2 zymoseptoria tritici - biorxiv · 2 26 abstract 27 the ability of fungal cells to undergo cell...

49
1 The role of vegetative cell fusions in the lifestyle of the wheat fungal pathogen 1 Zymoseptoria tritici 2 3 4 5 Carolina Sardinha Francisco 1* , Maria Manuela Zwyssig 1 , Javier Palma-Guerrero 1,2* 6 7 8 9 1 Plant Pathology Group, Institute of Integrative Biology, ETH Zürich, 8092 Zürich, 10 Switzerland 11 2 Take-all Group, Department of biointeractions and crop protection, Rothamsted 12 Research, Harpenden, Switzerland 13 14 15 16 *Corresponding authors 17 E-mail: [email protected] and javier.palma- 18 [email protected] 19 20 21 Keywords: cell-to-cell communication, anastomosis, vegetative growth, melanization, 22 pycnidial development, asexual reproduction 23 24 25 . CC-BY-ND 4.0 International license author/funder. It is made available under a The copyright holder for this preprint (which was not peer-reviewed) is the . https://doi.org/10.1101/2020.01.26.918797 doi: bioRxiv preprint

Upload: others

Post on 19-Apr-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

1

The role of vegetative cell fusions in the lifestyle of the wheat fungal pathogen 1

Zymoseptoria tritici 2

3

4

5

Carolina Sardinha Francisco1*, Maria Manuela Zwyssig1, Javier Palma-Guerrero1,2* 6

7

8

9

1Plant Pathology Group, Institute of Integrative Biology, ETH Zürich, 8092 Zürich, 10

Switzerland 11

2Take-all Group, Department of biointeractions and crop protection, Rothamsted 12

Research, Harpenden, Switzerland 13

14

15

16

*Corresponding authors 17

E-mail: [email protected] and javier.palma-18

[email protected] 19

20

21

Keywords: cell-to-cell communication, anastomosis, vegetative growth, melanization, 22

pycnidial development, asexual reproduction 23

24

25

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 2: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

2

Abstract 26

The ability of fungal cells to undergo cell fusion allows them to maximize their overall 27

fitness. In this study, we characterized the role of the so gene orthologous in 28

Zymoseptoria tritici and the biological contribution of vegetative cell fusions in the 29

lifestyle of this latent necrotrophic fungus. Firstly, we show that Z. tritici undergoes 30

self-fusion between distinct cellular structures and its mechanism is dependent on the 31

initial cell density. Next, the deletion of ZtSo resulted in the loss of cell-to-cell 32

communication affecting both hyphal and germlings fusion. We show that Z. tritici 33

mutants for MAP kinase-encoding ZtSlt2 (orthologous MAK-1) and ZtFus3 34

(orthologous MAK-2) genes also fail to undergo self-stimulation and self-fusion, 35

demonstrating the functional conservation of this signaling mechanism across 36

species. Additionally, the DZtSo mutant was severely impaired in melanization, which 37

leads us to identify a trade-off between fungal growth and melanization. Though it has 38

been proposed that So is a scaffold protein for MAP kinase genes from the CWI 39

pathway, its deletion did not affect the cell wall integrity of the fungus. Finally, we 40

demonstrated that anastomose is dispensable for pathogenicity, but essential for the 41

fruiting body development and its absence abolish the asexual reproduction of Z. tritici. 42

Taken together, our data show that ZtSo is required for fungal development, while 43

vegetative cell fusions are essential for fungal fitness. 44

45

46

47

48

49

50

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 3: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

3

Introduction 51

Communication is a ubiquitous primitive characteristic developed by all living species. 52

From high mammals to the simplest forms of life, the exchange of signals between 53

organisms have driven the evolution of the species (Endler, 1993, Ord & Garcia-Porta, 54

2012, Wilson, 1975). The ability to communicate effectively may affect mating, 55

predation, competition, dominance hierarchy, signal modalities, and survival (Endler, 56

1993, Gillam, 2011, Wilson, 1975). This complex mechanism starts when a given 57

organism (the sender) secretes in the environment a self-produced molecular signal 58

(the message) which alters the behavior of another organism (the receiver) (Endler, 59

1993, Wilson, 1975). Communication also happens at the cellular level. This so-called 60

“cell-to-cell communication” creates a complex signaling network that involves 61

different extracellular signals and distinct cell types that regulate manifold pathways 62

(van Gestel et al., 2012, Shrout et al., 2011, Fischer & Glass, 2019). Inter- and 63

intraspecies cell-to-cell communication has been widely studied in fungi to address 64

biological functions including the secretion of pheromones to attract the opposite 65

sexual partner; the production of quorum sensing molecules controlling the expression 66

of virulence factors and morphological changes and, the regulation of cell fusions 67

during vegetative growth (Cottier & Muhlschlegel, 2012, Bloemendal & Kuck, 2013, 68

Khang et al., 2010, Wongsuk et al., 2016, Fischer & Glass, 2019). 69

70

The fungal mycelium is formed by three integrated processes including hyphal 71

extension, branching and vegetative hyphal fusion - VHF (also known as anastomosis) 72

(Glass et al., 2000). In this last-mentioned process, two growing cells with identical 73

vegetative compatibility loci engage in cell-to-cell communication, known as “ping-74

pong self-signaling”, which is thought to involve the secretion of unknown diffusible 75

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 4: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

4

molecules and results in re-direct polarized hyphal growth toward each other. After 76

physical contact, the cell walls are remodeled, the plasma membranes fused and the 77

two interconnected cells exchange cytoplasm and organelles (Fleissner et al., 2009). 78

Whether the anastomosed individuals are vegetatively incompatible, the two fused 79

cells rapidly collapse following DNA degradation by programmed cell death or they are 80

severely inhibited in their growth (Saupe, 2000). Albeit non-self-anastomoses are 81

described (Roca et al., 2004, He et al., 1998), this might be a very rare event in nature. 82

It is widely accepted that mycelial network formed through VHF facilitates the intra-83

hyphal communication, translocation of water, nutrients, and signal molecules, which 84

improves the general homeostasis and the spatial expansion of the fungal colony 85

(Read et al., 2010, Hickey et al., 2002). In some pathogenic fungi, hyphal fusion is 86

required for pathogenicity and host adhesion (Craven et al., 2008, Prados Rosales & 87

Di Pietro, 2008). Fungal cell fusions can also occur between conidial cells. Conidium 88

is the asexual spore of many Ascomycetes and Basidiomycetes species. The process 89

of fusion between germinating conidia involves the formation and interaction of 90

specialized hyphae, called conidial anastomosis tube (CAT). CAT is thinner and 91

shorter than VHF, and its induction dependent on nutrient deprivation and initial cell 92

density (Roca et al., 2005a). It has been postulated that CATs improve colony 93

establishment, as well as they may increase the genetic variability by facilitating 94

heterokaryosis and parasexual recombination, especially for those fungal species 95

lacking sexual reproduction (Roca et al., 2005a). Evidence of gene and chromosome 96

transfer between intra- or inter-fungal species has been shown as a mechanism to 97

acquire pathogenicity or to broaden the host specificity (Mehrabi et al., 2011, Friesen 98

et al., 2006, Temporini & VanEtten, 2004). On the other hand, anastomose can also 99

be high-risk gambling by the acquisition of infectious cytoplasmic and genetic 100

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 5: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

5

elements, such as mycoviruses, selfish elements or debilitated organelles, that can 101

multiply uncontrolled and drain the energetic sources of the cell (Saupe, 2000, Biella 102

et al., 2002). 103

104

In the last decades, different studies about the molecular mechanisms underlying cell 105

fusion identified several mutants defective in anastomosis, revealing that fungal 106

communication and fusion are complex mechanisms, which encompasses 107

components of several signaling pathways (Fu et al., 2011, Xiang et al., 2002, 108

Aldabbous et al., 2010, Pandey et al., 2004, Fu et al., 2014, Fischer & Glass, 2019). 109

Hitherto, the best-characterized mutant is for the soft (so) gene of Neurospora crassa, 110

a gene from the mitogen-activated protein (MAP) kinase MAK-1 pathway, which 111

encodes a protein with unknown function (Fleissner et al., 2005). The So protein is 112

proposed to be the scaffold for cell wall sensors belonging to the MAP kinase cascade 113

from the cell wall integrity (CWI) signaling pathway (Fischer & Glass, 2019). 114

Furthermore, the so gene has been shown to have an essential role in the hyphal 115

anastomosis, presumably by regulating the secretion of an undefined chemoattractant 116

in an oscillatory manner with the MAK-2 from the MAP kinase signal response pathway 117

(Fleissner et al., 2009). Beyond N. crassa, the so gene has been characterized in 118

different filamentous fungi including model organisms, plant pathogens, and 119

endophytic fungi (Engh et al., 2007, Maruyama et al., 2010, Craven et al., 2008, 120

Prados Rosales & Di Pietro, 2008, Charlton et al., 2012). Though all so mutants fail to 121

undergo hyphal fusions, the distinct effects on pathogenicity reported by the deletion 122

of the so gene in different fungi suggests that the biological contribution of 123

anastomosis might depend on the infection strategies developed by different fungal 124

pathogen species. 125

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 6: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

6

126

Z. tritici is an apoplastic pathogen with a latent necrotrophic lifestyle and considered 127

the most damaging pathogen of wheat in Europe (Fones & Gurr, 2015). Hyphae 128

formed from either germinated ascospores (sexual spore), pycnidiospores (asexual 129

spores) or blastospores (asexual spores produced by budding) are essential for 130

penetrating wheat leaves through stomata and colonization of the apoplastic space. 131

After a long asymptomatic phase (which varies depending on the wheat genotype and 132

fungal strain combination), the onset of the necrotrophic phase is followed by the 133

appearance of lesions, disintegration of host tissue and formation of asexual fruiting 134

bodies. Though Z. tritici is among the top 10 most studied phytopathogens (Dean et 135

al., 2012), little is known about vegetative cell fusion in this organism. To date, it has 136

been shown that the deletion of the b-subunit of the heterotrimeric G protein - MgGpb1 137

or the ZtWor1, a transcriptional regulator of genes located downstream of the cyclic 138

adenosine monophosphate (cAMP) pathway, negatively regulate cell fusion producing 139

germ tubes that undergo extensive anastomosis (Mehrabi et al., 2009, Gohari et al., 140

2014). 141

142

In this study, we aimed to determine whether the ZtSo gene and vegetative cell fusions 143

play important biological roles in the lifestyle of Z. tritici. We showed that the ubiquitous 144

ability of Z. tritici to undergo self-fusion was disrupted by the deletion of ZtSo affecting 145

both hyphal and germling fusions. The characterization of mutants lacking the MAP 146

kinase-encoding genes ZtSlt2 (orthologous MAK-1) or ZtFus3 (orthologous MAK-2) 147

showed that the cell fusion of Z. tritici is also regulated by the ping-pong self-signaling 148

mechanism, demonstrating the functional conservation of this mechanism across 149

species. We found that ZtSo is required for vegetative growth and melanization, but 150

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 7: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

7

not to maintain the cellular integrity of the fungus. We discovered that anastomoses 151

are dispensable for pathogenicity, but they are essential for fruiting body development. 152

