contains a tandemly arrayed co-157

expressed cluster. In contrast, other adjacent genes are not usually up-regulated in the 158

same stage. The transcriptional independence of most neighboring genes in 159

Phytophthora contrasts with that of many eukaryotes, where genes are often organized 160

into co-expressed domains (40, 85, 94). The main mechanism for regulating 161

transcription during oomycete development may therefore not involve chromatin-level 162

effects, which in yeasts and metazoans can influence genes within a 5-50 kb region 163

(23). 164

One effect of high gene density in oomycetes appears to be a bias in gene 165

orientation. Within the gene-dense regions of P. infestans, only about 40% of loci are 166

transcribed divergently, i.e. head-to-head; this is much less than expected by random 167

chance, and less than in yeasts and plants. An example of genes in this orientation is 168

shown in Fig. 2C, where the intergenic region is only 301 nt. About 10% of such 169

closely-spaced genes show some degree of correlated or anti-correlated expression 170

(69). Oomycete promoters are therefore not only compact, but may overlap and 171

possibly antagonize their neighbors. 172

Intergenic distances are also small between the 3' ends of genes transcribed 173

towards each other, averaging 255 nt within the gene-dense regions of P. infestans 174

(69). Such regions must include nearby transcriptional terminators, a configuration 175

which may mechanistically efficient. This also may generate overlapping transcripts for 176

epigenetic regulation. A role for the latter is demonstrated in fungi and metazoans (16), 177

and it would be interesting to see if this also holds true for oomycetes. 178

The tight spacing of genes may also help stabilize genomes. Insertion of a 179

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

9

transposable element or any illegitimate recombination event within such a region would 180

likely be lethal and eliminated quickly from the population. In contrast, if intergenic 181

distances were large then insertions would be tolerable in the short-term, but might 182

foster rearrangements that would be catastrophic during sexual reproduction. The 183

selection for genome stability may counterbalance the normal propensity of 184

transposable elements to insert within regions of high transcriptional activity, where a 185

higher fraction of DNA is uncoiled and exposed (21). The same principle likely 186

underlies the expansion of the gene-poor regions of oomycete genomes, as these but 187

not the gene-dense regions could tolerate transposon-mediated expansion. 188

The small size of most intergenic regions can also be helpful for identifying 189

regulatory motifs through mutagenesis or bioinformatics (1, 69, 81, 93). For example, 190

by searching just upstream of open reading frames for over-represented motifs, over 191

100 putative transcription factor binding sites could be identified from Phytophthora, of 192

which many were verified by functional assays (69). Most were within 200 nt of the 193

transcription start, consistent with the compactness and potential simplicity of oomycete 194

promoters. Nevertheless, sequenced Phytophthora spp. are predicted to encode 195

between 543 and 669 transcription factors, similar to that of fungal phytopathogens 196

such as Fusarium graminearum and Magnaporthe oryzae (659 and 481, respectively; 197

(63)). H. arabidopsidis, however, has fewer predicted transcription factors at 369. This 198

could reflect its simplified life cycle compared to Phytophthora; while the latter form 199

sporangia that release zoospores that encyst and then form germ tubes, H. 200

arabidopsidis conidia germinate directly and lack the zoospore stage. 201

Most oomycete genes have few introns, although intron-rich genes are common. 202

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

10

Means of about 1.6 per gene are predicted for Phytophthora and Pythium, although a 203

lower number was reported for the subset of P. sojae genes with EST support (51, 77, 204

86). Many cases exist of "mis-splicing" in which unexcised introns or alternate junctions 205

would lead to premature translational termination (77). There is, however, only one 206

credible report of alternate splicing generating a new biological function. This comes 207

from Ps. cubensis, where the novel product had a new transporter activity (72). 208

209

Novel gene fusions. Oomycetes are unusually rich in proteins with novel combinations 210

of domains (59, 74). These have several predicted functions, which often involve 211

signaling. Two notable examples are separate families of phosphatidylinositol-4-212

phosphate kinases and aspartic proteases with transmembrane domains characteristic 213

of G protein-coupled receptors (43, 58). While their structures are mostly unconfirmed 214

by cDNA or proteomic analysis, most appear conserved throughout Phytophthora. 215

Protein kinases also have evolved domain combinations that are oomycete-specific or 216

seen only in closely related groups. One example found in Phytophthora, A. laibachii, 217

H. arabidopsidis, and S. parasitica is a MAP kinase with a N-terminal PAS signal 218

sensing fold (34). This combination does not appear in other eukaryotes including 219

diatoms or chromalveolates. Another unusual protein joins an N-terminal cyclin-220

dependent kinase with a C-terminal cyclin, which in other taxa are on separate proteins. 221

This unusual configuration is detected in most oomycetes but otherwise appears only in 222

apicomplexans and ciliates, which are chromalveolates. 223

Due to the tight spacing between oomycete genes, viable fusions can easily result 224

from the loss of a stop codon in an upstream gene. The diploidy of oomycetes provides 225

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

11

a setting in which innovative combinations of domains could persist for long periods 226

during which their benefits could be tested, or be eliminated by gene conversion during 227

asexual growth or replacement with the wild-type allele during sexual reproduction. 228

229

Genes for pathogenesis and host colonization. After an oomycete attaches to a 230

plant or animal host, several events are required for successful infection. The invader 231

must degrade host barriers such as cell walls, usually by secreting hydrolytic enzymes. 232

Nutrients must be acquired, which for necrotrophs such as Py. ultimum involves taking 233

molecules leaked from damaged host cells. In contrast, most biotrophs and 234

hemibiotrophs are assumed to acquire much of their nutrients through haustoria, which 235

are specialized hyphae that insert into a host cell while remaining external to its plasma 236

membrane. Movement through haustoria is presumably bidirectional; not only are 237

nutrients probably moved from the plant, but oomycete proteins move across the 238

interface to alter the physiology of the host including its immune system. It should be 239

noted that data on the role of oomycete haustoria in nutrient acquisition are limited (3), 240

although infection-specific transporters can be identified from microarray data. 241

Analyses of the Phytophthora and Py. ultimum genomes indicate they are largely 242

autotrophic and can use diverse molecules as substrates for metabolism. As described 243

later, obligate biotrophs are deficient in several pathways. Haustorial oomycetes (A. 244

laibachii, H. arabidopsidis, Phytophthora) have lost the gene for thiamine biosynthesis. 245

The gene is present in Py. ultimum and S. parasitica, which as a necrotrophs are not 246

thought to employ haustoria (84). While it seems that haustorial oomycetes have 247

chosen to take this vitamin from their host, this is not the case for rust fungi such as 248

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

12

Uromyces which makes their own thiamine in haustoria (78). 249

Most plant pathogenic oomycetes can not make sterols, due to the absence of the 250

relevant biosynthetic genes. An exception is the legume root pathogen Aphanomyces 251

euteiches, where a biosynthetic pathway was predicted from an expressed sequence 252

tag (EST) project and confirmed biochemically (53). The status of the sterol pathway 253

has been a topic of long-standing interest since such compounds are needed for sexual 254

reproduction. Also, while the agrochemical industry has found sterol biosynthetic 255

enzymes to be frequent fungicide targets, this has not been the case for oomycetes. It 256

follows that a practical application of genome sequencing is the identification of targets 257

for chemicals to defend against oomycete pathogens. 258

Plant defenses against oomycetes include producing enzymes that degrade the 259

pathogen's cell wall. Oomycete proteins inhibiting these were first identified from EST 260

projects, and later catalogued from whole genomes. P. infestans, for example, encodes 261

