+ All Categories
Home > Documents > A protease of the subtilase family negatively regulates ... · 2 2.9. transcriptional control in...

A protease of the subtilase family negatively regulates ... · 2 2.9. transcriptional control in...

Date post: 06-Jul-2020
Category:
Upload: others
View: 2 times
Download: 0 times
Share this document with a friend
332
En vue de l'obtention du Délivré par : Institut National Polytechnique de Toulouse (INP Toulouse) Discipline ou spécialité : Interactions plantes-microorganismes Présentée et soutenue par : le vendredi 12 février 2016 Titre : Unité de recherche : Ecole doctorale : A PROTEASE OF THE SUBTILASE FAMILY NEGATIVELY REGULATES PLANT DEFENCE THROUGH ITS INTERACTION WITH THE ARABIDOPSIS TRANSCRIPTION FACTOR AtMYB30 Sciences Ecologiques, Vétérinaires, Agronomiques et Bioingénieries (SEVAB) Laboratoire Interactions Plantes Microorganismes (LIPM) Directeur(s) de Thèse : MME SUSANA RIVAS Rapporteurs : M. PATRICK GALLOIS, UNIVERSITY OF MANCHESTER M. SEBASTIEN BAUD, INRA VERSAILLES GRIGNON Membre(s) du jury : 1 M. JEAN-PHILIPPE GALAUD, UNIVERSITE TOULOUSE 3, Président 2 Mme SUSANA RIVAS, INRA TOULOUSE, Membre 2 M. THOMAS KROJ, CIRAD MONTPELLIER, Membre
Transcript

En vue de l'obtention du

!"#$"%&$'!(')*+,-.(%/-$0'!('$"+)"+/(

Délivré par :

Institut National Polytechnique de Toulouse (INP Toulouse)

Discipline ou spécialité :

Interactions plantes-microorganismes

Présentée et soutenue par :

!"#$%&''&#()*+,%--

le vendredi 12 février 2016

Titre :

Unité de recherche :

Ecole doctorale :

A PROTEASE OF THE SUBTILASE FAMILY NEGATIVELY REGULATES

PLANT DEFENCE THROUGH ITS INTERACTION WITH THE

ARABIDOPSIS TRANSCRIPTION FACTOR AtMYB30

Sciences Ecologiques, Vétérinaires, Agronomiques et Bioingénieries (SEVAB)

Laboratoire Interactions Plantes Microorganismes (LIPM)

Directeur(s) de Thèse :

MME SUSANA RIVAS

Rapporteurs :

M. PATRICK GALLOIS, UNIVERSITY OF MANCHESTER

M. SEBASTIEN BAUD, INRA VERSAILLES GRIGNON

Membre(s) du jury :

1 M. JEAN-PHILIPPE GALAUD, UNIVERSITE TOULOUSE 3, Président

2 Mme SUSANA RIVAS, INRA TOULOUSE, Membre

2 M. THOMAS KROJ, CIRAD MONTPELLIER, Membre

Remerciements

Cette thèse a été réalisée au sein du Laboratoire des Interactions Plantes Microorganismes (LIPM) à

Toulouse.

Ma gratitude va à Sébastien Baud, Jean-Philippe Galaud, Patrick Gallois et Thomas Kroj pour avoir

aimablement accepté d’être jurés de thèse et pour leurs précieuses questions et réflexions qui, lors

de la soutenance, ont donné lieu à de riches discussions.

Je tiens à remercier Susana Rivas qui a remarquablement supervisé cette thèse et qui m’a prodigué

confiance, conseils et encouragements au cours de ces années passées au laboratoire.

Je souhaite exprimer mes remerciements très sincères à tous ceux (de la plateforme de microscopie,

du service de transgénèse et de l’équipe d’accueil) qui m’ont aidé à réaliser ce travail. Les résultats

présentés dans ce rapport sont aussi les leurs tant ils ont œuvré pour les obtenir.