In the absence of cell fusions, Z. tritici does not undergo asexual reproduction. These 153

findings illustrate the impact of ZtSo for fungal development and the importance of 154

vegetative cell fusions for fungal fitness. 155

156

Results 157

158

Cell fusion in Z. tritici allows the bidirectional transfer of cytoplasmic content, 159

but not nuclei exchange 160

We co-inoculated either blastospores or pycnidiospores of both 1E4GFP and 1E4mCh 161

fluorescent strains onto water agar (WA - 1% agar in water), a hyphal-inducing 162

medium, to investigate the ability of Z. tritici to undergo self-fusions. Fungal cells 163

expressing each fluorescent protein appeared in only one color-channel of the 164

fluorescence microscope (Fig. 1A), which ensures that the reciprocal cytoplasmic 165

streaming is explicitly due to cell fusions. Though Z. tritici produces blastospore and 166

pycnidiospores as asexual spores instead of conidium, we used the CAT terminology 167

to define the fusion between germinating spores. CATs formed between blastospore 168

or pycnidiospores germlings started after 4 hours of incubation, but are frequently 169

observed after 17 hours of incubation (Fig. 1B). Vegetative hyphal fusions (VHFs) from 170

germinated blastospores or pycnidiospores were noticed at 40 hours after incubation 171

(hai) (Fig. 1C and D). Multiple interconnections via fusion bridges were observed in all 172

tested morphotypes (Fig. 1B-D). The co-infection of wheat plants using either 173

blastospores or pycnidiospores of 1E4GFP and 1E4mCh strains also resulted in VHFs on 174

the wheat leaf surface (Fig. S1). Self-fusions and cytoplasmic mixing occurred in the 175

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 8: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

8

first 48 hai (Fig. S1A-B). It indicates that VHFs may play a role during the initial fungal 176

establishment and colonization of the host tissues. 177

178

To monitor the genetic exchange between genetically identical Z. tritici, we used the 179

IPO323 ZtHis1-ZtGFP strain. We looked for the presence of multinucleated cells 180

surrounding hyphal fusion points. None cell compartments containing more than one 181

nucleus, nor a nuclei exchange between the two interconnected hyphae were 182

observed (Fig. 2). After the fusion bridge formation, one of the neighbor’s nuclei 183

divides and migrate to occupy the new septal compartment. This finding suggests that 184

cell fusion in Z. tritici results only on cytoplasmic mixing, but not in heterokaryon 185

formation or parasexuality. 186

187

Orthologues of ZtSo are widely distributed within the Dothideomycetes 188

We identified the so orthologous (ZtSo) in the Z. tritici genome. The ZtSo gene 189

(Mycgr3G74194 or Zt09_7_00503) consist of 3,794 bp open reading frame and 190

encodes a polypeptide of 1,227 amino acid (Fig. S2A). Overall, proteins with varying 191

degrees of similarity to ZtSo were identified among several Dothideomycetes species 192

(n=20). We also found homology with other four fungal Classes, including 193

Sordariomycetes (n=4), Xylomycetes (n=1), Eurotiomycetes (n=3), and 194

Chaetothyriomycetes (n=1). The alignment of the amino acid sequences revealed 195

striking differences in homology among protein sequences (data not shown). 196

However, as earlier reported (Fleissner et al., 2005, Craven et al., 2008, Bork & Sudol, 197

1994), the WW protein-protein interaction domain is highly conserved (70% similarity) 198

within the fungal species analyzed. WW domain contains the motif PPLP and two 199

conserved tryptophan residues spaced 22 amino acid apart. Beyond the WW domain, 200

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 9: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

9

the Atrophin-1 superfamily and PhoD domains were also identified (Fig. S2B). 201

Comparison of ZtSo protein sequence with its orthologues in the most studied fungi 202

for anastomosis showed 53% identity with Epichloe festucae; 54% identity with 203

Neurospora crassa and Sordaria macrospora; 55% identity with Fusarium oxysporum; 204

60% identity with Aspergillus oryzae; and 63% identity with Alternaria brassicicola. 205

Phylogenetic analysis grouped the orthologues of the ZtSo gene onto three groups 206

based on fungal Classes (Dothideomycetes, Sordariomycetes, and 207

Chaetothyriomycetes together with Eurotiomycetes), independently whether they 208

were parasites, mutualists or saprotrophs (Fig. S2C). 209

210

The mutual attraction towards genetically identical fusion partners is regulated 211

by different genes from MAP kinase pathways 212

so, a gene from the CWI pathway has been shown to have an essential role in self-213

anastomosis. To determine whether the orthologous gene ZtSo plays the same role 214

in Z. tritici, we phenotyped the DZtKu70, DZtSo, and DZtSo-comp strains for the 215

presence of interconnected individuals through anastomoses during growth on WA. 216

Fusion bridges between blastospore germlings or filamentous hyphae were only 217

observed for those strains possessing the ZtSo gene. On the other hand, the DZtSo 218

mutant lost the ability to undergo vegetative fusions (Figs. 3A and B), unveiling the 219

contribution of ZtSo for anastomosis between genetically identical Z. tritici strains. 220

221

To ensure the failure of cytoplasmic mixing on those individuals lacking the ZtSo gene, 222

we mixed blastospores of each tested strain with 1E4GFP blastospores in a 1:1 ratio 223

and microscopically traced the spores up to 40 hours. Hyphal fusions and the 224

continuous streaming of cytoplasmic green fluorescence coming from the fusion with 225

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 10: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

10

the 1E4GFP strain were observed for all tested combinations (Figs. S3 and S4), except 226

for the mixture between DZtSo and 1E4GFP (Fig. S5). This confirmed that ZtSo is 227

needed in both fusion partners for a mutual recognition previous to the anastomosis. 228

229

To determine whether the MAK-1 and MAK-2 pathways are required for anastomoses 230

in Z. tritici, we incubated the knocked-out ZtSlt2 (orthologous to MAK-1) and ZtFus3 231

(orthologous to MAK-2) mutants on WA plates. Neither hyphal attraction nor hyphal 232

fusion was observed for both mutants, however, fusion bridges were frequently found 233

between IPO323 wild-type or DZtSlt2-complemented fungal cells (Fig. S6). These 234

results indicate the recruitment of both CWI and MAK-2 MAP kinase signaling 235

cascades for the regulation of the self-stimulation and self-fusion in Z. tritici. 236

237

Deletion of ZtSo has a differential impact on hyphal or blastospore growth 238

Z. tritici alters its growth morphology in detrimental to the nutritional conditions 239

(Francisco et al., 2019). To assess whether vegetative cell fusion affects fungal 240

growth, we determined the radial growth of DZtKu70, DZtSo, and DZtSo-comp on 241

different culture media. A slightly reduced growth was detected in the DZtSo colonies 242

compared to those possessing the ZtSo gene (Fig. S7A) when it’s grown on WA. On 243

average, the colony radii were 5.31 ± 0.12 for DZtKu70; 4.87 ± 0.10 for DZtSo; and 244

5.50 ± 0.09 for DZtSo-comp [radial growth (mm) ± standard error] (Fig. S7C). 245

Furthermore, DZtKu70 and DZtSo-comp formed colonies with highly hyphal dense 246

margins, while DZtSo exhibited only a few filamentations at the colony periphery (Fig. 247

S7B). No morphological differences were detected between blastospores of the tested 248

strains (Fig. S7D). In contrast, the DZtSo mutant grew significantly faster than the 249

DZtKu70 and DZtSo-comp in PDA, a nutrient-rich medium (Fig. S7E). Over time, the 250

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 11: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

11

relative growth rate of DZtKu70 and DZtSo-comp were, respectively, 32% and 16% 251

lower than the DZtSo mutant. Together, these results suggest that the reduced 252

mycelial expansion of the DZtSo mutant grown in a nutrient-poor medium is a 253

consequence of impairment of VHFs. On the other hand, the ZtSo gene may contribute 254

to the blastosporulation and its deletion promotes an imbalance of this mechanism in 255

Z. tritici. 256

257

ZtSo is required for melanization, but it is dispensable for stress sensitivity 258

No melanin accumulation was observed in DZtSo mutant colonies grown in vitro (Fig. 259

4), demonstrating the impact of ZtSo deletion for Z. tritici pigmentation. Next, we 260

postulated that the DZtSo mutant could either display cellular integrity defects and/or 261

being susceptible to environmental stresses. We tested nine different abiotic 262

stressors, such as temperature, oxidative, osmotic, cell wall, and cell membrane 263

stresses. Overall, no variability in stress response was noticed among the strains (Fig. 264

5). However, DZtSo formed slightly bigger colonies compared to those from DZtKu70 265

and DZtSo-comp, most likely to the increase of blastospore formation when its grown-266

on nutrient-rich media (Fig. S7E). Thus, we found no evidence that the ZtSo gene is 267

required for the maintenance of the cell wall integrity. Furthermore, our results 268

demonstrated that melanin does not act as “fungal armor” protecting Z. tritici against 269

the tested stresses. 270

271

Hyphal fusions are essential for the development of asexual fruiting bodies 272

We inoculated a susceptible wheat cultivar with the tested Z. tritici strains to assess 273

the biological role of vegetative cell fusion during the pathogen lifecycle in planta. 274

Typical symptoms caused by Z. tritici infections were visible after 9 days post-infection 275

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 12: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

12

(dpi) and the disease progression was similar among the plants inoculated with 276

DZtKu70, DZtSo or DZtSo-comp strains (Figs. 6A and S8). This indicates that ZtSo is 277

neither required for host penetration nor the asymptomatic or necrotrophic phases of 278

the fungus. The asexual fruiting bodies (pycnidia) were visible on plants inoculated 279

with those strains possessing the ZtSo gene at 14 days post-inoculation (dpi). In 280

contrast, plants infected with DZtSo never developed pycnidia. The failure to form the 281

asexual reproductive structure was also observed in other susceptible cultivars to Z. 282

tritici (Fig. S9). 283

284

To distinguish whether the absence of pycnidia formation would be either a 285

consequence of the lack of hyphal fusions or higher sensitivity of the DZtSo mutant to 286

the plant defenses, we used a wheat extract agar medium to induce pycnidia formation 287

in vitro. DZtSo produced only mycelial knots, the forerunner developmental stage of 288

mature pycnidia, but no asexual reproductive structures were further developed. 289

Unlike, the pycnidia-like structures formed by DZtKu70 or DZtSo-comp strains were 290

exuding a whitish liquid similar to the oozed cirrhus-containing pycnidiospores 291

observed in planta (Fig. 6B). 292

293

Therefore, we next postulated that DZtSo has a similar colonization pattern than the 294

wild-type strain, including accumulation of hyphae in the sub-stomatal cavity, but the 295

lack of cell fusion would obstruct the pycnidial development. We monitored the wheat 296

plants infected with 1E4GFP or 1E4GFPDZtSo strains using confocal microscopy up to 297

12 dpi. At the earlier evaluated stages of the plant infection, we observed mainly host 298

penetration, initial intercellular hyphal extension and sub-stomatal colonization (Fig. 299

S10 – 6 and 7 dpi). None difference was noticed in fungal development, however, 300

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 13: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

13

hyphal fusions established during epiphytic host colonization were only found for 301

1E4GFP strain. At 8 and 9 dpi, we observed that the first intercellular hyphae 302

surrounding the stomatal guard cells produced specialized knots from where 303

secondary hyphae emerge and germinate (Figs. 7 and S10). These secondary hyphae 304

fuse with other nearby hyphae, creating an interconnected network in the sub-stomatal 305

cavity (e.g. for the 1E4GFP strain) or it keeps extending as individual hypha (e.g. for the 306

1E4GFPDZtSo mutant) (Figs. 7 and S10). The combination of sub-stomatal hyphal 307

accumulation and anastomoses generates the mature pycnidium, which later supports 308

the asexual reproduction of Z. tritici. On the other hand, the lack of anastomosis stops 309

the development of the pycnidium and, consequently, impair the asexual cycle of the 310

fungus. Thus, we concluded that hyphal fusions are crucial for the pycnidial 311

development and the disturbance of this mechanism ceases the asexual reproduction 312

of Z. tritici. 313

314

Discussion 315

Cell-to-cell communication regulates a myriad of biological process that drive fungal 316

development and ecological diversifications. The sophisticated evolution in the fungal 317

language allowed a fine-tune coordination of signal senders and receivers and, 318

consequently, the regulation of complex signaling networks. Here, we explored the 319

functional relationship of a gene involved in cell-to-cell communication and its 320

biological contribution to the development and fitness of a filamentous fungus plant 321

pathogen. 322

323

Vegetative cell fusion is one of the most important cellular developmental processes 324

of a mycelial fungal colony (Glass et al., 2000, Simonin et al., 2012). The cytoplasmic 325