38 secreted Kazal-like and cystatin protease inhibitors, some of which have been shown 262

to suppress host apoplastic serine and cysteine proteases, respectively. These 263

inhibitor-like genes are present in moderately reduced numbers in the other 264

Phytophthora spp. and Py. ultimum (51). The phytopathogenic oomycetes also make 265

inhibitors of plant β-1,3-glucanases (15). 266

Oomycetes themselves use proteases and other host-degrading enzymes for 267

offense, similar to other phytopathogens. Many have similar numbers in Phytophthora 268

and Py. ultimum, such as cysteine proteases which average 37 in Phytophthora versus 269

42 in Py. ultimum. Others vary widely, such as cutinases and pectin esterases which 270

have about 7-12 copies in each Phytophthora but none in Py. ultimum. This may be 271

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

13

because the latter primarily infects young non-suberized tissue while Phytophthora can 272

colonize plant parts with thick cuticles. In contrast, lipases and subtilisin-related serine 273

and aspartyl proteases are more common in Py. ultimum. As would be expected the 274

opportunistic fish pathogen S. parasitica is largely deficient in enzymes for degrading 275

plant cell walls, although its genome is annotated with two polygalacturonases. The 276

latter might be maintained for saprophytic growth on plant debris. 277

Whether an oomycete consumes plants or animals, host attachment is required for 278

infection; binding to a growth substrate must also be useful for saprophytes. 279

Oomycetes express several factors that may act in this role, such as glue-like proteins 280

released by zoospores, mucins, thrombospondin repeat proteins, jacalin domain 281

proteins, and cellulose-binding proteins including CBEL which genome data indicate are 282

made throughout oomycetes (25, 31, 68, 84). Interestingly, CBELs contain a 283

PAN/Apple domain also used by parasitic apicomplexans (also chromalveolates) to bind 284

their hosts. Recent discoveries in the Py. ultimum and Phytophthora genomes also 285

include cadherins, which play adhesion and secretory roles in animals (51). Except for 286

CBELs, most of these lack proved roles in pathogenesis, and it should be noted that 287

many oomycetes form adherent gametangia during sexual development which might 288

also use such proteins (37). 289

290

Fast-evolving families of effectors. Fungal, bacterial, and oomycete pathogens 291

secrete proteins collectively known as effectors that modify their hosts. Some proteins 292

described above are often also categorized as effectors, but much recent work has 293

focused on families unique to oomycetes especially the RXLRs and Crinklers/CRNs 294

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

14

(41). The CRNs are named after a "crinkling" or necrotic phenotype that many cause 295

when overexpressed in plants, and RXLR after an N-terminal motif (X is any amino 296

acid) involved in uptake by host cells. Many RXLRs suppress plant immunity, such as 297

Avr3a and AvrBlb2 which respectively stabilize an E3 ubiquitin-ligase used in plant 298

defense, and block transport of a plant protease (10). The names of these two effectors 299

indicate that they sometimes act as avirulence factors, which means that some plants 300

encode resistance or R proteins that recognize them and respond with a cell death 301

response (33). These proteins, but not necessarily all RXLRs, also contribute to 302

pathogenesis against plant genotypes lacking the cognate R genes. The reader is 303

directed to recent reviews for more details of their activities (10, 79). 304

RXLRs in particular represent a success story for bioinformatics. The family was 305

recognized because of a short N-terminal motif shared between cloned avirulence 306

proteins, as the rest of the proteins are very dissimilar. A conventional N-terminal signal 307

peptide is required, and a nearby dEER motif often associates with RXLR. Early 308

studies used that information to identify 563, 374, and 396 such proteins in P. infestans, 309

P. ramorum, P. sojae, and then 134 in H. arabidopsidis. It was later observed that the 310

original RXLR consensus may be too restrictive. For example, Ps. cubensis employs 311

RXLQ and A. laibachii both RXLR and RXLQ (45, 82). 312

Identification of the RXLR motif stimulated searches of secretomes for conserved 313

motifs and other hallmarks of host-translocated effectors. The CRNs were shown to 314

have a LXLFLAK motif, Albugo spp. contained a family defined by a CHXC, and Py. 315

ultimum was predicted to encode a group with YXSL[RK] (45, 51, 73). Some of these 316

are not widely distributed, such as the CHXCs which have 29 members in A. laibachii 317

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

15

but much fewer in the other oomycetes. It should be stressed that outside of A. laibachii 318

and P. infestans where functional testing of the CHXCs and CRNs was done, these are 319

just candidate effectors. 320

When examined at a genome-wide level it is clear that effectors are quite variable 321

within and between species. As noted before, their genes tend to reside in gene-poor 322

regions where DNA repeats may contribute to changes in family size, copy number, and 323

sequence. Some RXLR genes for example exist in tandem arrays of near-identical 324

copies, which vary in number in different strains (18). Comparisons of P. infestans with 325

four close relatives that infect different plants found that RXLR genes and CRN genes 326

were highly diversified, often through recombination within C-terminal domains which 327

may confer new activities for adaptation to their hosts (65). Examinations of genome or 328

EST data from other species have also identified many non-synonymous changes that 329

suggest an "arms race" between the pathogens and host resistance proteins that 330

interact with the effectors (4, 45, 65, 73). 331

Not all oomycete pathogens employ RXLRs. While found in most sequenced 332

peronosporalean genomes, RXLRs appear absent from Py. ultimum. This probably 333

relates to its necrotrophic lifestyle (in contrast biotrophs and hemibiotrophs need 334

effectors as part of stealthy infection strategies) and its broad host range (distinct 335

RXLRs are probably needed for different hosts). Only a few candidates were found in 336

the S. parasitica genome and in ESTs from the mycoparasite Pythium oligandrum, and 337

none from ESTs of the basal oomycete and algal pathogen Eurychasma dicksonii and 338

the legume root pathogen A. eutiches (25, 28, 32). However, inferences from EST 339

studies are not definitive. For example, sterol-carrier proteins known as elicitins are 340

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

16

predicted from the S. parasitica genome sequence but were not detected in ESTs (84). 341

Another effector group of note is the NLP family, which encode a phytotoxin active 342

against dicots (26). Like those above it is diversified in copy number and sequence 343

throughout the Oomycota. Unlike the others, NLPs represent an ancient family with 344

members in bacteria and fungi. Counterintuitively, the necrotroph Py. ultimum has 345

about one-sixth the number of NLP genes than the hemibiotrophic Phytophthora spp., 7 346

vs. about 45 excluding pseudogenes. A recent study in P. sojae, however, showed that 347

only half of its NLP genes were expressed and only eight had necrosis-inducing activity 348

(17). This means that a reduction in gene number does not necessarily translate to 349

reduced protein levels, which is an important warning to those making inferences from 350

genome analyses without functional data. 351

352

Adaptations for obligate parasitism. Downy mildews and white rusts are not 353

culturable on artificial media and seem to have evolved to be wholly dependent on their 354

hosts. This occurred multiple times in oomycete history, and genome studies suggest 355

an association with defects in metabolism (6, 45). Both A. laibachii and H. arabidopsidis 356

lack nitrite and nitrate reductase and thus are limited in their use of inorganic nitrogen. 357

A. laibachii but not H. arabidopsidis has also lost genes for making the molybdenum 358

cofactor used by nitrate reductase, and another cofactor-dependent enzyme, sulfite 359

oxidase, which generates ATP from sulfite. Both also lack sulfite reductase. These 360

enzymes are found in the other sequenced oomycetes such as Phytophthora, but 361

interestingly are also missing in Plasmodium, an obligately pathogenic chromaveolate 362

(24). Knowing these defects might lead to the artificial culture of downy mildews and 363