Ma reconnaissance va également à mon comité de thèse, Renier van der Hoorn et Manuel Piňeiro,

pour leurs suggestions et leurs conseils pertinents sur ce projet.

Je remercie mes professeurs de l’Université d’Albi, de l’Université d’Algarve au Portugal et de

l’Université Toulouse III pour avoir suscité mon intérêt pour la biologie.

1

Table of Contents

__________________________________________________________________________

ABBREVIATIONS 4

FIGURE LIST 7

TABLE LIST 10

INTRODUCTION 11

1. PLANT-PATHOGEN INTERACTIONS AND PLANT IMMUNITY 12

1.1. PLANT-PATHOGEN INTERACTIONS 12

1.2. PLANT DEFENCE MECHANISMS: A MOLECULAR BATTLEFIELD 14

1.2.1. Constitutive defences 14

1.2.2. Inducible defences 16

· Pathogen-Triggered Immunity (PTI) 17

· Effector-Triggered Susceptibility (ETS) 20

· Effector-Triggered immunity (ETI) 22

1.2.3. Signalling events during plant defence responses 24

1.3. PLANT PROTEASES: ROLES IN LIFE AND DEATH DURING PLANT DEFENCE SIGNALLING 27

1.3.1. Aspartic proteases 29

1.3.2. Cysteine proteases 30

1.3.3. Metalloproteases 33

1.3.4. Serine proteases 33

· Carboxypeptidase-like proteases 34

· Subtilisin-like proteases or Subtilases 34

2. TRANSCRIPTIONAL REGULATION OF PLANT DEFENCE RESPONSES 41

2.1. AP2/EREBP TFS 42

2.2. BHLH TFS 43

2.3. BZIP TFS 43

2.4. BBX TFS 43

2.5. NAC TFS 44

2.6. WHIRLY TFS 44

2.7. WRKY TFS 44

2.8. MYB TFS 45

2.8.1. DNA MYB Binding Sites (MBSs) 46

2.8.2. Classification of MYB TFs 47

· 1R-MYB 47

· 2R-MYB (or R2R3-MYB) 47

· 3R-MYB (or R1R2R3-MYB) 48

· 4R-MYB 48

2.8.3. Functions of MYB TFs 49

2

2.9. TRANSCRIPTIONAL CONTROL IN PLANT DEFENCE (REVIEW ARTICLE) 51

3. ATMYB30 A POSITIVE REGULATOR OF THE HR IN A. THALIANA 52

3.1. IDENTIFICATION OF ATMYB30 52

3.2. EXPRESSION AND FUNCTION OF ATMYB30 52

3.3. HORMONAL CONTROL OF THE ATMYB30-MEDIATED HR 54

3.4. TRANSCRIPTIONAL TARGETS OF ATMYB30 55

3.5. REGULATION OF ATMYB30 56

3.5.1. Post-transcriptional regulation of AtMYB30 56

3.5.2. Post translational modification of AtMYB30 57

3.5.3. Regulation of AtMYB30 activity through protein-protein interactions 58

SCIENTIFIC CONTEXT OF THE PHD PROJECT 62

OBJECTIVES OF THE PHD PROJECT 63

RESULTS 64

A PROTEASE OF THE SUBTILASE FAMILY NEGATIVELY REGULATES PLANT DEFENCE THROUGH ITS INTERACTION WITH THE

ARABIDOPSIS TRANSCRIPTION FACTOR ATMYB30 65

Previous results: Identification of AtSBT5.2 as a new AtMYB30 interacting partner. 65