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 14: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

14

continuity generated by cell fusion provides adaptative advantages to the 326

interconnected mycelial network essentially for resource sharing and introgression of 327

genetic material (Simonin et al., 2012, Roca et al., 2005b, Mehrabi et al., 2011). CATs 328

forming at earlier stages of the vegetative growth have been documented for several 329

filamentous fungi (Roca et al., 2005a). We observed that these specialized fusion 330

bridges require a certain cell density, indicating that Z. tritici may induce CATs at a 331

critical concentration of a self-produced extracellular molecule, most likely a quorum-332

sensing molecule. A CAT inducer signal based on a quorum sensing was previously 333

proposed for N. crassa and Venturia inaequalis (Roca et al., 2005b, Read et al., 2012), 334

but the nature of this molecule and its receptor on the cell remain unknown. On the 335

other hand, the induction of VHF in Z. tritici does not seem to be associated with a 336

quorum-sensing because anastomoses were frequently observed at lower initial cell 337

concentration and thus, VHF may be induced by some other environmental signal. For 338

instance, VHFs were pronounced in nutrient-limited conditions in Z. tritici, as observed 339

for the causal agent of anthracnose disease, Colletotrichum lindemuthianum (Ishikawa 340

et al., 2010). Plant pathogens typically experience nutrient limitations while growing 341

on leaves and the perception of a nutrient-limited environment may act as a stimulus 342

to induce cell fusion in foliar plant pathogens. Both germinating blastospores or 343

pycnidiospores of Z. tritici undergo CAT and VHF in vitro and in planta, supporting our 344

previous observation (Francisco et al., 2019). Thus, we demonstrated that vegetative 345

cell fusions occur independently of the morphotype and it may have an important 346

contribution to the lifestyle of Z. tritici during the plant-pathogen interaction. 347

348

Unlike the majority of fungal species that form a multinuclear hyphal network (Roper 349

et al., 2011), Z. tritici has only one nucleus per septal compartment (Kilaru et al., 2017). 350

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 15: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

15

We used a GFP-tagged nucleus strain, to track the nuclei movements between two Z. 351

tritici fused partners. Nuclei transfer between two encountering hyphae has been 352

described for different fungal species (Chagnon, 2014, Mehrabi et al., 2011, Roper et 353

al., 2011). The consequences of genetic exchange include the formation of viable 354

heterokaryons (Roper et al., 2011) and the risk to introduce pathogenic elements or 355

virulence genes (Biella et al., 2002, Goddard & Burt, 1999, Friesen et al., 2006). This 356

raises the question of whether heterokaryon formation could happen in Z. tritici after 357

cell fusions. We showed that after the fusion bridge formation connecting the two 358

identical Z. tritici cells, the new cell compartment is occupied by a migrating nucleus 359

coming from a neighboring nucleus-divided by mitosis. Heterokaryon cells were never 360

observed neither near nor far from the anastomosis point, indicating that Z. tritici may 361

have evolved to limit the spreading of genetic elements and restricting the formation 362

of heterokaryons and the parasexual cycle. However, we do not discard the possibility 363

that exchanges of DNA fragments or pathogenic elements may affect the fitness and 364

evolution of this fungus. For instance, transposable elements (TEs) are extraordinary 365

generators of fungal diversity and versatility (Mat Razali et al., 2019, Castanera et al., 366

2016). These selfish elements, in particular, the DNA Transposon (or Class II TE), that 367

represents 14.6% of the repetitive DNA in Z. tritici (Dhillon et al., 2014) and uses the 368

transposase activity to “cut and paste” genome sequences, could mediate gene or 369

chromosome horizontal transfers. Notwithstanding, this mechanism was proposed for 370

the ToxA neighboring type II hAT-like transposase gene horizontally transferred 371

between three wheat pathogen-related species (McDonald et al., 2019). Likewise, the 372

mycoviruses existing within the cytoplasm can also be horizontally transmitted during 373

anastomosis (Ihrmark et al., 2002, Pearson et al., 2009). Z. tritici is known to host 374

double-stranded RNA (dsRNA) mycoviruses species (Zelikovitch et al., 1990, Kema 375

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 16: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

16

et al., 2008), including a homologous dsRNA hypovirus identified in Fusarium 376

graminearum and associated with altered fungal growth, pigmentation and reduced 377

virulence (Chu et al., 2002). Hypovirulence associated with dsRNA mycoviruses has 378

been reported for several plant pathogen fungi (Nuss, 2005), but the impact of these 379

obligate parasites on Z. tritici biology remain broadly unknown. Here, we propose that 380

cell fusions may generate non-conventional possibilities to contribute to the high 381

genetic diversity observed in Z. tritici isolates. Whether these genetic elements or 382

mycoviruses can be acquired by anastomosis and impact Z. tritici diversification will 383

be a subject of futures studies. 384

385

In the last decades, the identification of fusion defective mutants contributed to the 386

understanding of the molecular mechanisms underlying cell communication and 387

fusion, especially by unveiling the interplay of the mitogen-activated protein (MAP) 388

kinase pathways during chemotropic interaction (Fleissner et al., 2009). MAP kinase 389

pathways are involved in extracellular signal perception and regulation of diverse 390

genes essential for mating, filamentation, pathogenicity, cell integrity and stress 391

response (Deng et al., 2018, Zhao et al., 2007, Pandey et al., 2004, Maddi et al., 2012, 392

Leng & Zhong, 2015, Hagiwara et al., 2016) and thus, the deletion of MAP kinase-393

related genes result in pleiotropic phenotypes due to the interaction in multiple 394

biological processes. Though fungal communication and fusion require the regulation 395

of several genes (Fischer & Glass, 2019), the crosstalk between cell wall integrity 396

(CWI) and MAK-2 signal response, two conserved MAP kinase signaling pathways, 397

are essential to produce, secrete and sense the chemoattractant molecule produced 398

during cell fusion (Fleissner et al., 2009). The deletion of MAK-1 or MAK-2 genes 399

disrupts the signaling cascade affecting the self-anastomosis in filamentous fungi. We 400

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 17: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

17

used the ZtSlt2 (MAK-1) and ZtFus3 (MAK-2) Z. tritici orthologues to demonstrate the 401

functional conservation of these MAP kinase pathways in this fungus. Hitherto, ZtSlt2 402

and ZtFus3 were known as essential genes for the pathogenicity of Z. tritici, regulating 403

invasive growth and host penetration (Mehrabi et al., 2006, Cousin et al., 2006). Here, 404

we showed that both DZtSlt2 and DZtFus3 lost their self-stimulation and were unable 405

to undergo anastomosis. Our results indicate that cell fusions in Z. tritici follow the 406

ping-pong self-signaling mechanism described for N. crassa (Fleissner et al., 2009, 407

Read et al., 2009), where the signal sending and receiving is coordinated by genes-408

associated with the CWI and MAK-2 pathways. Besides, this is the first report of CWI 409

and MAK-2 pathways regulating cell communication in Z. tritici. 410

411

To evaluate the impact of vegetative cell fusion for the biology of a latent necrotrophic 412

fungus, we used ZtSo orthologous of N. crassa so. In contrast to the MAP kinase-413

related genes, the characterization of so orthologues results in less pleiotropic 414

phenotypes (Fleissner et al., 2005, Prados Rosales & Di Pietro, 2008, Craven et al., 415

2008). Filamentous fungus lacking so gene is impaired in self-anastomosis (Craven et 416

al., 2008, Prados Rosales & Di Pietro, 2008, Charlton et al., 2012, Fleissner et al., 417

2005), including Z. tritici, in which we showed that the deletion of ZtSo abolishes the 418

vegetative cell fusion of this fungus. We characterized the impact of this fusion defect 419

during different developmental stages of Z. tritici. For example, while the fusion 420

competent individuals have an advantageous growth on WA medium promoting larger 421

colonies with dense hyphal borders, the DZtSo mutant exhibited an asymmetrical 422

hyphal growth extension. It has been previously demonstrated that the direction of 423

nutrient distribution occurs mostly from the central part of mycelium to outwards and 424

its streaming speed is driven by the anastomosis (Simonin et al., 2012). Thus, the 425

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 18: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

18

repression of colony extension observed for the DZtSo mutant grown on a nutrient-426

limited environment may be a consequence of the irregular distribution of cytoplasmic 427

content, including this limited-food resource, throughout the mycelial colony. 428

Surprisingly, DZtSo underwent extensive production of blastospores, generating larger 429

yeast-like colonies than the fusion competent individuals in a nutrient-rich 430

environment. This result suggests that the increased growth rate observed for the 431

fusion-defective mutant may reflect the role of the ZtSo gene on the vegetative growth, 432

in this case, the boost of the blastosporulation mechanism, probably due to an 433

interplay of different signaling pathways. The genetic relationship between the 434

derepression of blastosporulation and the lack of ZtSo gene remains to be elucidated. 435

436

We showed that the DZtSo does not accumulate melanin, resulting in whiter and larger 437

colonies than those formed by DZtKu70 and DZtSo-comp strains. These findings 438

illustrate how melanization drove pathogen adaptation to a trade-off between energy 439

cost for pigment production and fungal growth. The deleterious effect on Z. tritici 440

growth caused by a higher accumulation of melanin was previously reported for this 441

fungus (Krishnan et al., 2018). Melanins are dark-pigmented secondary metabolites 442

often associated with the fungal cell walls. Though fungi can produce different kinds 443

of melanins, hitherto, it has been suggested that melanization of Z. tritici is controlled 444

only by the polyketide synthase (PKS) gene cluster containing catalytic enzymes and 445

transcription regulators of the 1,8-dihydroxynaphthalene (DHN) melanin 446

(Lendenmann et al., 2014, Krishnan et al., 2018). The CWI pathway, to which ZtSo 447

belongs, is the central signaling cascade regulating diverse biological processes, 448

including the production of secondary metabolites (Levin, 2005, Valiante, 2017, Park 449

et al., 2008). It has been demonstrated to plant-pathogenic fungi that the deletion of 450

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 19: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

19

CWI-associated genes inhibited pigmentation by reducing the expression of DHN 451

melanin biosynthetic genes (Yago et al., 2011, Liu et al., 2011, Valiante et al., 2015). 452

Thus, we speculate that the CWI signaling pathway regulates the PKS-encoding 453

genes leading to the melanin accumulation in Z. tritici. The deletion of ZtSo may impair 454

the CWI-regulatory cascade and, consequently, the regulation of DHN-melanin 455

production, resulting in the lack of pigmentation observed for both DZtSo and DZtSlt2 456

(Mehrabi et al., 2006) mutants from the CWI pathway. On the other hand, so is also 457

described for its putative function in the secretion of internal vesicles transporting the 458

chemoattractant molecule at the growing cell tip (Fleissner et al., 2005, Fleissner & 459

Herzog, 2016). Considering that different studies have demonstrated that fungal 460

melanin may be synthesized in internal vesicles and transported to the cell wall 461

(Rodrigues et al., 2008, Silva et al., 2014), we propose an alternative hypothesis to 462

explain the impaired melanization by DZtSo mutant. The disruption of ZtSo may affect 463

the transport of internal vesicles carrying the pigment to the cell wall or the laccase 464

oxidase enzyme required to polymerize 1,8-DHN to form the DHN-melanin polymer 465

(Plonka & Grabacka, 2006), resulting in a non-melanized strain. Both hypotheses 466

require further investigations. 467

468

Melanin has also been postulated to contribute to fungal protection against fungicide 469

and environmental stresses in Z. tritici (Krishnan et al., 2018, Lendenmann et al., 470