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

17

white rusts, however this may be an oversimplification since the species may have also 364

evolved to use plant signals to regulate their metabolism and development. 365

A. laibachii and H. arabidopsidis also encode reduced numbers of many proteins 366

involved in pathogenesis, including enzymes for degrading plant cell walls. Pectate 367

lyase, for example, is encoded by 30 genes in P. infestans but only 8 and 1 in A. 368

laibachii and H. arabidopsidis, respectively. Reductions are less dramatic for other wall 369

hydrolases, which may be used to form haustoria or remodel the pathogen's own walls. 370

The two species also have a reduced complement of CRNs and RXLRs. A. laibachii 371

and H. arabidopsidis encode only 3 and 20 CRNs compared to 196 for P. infestans. 372

CRNs induce host necrosis, so this is consistent with a greater reliance on biotrophy. 373

Genes for RXLR/Q proteins, which are encoded by >300 genes in each Phytophthora, 374

only number 49 in A. laibachii and 115 in H. arabidopsidis (note that A. laibachii also 375

has 29 of the CHXC proteins). The reduction may be explained by the fact that these 376

species need to interact with just one type of host. It is interesting to speculate whether 377

the loss of effectors was an adaptation to a narrowed host range, or its cause. It is also 378

possible that since these species inflict less damage to host cells than do 379

hemibiotrophs, fewer defenses against the plant immune response are required. 380

Gene content is also reduced in H. arabidopsidis due to loss of the zoospore 381

pathway. This includes at least 200 genes encoding structural components of flagella 382

along with >300 others with roles in zoospore assembly, chemotaxis, and encystment. 383

Fossilized remnants of many of these genes can be detected within the genome, but 384

most appear to have disappeared through slow mutation and genome contraction 385

through recombination. Albugo has however maintained the pathway, as have many 386

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

18

other downy mildews. Curiously, genes for flagella are retained in Py. ultimum despite 387

the apparent rarity of that stage in nature (51). 388

389

Transposable elements (TEs) and their impacts. Oomycetes contain diverse 390

populations of retrotransposon and DNA transposon-like sequences. Their abilities to 391

self-replicate and form recombination-stimulating repeats shape the structure of 392

genomes and induce variation. As noted above, TEs appear to have particularly 393

influenced the gene-sparse regions of oomycetes. Within gene-dense regions where 394

gene order is fairly conserved between species, sequences resembling TEs often reside 395

where microsynteny is disrupted. 396

As in many eukaryotes the most abundant oomycete retrotransposons belong to the 397

long terminal repeat (LTR)-containing gypsy and copia families, with non-LTR LINE 398

elements also present. Most are degenerate and incapable of movement. Genome 399

size in Phytophthora correlates with the number of retroelements. P. infestans contains 400

about 10,000 gypsy and 4,000 copia-like elements which comprise 37% of the genome, 401

while P. sojae and P. ramorum contain only 3,000 and 2,400 in total (31). In contrast, in 402

A. laibachii these only represent 9% of the genome and copia outnumbers gypsy by 3:1, 403

and the two make up only 2% of Py. ultimum chromosomes (45, 51). 404

The ratio of DNA transposons to retroelements is strikingly different between 405

species. For example, these comprise 10% and 37% of the P. infestans genome, 406

respectively, and 12% and 9% of A. laibachii. A massive expansion of retrotransposons 407

and not DNA elements thus primarily accounts for the larger Phytophthora genomes. 408

Nevertheless a diverse population of DNA transposons exists in Phytophthora, including 409

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

19

those that replicate by cut and paste, rolling circle, and self-synthesizing mechanisms. 410

Listed in declining order of copy number in P. infestans, ranging from about 2000 to 8, 411

these represent piggyBac, helitron, hAT, crypton, Tc1/mariner/pogo, MuDR/foldback, 412

Sola, Maverick, and PIF/harbinger families (20, 31, 87). Most are more abundant in P. 413

infestans, particularly piggyBac and helitrons. P. infestans has 273 helitrons of which 414

13 appear intact, which is more than other eukaryotes. P. ramorum and P. sojae 415

contain about one-tenth the number, none being functional. 416

Most of the DNA transposons have representatives in A. laibachii, H. arabidopsidis, 417

and Py. ultimum which suggests they are ancient members of the oomycete lineage. 418

Some such as hAT and Tc1/mariner may have been acquired after Stramenopiles 419

diverged from other eukaryotes as they are reportedly absent from the diatom 420

Thalassiosira pseudonana. The history of the cryptons is also interesting. Two of the 421

six major subfamilies described in eukaryotes, F and S, are throughout oomycetes while 422

diatoms only contain cryptonS. CryptonF in contrast is in Phytophthora but not A. 423

laibachii, H. arabidopsidis, or Py. ultimum. Outside of Phytophthora cryptonF has only 424

been found in ascomycetes, which suggests horizontal transfer (46). 425

There is evidence for some movement of TEs. One P. infestans gene was found to 426

contain an inserted gypsy element, including host target site duplications (31). hAT, 427

helitron, and PIF elements may also have "captured" Phytophthora genes and copied 428

them through the genome, based on the detection of their inverted terminal repeats with 429

target site duplications flanking adjoining transposase/replicase and host genes (87). 430

The latter encode SET domain proteins and AdoMet-dependent methyltransferases, 431

which interestingly are both involved in epigenetic regulation, an ABC transporter, 432

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

20

cysteine protease, and transglutaminase. 433

434

TEs, epigenetics, and transcriptional polymorphisms. Although few TEs are active 435

in oomycetes, many are transcribed (35). Others are silenced based on the detection in 436

P. infestans of small interfering RNAs or siRNA (90). That oomycetes have an active 437

epigenetic system was also shown by RNA interference studies, where silencing was 438

associated with siRNA and heterochromatinization that spread from the target locus (2, 439

39, 89). In non-oomycetes, TEs influence nearby genes by recruiting heterochromatin 440

modifications that repress transcription, or insulating genes from such effects (52). 441

Since a TE is within 2 kb over half of effector genes in P. infestans, their influence on 442

expression should be considered. 443

In fact, certain P. infestans and P. sojae RXLR genes are known to be 444

transcriptionally inactive, with virulent and avirulent alleles distinguished by differences 445

in expression rather than protein sequence (18, 27). Some alleles also vary 446

quantitatively in mRNA level (76). Promoter mutations were associated with one case 447

of RXLR silencing, but others could involve epigenetic events. Both TEs and 448

epigenetics may relate to the phenotypic variability observed within Phytophthora spp., 449

including reports of strains with unstable avirulence phenotypes or losing pathogenicity. 450

In P. ramorum, isolates with unusual colony morphologies and an early senescence 451

phenotype were observed to have higher levels of expression of several copia elements 452

(42), Although causality was not established, this may be a sign of epigenetic flux 453

within their genomes. The contribution of expression level polymorphisms to oomycete 454

biology, due to promoter differences or epialleles, is just starting to be appreciated. 455

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

21

It should be remembered that mobile elements and epigenetic phenomena are not 456

the only mechanisms driving change within oomycete genomes. Changes in ploidy, 457

gene conversion, and loss of heterozygosity are reported in several species (11, 30, 458

50). Unstable supernumerary chromosomes have also been described in Pythium (56). 459

460

Horizontal gene transfer (HGT). Prior to genome sequencing it was known that some 461

fungal and oomycete genes, such as the wall-degrading polygalacturonases, were 462

similar suggesting convergent evolution or HGT (83). As noted above the NLP 463

phytotoxins were also seen as HGT candidates (26). Careful phylogenetic analyses 464

using whole-genome data now provide strong support for HGT. Transfers from fungi to 465

oomycetes appear to have involved about 20 types of genes (67). About two-thirds 466

were predicted to be secreted and/or have potential roles in pathogenesis including 467

lipases, carbohydrate depolymerizing enzymes, sugar and nitrogenous base 468

transporters, and enzymes to degrade plant defense compounds. Since transfers from 469

fungi to oomycetes were much more common than the reverse, it was suggested that 470

plant pathogenesis developed in oomycetes more recently than in fungi. The same 471

study also reported very few cases of HGT for any gene in A. laibachii or S. parasitica 472