1. CHARACTERIZATION OF ATSBT5.2 66

1.1. AtSBT5.2 is alternative spliced and encodes two distinct isoforms. 66

1.2. AtSBT5.2(a) is a secreted protein whereas AtSBT5.2(b) is intracellular. 67

1.3. AtSBT5.2(a), but not AtSBT5.2(b), is glycosylated in planta. 68

1.4. AtSBT5.2(a) is an active serine protease. 70

2. CHARACTERIZATION OF THE INTERACTION BETWEEN ATMYB30 AND ATSBT5.2 73

2.1. Neither AtSBT5.2(a) nor AtSBT5.2(b) affect AtMYB30 protein accumulation in planta. 73

2.2. AtSBT5.2(b), but not AtSBT5.2(a), colocalizes with AtMYB30 in planta. 73

2.3. AtSBT5.2(b), but not (a), interacts with AtMYB30 in planta. 74

2.4. The AtSBT5.2(b)-AtMYB30 interaction is specific and mediated through AtSBT5.2(b) C-terminal

domain 75

2.5. AtSBT5.2(b) localization in intracellular vesicles is mediated through its N-terminal domain 76

3. FUNCTIONAL ANALYSIS OF ATSBT5.2 IN PLANT DEFENCE 78

3.1. AtSBT5.2 negatively regulates plant defence and HR. 78

3.2. AtSBT5.2 controls the HR via AtMYB30. 80

3.3. AtSBT5.2(b), but not AtSBT5.2(a), negatively regulates defence-associated cell death responses.

81

3.4. Analysis of AtSBT5.2 expression after bacterial inoculation. 81

DISCUSSION 83

1. AS, AN EMERGING REGULATORY MECHANISM OF PLANT DEFENCE 84

2. REGULATION OF ATSBT5.2 FUNCTION THROUGH AS 87

2.1. AS AFFECTS THE SUBCELLULAR LOCALIZATION OF RESULTING ATSBT5.2 PROTEIN VARIANTS 87

3

2.1.1. AtSBT5.2(b) localizes to endosomes 88

2.1.2. Endosomes as important sites for regulation of defence signalling 91

2.2. AS AFFECTS THE GLYCOSYLATION STATUS OF RESULTING ATSBT5.2 PROTEIN VARIANTS 93

2.3. AS APPEARS TO AFFECT THE CATALYTIC ACTIVITY OF RESULTING ATSBT5.2 PROTEIN VARIANTS 94

3. THE APOPLAST AS A PRIVILEGED SITE FOR ANTI-MICROBIAL DEFENCE 98

4. NUCLEAR EXCLUSION THROUGH INTERACTION WITH ATSBT5.2(B): A NEW REGULATORY MECHANISM OF

ATMYB30 ACTIVITY 100

MATERIALS AND METHODS 103

OTHER RESULTS 113

REFERENCES 115

4

Abbreviations A, Ala Alanine A Aa Alternaria alternataf.sp. lycopersici Ab Alternaria brassicicola ABA Abscisic Acid ABP Activity-Based Profiling AD Activation Domain AP2/EREBP APETALA2/Ethylene Responsive Element Binding ARF Auxin-Response Factor At Arabidopsis thaliana AtPLA2 Phospholipase A2 of Arabidopsis thaliana Atu Agrobacterium tumefaciens As Avena sativa (oat) AS Alternative Splicing BBX B-box protein B Bc Botrytis cinerea BD Binding Domain Be Botrytis elliptica bHLH basic Helix-Loop-Helix BR Brassinosteroid BRI1 Brassinosteroid Insensitive 1 BRS1 Brassinosteroid Insensitive Suppressor 1 Bt Botrytis tulipae bZIP basic Domain Leucine Zipper Ca Capsicum annuum (Pepper) C CC Coiled-Coil Cd Colletotrichum destructivum CDPK Calcium-Dependent Protein Kinase CEV Citrus Exocortis Viroid Cf Cladosporium fulvum CFP Cyan Fluorescent Protein CK Cytokinin Cv Cochliobolus victoriae D, Asp Aspartic Acid D DAMPs Damage-Associated Molecular Patterns Ea Erwinia amylovora E Ec Erysiphe cruciferarum EE Early endosome EF-Tu Elongation Factor Thermo-unstable ECM Extracellular Matrix Endo H Endoglycosidase H ER Endoplasmic Reticulum ET Ethylene ETI Effector-Triggered Immunity ETS Effector- Triggered Susceptibility FLS2 Flagellin-Sensing 2 F FRET-FLIM Förster Resonance Energy Transfer-Fluorescence Lifetime Imaging Fs Fusarium solani