2014) and its regulation depends on environmental cues and colony development 471

(Lendenmann et al., 2014), though not yet completely understood. We used nine 472

different cellular stressors to evaluate whether (i) the defect in melanin accumulation 473

or (ii) the deletion of ZtSo, the scaffold protein for the MAP kinase genes from the CWI 474

pathway, would affect pathogen stress tolerance. Since the non-melanized DZtSo 475

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 20: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

20

mutant displayed the same degree of stress sensitivity than DZtKu70 and DZtSo-comp 476

strains, we concluded that ZtSo does not act as a scaffold protein for all CWI pathway 477

function. This observation was previously described for the model fungus Sordaria 478

macrospora, where the PRO40 (orthologous to ZtSo) operates as a scaffold for the 479

CWI-encoding genes during fungal development, hyphal fusion, and stress response, 480

but not for growth under cell-wall stress agents (Teichert et al., 2014). Our results 481

showed that melanin accumulation of Z. tritici does not promote an advantage of fungal 482

survival in harsh environments, at least for the stressful conditions tested in this study. 483

484

We demonstrated that VHFs are dispensable for the pathogenicity of Z. tritici. The 485

fusion-defective DZtSo mutant displayed a similar host damage progression than 486

those individuals possessing the gene. This finding exemplifies the distinct effects of 487

cell fusions on fungal pathogenicity. For instance, for the soil-borne Fusarium 488

oxysporum, VHF-impaired mutants exhibited only a slightly reduced virulence, 489

whereas, for the necrotrophic plant pathogen A. alternata, VHFs are necessary for the 490

full virulence of the fungus (Prados Rosales & Di Pietro, 2008, Craven et al., 2008). 491

Though the deletion of ZtSo is not essential for host penetration, colonization or for 492

the onset of the necrotrophic phase per se, it is during hyphal accumulation in the sub-493

stomatal cavity that the fusion defect impacts Z. tritici fitness. As far as we know, this 494

is the first time that pycnidial development has been microscopically detailed in a 495

filamentous fungus. We showed that the first intercellular hyphae surrounding the 496

stomatal guard cells produced specialized knots from where secondary hyphae 497

emerge and germinate to fuse with other adjacent hyphae. Consequently, this 498

preliminary hyphal network creates the basis for a symphogenous development that 499

builds the concave-shaped pycnidial wall of the mature pycnidium. On the other hand, 500

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 21: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

21

the inability to undergo anastomosis ceases the development of the asexual fruiting 501

bodies of the fungus, and therefore, abolish fungal reproduction. The accumulation of 502

hyphae observed in the sub-stomata chamber by the DZtSo mutant confirms that there 503

is a specific signal that triggers sub-stomatal hyphal aggregation, which is independent 504

of the chemoattractant molecule secreted by the fungus to induce hyphal fusion. 505

Regarding the sexual phase, we did not investigate the impact of the DZtSo for the 506

sexual cycle, because is not yet feasible to generate in vitro crosses for Z. tritici. 507

Fleissner et al. (2005) demonstrated that the deletion of so in N. crassa affects female 508

fertilization, which blocks the sexual reproduction of the mutant. Further experiments 509

need to be performed to address this question for Z. tritici, but we believe that the ZtSo 510

gene would also play a crucial role in the sexual reproduction of this pathogen. 511

512

In summary, the characterization of ZtSo gene demonstrated its fundamental role in 513

fungal biology. Beyond the impact of ZtSo for self-stimulation and self-fusion, we show 514

the contribution of this gene for fungal development, including a negative impact on 515

hyphal extension, derepressing of blastosporulation and impairment in melanization. 516

Besides, we demonstrated that VHFs are dispensable for pathogenicity, but essential 517

for pycnidial development. Taken together, our data illustrate how cell fusion affects 518

Z. tritici fitness and provide a powerful new gene target to control Septoria tritici blotch 519

(STB) disease. 520

521

Experimental Procedures 522

523

Strains and growth conditions 524

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 22: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

22

The Swiss Z. tritici strain ST99CH_1E4 (abbreviated as 1E4), described by Zhan et 525

al. (2002), and mutant lines derived from this strain, as well as the 1E4 strain 526

expressing cytoplasmic GFP (1E4GFP) or mCherry (1E4mCh), were used in this study. 527

The knocked-out DZtSlt2 (Mehrabi et al., 2006) and DZtFus3 (Cousin et al., 2006) 528

mutants were provided by Marc-Henri Lebrun (National Institute of Agricultural 529

Research – INRA, France).Because the MAP kinase mutants were generated in the 530

genetic background of IPO323 (Kema & van Silfhout, 1997), this strain was also used. 531

Routinely, Z. tritici was cultivated on yeast-sucrose broth (YSB) medium (10 g/L yeast 532

extract, 10 g/L sucrose, 50 µg/mL kanamycin sulfate; pH 6.8). Each strain was stored 533

in glycerol at -80°C until required and then recovered in YSB medium incubated at 534

18°C for four days. 535

536

Plant infection to obtain fluorescent pycnidiospores 537

Wheat seedlings from the susceptible wheat cultivar Drifter were grown for 16 days in 538

the greenhouse at 18°C (day) and 15°C (night) with a 16h photoperiod and 70% 539

humidity. Blastospore suspensions of 1E4GFP or 1E4mCh were obtained after four days 540

of growth in YSB medium. Spore suspensions were adjusted to a final concentration 541

of 106 blastospores/mL in 30 mL of sterile water supplemented with 0.1% (v/v) Tween 542

and applied to run-off using a sprayer and the plants were kept for three days in sealed 543

plastic bags, followed by 21 days in a greenhouse. Leaves with pycnidia were 544

harvested and transferred to a 50 ml Falcon tube containing sterile water and gently 545

agitated to harvest the pycnidiospores. Fluorescent pycnidiospores were used to 546

assess the cell fusion events during in vitro and in vivo growth. Pycnidiospore 547

suspension-tagged GFP or mCherry were also adjusted to a final concentration of 106 548

pycnidiospores/mL and a new batch of plants were inoculated as described above. 549

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 23: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

23

Plants co-infected by both 1E4GFP and 1E4mCh strains were used to observe VHFs on 550

the wheat leaf surface. 551

552

Characterization of cell fusion events in vitro and in vivo 553

The ability of Z. tritici to undergo cell fusions was evaluated using blastospores and 554

pycnidiospores of 1E4GFP and 1E4mCh. Cell concentrations were adjusted to 3.3x107 555

blastospores/mL or 3.3x106 blastospores/mL to induce CATs or VHFs, respectively. 556

300 µL of each morphotype and fluorescence was plated on water agar (WA) to create 557

a ratio of 1:1 and to provide a final concentration of 107 blastospores/mL or 106 558

blastospores/mL. A section of about 1 cm2 of agar was aseptically cut and placed on 559

a microscope slide. The mixing of both cytoplasms content through CATs or VHFs 560

was checked up to 40 hai using a Leica DM2500 fluorescent microscope with LAS 561

v.4.6.0 software. GFP excitation and emission was at 480/40 nm and 527/30 nm, 562

respectively, whereas, mCherry was excited at 580/20 nm and detected at 632/60 nm. 563

564

VHFs during spore germination on wheat leaf surface were obtained via confocal 565

images using a Zeiss LSM 780 inverted laser-scanning microscope with ZEN Black 566

2012 software. An argon laser at 500 nm was used to excite GFP fluorescence and 567

chloroplast autofluorescence, while mCherry excitation was at 588 nm. The emission 568

wavelength was 490-535, 624-682, and 590-610 nm for GFP, chloroplast 569

autofluorescence, and mCherry, respectively. Plants co-inoculated with blastospores 570

or pycnidiospores of 1E4GFP and 1E4mCh strains were checked up to 48 hai. 571

572

Genetic exchange during anastomosis 573

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 24: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

24

To determine whether nuclei exchange happens during VHFs in Z. tritici, we used the 574

IPO323 ZtHis1-ZtGFP strain (Kilaru et al., 2017), which has the GFP as a fluorescent 575

marker labeling the nucleus. Blastospores of the IPO323 ZtHis1-ZtGFP strain were 576

plated on WA plates at a final concentration of 106 blastospores/mL. Spores were 577

monitored up to 72 hai by light and fluorescent microscope. 578

579

Orthologues identification, protein alignment and phylogenetic analysis 580

The so gene sequence from Neurospora crassa (XM_958983.3) was blasted against 581

the Z. tritici genome (https://genome.jgi.doe.gov/Mycgr3/Mycgr3.home.html) to 582

identify its orthologous in this fungus. The Z. tritici So orthologous protein sequence 583

was used for a Blastp analysis against the NCBI database (National Center of 584

Biotechnology Information). Blastp searches at expected value homology cut-off of 1e-585

10 were included as positive. A dataset containing So orthologues proteins of different 586

Ascomycete species were used for phylogenetic analysis. Three members of 587

Basidiomycetes were used as outgroup. Protein sequences were aligned using 588

AliView program (Larsson, 2014). The best-fit model of amino acid evolution was the 589

LG+G, determined by Mega6 software (Tamura et al., 2013). Amino acid sequences 590

were aligned using Muscle followed by maximum likelihood phylogeny reconstruction 591

using 1,000 bootstraps and performed with the software Mega6 (Tamura et al., 2013). 592

593

Plasmid constructions and transformations 594

All PCRs for cloning procedures were performed using NEB Phusion polymerase 595

(New England Biolabs). Primers used for cloning, sequencing and knock-out 596

confirmations are listed in Supporting Information Table S1. DNA assemblies were 597

conducted with the In-Fusion HD Cloning Kit (Takara BIO) following the 598

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 25: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

25

manufacturer’s instructions. To increase the homologous recombination efficiency, we 599

first inactivated the ZtKu70 (Mycgr3G85040 or Zt09_3_00215) gene in the1E4 strain 600

using the plasmid pGEN-YR-DZtKu70 (Sidhu et al., 2015), containing a geneticin 601

resistance gene cassette (also known as G418), as a selectable marker. To disrupt 602

the Z. tritici so gene (ZtSo), 1 Kb size of both flanking regions were amplified from the 603

1E4 genomic DNA. The hygromycin resistance gene cassette (hph), used as a 604

selective marker, was amplified from pES6 plasmid (obtained from E. H. Stukenbrock, 605

Kiel University, unpublished). The pES1 plasmid (obtained from E. H. Stukenbrock, 606

Kiel University, unpublished) was digested with KpnI and SbfI for plasmid linearization, 607

and three fragments were assembled, which resulted in the pES1-DZtSo. To 608

reintroduce the ZtSo gene into the DZtSo mutant strain, we used the plasmid pES1. 609

We amplified the nourseothricin resistance gene cassette (nat) from pES43 plasmid 610

(obtained from E. H. Stukenbrock, Kiel University, unpublished) to be used as a 611

selectable marker. ZtSo gene containing 1 Kb size of each flank region and the nat 612

resistance gene cassette was assembled into pES1, resulting in pES1-DZtSo-comp 613

that allowed to introduce the ZtSo gene into its native location (Fig. S11). We also 614

used the pES1-DZtSo plasmid to knock-out the ZtSo gene in the 1E4GFP genome 615

background, enabling the visualization of the GFP-tagged mutant during host 616

infection. 617

618

Plasmids were transformed into E. coli NEB 5-alpha by heat shock transformation for 619

plasmid propagations, followed by plasmid miniprep using QIAprep Spin Miniprep 620

(Qiagen) according to manufacturer’s instructions. Successful plasmid constructions 621

were confirmed by Sanger sequencing before been transformed into Agrobacterium 622

tumefaciens strain AGL1 cells by electroporation. 623

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 26: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