(67), and it is interesting to speculate that the larger TE-rich Phytophthora genomes are 473

more tolerant of recombination with foreign DNA. 474

Oomycetes also have obtained genes not related directly to pathogenesis by HGT, 475

including several involved in cofactor metabolism and amino acid or lipopolysaccharide 476

biosynthesis (91). Glucokinases are another example, with different genes coming from 477

plants and bacteroidales (38, 91). The retention of similar genes from two origins is not 478

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

22

very common in eukaryotes, but these may function in distinct cellular locations. 479

480

Uncertain support for chromalveolate hypothesis. In one evolutionary model, the 481

Chromalveolata are a major eukaryotic group derived from a biflagellated cell with a red 482

algal endosymbiont (44). Chromalveolates are proposed to include about half of known 483

protists and algae, mainly cryptophytes, alveolates, haptophytes, and stramenopiles 484

(the latter including various classes of golden-brown or brownish-green algae, diatoms, 485

and oomycetes). The diatom and algal lineages are thought to have retained the red 486

alga as a photosynthetic plastid, with oomycetes losing the endosymbiont but retaining 487

some genes (54). Studies of the first sequenced oomycetes, P. ramorum and P. sojae, 488

reported 855 genes of likely red algal origin based on sequence similarity (86). Combined 489

with the finding that chromalveolate and oomycete proteins share some novel domain 490

combinations, this was taken by some to support the model (44, 59). However, 491

endosymbiotic acquisition is hard to distinguish from other events including HGT. 492

More recent analyses may be leading to a consensus for a polyphyletic origin or 493

falsification of the Chromalveolata. Diatom and oomycete genes having affinity with red 494

alga have limited overlap, so some suggested that separate endosymbiotic events 495

occurred before and after they diverged from their common ancestor (5, 57, 62, 80). 496

Support for endosymbiosis in the diatom lineage appears solid, but its occurrence prior 497

to their divergence from oomycetes was challenged by a study that argued that other 498

evolutionary events provide better explanations for the presence of red algal-like genes 499

in oomycetes (80). 500

Regardless of the validity of the chromalveolate hypothesis, several data indicate 501

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

23

that stramenopiles have a monophyletic origin, including a recent analysis of 108 502

concatenated nuclear proteins (5). Oomycetes and diatoms also clearly contain 503

common sets of genes of plant ("green genes") and bacterial origin (9, 61). These 504

include many membrane transporters associated with nutrient acquisition or metabolism 505

with roles at the environmental interface or in organelles. Examples include 506

transporters for taking up nitrate, phosphate, or sulfate, and moving ADP/ATP across 507

mitochondrial membranes (12). 508

509

CONCLUSIONS AND OUTLOOK 510

Genomics has provided insight into the biology of oomycetes and paved the way for 511

future experimentation, but also raises new questions. The species sequenced so far 512

include agents of globally important diseases such as Phytophthora and Pythium, and 513

pathogens of a model plant. Phytophthora is also a good subject since it is amenable to 514

transformation, enabling functional studies of genes (36). However, data from more 515

species need to be made available publicly to help researchers understand the 516

foundations of pathogenesis and relationships between orders. These should include 517

both nonpathogens and pathogens from additional clades. These need not be limited to 518

species that are culturable in the lab, based on the success with the obligate pathogens 519

A. laibachii and H. arabidopsidis for which spores were taken from infected leaves for 520

analysis. This should also be feasible for the graminicolous (grass and grain-infecting) 521

downy mildews which are of major economic importance. Basal genera such as 522

Eurychasma or Haptoglossa, which include algal and nematode pathogens, are also 523

obligate pathogens but grow intracellularly with few external structures (8). Perhaps 524

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

24

sufficient DNA could be obtained from zoospores, or pathogen sequences separated 525

from host DNA in silico. 526

The extant data indicate that oomycete genomes have had dynamic histories 527

colored by expansions/contractions, birth/death events, gene fusions, and HGT. The 528

acquisition of bacterial, fungal, and organellar genes has given oomycetes unique 529

capabilities, enabling them to occupy multiple environmental niches. Variation within 530

species is also just starting to be understood; one must wonder what pressures are 531

imposed by environmental changes and agronomic practices such as agrochemical 532

treatment, which may select for alterations in coding and regulatory sequences and may 533

also have epigenetic effects. Several Phytophthora spp. appear to be recent 534

interspecific hybrids (48), perhaps fostered by man's movement of plant material, and it 535

would be of interest to see how their genomes evolve. 536

Evidence of adaptations are evident in the genomes of the obligately biotrophic 537

oomycetes, which are partially condensed. They have not undergone the dramatic 538

reductions of obligate pathogens like Giardia and Plasmodium (24, 60), perhaps since 539

plants provide a less stable environment than an animal host. The benefits of genome 540

reduction in H. arabidopsidis due to loss of the zoospore might seem obvious but 541

neither Albugo or Py. ultimum (which rarely if ever makes zoospores) has chosen this 542

path. While much remains to be learned about the ecological and environmental 543

pressures that influence the evolution of oomycetes, it is clear that nature provides 544

niches for both streamlined and expanded genomes. 545

546

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

25

ACKNOWLEDGMENTS 547

This work was supported by grants from the United States Department of Agriculture-548

National Institute of Food and Agriculture, and National Science Foundation. 549

550

REFERENCES 551

1. Ah Fong A, Xiang Q, Judelson HS. 2007. Architecture of the sporulation-specific 552

Cdc14 promoter from the oomycete Phytophthora infestans. Eukaryot. Cell 553

6:2222-2230. 554

2. Ah-Fong AM, Bormann-Chung CA, Judelson HS. 2008. Optimization of 555

transgene-mediated silencing in Phytophthora infestans and its association with 556

small-interfering RNAs. Fungal Genet. Biol. 45:1197-1205. 557

3. Andrews JH. 1975. Distribution of label from 3H-glucose and 3H-leucine in lettuce 558

cotyledons during the early stages of infection with Bremia lactucae. Can. J. Bot. 559

53:1103-1115. 560

4. As-sadi F, Carrere S, Gascuel Q, Hourlier T, Rengel D, Le Paslier MC, Bordat 561

A, Boniface MC, Brunel D, Gouzy J, Godiard L, Vincourt P. 2011. 562

Transcriptomic analysis of the interaction between Helianthus annuus and its 563

obligate parasite Plasmopara halstedii shows single nucleotide polymorphisms in 564

CRN sequences. BMC Genomics 12:498. 565

5. Baurain D, Brinkmann H, Petersen J, Rodriguez-Ezpeleta N, Stechmann A, 566

Demoulin V, Roger AJ, Burger G, Lang BF, Philippe H. 2010. Phylogenomic 567

evidence for separate acquisition of plastids in cryptophytes, haptophytes, and 568

stramenopiles. Molec. Biol. Evol. 27:1698-1709. 569

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

26

6. Baxter L, Tripathy S, Ishaque N, Boot N, Cabral A, Kemen E, Thines M, Ah-570

Fong A, Anderson R, Badejoko W, Bittner-Eddy P, Boore JL, Chibucos MC, 571

Coates M, Dehal P, Delehaunty K, Dong S, Downton P, Dumas B, Fabro G, 572

Fronick C, Fuerstenberg SI, Fulton L, Gaulin E, Govers F, Hughes L, 573

Humphray S, Jiang RH, Judelson H, Kamoun S, Kyung K, Meijer H, Minx P, 574

Morris P, Nelson J, Phuntumart V, Qutob D, Rehmany A, Rougon-Cardoso 575

A, Ryden P, Torto-Alalibo T, Studholme D, Wang Y, Win J, Wood J, Clifton 576

SW, Rogers J, Van den Ackerveken G, Jones JD, McDowell JM, Beynon J, 577

Tyler BM. 2010. Signatures of adaptation to obligate biotrophy in the 578

Hyaloperonospora arabidopsidis genome. Science 330:1549-1551. 579

7. Beakes GW, Glockling SL, Sekimoto S. 2011. The evolutionary phylogeny of the 580

oomycete "fungi". Protoplasma 249: 3-19. 581

8. Beakes GW, Sekimoto S. 2009. The evolutionary phylogeny of oomycetes-insights 582

gained from studies of holocarpic parasites of algae and vertebrates, p. 1-24. In 583