5

GA Giberellic Acid G Gc Golovinomyces cichoracearum GFP Green Fluorescent Protein Gm Glycin max (Soybean) H, His Histidine H HA Hemagglutinin (HA)-epitope tag Ha Hyaloperonospora arabidopsis HLH Helix-Loop-Helix HR Hypersensitive Response hrp HR and Pathogenicity IF Intercellular Fluid I ISR Induced Systemic Resistance JA Jasmonic Acid J Le Lycopersicon esculentum (Tomato) L LE Late Endosom LRR Leucine-Rich Repeat lsd lesion simulating disease MAMP Microbe-Associated Molecular Pattern M MAPK Mitogen-Activated Protein Kinase MBSs MYB Binding Sites Md Malus domestica (Apple tree) Me Manihot esculenta (Cassava) MIEL AtMYB30-Interacting E3 Ligase Mo Magnaporthe oryzae MS Murashige and Skoog MYB Myeloblastom N, Asn Arginine N NAC NAM (No Apical Meristem), ATAF (Arabidopsis thaliana transcription

Activation Factor) and CUC2 (Cup-Shaped Cotyledon) NBS Nucleotide-Binding Site NLR Nucleotide-binding Oligomerization Domain-Like Receptor NLS Nuclear Localization Signal Nb Nicotiana benthamiana Nt Nicotiana tabacum Nu Nicotiana umbratica Os Oriza sativa (rice) O p35S The cauliflower mosaic virus promoter P PA Protease associated domain PAMP Pathogen-Associated Molecular Pattern PCD Programmed Cell Death PCR Polymerase Chain Reaction PD Prodomain PGNase F Peptide N-glycosidase F PGSs Putative N-glycosylation sites Phs Phytophthora sojae Pi Phytophthora infestans

6

PR Pathogenesis-Related PRR Pattern-Recognition Receptor Ps Pseudomonas syringae Pst Pseudomonas syringuae pv. tomato Pt Puccinia striiformis f. sp. tritici PTI PAMP-Triggered Immunity PTM Post Translational Modification Pv Plasmopara viticola pv. Pathovar RFP Red Fluorescent Protein R RLCK Receptor-Like Cytoplasmic Kinase RLK Receptor-Like Kinase RLP Receptor-Like Protein ROS Reactive Oxygen Species Rs Ralstonia solanacearum S, Ser Serine S SA Salicylic Acid SAR Systemic Acquired Resistance SBT Subtilase SCF Skp1, Cullin, F-box-type ligase Sf Spodoptera frugiperda Sl Solanum lycopersicum (Tomato) St Solanum tuberosum (Potato) SUMO Small Ubiquitin-like Modifier T2SS Type II Secretion System T T35S The cauliflower mosaic virus terminator T3SS Type III Secretion System Ta Triticum aestivum (Wheat) TAD Transcription Activation Domain TAL Transcription Activator-Like TF Transcription Factor TGN Trans Golgi Network TIR Toll-Interleukin1 Receptor TMD Transmembrane Domain TMV Tobacco Mosaic Virus Tn Trichoplusia ni Ub Ubiquitine U UPS Ub/Proteasome System UTR Untranslated Region VLCFA Very Long Chain Fatty Acid V Vv Vitis vinifera (Grapevine) Xcc Xanthomonas campestris pv. campestris X Xcv Xanthomonas campestris pv. vesicatoria Xoo Xanthomonas oryzae pv. oryzae Y2H Yeast Two-Hybrid Y YFPv Yellow Fluorescent Protein venus Zm Zea mays (Maize) Z

7

Figure list

Figure 1. Disease symptoms on Arabidopsis leaves caused by pathogens (From Pieterse et al., 2009).