26

624

Z. tritici 1E4 strain was transformed by Agrobacterium tumefaciens-mediated 625

transformation (ATMT) according to Meile et al. (2018). The knock-out of the target 626

genes was verified by a PCR-based approach using a forward primer specific to the 627

upstream sequence of the disrupted gene and a reverse primer specific to bind in the 628

resistance cassette (Table S1). We determined the copy number of the transgene by 629

quantitative PCR (qPCR) on genomic DNA extracted with the DNeasy Plant Mini Kit 630

(Qiagen). We used as qPCR target gene the selection marker and the TFIIIC1 or 18S, 631

as reference genes (Table S1). Lines with a single insertion were selected for further 632

experiments. 633

634

Phenotypic characterizations 635

For all phenotypic analyses, DZtKu70 was considered the wild-type (WT) strain. To 636

pinpoint the role of the ZtSo gene on the vegetative cell fusion, blastospore suspension 637

of DZtKu70, DZtSo, and DZtSo-comp was added to a final concentration of 106 638

blastospores/mL or 107 blastospores/mL into WA and incubated at 18°C to induce 639

CATs or VHFs, respectively. For the MAP kinase DSlt2 and DFus3, IPO323 and DSlt2-640

complemented strains, WA plates were inoculated only at a final concentration of 106 641

blastospores/mL. Cell fusion events were monitored up to 40 hai by light microscopy. 642

Because fusion bridges were not observed between individuals lacking ZtSo, we 643

mixed 150 µL of 106 blastospores/mL of DZtKu70, DZtSo, or DZtSo-comp strains with 644

the same concentration of 1E4GFP blastospores in a ratio of 1:1 to confirm the failure 645

of cytoplasm streaming. At least 50 spores of each sample combination were 646

monitored by light and fluorescent microscopy.. 647

648

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 27: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

27

To test for altered fungal growth, we used PDA (39 g/L potato dextrose agar, 50 µg/mL 649

kanamycin sulfate) and WA media to induce blastospore and hyphal growth, 650

respectively. 200 µL of spore suspension of DZtKu70, DZtSo, and DZtSo-comp were 651

plated at a final concentration of 2x102 blastospores/mL on each aforementioned 652

media and incubated at 18°C. At least 40 colonies formed in five independent PDA 653

plates were photographed from the bottom using a standardized camera setting 654

(Lendenmann et al., 2014) at 8, 9, 10, 11, 13, and 15 days post-incubation (dpi). Digital 655

images were processed using a macro developed in the ImageJ software (Schneider 656

et al., 2012), which scores the area of individual colonies in the images. Fungal growth 657

was obtained by converting the colony area into radial growth (millimeter (mm)) based 658

on the formula 𝒓 = #𝑨𝝅 , for each strain. Radial growth values were plotted in a boxplot 659

graphic using the ggplot2 package from R (Wickham, 2009). Analysis of variance 660

(ANOVA) was performed to determine the differences in fungal growth among the 661

strains using the agricolae package in R (Mendiburu, 2015). The radial growth rate 662

(mm/day) for each strain was measured by plotting the colony radius over time, which 663

fitted to a linear model (Pearson’s correlation coefficient value (r2 ≥ 0.98)). Relative 664

growth rate was calculated by dividing the slope of the regression line of DZtSo by the 665

slope of DZtKu70 or DZtSo-comp strains. 666

667

Because mycelial growth on the WA plate exhibits a poor color contrast, it was not 668

possible to use the macro to estimate mycelial area. Thus, we manually calculated the 669

mycelial diameter from digital images of at least 40 colonies formed in five independent 670

Petri dishes at 15 dpi and using ImageJ software (Schneider et al., 2012). The mycelial 671

diameter values were divided by two to generate the radial growth (mm) values. 672

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 28: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

28

673

The degree of melanization of each tested Z. tritici strain was estimated from at least 674

40 colonies formed on PDA plates. We used a macro developed in the ImageJ 675

(Schneider et al., 2012), which scores the mean gray value of each individual colony. 676

Gray values range from 0 to 255, with 0 representing black and 255 representing 677

white. The mean gray values of each strain over time were plotted in a boxplot using 678

ggplot2 package from R (Wickham, 2009). 679

680

The impact on the cell integrity caused by the deletion of ZtSo was verified by exposing 681

blastospores of DZtKu70, DZtSo, and DZtSo-comp to nine different stress conditions, 682

including different temperatures (18°C and 27°C), oxidative stress (0.5, and 1 mM of 683

hydrogen peroxide – H2O2), osmotic stress (1M sodium chloride - NaCl and 1M 684

sorbitol), cell wall stress (2 mg/mL congo red and 10 µg/mL calcofluor white - CFW), 685

and plasma membrane stress (0.01% sodium dodecyl sulfate – SDS). Spore 686

suspensions of each strain were serial diluted to 4x104, 4x105, 4x106, and 4x107 687

blastospores/mL and drops of 3.5 µL were plated on five independent PDA plates 688

containing the mentioned stresses and incubated at 18°C. Colony phenotypes were 689

assessed by digital images taken at 5 dpi. 690

691

Virulence assay and pycnidia formation in vitro 692

We used five different winter cultivars of wheat (Triticum aestivum L.) based on their 693

susceptibility or resistance to Z. tritici, as described in the Swiss granum website 694

(https://www.swissgranum.ch/documents/741931/1152834/LES_Winterweizen_2020695

.pdf/e624760c-8329-e11d-7e53-afbec9146156). Drifter, Claro and Runal are 696

classified as susceptible cultivars. Arina is considered intermediate and Titlis and 697

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 29: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

29

Camedo are described as resistant cultivars to Z. tritici. Seeding, greenhouse and 698

plant growth conditions, inoculum preparation and plant inoculation followed the same 699

procedures described by Meile et al. (2018). To estimate the percentage of leaf-700

covered by lesions (PLACL) and pycnidia formation, we harvested the second leaves 701

of Drifter plants inoculated with DZtKu70, DZtSo or DZtSo-comp strains at 8, 10, 11, 702

12, 14, 16, and 21 dpi. For the rest of the cultivars, infected leaves were harvested 703

only at 14 and 21 dpi to check for pycnidia formation. Leaves were mounted on a 704

paper sheet, scanned with a flatbed scanner, and analyzed using automated image 705

analysis (Stewart et al., 2016). Data analysis and plotting were performed using 706

ggplot2 package (Wickham, 2009). 707

708

The defect in pycnidia formation was confirmed by plating blastospores of DZtKu70, 709

DZtSo and DZtSo-comp strains onto wheat extract agar medium (50 g/L blended 21 710

days-old wheat leaves cv. Drifter, 10 g/L agar) and incubated under UV-A light (16:8 711

light:dark cycle) up to 40 days at 18°C. 712

713

Confocal laser-scanning microscopy of infected wheat leaves 714

To assess the impact of ZtSo deletion on fungal fitness during host colonization, we 715

inoculated wheat plants with 1E4GFP and other two independent GFP-tagged DZtSo 716

mutants. Infected leaves were harvested at 6, 7, 8, 9, 10, 11, and 12 dpi and checked 717

for developmental stages of asexual fruiting bodies. Microscopy was conducted using 718

Zeiss LSM 780 inverted laser-scanning microscope with ZEN Black 2012 software. An 719

argon laser at 500 nm was used to exited GFP fluorescence and chloroplast 720

autofluorescence with an emission wavelength of 490-535 nm and 624-682 nm, 721

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 30: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

30

respectively. Analyses, visualization, and processing of image z-stacks were 722

performed using ImageJ software (Schneider et al., 2012). 723

724

ACKNOWNLEDGEMENTS 725

This study was financed in part by the Coordenação de Aperfeiçoamento de Pessoal 726

de Nível Superior – Brasil (CAPES) – Finance Code 001. We acknowledge Marc-Henri 727

Lebrun from the National Institute of Agricultural Research (INRA) in France for kindly 728

providing the MAP kinase mutants used in this study. We thank Dominik Vetsch for 729

his help during the plant assays using different wheat cultivars. 730

731

REFERENCES 732

Aldabbous, M. S., Roca, M. G., Stout, A., Huang, I. C., Read, N. D. and Free, S. J. (2010) The ham-5, rcm-1 and 733 rco-1 genes regulate hyphal fusion in Neurospora crassa. Microbiology, 156, 2621-2629. 734

Biella, S., Smith, M. L., Aist, J. R., Cortesi, P. and Milgroom, M. G. (2002) Programmed cell death correlates 735 with virus transmission in a filamentous fungus. Proc Biol Sci, 269, 2269-2276. 736

Bloemendal, S. and Kuck, U. (2013) Cell-to-cell communication in plants, animals, and fungi: a comparative 737 review. Naturwissenschaften, 100, 3-19. 738

Bork, P. and Sudol, M. (1994) The WW domain - a signalling site in dystrophin? Trends in Biochemical Science, 739 19, 531-533. 740

Castanera, R., Lopez-Varas, L., Borgognone, A., LaButti, K., Lapidus, A., Schmutz, J., et al. (2016) Transposable 741 Elements versus the Fungal Genome: Impact on Whole-Genome Architecture and Transcriptional 742 Profiles. PLoS Genet, 12, e1006108. 743

Chagnon, P. L. (2014) Ecological and evolutionary implications of hyphal anastomosis in arbuscular mycorrhizal 744 fungi. FEMS Microbiol Ecol, 88, 437-444. 745

Charlton, N. D., Shoji, J. Y., Ghimire, S. R., Nakashima, J. and Craven, K. D. (2012) Deletion of the fungal gene 746 soft disrupts mutualistic symbiosis between the grass endophyte Epichloe festucae and the host plant. 747 Eukaryot Cell, 11, 1463-1471. 748

Chu, Y. M., Jeon, J. J., Yea, S. J., Kim, Y. H., Yun, S. H., Lee, Y. W., et al. (2002) Double-stranded RNA 749 mycovirus from Fusarium graminearum. Appl Environ Microbiol, 68, 2529-2534. 750

Cottier, F. and Muhlschlegel, F. A. (2012) Communication in fungi. Int J Microbiol, 2012, 351832. 751 Cousin, A., Mehrabi, R., Guilleroux, M., Dufresne, M., T, V. D. L., Waalwijk, C., et al. (2006) The MAP kinase-752

encoding gene MgFus3 of the non-appressorium phytopathogen Mycosphaerella graminicola is required 753 for penetration and in vitro pycnidia formation. Mol Plant Pathol, 7, 269-278. 754

Craven, K. D., Velez, H., Cho, Y., Lawrence, C. B. and Mitchell, T. K. (2008) Anastomosis is required for 755 virulence of the fungal necrotroph Alternaria brassicicola. Eukaryot Cell, 7, 675-683. 756

Dean, R., Van Kan, J. A., Pretorius, Z. A., Hammond-Kosack, K. E., Di Pietro, A., Spanu, P. D., et al. (2012) The 757 Top 10 fungal pathogens in molecular plant pathology. Mol Plant Pathol, 13, 414-430. 758

Deng, Y. Z., Zhang, B., Chang, C., Wang, Y., Lu, S., Sun, S., et al. (2018) The MAP Kinase SsKpp2 Is Required 759 for Mating/Filamentation in Sporisorium scitamineum. Front Microbiol, 9, 2555. 760

Dhillon, B., Gill, N., Hamelin, R. C. and Goodwin, S. B. (2014) The landscape of transposable elements in the 761 finished genome of the fungal wheat pathogen Mycosphaerella graminicola. BMC Genomics, 15, 1132. 762

Endler, J. A. (1993) Some general comments on the evolution and design of animal communication systems. 763 Philosophical Transactions of the Royal Society B, 340, 215-225. 764

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 31: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

31

Engh, I., Wurtz, C., Witzel-Schlomp, K., Zhang, H. Y., Hoff, B., Nowrousian, M., et al. (2007) The WW domain 765 protein PRO40 is required for fungal fertility and associates with Woronin bodies. Eukaryot Cell, 6, 831-766 843. 767