K. Lamour, and S. Kamoun (eds.), Oomycete Genetics and Genomics: Diversity, 584

Interactions, and Research Tools. John Wiley and Sons, Hoboken. 585

9. Bowler C, Allen AE, Badger JH, Grimwood J, Jabbari K, Kuo A, Maheswari U, 586

Martens C, Maumus F, Otillar RP, Rayko E, Salamov A, Vandepoele K, 587

Beszteri B, Gruber A, Heijde M, Katinka M, Mock T, Valentin K, Verret F, 588

Berges JA, Brownlee C, Cadoret JP, Chiovitti A, Choi CJ, Coesel S, De 589

Martino A, Detter JC, Durkin C, Falciatore A, Fournet J, Haruta M, Huysman 590

MJ, Jenkins BD, Jiroutova K, Jorgensen RE, Joubert Y, Kaplan A, Kroger N, 591

Kroth PG, La Roche J, Lindquist E, Lommer M, Martin-Jezequel V, Lopez 592

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

27

PJ, Lucas S, Mangogna M, McGinnis K, Medlin LK, Montsant A, Oudot-Le 593

Secq MP, Napoli C, Obornik M, Parker MS, Petit JL, Porcel BM, Poulsen N, 594

Robison M, Rychlewski L, Rynearson TA, Schmutz J, Shapiro H, Siaut M, 595

Stanley M, Sussman MR, Taylor AR, Vardi A, von Dassow P, Vyverman W, 596

Willis A, Wyrwicz LS, Rokhsar DS, Weissenbach J, Armbrust EV, Green BR, 597

Van de Peer Y, Grigoriev IV. 2008. The Phaeodactylum genome reveals the 598

evolutionary history of diatom genomes. Nature 456:239-244. 599

10. Bozkurt TO, Schornack S, Banfield MJ, Kamoun S. 2012. Oomycetes, effectors, 600

and all that jazz. Curr. Opin. Plant Biol. 15:483-492. 601

11. Chamnanpunt J, Shan WX, Tyler Brett M. 2001. High frequency mitotic gene 602

conversion in genetic hybrids of the oomycete Phytophthora sojae. Proc. Natl. 603

Acad. Sci. USA 98:14530-14535. 604

12. Chan CX, Reyes-Prieto A, Bhattacharya D. 2011. Red and green algal origin of 605

diatom membrane transporters: insights into environmental adaptation and cell 606

evolution. PLoS One 6:e29138. 607

13. Chen C, Gentles AJ, Jurka J, Karlin S. 2002. Genes, pseudogenes, and Alu 608

sequence organization across human chromosomes 21 and 22. Proc Natl Acad 609

Sci SA 99:2930-2935. 610

14. Cvitanich C, Salcido M, Judelson HS. 2006. Concerted evolution of a tandemly 611

arrayed family of mating-specific genes in Phytophthora analyzed through inter- 612

and intraspecific comparisions. Molec. Genet. Genom. 275:169-184. 613

15. Damasceno CM, Bishop JG, Ripoll DR, Win J, Kamoun S, Rose JK. 2008. 614

Structure of the glucanase inhibitor protein (GIP) family from Phytophthora 615

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

28

species suggests coevolution with plant endo-beta-1,3-glucanases. Molec. Plant 616

Microbe Interact. 21:820-830. 617

16. Donaldson ME, Saville BJ. 2012. Natural antisense transcripts in fungi. Molec. 618

Microbiol. 85:405-417. 619

17. Dong S, Kong G, Qutob D, Yu X, Tang J, Kang J, Dai T, Wang H, Gijzen M, 620

Wang Y. 2012. The NLP toxin family in Phytophthora sojae includes rapidly 621

evolving groups that lack necrosis-inducing activity. Molec. Plant Microbe 622

Interact. 25:896-909. 623

18. Dong S, Yu D, Cui L, Qutob D, Tedman-Jones J, Kale SD, Tyler BM, Wang Y, 624

Gijzen M. 2011. Sequence variants of the Phytophthora sojae RXLR effector 625

Avr3a/5 are differentially recognized by Rps3a and Rps5 in soybean. PLoS One 626

6:e20172. 627

19. Erwin DC, Ribeiro OK. 1996. Phytophthora diseases worldwide. APS Press, St. 628

Paul, Minn. 629

20. Feschotte C, Pritham EJ. 2007. DNA transposons and the evolution of eukaryotic 630

genomes. Ann. Rev. Genet. 41:331-368. 631

21. Fontanillas P, Hartl DL, Reuter M. 2007. Genome organization and gene 632

expression shape the transposable element distribution in the Drosophila 633

melanogaster euchromatin. PLoS Genet 3:e210. 634

22. Fry WE. 2008. Phytophthora infestans: the plant (and R gene) destroyer. Molec. 635

Plant Pathol. 9:385-402. 636

23. Fukuoka Y, Inaoka H, Kohane IS. 2004. Inter-species differences of co-expression 637

of neighboring genes in eukaryotic genomes. BMC Genomics 5:4. 638

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

29

24. Gardner MJ, Hall N, Fung E, White O, Berriman M, Hyman RW, Carlton JM, 639

Pain A, Nelson KE, Bowman S, Paulsen IT, James K, Eisen JA, Rutherford 640

K, Salzberg SL, Craig A, Kyes S, Chan MS, Nene V, Shallom SJ, Suh B, 641

Peterson J, Angiuoli S, Pertea M, Allen J, Selengut J, Haft D, Mather MW, 642

Vaidya AB, Martin DM, Fairlamb AH, Fraunholz MJ, Roos DS, Ralph SA, 643

McFadden GI, Cummings LM, Subramanian GM, Mungall C, Venter JC, 644

Carucci DJ, Hoffman SL, Newbold C, Davis RW, Fraser CM, Barrell B. 2002. 645

Genome sequence of the human malaria parasite Plasmodium falciparum. 646

Nature 419:498-511. 647

25. Gaulin E, Madoui MA, Bottin A, Jacquet C, Mathe C, Couloux A, Wincker P, 648

Dumas B. 2008. Transcriptome of Aphanomyces euteiches: new oomycete 649

putative pathogenicity factors and metabolic pathways. PLoS One 3:e1723. 650

26. Gijzen M, Nurnberger T. 2006. Nep1-like proteins from plant pathogens: 651

recruitment and diversification of the NPP1 domain across taxa. Phytochemistry 652

67:1800-1807. 653

27. Gilroy EM, Breen S, Whisson SC, Squires J, Hein I, Kaczmarek M, Turnbull D, 654

Boevink PC, Lokossou A, Cano LM, Morales J, Avrova AO, Pritchard L, 655

Randall E, Lees A, Govers F, van West P, Kamoun S, Vleeshouwers VG, 656

Cooke DE, Birch PR. 2011. Presence/absence, differential expression and 657

sequence polymorphisms between PiAVR2 and PiAVR2-like in Phytophthora 658

infestans determine virulence on R2 plants. New Phytol. 191:763-776. 659

28. Grenville-Briggs L, Gachon CM, Strittmatter M, Sterck L, Kupper FC, van West 660

P. 2011. A molecular insight into algal-oomycete warfare: cDNA analysis of 661

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

30

Ectocarpus siliculosus infected with the basal oomycete Eurychasma dicksonii. 662