Figure 2. Overview on the various types of interaction (Adapted from Nürnberger et al., 2004). Figure 3. The zigzag model illustrates the quantitative output of the plant immune system (Adapted from Jones and Danggl, 2006 and from Pieterse et al., 2009).

Figure 4. Schematic representation of systemically induced immune responses (Adapted from Pieterse et al., 2009).

Figure 5. Plant PRRs and their signalling adapters. Figure 6. Examples of plant targets of bacterial type III effector proteins (From Deslandes and Rivas, 2014).

Figure 7. Major families of R proteins. Figure 8. Model of integrated decoys in NLR protein pairs (From Cesari et al., 2014).

Figure 9. Major signalling mechanisms in plant defence (From Bigeard et al., 2015).

Figure 10. Classic model established for the hormonal control of the plant defence (Addapted from David De Vleesschauwer et al., 2013). Figure 11. Development of the hypersensitive response (HR) on tobacco leaf in response to Pseudomonas syringae pv. tomato DC3000 (http://www.sidthomas.net/images/hypersensitive.jpg).

Figure 12. Cleavage mechanisms of the four major catalytic classes of proteases (From van der Hoorn, 2008).

Figure 13. Classification and number of the catalytic types of Arabidopsis proteases (From van der Hoorn and Jones, 2004).

Figure 14. Protein structure of proteases.

Figure 15. Phylogenetic tree of Arabidospsis subtilases (From Rautengarten et al., 2005).

Figure 16. Plant MYB transcription factor classes (Adapted from Dubos et al., 2010).

Figure 17. Schematic illustration of different MYB protein classes and their functions (From Ambawat et al., 2013). Figure 18. Schematic representation of the relationships between the different Arabidopsis R2R3-MYB subgroups (From Dubos et al., 2010).

Figure 19. Schematic representation of the AtMYB30 protein.

8

Figure 20. Analysis of AtMYB30 expression in Arabidopsis upon bacterial infection (From Daniel et al., 1999). Figure 21. Overexpression of AtMYB30 in tobacco leads to accelerated HR in response to inoculation with different pathogens (From Vailleau et al., 2002).

Figure 22. AtMYB30 modulates the expression of very long chain fatty acid (VLCFA)-related genes after bacterial inoculation (From Raffaele et al., 2008).

Figure 23. Schematic overview of metabolic pathways regulated by AtMYB30 during the incompatible interaction between Arabidopsis and avirulent bacterial pathogens (Adapted from Raffaele et al., 2008).

Figure 24. Simplified model for the simultaneous regulation of AtMYB30-mediated HR cell death through interaction with AtsPLA

2−α and MIEL1 (Adapted from Raffaele and Rivas,

2013). Figure 25. Interaction between AtMYB30 and AtSBT5.2 in yeast.

Figure 26. AtSBT5.2 is alternatively spliced.

Figure 27. Sequence alignment of AtSBT5.2(a) and AtSBT5.2(b) proteins.

Figure 28. Subcellular localization studies show that AtST5.2(a) is secreted whereas

AtSBT5.2(b) is intracellular.

Figure 29. Intercellular fluid isolation confirms that AtSBT5.2(a) is secreted whereas AtSBT5.2(b) is intracellular.

Figure 30. Schematic representation of AtSBT5.2(a) and AtSBT5.1 protein sequences.

Figure 31. Sequence alignment of AtSBT5.2(a) and AtSBT5.1 proteins. Figure 32. AtSBT5.2(a), but not AtSBT5.2(b), is glycosylated in planta. Figure 33. All PGS in AtSBT5.2(a) are used for glycosylation in planta.

Figure 34. AtSBT5.2(a) self cleaves in planta.

Figure 35. AtSBT5.2(a) is an active serine hydrolase.