Fischer, M. S. and Glass, N. L. (2019) Communicate and Fuse: How Filamentous Fungi Establish and Maintain 768 an Interconnected Mycelial Network. Front Microbiol, 10, 619. 769

Fleissner, A. and Herzog, S. (2016) Signal exchange and integration during self-fusion in filamentous fungi. Semin 770 Cell Dev Biol, 57, 76-83. 771

Fleissner, A., Leeder, A. C., Roca, M. G., Read, N. D. and Glass, N. L. (2009) Oscillatory recruitment of signaling 772 proteins to cell tips promotes coordinated behavior during cell fusion. Proc Natl Acad Sci U S A, 106, 773 19387-19392. 774

Fleissner, A., Sarkar, S., Jacobson, D. J., Roca, M. G., Read, N. D. and Glass, N. L. (2005) The so locus is required 775 for vegetative cell fusion and postfertilization events in Neurospora crassa. Eukaryot Cell, 4, 920-930. 776

Fones, H. and Gurr, S. (2015) The impact of Septoria tritici Blotch disease on wheat: An EU perspective. Fungal 777 Genet Biol, 79, 3-7. 778

Francisco, C. S., Ma, X., Zwyssig, M. M., McDonald, B. A. and Palma-Guerrero, J. (2019) Morphological 779 changes in response to environmental stresses in the fungal plant pathogen Zymoseptoria tritici. Sci Rep, 780 9, 9642. 781

Friesen, T. L., Stukenbrock, E. H., Liu, Z., Meinhardt, S., Ling, H., Faris, J. D., et al. (2006) Emergence of a new 782 disease as a result of interspecific virulence gene transfer. Nat Genet, 38, 953-956. 783

Fu, C., Ao, J., Dettmann, A., Seiler, S. and Free, S. J. (2014) Characterization of the Neurospora crassa cell fusion 784 proteins, HAM-6, HAM-7, HAM-8, HAM-9, HAM-10, AMPH-1 and WHI-2. PLoS One, 9, e107773. 785

Fu, C., Iyer, P., Herkal, A., Abdullah, J., Stout, A. and Free, S. J. (2011) Identification and characterization of 786 genes required for cell-to-cell fusion in Neurospora crassa. Eukaryot Cell, 10, 1100-1109. 787

Gillam, E. (2011) An introduction to animal communication. Nature Education Knowledge 3, 70. 788 Glass, N. L., Jacobson, D. J. and Shiu, P. K. T. (2000) The genetics of hyphal fusion and vegetative incompatibility 789

in filamentous ascomycete fungi. Annual Review of Genetics, 34, 165-186. 790 Goddard, M. R. and Burt, A. (1999) Recurrent invasion and extinction of a selfish gene. PNAS, 96, 13880-13885. 791 Gohari, A. M., Mehrabi, R., Robert, O., Ince, I. A., Boeren, S., Schuster, M., et al. (2014) Molecular 792

characterization and functional analyses of ZtWor1, a transcriptional regulator of the fungal wheat 793 pathogen Zymoseptoria tritici. Mol Plant Pathol, 15, 394-405. 794

Hagiwara, D., Sakamoto, K., Abe, K. and Gomi, K. (2016) Signaling pathways for stress responses and adaptation 795 in Aspergillus species: stress biology in the post-genomic era. Biosci Biotechnol Biochem, 80, 1667-796 1680. 797

He, C., Rusu, A. G., Poplawski, A. M., Irwin, J. A. G. and Manners, J. M. (1998) Transfer of a Supernumerary 798 Chromosome Between Vegetatively Incompatible Biotypes of the Fungus Colletotrichum 799 gloeosporioides. Genetics Society of America, 150, 1459-1466. 800

Hickey, P. C., Jacobson, D. J., Read, N. D. and Glass, N. L. (2002) Live-cell imaging of vegetative hyphal fusion 801 in Neurospora crassa. Fungal Genetics and Biology, 37, 109-119. 802

Ihrmark, K., Johannesson, H., Stenström, E. and Stenlid, J. (2002) Transmission of double-stranded RNA in 803 Heterobasidion annosum. Fungal Genetics and Biology, 36, 147-154. 804

Ishikawa, F. H., Souza, E. A., Read, N. D. and Roca, M. G. (2010) Live-cell imaging of conidial fusion in the 805 bean pathogen, Colletotrichum lindemuthianum. Fungal Biol, 114, 2-9. 806

Kema, G. H. and van Silfhout, C. H. (1997) Genetic Variation for Virulence and Resistance in the Wheat-807 Mycosphaerella graminicola Pathosystem III. Comparative Seedling and Adult Plant Experiments. 808 Phytopathology, 87, 266-272. 809

Kema, G. H. J., van der Lee, T. A. J., Mendes, O., Verstappen, E. C. P., Lankhorst, R. K., Sandbrink, H., et al. 810 (2008) Large-Scale Gene Discovery in the Septoria Tritici Blotch Fungus Mycosphaerella graminicola 811 with a Focus on In Planta Expression. Molecular Plant-Microbe Interactions, 21, 1249-1260. 812

Khang, C. H., Berruyer, R., Giraldo, M. C., Kankanala, P., Park, S. Y., Czymmek, K., et al. (2010) Translocation 813 of Magnaporthe oryzae effectors into rice cells and their subsequent cell-to-cell movement. Plant Cell, 814 22, 1388-1403. 815

Kilaru, S., Schuster, M., Ma, W. and Steinberg, G. (2017) Fluorescent markers of various organelles in the wheat 816 pathogen Zymoseptoria tritici. Fungal Genet Biol, 105, 16-27. 817

Krishnan, P., Meile, L., Plissonneau, C., Ma, X., Hartmann, F. E., Croll, D., et al. (2018) Transposable element 818 insertions shape gene regulation and melanin production in a fungal pathogen of wheat. BMC Biol, 16, 819 78. 820

Larsson, A. (2014) AliView: a fast and lightweight alignment viewer and editor for large datasets. Bioinformatics, 821 30, 3276-3278. 822

Lendenmann, M. H., Croll, D., Stewart, E. L. and McDonald, B. A. (2014) Quantitative trait locus mapping of 823 melanization in the plant pathogenic fungus Zymoseptoria tritici. G3 (Bethesda), 4, 2519-2533. 824

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 32: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

32

Leng, Y. and Zhong, S. (2015) The Role of Mitogen-Activated Protein (MAP) Kinase Signaling Components in 825 the Fungal Development, Stress Response and Virulence of the Fungal Cereal Pathogen Bipolaris 826 sorokiniana. PLoS One, 10, e0128291. 827

Levin, D. E. (2005) Cell wall integrity signaling in Saccharomyces cerevisiae. Microbiol Mol Biol Rev, 69, 262-828 291. 829

Liu, W., Soulie, M. C., Perrino, C. and Fillinger, S. (2011) The osmosensing signal transduction pathway from 830 Botrytis cinerea regulates cell wall integrity and MAP kinase pathways control melanin biosynthesis 831 with influence of light. Fungal Genet Biol, 48, 377-387. 832

Maddi, A., Dettman, A., Fu, C., Seiler, S. and Free, S. J. (2012) WSC-1 and HAM-7 are MAK-1 MAP kinase 833 pathway sensors required for cell wall integrity and hyphal fusion in Neurospora crassa. PLoS One, 7, 834 e42374. 835

Maruyama, J., Escano, C. S. and Kitamoto, K. (2010) AoSO protein accumulates at the septal pore in response to 836 various stresses in the filamentous fungus Aspergillus oryzae. Biochem Biophys Res Commun, 391, 868-837 873. 838

Mat Razali, N., Cheah, B. H. and Nadarajah, K. (2019) Transposable Elements Adaptive Role in Genome 839 Plasticity, Pathogenicity and Evolution in Fungal Phytopathogens. Int J Mol Sci, 20. 840

McDonald, M. C., Taranto, A. P., Hill, E., Schwessinger, B., Liu, Z., Simpfendorfer, S., et al. (2019) Transposon-841 Mediated Horizontal Transfer of the Host-Specific Virulence Protein ToxA between Three Fungal Wheat 842 Pathogens. MBio, 10. 843

Mehrabi, R., Bahkali, A. H., Abd-Elsalam, K. A., Moslem, M., Ben M'barek, S., Gohari, A. M., et al. (2011) 844 Horizontal gene and chromosome transfer in plant pathogenic fungi affecting host range. FEMS 845 Microbiol Rev, 35, 542-554. 846

Mehrabi, R., Ben M'Barek, S., van der Lee, T. A., Waalwijk, C., de Wit, P. J. and Kema, G. H. (2009) G(alpha) 847 and Gbeta proteins regulate the cyclic AMP pathway that is required for development and pathogenicity 848 of the phytopathogen Mycosphaerella graminicola. Eukaryot Cell, 8, 1001-1013. 849

Mehrabi, R., van der Lee, T., Waalwijk, C. and Kema, G. H. (2006) MgSlt2, a cellular integrity MAP kinase gene 850 of the fungal wheat pathogen Mycosphaerella graminicola, is dispensable for penetration but essential 851 for invasive growth. Molecular Plant-Microbe Interactions, 19, 389-398. 852

Meile, L., Croll, D., Brunner, P. C., Plissonneau, C., Hartmann, F. E., McDonald, B. A., et al. (2018) A fungal 853 avirulence factor encoded in a highly plastic genomic region triggers partial resistance to septoria tritici 854 blotch. New Phytol. 855

Mendiburu, F. D. (2015) Agricolae: Statistical Procedures for Agricultural Research. R Package Version 1.2-3. 856 Nuss, D. L. (2005) Hypovirulence: mycoviruses at the fungal-plant interface. Nat Rev Microbiol, 3, 632-642. 857 Ord, T. J. and Garcia-Porta, J. (2012) Is sociality required for the evolution of communicative complexity? 858

Evidence weighed against alternative hypotheses in diverse taxonomic groups. Philos Trans R Soc Lond 859 B Biol Sci, 367, 1811-1828. 860

Pandey, A., Roca, M. G., Read, N. D. and Glass, N. L. (2004) Role of a mitogen-activated protein kinase pathway 861 during conidial germination and hyphal fusion in Neurospora crassa. Eukaryot Cell, 3, 348-358. 862

Park, G., Pan, S. and Borkovich, K. A. (2008) Mitogen-activated protein kinase cascade required for regulation 863 of development and secondary metabolism in Neurospora crassa. Eukaryot Cell, 7, 2113-2122. 864

Pearson, M. N., Beever, R. E., Boine, B. and Arthur, K. (2009) Mycoviruses of filamentous fungi and their 865 relevance to plant pathology. Mol Plant Pathol, 10, 115-128. 866

Plonka, P. M. and Grabacka, M. (2006) Melanin synthesis in microorganisms — biotechnological and medical 867 aspects. Acta Biochimica Polonica, 53, 429-443. 868

Prados Rosales, R. C. and Di Pietro, A. (2008) Vegetative Hyphal Fusion Is Not Essential for Plant Infection by 869 Fusarium oxysporum. Eukaryotic Cell, 7, 162. 870

Read, N. D., Fleissner, A., Roca, M. G. and Glass, N. L. (2010) Hyphal Fusion. In: Cellular abd Molecular 871 Biology of Filamentous Fungi. (Borkovich, K. A. and Edolle, D., eds.). Washington DC: American 872 Society of Microbiology, pp. 260-273. 873

Read, N. D., Goryachev, A. B. and Lichius, A. (2012) The mechanistic basis of self-fusion between conidial 874 anastomosis tubes during fungal colony initiation. Fungal Biology Reviews, 26, 1-11. 875