PLoS One 6:e24500. 663

29. Grunwald NJ, Garbelotto M, Goss EM, Heungens K, Prospero S. 2012. 664

Emergence of the sudden oak death pathogen Phytophthora ramorum. Trends 665

Microbiol. 20:131-138. 666

30. Gu W, Spielman LJ, Matuszak JMM, Aist JR, Bayles CJ, Fry WE. 1993. 667

Measurement of nuclear DNA contents of Mexican isolates of Phytophthora 668

infestans. Mycol. Res. 97:857-860. 669

31. Haas BJ, Kamoun S, Zody MC, Jiang RH, Handsaker RE, Cano LM, Grabherr 670

M, Kodira CD, Raffaele S, Torto-Alalibo T, Bozkurt TO, Ah-Fong AM, 671

Alvarado L, Anderson VL, Armstrong MR, Avrova A, Baxter L, Beynon J, 672

Boevink PC, Bollmann SR, Bos JI, Bulone V, Cai G, Cakir C, Carrington JC, 673

Chawner M, Conti L, Costanzo S, Ewan R, Fahlgren N, Fischbach MA, 674

Fugelstad J, Gilroy EM, Gnerre S, Green PJ, Grenville-Briggs LJ, Griffith J, 675

Grunwald NJ, Horn K, Horner NR, Hu CH, Huitema E, Jeong DH, Jones AM, 676

Jones JD, Jones RW, Karlsson EK, Kunjeti SG, Lamour K, Liu Z, Ma L, 677

Maclean D, Chibucos MC, McDonald H, McWalters J, Meijer HJ, Morgan W, 678

Morris PF, Munro CA, O'Neill K, Ospina-Giraldo M, Pinzon A, Pritchard L, 679

Ramsahoye B, Ren Q, Restrepo S, Roy S, Sadanandom A, Savidor A, 680

Schornack S, Schwartz DC, Schumann UD, Schwessinger B, Seyer L, 681

Sharpe T, Silvar C, Song J, Studholme DJ, Sykes S, Thines M, van de 682

Vondervoort PJ, Phuntumart V, Wawra S, Weide R, Win J, Young C, Zhou S, 683

Fry W, Meyers BC, van West P, Ristaino J, Govers F, Birch PR, Whisson 684

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

31

SC, Judelson HS, Nusbaum C. 2009. Genome sequence and analysis of the 685

Irish potato famine pathogen Phytophthora infestans. Nature 461:393-398. 686

32. Horner NR, Grenville-Briggs LJ, van West P. 2011. The oomycete Pythium 687

oligandrum expresses putative effectors during mycoparasitism of Phytophthora 688

infestans and is amenable to transformation. Fungal Biol. 116:24-41. 689

33. Jones JD, Dangl JL. 2006. The plant immune system. Nature 444:323-329. 690

34. Judelson HS, Ah-Fong AM. 2010. The kinome of Phytophthora infestans reveals 691

oomycete-specific innovations and links to other taxonomic groups. BMC 692

Genomics 11:700. 693

35. Judelson HS, Ah-Fong AM, Aux G, Avrova AO, Bruce C, Cakir C, da Cunha L, 694

Grenville-Briggs L, Latijnhouwers M, Ligterink W, Meijer HJ, Roberts S, 695

Thurber CS, Whisson SC, Birch PR, Govers F, Kamoun S, van West P, 696

Windass J. 2008. Gene expression profiling during asexual development of the 697

late blight pathogen Phytophthora infestans reveals a highly dynamic 698

transcriptome. Molec. Plant Microbe Interact. 21:433-447. 699

36. Judelson HS, Ah-Fong AMV. 2009. Progress and challenges in oomycete 700

transformation, p. 435-454. In K. Lamour, and S. Kamoun (eds.), Oomycete 701

Genetics and Genomics. Wiley. 702

37. Judelson HS, Blanco FA. 2005. The spores of Phytophthora: weapons of the plant 703

destroyer. Nature Micro. Rev. 3:47-58. 704

38. Judelson HS, Narayan RD, Tani S. 2009. Metabolic adaptation of Phytophthora 705

infestans during growth on leaves, tubers, and artificial media. Molec. Plant 706

Pathol. 10:843-855. 707

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

32

39. Judelson HS, Tani S. 2007. Transgene-induced silencing of the zoosporogenesis-708

specific PiNIFC gene cluster of Phytophthora infestans involves chromatin 709

alterations. Eukaryot. Cell 6:1200-1209. 710

40. Kalmykova AI, Nurminsky DI, Ryzhov DV, Shevelyov YY. 2005. Regulated 711

chromatin domain comprising cluster of co-expressed genes in Drosophila 712

melanogaster. Nucl. Acids Res. 33:1435-1444. 713

41. Kamoun S. 2006. A Catalogue of the effector secretome of plant pathogenic 714

oomycetes. Ann. Rev. Phytopathol. 44:41-60. 715

42. Kasuga T, Kozanitas M, Bui M, Huberli D, Rizzo DM, Garbelotto M. 2012. 716

Phenotypic diversification is associated with host-induced transposon 717

derepression in the sudden oak death pathogen Phytophthora ramorum. PloS 718

One 7:e34728. 719

43. Kay J, Meijer HJ, ten Have A, van Kan JA. 2011. The aspartic proteinase family of 720

three Phytophthora species. BMC Genomics 12:254. 721

44. Keeling PJ. 2009. Chromalveolates and the evolution of plastids by secondary 722

endosymbiosis. J Eukaryot. Microbiol. 56:1-8. 723

45. Kemen E, Gardiner A, Schultz-Larsen T, Kemen AC, Balmuth AL, Robert-724

Seilaniantz A, Bailey K, Holub E, Studholme DJ, Maclean D, Jones JD. 2011. 725

Gene gain and loss during evolution of obligate parasitism in the white rust 726

pathogen of Arabidopsis thaliana. PLoS Biol 9:e1001094. 727

46. Kojima KK, Jurka J. 2011. Crypton transposons: identification of new diverse 728

families and ancient domestication events. Mob. DNA 2:12. 729

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

33

47. Kristiansson E, Thorsen M, Tamas MJ, Nerman O. 2009. Evolutionary forces act 730

on promoter length: identification of enriched cis-regulatory elements. Molec. 731

Biol. Evol. 26:1299-1307. 732

48. Kroon LP, Brouwer H, de Cock AW, Govers F. 2012. The genus Phytophthora 733

anno 2012. Phytopathology 102:348-364. 734

49. Kunjeti SG, Evans TA, Marsh AG, Gregory NF, Kunjeti S, Meyers BC, 735

Kalavacharla VS, Donofrio NM. 2011. RNA-Seq reveals infection-related global 736

gene changes in Phytophthora phaseoli, the causal agent of lima bean downy 737

mildew. Molec. Plant Pathol. 13:454-466. 738

50. Lamour K, Mudge J, Gobena D, Hurtado-Gonzales OP, Schmutz J, Kuo A, 739

Miller NA, Rice BJ, Raffaele S, Cano L, Bharti AK, Donahoo RS, Finley SL, 740

Huitema E, Hulvey J, Platt D, Salamov A, Savidor A, Sharma R, Stam R, 741

Storey D, Thines M, Win J, Haas B, Dinwiddie D, Jenkins J, Knight J, 742

Affourtit J, Han CS, Chertkov O, Lindquist EA, Detter C, Grigoriev IV, 743

Kamoun S, Kingsmore SF. 2012. Genome sequencing and mapping reveal loss 744

of heterozygosity as a mechanism for rapid adaptation in the vegetable pathogen 745