Figure 36. Mutation of some glycosylated residues affects the catalytic activity of AtSBT5.2(a).

Figure 37. Neither AtSBT5.2(a) nor AtSBT5.2(b) affect AtMYB30 protein accumulation in

planta.

Figure 38. AtSBT5.2(b), but not AtSBT5.2(a), colocalises with and retains AtMYB30 outside

the nucleus.

9

Figure 39. AtSBT5.2(b)-mediated retention of AtMYB30 outside the nucleus is independent

of C-terminal tagging of the subtilase.

Figure 40. AtSBT5.2(b) does not affect AtMYB123 nuclear localization.

Figure 41. AtMYB30 and AtMYB123 colocalize in nuclei with both AtSBT5.2362-730

and

AtSBT5.1405-780

.

Figure 42. Sequence alignment of AtSBT5.2(b)362-730

and AtSBT5.1405-780

proteins.

Figure 43. AtMYB30 localization in intracellular vesicles is AtSBT5.2(b) N-terminal domain-

dependant.

Figure 44. Genetic analysis of AtSBT5.2 and AtSBT5.1 Arabidopsis mutant lines.

Figure 45. AtSBT5.2 negatively regulates HR development and plant resistance to bacterial inoculation.

Figure 46. AtSBT5.2 is a negative regulator of AtMYB30-mediated HR cell death.

Figure 47. Characterization of AtSBT5.2(a) and AtSBT5.2(b) overexpressing Arapidopsis

lines.

Figure 48. AtSBT5.2(b), but not AtSBT5.2(a), negatively regulates defence-related HR cell

death.

Figure 49. AtSBT5.2(b) and AtMYB30 expression profiles and induction rates overlap during

infection with avirulent HR-inducing bacteria.

Figure 50. Predicted effects of AS on the proteins encoded by AtSBT2.2, AtSBT3.6,

AtSBT4.11 and AtSBT4.12 splice variants.

10

Table list

Table 1. Genetic model of the gene for gene theory (From Flor, 1971).

Table 2. Role of proteases in plant defence.

Table 3. FRET-FLIM analysis shows that AtMYB30 physically interacts with AtSBT5.2(b) in

N. benthamiana epidermal cells.

Supplemental Information Table 1. Oligonucleotide primers used in this study.

11

Introduction

__________________________________________________________________________

Figure 1. Disease symptoms on Arabidopsis leaves caused by pathogens (From

Pieterse et al., 2009). Disease symptoms on Arabidopsis leaves caused by the necrotrophic fungus Botrytis cinerea (left),

the biotrophic oomycete Hyaloperonospora arabidopsidis (center) and the hemibiotrophic

bacterium Pseudomonas syringae (right). Photos: Hans van Pelt.

12

1. Plant-pathogen interactions and plant immunity

1.1. Plant-pathogen interactions

Plants are primary producers and therefore a source of nutrients for many organisms

(Cardinale et al., 2011). To adapt to their habitat and maximize their chances of survival,

plants have developed both root and aerial systems, which, in turn, increases the range of

organisms that they can encounter. Many of these organisms, such as plant growth-

promoting rhizobacteria, are beneficial to the plant (Gopal et al., 2013) whereas other

organisms, including phytopathogenic insects, viruses, bacteria, nematodes, fungi, and

oomycetes, have a detrimental effect on plant long-term survival (Dangl and Jones, 2001).

According to their lifestyles, phytopathogens are classified into broad categories (Figure 1):

(i) necrotrophs that kill the host, often through the production of phytotoxins, before

parasitising it, extract nutrients from the cells and then live on dead tissue (such as the fungal

pathogen Botrytis cinerea usually called grey mould) and (ii) biotrophs that obtain nutrients

from living cells, commonly through specialized feeding structures (haustoria) that invaginate

the host cell without disrupting it and require a living host to continue their life cycle (such as

the oomycete pathogen Hyaloperonospora arabidopsis). Hemibiotrophs are microbes that

require a living host initially, but kill it at later stages of infection (such as the bacterial

pathogens Pseudomonas syringae or Xanthomonas campestris) (Glazebrook, 2005).