Read, N. D., Lichius, A., Shoji, J. Y. and Goryachev, A. B. (2009) Self-signalling and self-fusion in filamentous 876 fungi. Curr Opin Microbiol, 12, 608-615. 877

Roca, G. M., Read, N. D. and Wheals, A. E. (2005a) Conidial anastomosis tubes in filamentous fungi. FEMS 878 Microbiol Lett, 249, 191-198. 879

Roca, M. G., Arlt, J., Jeffree, C. E. and Read, N. D. (2005b) Cell biology of conidial anastomosis tubes in 880 Neurospora crassa. Eukaryot Cell, 4, 911-919. 881

Roca, M. G., Davide, L. C., Davide, L. M., Mendes-Costa, M. C., Schwan, R. F. and Wheals, A. E. (2004) Conidial 882 anastomosis fusion between Colletotrichum species. Mycol Res, 108, 1320-1326. 883

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 33: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

33

Rodrigues, M. L., Nakayasu, E. S., Oliveira, D. L., Nimrichter, L., Nosanchuk, J. D., Almeida, I. C., et al. (2008) 884 Extracellular vesicles produced by Cryptococcus neoformans contain protein components associated 885 with virulence. Eukaryot Cell, 7, 58-67. 886

Roper, M., Ellison, C., Taylor, J. W. and Glass, N. L. (2011) Nuclear and genome dynamics in multinucleate 887 ascomycete fungi. Curr Biol, 21, R786-793. 888

Saupe, S. J. (2000) Molecular Genetics of Heterokaryon Incompatibility in Filamentous Ascomycetes. 889 Microbiology and Molecular Biology Reviews, 64, 489-502. 890

Schneider, C. A., Rasband, W. S. and Eliceiri, K. W. (2012) NIH Image to ImageJ: 25 years of image analysis. 891 Nature Methods, 9, 671. 892

Shrout, J. D., Tolker-Nielsen, T., Givskov, M. and Parsek, M. R. (2011) The contribution of cell-cell signaling 893 and motility to bacterial biofilm formation. MRS Bull, 36, 367-373. 894

Sidhu, Y. S., Cairns, T. C., Chaudhari, Y. K., Usher, J., Talbot, N. J., Studholme, D. J., et al. (2015) Exploitation 895 of sulfonylurea resistance marker and non-homologous end joining mutants for functional analysis in 896 Zymoseptoria tritici. Fungal Genet Biol, 79, 102-109. 897

Silva, B. M., Prados-Rosales, R., Espadas-Moreno, J., Wolf, J. M., Luque-Garcia, J. L., Goncalves, T., et al. 898 (2014) Characterization of Alternaria infectoria extracellular vesicles. Med Mycol, 52, 202-210. 899

Simonin, A., Palma-Guerrero, J., Fricker, M. and Glass, N. L. (2012) Physiological significance of network 900 organization in fungi. Eukaryot Cell, 11, 1345-1352. 901

Stewart, E. L., Hagerty, C. H., Mikaberidze, A., Mundt, C. C., Zhong, Z. and McDonald, B. A. (2016) An 902 Improved Method for Measuring Quantitative Resistance to the Wheat Pathogen Zymoseptoria tritici 903 Using High-Throughput Automated Image Analysis. Phytopathology, 106, 782-788. 904

Tamura, K., Stecher, G., Peterson, D., Filipski, A. and Kumar, S. (2013) MEGA6: Molecular Evolutionary 905 Genetics Analysis version 6.0. Mol Biol Evol, 30, 2725-2729. 906

Teichert, I., Steffens, E. K., Schnass, N., Franzel, B., Krisp, C., Wolters, D. A., et al. (2014) PRO40 is a scaffold 907 protein of the cell wall integrity pathway, linking the MAP kinase module to the upstream activator 908 protein kinase C. PLoS Genet, 10, e1004582. 909

Temporini, E. D. and VanEtten, H. D. (2004) An analysis of the phylogenetic distribution of the pea pathogenicity 910 genes of Nectria haematococca MPVI supports the hypothesis of their origin by horizontal transfer and 911 uncovers a potentially new pathogen of garden pea: Neocosmospora boniensis. Curr Genet, 46, 29-36. 912

Valiante, V. (2017) The Cell Wall Integrity Signaling Pathway and Its Involvement in Secondary Metabolite 913 Production. J Fungi (Basel), 3. 914

Valiante, V., Macheleidt, J., Foge, M. and Brakhage, A. A. (2015) The Aspergillus fumigatus cell wall integrity 915 signaling pathway: drug target, compensatory pathways, and virulence. Front Microbiol, 6, 325. 916

van Gestel, J., Nowak, M. A. and Tarnita, C. E. (2012) The evolution of cell-to-cell communication in a 917 sporulating bacterium. PLoS Comput Biol, 8, e1002818. 918

Wickham, H. (2009) ggplot2: Elegant Graphics for Data Analysis. New York: Use R. Springer-Verlag. 919 Wilson, E. O. (1975) Sociobiology: the new synthesis. Harvard University Press. 920 Wongsuk, T., Pumeesat, P. and Luplertlop, N. (2016) Fungal quorum sensing molecules: Role in fungal 921

morphogenesis and pathogenicity. J Basic Microbiol, 56, 440-447. 922 Xiang, Q., Rasmussen, C. and Glass, N. L. (2002) The ham-2 Locus, Encoding a Putative Transmembrane Protein, 923

Is Required for Hyphal Fusion in Neurospora crassa. Genetics, 160, 169-180. 924 Yago, J. I., Lin, C. H. and Chung, K. R. (2011) The SLT2 mitogen-activated protein kinase-mediated signalling 925

pathway governs conidiation, morphogenesis, fungal virulence and production of toxin and melanin in 926 the tangerine pathotype of Alternaria alternata. Mol Plant Pathol, 12, 653-665. 927

Zelikovitch, N., Eyal, Z., Ben-Zyi, B. and Koltin, Y. (1990) Double-stranded RNA mycoviruses in Septoria tritici. 928 Mycol Res, 94, 590-594. 929

Zhan, J., Kema, G. H., Waalwijk, C. and McDonald, B. A. (2002) Distribution of mating type alleles in the wheat 930 pathogen Mycosphaerella graminicola over spatial scales from lesions to continents. Fungal Genetics 931 and Biology, 36, 128-136. 932

Zhao, X., Mehrabi, R. and Xu, J. R. (2007) Mitogen-activated protein kinase pathways and fungal pathogenesis. 933 Eukaryot Cell, 6, 1701-1714. 934

935

936

Figure Legends 937

938

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 34: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

34

Figure 1. Vegetative cell fusions of Zymoseptoria tritici. Blastospores or 939

pycnidiospores expressing either the cytoplasmic green fluorescent protein (GFP) or 940

red-fluorescent protein (mCherry or mCh) were co-inoculated on water agar (WA) 941

plates in a 1:1 ratio. (A) Blastospores-tagged GFP or mCh appeared in only one-color 942

channel of the fluorescence microscope. (B) High initial cell concentration (1x107 943

blastospore/mL) induces conidial anastomosis tubes (CATs) that were frequently 944

observed after 17 hours of incubation. (C-D) Vegetative hyphal fusions (VHFs) from 945

germinating blastospores or pycnidiospores are induced at lower initial cell 946

concentration (1x106 blastospore/mL) and were noticed after 40 hours of incubation, 947

respectively. Black arrows or white triangles point to the CATs or VHFs, respectively, 948

which resulted in the mixture of both GFP and mCh fluorescent proteins due to the 949

exchange of cytoplasm content between the two fused cells. White asterisks point to 950

the fusion bridges formed from individuals expressing the same fluorescent proteins. 951

952

Figure 2. Hyphal fusion does not lead to the generation of heterokaryon individuals in 953

Zymoseptoria tritici. ZtHis1-ZtGFP strain was used to monitor the nuclei movement 954

during hyphal fusion. None septum containing more than one nucleus was observed, 955

neither the genetic exchange between the two fused hyphal individuals. Black 956

triangles indicate the septal compartment formed in the fusion bridges containing only 957

one nucleus. 958

959

Figure 3. ZtSo is required for vegetative cell fusion in Zymoseptoria tritici. (A) At initial 960

higher cell density – an inducing condition for CATs, fusion bridges between 961

blastospore germlings were only observed for DZtKu70 and DZtSo-comp strains. Black 962

arrows indicate the CATs. (B) At initial lower cell density – an inducing condition for 963

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 35: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

35

VHFs, anastomoses were noticed only for those strain possessing the ZtSo gene. 964

CATs or VHFs were never formed between DZtSo mutant spores. White triangles point 965

to the fusion bridges between two fused hyphal cells. 966

967

Figure 4. Disruption of ZtSo strongly affects the melanization of Zymoseptoria tritici. 968

The fusion defective mutant was significantly less melanized than the DZtKu70 and 969

DZtSo-comp strains, which exhibited higher melanin accumulation over time. Bars 970

represent standard errors of the mean gray value on at least 40 colonies. Different 971

letters on the top of the bars indicate a significant difference among the tested strains 972

according to the Analysis of Variance (ANOVA). The notch displays a 95% confidence 973

interval of the median. Open circles represent the outlier values of each strain. Pictures 974

shown below the bar plot represent the melanization level of DZtKu70, DZtSo and 975

DZtSo-comp strains. Gray value scale (0 = black and 255 = white) is shown on the left. 976

977

Figure 5. ZtSo is dispensable for the cellular integrity. A serial dilution of blastospore 978

suspension from DZtKu70, DZtSo, and DZtSo-comp strains were exposed to nine 979

different stress conditions, including different temperatures (18°C and 27°C), oxidative 980

stress (0.5 and 1 mM of hydrogen peroxide - H2O2), osmotic stress (1M sodium 981

chloride - NaCl and 1M sorbitol), cell wall stress (2 mg/mL congo red and 10 µg/mL 982

calcofluor white - CFW), and plasma membrane stress (0.01% sodium dodecyl sulfate 983

- SDS) for five days. Besides the impairment in melanization exhibited by the DZtSo 984

mutant, all tested strains do not vary in their tolerance to different cellular stressor 985

agents. 986

987

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 36: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

36

Figure 6. The biological role of vegetative hyphal fusions for the disease progress and 988

development of the asexual reproductive structures of Zymoseptoria tritici. (A) 989

Susceptible wheat cv. Drifter inoculated with DZtKu70, DZtSo, or DZtSo-comp strains 990

were evaluated up to 21 days post-infection (dpi). All tested strains exhibited similar 991

disease progress. The onset of the necrotrophic phase started at 11 dpi, whereas the 992

necrotic lesions continued to expand across the leaves over time, resulting in the 993

collapse of the host tissue at 21 dpi. Pre-pycnidia were noticed at 12 dpi, and the 994

mature pycnidia are prominent after 14 days of incubation. However, neither pre-995

pycnidia nor mature pycnidia were formed in the wheat plants inoculated with the 996

fusion defective DZtSo mutant. (B) Blastospores of DZtKu70, DZtSo, and DZtSo-comp 997

strains were inoculated on wheat extract agar plates at 18°C and incubated under UV-998

A light (16:8 light:dark cycle). After 20 days of incubation, DZtKu70 and DZtSo-comp 999

strains produced brown pycnidia-like structures exuding a whitish liquid similar to the 1000

oozed cirrhus-containing pycnidiospores spores observed for Z. tritici-infected wheat 1001

plants. In contrast, DZtSo mutant formed only mycelial knots, which are aggregates of 1002

filamentous hyphae and the precursor developmental stage of pycnidium, but the 1003

mature asexual fruiting bodies were never formed. 1004

1005

Figure 7. Schematic demonstration of pycnidial development during plant infection by 1006

a filamentous fungus. Susceptible wheat cv. Drifter was inoculated with the fluorescent 1007