Phytophthora capsici. Molec. Plant Microbe Interact. 746

http://dx.doi.org/10.1094/MPMI-02-12-0028-R. 747

51. Levesque CA, Brouwer H, Cano L, Hamilton JP, Holt C, Huitema E, Raffaele S, 748

Robideau GP, Thines M, Win J, Zerillo MM, Beakes GW, Boore JL, Busam 749

D, Dumas B, Ferriera S, Fuerstenberg SI, Gachon CM, Gaulin E, Govers F, 750

Grenville-Briggs L, Horner N, Hostetler J, Jiang RH, Johnson J, Krajaejun T, 751

Lin H, Meijer HJ, Moore B, Morris P, Phuntmart V, Puiu D, Shetty J, Stajich 752

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

34

JE, Tripathy S, Wawra S, van West P, Whitty BR, Coutinho PM, Henrissat B, 753

Martin F, Thomas PD, Tyler BM, De Vries RP, Kamoun S, Yandell M, 754

Tisserat N, Buell CR. 2010. Genome sequence of the necrotrophic plant 755

pathogen Pythium ultimum reveals original pathogenicity mechanisms and 756

effector repertoire. Genome Biol. 11:R73. 757

52. Lisch D, Bennetzen JL. 2011. Transposable element origins of epigenetic gene 758

regulation. Curr. Opin. Plant Biol. 14:156-161. 759

53. Madoui MA, Bertrand-Michel J, Gaulin E, Dumas B. 2009. Sterol metabolism in 760

the oomycete Aphanomyces euteiches, a legume root pathogen. New Phytol. 761

183:291-300. 762

54. Martens C, Vandepoele K, Van de Peer Y. 2008. Whole-genome analysis reveals 763

molecular innovations and evolutionary transitions in chromalveolate species. 764

Proc. Natl. Acad. Sci. USA 105:3427-3432. 765

55. Martin F. 1995. Meiotic instability of Pythium sylvaticum as demonstrated by 766

inheritance of nuclear markers and karyotype analysis. Genetics 139:1233-1246. 767

56. Martin FM. 1995. Electrophoretic karyotype polymorphisms in the genus Pythium. 768

Mycologia 87:333-353. 769

57. Maruyama S, Matsuzaki M, Misawa K, Nozaki H. 2009. Cyanobacterial 770

contribution to the genomes of the plastid-lacking protists. BMC Evol Biol 9:197. 771

58. Meijer HJ, Govers F. 2006. Genomewide analysis of phospholipid signaling genes 772

in Phytophthora spp.: novelties and a missing link. Molec. Plant Microbe Interact. 773

19:1337-1347. 774

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

35

59. Morris PF, Schlosser LR, Onasch KD, Wittenschlaeger T, Austin R, Provart N. 775

2009. Multiple horizontal gene transfer events and domain fusions have created 776

novel regulatory and metabolic networks in the oomycete genome. PLoS One 777

4:e6133. 778

60. Morrison HG, McArthur AG, Gillin FD, Aley SB, Adam RD, Olsen GJ, Best AA, 779

Cande WZ, Chen F, Cipriano MJ, Davids BJ, Dawson SC, Elmendorf HG, 780

Hehl AB, Holder ME, Huse SM, Kim UU, Lasek-Nesselquist E, Manning G, 781

Nigam A, Nixon JE, Palm D, Passamaneck NE, Prabhu A, Reich CI, Reiner 782

DS, Samuelson J, Svard SG, Sogin ML. 2007. Genomic minimalism in the 783

early diverging intestinal parasite Giardia lamblia. Science 317:1921-1926. 784

61. Moustafa A, Beszteri B, Maier UG, Bowler C, Valentin K, Bhattacharya D. 2009. 785

Genomic footprints of a cryptic plastid endosymbiosis in diatoms. Science 786

324:1724-1726. 787

62. Parfrey LW, Grant J, Tekle YI, Lasek-Nesselquist E, Morrison HG, Sogin ML, 788

Patterson DJ, Katz LA. 2010. Broadly sampled multigene analyses yield a well-789

resolved eukaryotic tree of life. Syst. Biol. 59:518-533. 790

63. Park J, Park J, Jang S, Kim S, Kong S, Choi J, Ahn K, Kim J, Lee S, Kim S, 791

Park B, Jung K, Kim S, Kang S, Lee YH. 2008. FTFD: an informatics pipeline 792

supporting phylogenomic analysis of fungal transcription factors. Bioinformatics 793

24:1024-1025. 794

64. Ploch S, Thines M. 2011. Obligate biotrophic pathogens of the genus Albugo are 795

widespread as asymptomatic endophytes in natural populations of Brassicaceae. 796

Molec. Ecol. 20:3692-3699. 797

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

36

65. Raffaele S, Farrer RA, Cano LM, Studholme DJ, MacLean D, Thines M, Jiang 798

RH, Zody MC, Kunjeti SG, Donofrio NM, Meyers BC, Nusbaum C, Kamoun 799

S. 2010. Genome evolution following host jumps in the Irish potato famine 800

pathogen lineage. Science 330:1540-1543. 801

66. Raffaele S, Win J, Cano LM, Kamoun S. 2010. Analyses of genome architecture 802

and gene expression reveal novel candidate virulence factors in the secretome of 803

Phytophthora infestans. BMC Genomics 11:637. 804

67. Richards TA, Soanes DM, Jones MD, Vasieva O, Leonard G, Paszkiewicz K, 805

Foster PG, Hall N, Talbot NJ. 2011. Horizontal gene transfer facilitated the 806

evolution of plant parasitic mechanisms in the oomycetes. Proc. Natl. Acad. Sci. 807

USA 108:15258-15263. 808

68. Robold AV, Hardham AR. 2005. During attachment Phytophthora spores secrete 809

proteins containing thrombospondin type 1 repeats. Curr Genet 47:307-315. 810

69. Roy S. 2011. A multidisciplinary approach for identifying stage-specific transcription 811

factor binding sites in the Irish potato famine pathogen, Phytophthora Infestans. 812

Ph. D. thesis, University of California, Riverside. 813

70. Runge F, Telle S, Ploch S, Savory E, Day B, Sharma R, Thines M. 2011. The 814

inclusion of downy mildews in a multi-locus-dataset and its reanalysis reveals a 815

high degree of paraphyly in Phytophthora. IMA Fungus 2:163-171. 816

71. Savory EA, Adhikari BN, Hamilton JP, Vaillancourt B, Buell CR, Day B. 2012. 817

mRNA-Seq analysis of the Pseudoperonospora cubensis transcriptome during 818

cucumber (Cucumis sativus L.) infection. PLoS One 7:e35796. 819

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

37

72. Savory EA, Zou C, Adhikari BN, Hamilton JP, Buell CR, Shiu SH, Day B. 2012. 820

Alternative splicing of a multi-drug transporter from Pseudoperonospora cubensis 821

generates an RXLR effector protein that elicits a rapid cell death. PLoS One 822

7:e34701. 823

73. Schornack S, van Damme M, Bozkurt TO, Cano LM, Smoker M, Thines M, 824

Gaulin E, Kamoun S, Huitema E. 2010. Ancient class of translocated oomycete 825

effectors targets the host nucleus. Proc. Natl. Acad. Sci. USA 107:17421-17426. 826