Bacteria and fungi adopt biotrophic, hemibiotrophic or necrotrophic modes of infection while

viruses are ideal biotrophs (although viral infection can eventually result in host cell death)

(Dangl and Jones, 2001).

Although plants must thus face the diversity of aggressive biotic agents, over-investing in

defence in the absence of infection can be just as detrimental to survival as disease (Brown,

2003). Indeed, plant resistance to disease is a costly response, closely connected to plant

physiological and developmental processes (Lozano-Durán et al., 2013, Fan et al., 2014,

Figure 2. Overview on the various types of interaction (Adapted from Nürnberger et

al., 2004).

Non host resistance/immunity

No pathogen differentiation on the

plant

Sufficient preformed defense

Compatible interaction

Incompatible interactions

Susceptibility/disease

Pathogen propagation on the plant

Insufficient preformed/inducible

defense

Host resistance/immunity

No pathogen propagation on the

plant

Race/cultivar-specific resistance

13

Malinovsky et al., 2014). In agreement, mutants with constitutively active defence responses

often present reduced growth and low fertility (Lorrain et al., 2003).

In parallel, the establishment of a parasitic relationship is dependent on the response of the

plant under attack. Indeed, to adapt to their hostile environment, plants have evolved

sophisticated mechanisms of protection to counteract constant pathogen attacks. Some of

these mechanisms are efficient against a broad range of pathogens, while others are limited

to specific pathogens. These mechanisms are based on an efficient immune system that

depends on cell-autonomous events, and on the ability to develop systemic signals from the

site of infection. Therefore, in host-pathogen relationships severe epidemics of disease

remain the exception rather than the rule (Burdon, 1987).

Interactions of pathogens with plants can either be compatible or incompatible. A

compatible interaction occurs when the pathogen infects a susceptible or a tolerant host

plant. In this case, the plant reacts more or less effectively to this aggression and the

severity of symptoms is variable. Symptoms of disease include death and destruction of host

tissue, wilting, abnormal growth and differentiation and discolouration of host tissue (Dangl

and Jones, 2001). If the plant keeps the ability to grow, the plant is tolerant. Otherwise, if the

plant develops a disease that alters its development, the plant is susceptible. Defence

mechanisms are triggered but in a manner that is too slow and/or too weak for the plant to

survive. The pathogen is qualified as virulent; it multiplies actively within the plant and

appears to be able to suppress the resistance mechanisms of the host (Nürnberger et al.,

2004) (Figure 2). For a biotroph to form a successful infection, it must establish a basic

compatibility with its host. The pathogen may also produce compatibility factors that delay,

avoid or negate recognition by a normally resistant host plant.

An incompatible interaction occurs when the pathogen encounters a non-host plant (non-

host resistance) or a resistant host plant (cultivar-specific resistance or host

resistance). In both cases, failure of the pathogen to invade host cells will prevent it from

14

colonising the host and the plant will be named resistant. The pathogen qualifies as

avirulent as it loses its pathogenicity. Resistant hosts prevent or slow the development and

reproduction of the majority of pathogens that they come into contact with. Resistance can

be expressed at many stages in the infection process, from inhibition of germination and

penetration, to restriction of colony development after entry. In the case of a resistant plant

two situations are possible. When an entire plant species is resistant to all races of a

microorganism that is pathogenic to other plant species, resistance is known as non-host

resistance (Senthil-Kumar and Mysore, 2013). Non-host resistance, therefore, is the most

common form of disease resistance exhibited by plants. Such broad-spectrum resistance

contrasts with host resistance, which is displayed by plant genotypes of susceptible host

species against a specific pathogen agent (Hammond-Kosack and Jones, 1997) (Figure 2).