Zymoseptoria tritici 1E4GFP (wild-type) and DZtSoGFP strains and monitored by confocal 1008

microscopy at different days post-infection (dpi). For earlier time points (6, 7 and 8 1009

dpi), please see Fig.S10. The first intracellular hyphae surrounding the stomatal guard 1010

cells produced specialized knots from where secondary hyphae emerge and 1011

germinate. After 9 days of infection, these secondary hyphae from the 1E4GFP strain 1012

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 37: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

37

fuse with another nearby hypha (represented by black circles), creating an 1013

interconnected network in the sub-stomatal cavity. The combination of sub-stomatal 1014

hyphal accumulation and anastomoses generates the pre-pycnidium at 12 days, which 1015

later supports the asexual reproduction of Z. tritici. Unlike, the filamentous hyphae 1016

from the DZtSoGFP mutant kept extending as individual hyphae and none fusion point 1017

was observed. The lack of anastomosis stops the developmental process of the 1018

pycnidium and, consequently, impair the asexual cycle of Z. tritici. 1019

1020

Supporting Information 1021

1022

Table S1. List of primers used in this study for cloning, sequencing and knock-out 1023

mutants confirmation. 1024

1025

Figure S1. Vegetative hyphal fusion occurs during epiphytic growth on wheat leaves. 1026

Co-infection of wheat plants with blastospores (A) or pycnidiospores (B) expressing 1027

either the cytoplasmic green fluorescent protein (GFP) or the red-fluorescent protein 1028

(mCherry) resulted in hyphal fusions and cytoplasm streaming of both fluorescent 1029

proteins after 48 hours of infection. Hyphal fusion during epiphytic colonization may 1030

assist the fungus to create an interconnected network supporting its establishment on 1031

the leaf surface before host penetration. 1032

1033

Figure S2. Scheme showing ZtSo gene and protein sequences and its phylogenetic 1034

relationship among Ascomycete species. (A) Illustration demonstrates the ZtSo gene 1035

locus and its protein sequencing containing the Atrophin 1, WW and PhoD as protein 1036

domains. (B) The alignment of the WW protein-protein interaction domain including 1037

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 38: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

38

the PPLP motif of 41 different fungal species. Red boxes surround the two conserved 1038

tryptophan residues spaced 22 amino acids apart. (C) Phylogenetic analysis grouped 1039

the orthologues of the ZtSo gene onto three groups based on fungal Classes 1040

(Dothideomycetes, Sordariomycetes, and Chaetothyriomycetes together with 1041

Eurotiomycetes), independently whether they were parasites, mutualists or 1042

saprotrophs. Three members of Basidiomycetes were used as an outgroup. 1043

1044

Figure S3. Cytoplasmic streaming between DZtKu70 and the GFP-tagged 1E4 strain. 1045

Blastospores of DZtKu70 and 1E4(GFP) were co-inoculated on water agar (WA) plates, 1046

a hyphal fusion-inducing condition. After 40 hours of incubation, fusion bridges were 1047

observed between DZtKu70 and 1E4(GFP) strains. The detection of the green 1048

fluorescent protein in the cytoplasm of the recipient hypha DZtKu70 confirms the 1049

cytoplasmic streaming between the two fused individuals (panel 1). Black asterisk 1050

points to the non-fluorescent DZtKu70 spore before to hyphal fusion. White triangle 1051

indicates the fusion point between the DZtKu70 and 1E4(GFP) strains. 1052

1053

Figure S4. Cytoplasmic streaming between DZtSo-comp and GFP-tagged 1E4 strain. 1054

Blastospores of DZtSo-comp and 1E4(GFP) were co-inoculated on water agar (WA) 1055

plates, a hyphal fusion-inducing condition. After 40 hours of incubation, fusion bridges 1056

were observed between DZtSo-comp and 1E4(GFP) strains. The detection of the green 1057

fluorescent protein in the cytoplasm of the recipient hyphae DZtSo-comp confirm the 1058

cytoplasmic streaming between the fused individuals (panel 1). Black asterisks point 1059

to the non-fluorescent DZtSo-comp spore before to hyphal fusion. White triangles 1060

indicate the fusion points between the DZtSo-comp and 1E4(GFP) strains. 1061

1062

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 39: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

39

Figure S5. Co-inoculation of DZtSo and GFP-tagged 1E4 strain confirms the failure of 1063

the DZtSo mutant to undergo hyphal fusion. Blastospores of DZtSo and 1E4(GFP) were 1064

co-inoculated on water agar (WA) plates, a hyphal fusion-inducing condition. After 40 1065

hours of incubation, fusion bridges were only observed between 1E4(GFP) germinating 1066

spores (panel 1). Fluorescent green protein was never detected on the cytoplasm of 1067

DZtSo cells (panels 1, 2 and 3). The filamentous hyphae of DZtSo mutant could grow 1068

in parallel with those hyphae from the 1E4(GFP), but they never undergo hyphal fusion 1069

(panels 2 and 3), demonstrating that ZtSo is required in both fusion partners for the 1070

establishment of the fungal communication required for perception and attraction 1071

before fusion. Black asterisks point to the non-fluorescent DZtSo spores. 1072

1073

Figure S6. MAP kinase-encoding ZtSlt2 and ZtFus3 genes are required for 1074

anastomosis in Zymoseptoria tritici. Hyphal fusions were regularly found for the wild-1075

type strain (IPO323). However, deletion of ZtSlt2 (orthologous to MAK-1) or ZtFus3 1076

(orthologous to MAK-2) resulted in fusion-defective mutants, probably due to the 1077

disruption of the oscillatory recruitment of both MAP kinase modules required for cell-1078

to-cell communication and fusion, as described for Neurospora crassa (Fleissner et 1079

al., 2009). The defective phenotype was restored in the complemented DZtSlt2-comp 1080

strain. White triangles point to self-fusion events. 1081

1082

Figure S7. Effect of ZtSo deletion on the vegetative fungal growth. (A) The DZtSo 1083

mutant exhibited a similar colony morphology than DZtKu70 and DZtSo-comp strains. 1084

(B) Light microscopy of colony edges showed a dense hyphal-thickened margin for 1085

DZtKu70 and DZtSo-comp colonies, whereas the DZtSo mutant exhibited only a few 1086

filamentous at the colony periphery. Dashed squares point to the localization of 1087

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 40: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

40

microscope images. (C) Thought none morphological differences were observed for 1088

the tested Zymoseptoria tritici strains, the fusion defective DZtSo mutant had a slight, 1089

but significant reduction of its radial growth (mm) compared to DZtKu70 and DZtSo-1090

comp, when grown on a nutrient-limited medium. At least 40 colonies of each tested 1091

strains were evaluated. Two and three stars indicate a p-value <0.005 and <0.0005, 1092

respectively. (D) None morphological differences were noticed for the blastospores 1093

produced by DZtKu70, DZtSo or DZtSo-comp strains. (E) DZtSo mutant grew faster 1094

and had greater radial growth overtime than the DZtKu70 and DZtSo-comp strains, 1095

when incubated on a nutrient-rich medium. Bars represent standard errors of the radial 1096

growth (mm) of at least 40 colonies. Different letters on the top of the bars indicate a 1097

significant difference among the tested strains according to the Analysis of Variance 1098

(ANOVA). The notch displays a 95% confidence interval of the median. Open circles 1099

represent the outlier values of each strain. 1100

1101

Figure S8. Percentage of leaves cover by lesions (PLACL) produced by DZtKu70, 1102

DZtSo, and DZtSo-comp strains. The second leaves of the wheat plants cv. Drifter 1103

infected with the Zymoseptoria tritici strains were harvested at different days post-1104

inoculation (dpi). Bars represent standard errors of PLACL values of at least six 1105

infected leaves. Different letters on the top of the bars indicate significant difference 1106

among the tested strains according to the Analysis of Variance (ANOVA). Black circles 1107

represent the outlier data points. 1108

1109

Figure S9. DZtSo and DZtKu70 strains does not vary in pathogenicity, except for the 1110

failure of DZtSo mutant to undergo asexual reproduction. Five different winter cultivars 1111

were infected with DZtKu70 or DZtSo strains and evaluated at 14- and 21 days post-1112

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 41: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

41

inoculation (dpi) for host damage and pathogen reproduction, respectively. Runal and 1113

Claro are classified as susceptible cultivars. Arina is considered intermediate, whereas 1114

Titlis and Camedo are described as resistant cultivars to Z. tritici. On the left panel, 1115

DZtKu70 and DZtSo showed a comparable host damage for the Runal, Claro and Titlis 1116

cultivars at 14 dpi. Both strains were avirulent in Arina and Camedo. On the right panel, 1117

the asexual reproductive structures were observed within the necrosis of those plants 1118

inoculated with the DZtKu70 strain at 21 dpi. None pycnidium was observed for plants 1119

inoculated with the DZtSo mutant, demonstrating that the failure to undergo asexual 1120

reproduction is associated with the disruption of ZtSo per se than a cultivar-specific 1121

interaction. 1122

1123

Figure S10. Schematic demonstration of hyphal penetration, substomatal colonization 1124

and initial stages of pycnidial development. Susceptible wheat cv. Drifter was 1125

inoculated with the fluorescent Zymoseptoria tritici 1E4GFP (wild-type) and DZtSoGFP 1126

strains and monitored by confocal microscopy at different days post-infection (dpi). 1127

For later time points (9 and 12 dpi), please see Fig.7. At 6 dpi, the epiphytic filamentous 1128

hyphae penetrate the stomata and enter first in the substomatal cavity. At 7 dpi, initiate 1129

the intracellular hyphal colonization, which the filamentous surrounding the stomatal 1130

guard cells produce specialized knots from where secondary hyphae emerge and 1131

germinate. Up to this point, none morphological difference for hyphal extension and 1132

intracellular hyphal colonization was noticed between 1E4GFP and DZtSoGFP strains. At 1133

8 dpi, the secondary hyphae from 1E4GFP strain fuse with another nearby hypha 1134

(represented by black circles), creating an interconnected network in the sub-stomatal 1135

cavity. Unlike, the secondary hyphae from DZtSoGFP kept extending as individual 1136

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 42: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

42

filamentous and none indication of anastomosis was observed until this developmental 1137

stage. 1138

1139

Figure S11. Description of functional characterizations performed in this study. (A) 1140

ZtKu70 gene was disrupted by the geneticin resistance cassette (gen) via homologous 1141

recombination in the 1E4 wild-type (WT) strains and generating the 1E4DZtKu70 1142

mutant. Next, the ZtSo gene was knocked-out by the hygromycin resistance cassette 1143

(hph) via homologous recombination in the genetic background of the 1E4DZtKu70 1144

mutant, generating the double mutant 1E4DZtKu70DZtSo. Later, the ZtSo gene was 1145

re-introduced in locus in the previous double mutant 1E4DZtKu70DZtSo using the 1146

selective nourseothricin resistance gene (nat) and generating the 1E4DZtKu70DZtSo-1147

comp mutant. (B) Agarose gel shows the PCR fragments at the expected sizes of 1148

3’954 base pairs (bp) or 1’626 bp, confirming the presence of the ZtSo native gene or 1149

the hygromycin resistance gene, respectively. (C) Agarose gel shows a PCR fragment 1150

of 286 bp confirming the presence of the nourseothricin resistance gene only in the 1151

1E4DZtKu70DZtSo-comp mutant. 1152

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 43: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 44: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 45: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 46: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 47: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 48: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint

Page 49: 2 Zymoseptoria tritici - bioRxiv · 2 26 Abstract 27 The ability of fungal cells to undergo cell fusion allows them to maximize their overall 28 fitness. In this study, we characterized

.CC-BY-ND 4.0 International licenseauthor/funder. It is made available under aThe copyright holder for this preprint (which was not peer-reviewed) is the. https://doi.org/10.1101/2020.01.26.918797doi: bioRxiv preprint