74. Seidl MF, Van den Ackerveken G, Govers F, Snel B. 2011. A domain-centric 827

analysis of oomycete plant pathogen genomes reveals unique protein 828

organization. Plant Physiol. 155:628-644. 829

75. Seidl MF, Van den Ackerveken G, Govers F, Snel B. 2012. Reconstruction of 830

oomycete genome evolution identifies differences in evolutionary trajectories 831

leading to present-day large gene families. Genome Biol. Evol. 4:199-211. 832

76. Shan W, Cao M, Leung D, Tyler BM. 2004. The Avr1b locus of Phytophthora sojae 833

encodes an elicitor and a regulator required for avirulence on soybean plants 834

carrying resistance gene Rps1b. Molec. Plant Microbe Interact. 17:394-403. 835

77. Shen D, Ye W, Dong S, Wang Y, Dou D. 2011. Characterization of intronic 836

structures and alternative splicing in Phytophthora sojae by comparative analysis 837

of expressed sequence tags and genomic sequences. Can. J. Microbiol. 57:84-838

90. 839

78. Sohn J, Voegele RT, Mendgen K, Hahn M. 2000. High level activation of vitamin 840

B1 biosynthesis genes in haustoria of the rust fungus Uromyces fabae. Molec. 841

Plant Microbe Interact. 13:629-636. 842

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

38

79. Stassen J, Van den Ackerveken G. 2011. How do oomycete effectors interfere 843

with plant life? Curr. Opin. Plant Biol. 14:407-414. 844

80. Stiller JW, Huang J, Ding Q, Tian J, Goodwillie C. 2009. Are algal genes in 845

nonphotosynthetic protists evidence of historical plastid endosymbioses? BMC 846

Genomics 10:484. 847

81. Tani S, Judelson HS. 2006. Activation of zoosporogenesis-specific genes in 848

Phytophthora infestans involves a 7-nucleotide promoter motif and cold-induced 849

membrane rigidity. Eukaryot. Cell 5:745-752. 850

82. Tian M, Win J, Savory E, Burkhardt A, Held M, Brandizzi F, Day B. 2011. 454 851

Genome sequencing of Pseudoperonospora cubensis reveals effector proteins 852

with a QXLR translocation motif. Mol Plant Microbe Interact 24:543-553. 853

83. Torto TA, Rauser L, Kamoun S. 2002. The pipg1 gene of the oomycete 854

Phytophthora infestans encodes a fungal-like endopolygalacturonase. Curr. 855

Genet. 40:385-390. 856

84. Torto-Alalibo T, Tian M, Gajendran K, Waugh ME, van West P, Kamoun S. 857

2005. Expressed sequence tags from the oomycete fish pathogen Saprolegnia 858

parasitica reveal putative virulence factors. BMC Microbiol 5:46. 859

85. Tsai HK, Huang PY, Kao CY, Wang D. 2009. Co-Expression of neighboring genes 860

in the zebrafish (Danio rerio) genome. Int. J. Mol. Sci. 10:3658-3670. 861

86. Tyler BM, Tripathy S, Zhang X, Dehal P, Jiang RH, Aerts A, Arredondo FD, 862

Baxter L, Bensasson D, Beynon JL, Chapman J, Damasceno CM, Dorrance 863

AE, Dou D, Dickerman AW, Dubchak IL, Garbelotto M, Gijzen M, Gordon 864

SG, Govers F, Grunwald NJ, Huang W, Ivors KL, Jones RW, Kamoun S, 865

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

39

Krampis K, Lamour KH, Lee MK, McDonald WH, Medina M, Meijer HJ, 866

Nordberg EK, Maclean DJ, Ospina-Giraldo MD, Morris PF, Phuntumart V, 867

Putnam NH, Rash S, Rose JK, Sakihama Y, Salamov AA, Savidor A, 868

Scheuring CF, Smith BM, Sobral BW, Terry A, Torto-Alalibo TA, Win J, Xu Z, 869

Zhang H, Grigoriev IV, Rokhsar DS, Boore JL. 2006. Phytophthora genome 870

sequences uncover evolutionary origins and mechanisms of pathogenesis. 871

Science 313:1261-1266. 872

87. Vadnagara K. 2010. A tale of three phytopathogens: impact of transposable 873

elements on genome evolution. M. S. Thesis, University of Texas, Arlington. 874

88. van West P. 2006. Saprolegnia parasitica, an oomycete pathogen with a fishy 875

appetite: new challenges for an old problem. Mycologist 20:99-104. 876

89. van West P, Shepherd SJ, Walker CA, Li S, Appiah AA, Grenville-Briggs LJ, 877

Govers F, Gow NA. 2008. Internuclear gene silencing in Phytophthora infestans 878

is established through chromatin remodelling. Microbiology 154:1482-1490. 879

90. Vetukuri RR, Tian Z, Avrova AO, Savenkov EI, Dixelius C, Whisson SC. 2011. 880

Silencing of the PiAvr3a effector-encoding gene from Phytophthora infestans by 881

transcriptional fusion to a short interspersed element. Fungal Biol. 115:1225-882

1233. 883

91. Whitaker JW, McConkey GA, Westhead DR. 2009. The transferome of metabolic 884

genes explored: analysis of the horizontal transfer of enzyme encoding genes in 885

unicellular eukaryotes. Genome Biol 10:R36. 886

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

40

92. Xiang Q, Judelson HS. 2010. Myb transcription factors in the oomycete 887

Phytophthora with novel diversified DNA-binding domains and developmental 888

stage-specific expression. Gene 453:1-8. 889

93. Xiang Q, Roy S, Kim KS, Judelson HS. 2009. A motif within a complex promoter 890

from the oomycete Phytophthora infestans determines transcription during an 891

intermediate stage of sporulation. Fungal Genet. Biol. 46:400-409. 892

94. Zhan S, Horrocks J, Lukens LN. 2006. Islands of co-expressed neighboring genes 893

in Arabidopsis thaliana suggest higher-order chromosome domains. Plant J. 894

45:3347-3357. 895

896

897

898

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

41

FIGURE LEGENDS 899

Fig. 1. Evolutionary history of oomycetes, lifestyles, and genome sequencing overview. 900

The main part of the figure portrays phylogenetic relationships between the major 901

orders, based on 28S rRNA sequences. Orders comprising the peronosporalean group 902

(~1000 species), saprolegnians (~500 species), and basal clades are noted (8). While 903

Phytophthora was traditionally considered as part of the Pythiales, it is now recognized 904

to belong to the Peronosporales along with the downy mildews (70). Species with 905

publicly released genome data are listed along with their predicted gene number, 906

genome size, and repetitive DNA content if known; values are based on the most recent 907

publication or data on web sites. The phylogram in the lower right shows the position of 908

oomycetes compared to other eukaryotes. 909

910

Fig. 2. Structural and transcriptional landscape of the P. infestans genome. (A) 911

Expression patterns of genes in a portion of a chromosome. Genes with >2-fold higher 912

mRNAs in one life-stage versus average are colored based when expression is highest, 913

according the key in the upper right (HY, hyphae; SP, sporangia; CLSP, sporangia 914

cleaving into zoospores; ZO, swimming zoospores; GC, zoospore cysts germinating 915

and making infection structures, e.g. appressoria). Genes not changing are in dark 916

grey, at half-height. Note that the scaffold is split into two portions, and the horizontal 917

axis represents gene order and not distance. (B) Representative gene-dense and 918

gene-sparse regions. Shown are gene orientations, expression patterns as in panel A, 919

and flanking DNA transposon (T) or retroelement-like sequences (R). The two regions 920

are separated by 450 kb, which is also gene-sparse. (C) Example of closely spaced 921

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

42

opposing promoters from genes with strong EST support from panel B. The 301 nt 922

intergenic region includes a combined 142 nt of 5' UTRs, and has five predicted 923

transcription factor binding sites. The genes have a similar configuration in P. sojae. 924

925

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from

on August 21, 2018 by guest

http://ec.asm.org/

Dow

nloaded from