Having introduced the bases of the interaction between plants and pathogens, the next

section provides an overview of the co-evolutionary forces and the molecular mechanisms

that determine the outcome of this interaction.

1.2. Plant defence mechanisms: a molecular battlefield

Plants, unlike mammals, lack mobile defender cells and a somatic adaptive immune system.

Instead, they rely entirely on their innate immunity and on systemic signals originating from

infection sites (Dangl et al., 2013). Indeed, in response to pathogen attack, plants have

developed complex, multilayered signalling and defence mechanisms to protect themselves.

Defence barriers and mechanisms used by plants represent a co-ordinated network of

molecular, cellular and tissue-based responses that can be classified into constitutive

(passive) and inducible (active) defences.

1.2.1. Constitutive defences

Pathogen initial invasion can be primarily prevented by preformed physical and/or chemical

barriers called constitutive defences. Physical barriers largely involve properties of the plant

15

surface such as the thickness of the cuticle of leaves, cuticular lipid, wax layers or the size of

stomatal pores. The cuticle is a layer coating the outer surface of epidermal cells of organs of

the aerial part of the plants and also present within seed coats (Serrano et al., 2014).

Composed of an insoluble cutin polymer matrix and interspersed with waxes (epicuticular

and intracuticular lipids), the cuticle protects from desiccation and acts as a mechanical

barrier against various abiotic and biotic stress, such as UV radiation and pathogens

(Serrano et al., 2014). In addition to its hydrophobic surface, a vertical leaf orientation can

also add to plant resistance, by preventing the formation of moisture films on the leaf surface,

thus inhibiting infection by pathogens reliant on water for motility. Some plants present a very

thick cuticle and bark, if present; can also prevent the entry of pathogens (Reina-Pinto and

Yephremov, 2009). Therefore, to enter inside the plant, microorganisms must force these

barriers or use wounds or natural openings such as stomata or hydathodes, naturally used

for gas exchange or water-excreting, respectively (Schwab et al., 2015). Stomatal aperture is

driven by specialized plant cells called guard cells (Assmann and Shimazaki, 1999) that

control openings by turgor pressure. This aperture is very tightly controlled by several plant

hormones, most notably abscisic acid (ABA) (Pillitteri and Dong, 2013). In the presence of

pathogenic bacteria and fungi, stomata close rapidly to prevent microbial entry (Gudesblat et

al., 2009). Hydathodes are specialized stomata. Since aperture of hydathodes is not

controlled by the plant, they particularly serve as pathogen entry points (Gu et al., 2013)

especially after a guttation period when the water droplet is sucked back into the plant.

If a microorganism reaches the intercellular space, called apoplast, it must face the plant cell

wall before entering the cell. Although the composition and structure of the cell wall differ

significantly in the relative amounts of its compounds among plant lineages, plant cell walls

are composed of a complex network of polysaccharides, including cellulose microfibrils

embedded in a matrix of pectin, hemicelluloses, lignin, and structural proteins (Loqué et al.,

2015). This composition serves as the plant exoskeleton providing mechanical support but it

also serves as a physical barrier, important for resistance to pathogens. Evidence of the role

Phase 1: PTI Phase 2: ETS Phase 3: ETI

Figure 3. The zigzag model illustrates the quantitative output of the plant immune

system (From Jones and Dangl, 2006 and Pieterse et al., 2009). (A) The zigzag model illustrates the intensity of the plant defence responses in place upon

pathogen interaction. In phase 1, plants detect microbial/pathogen-associated molecular

patterns (MAMPs/PAMPs, red diamonds) via Pattern Recognition Receptors (PRRs) to trigger

PAMP-triggered immunity (PTI). In phase 2, successful pathogens deliver effectors that

interfere with PTI, or otherwise enable pathogen nutrition and dispersal, resulting in Effector-

Triggered Susceptibility (ETS). In phase 3, one effector (indicated in red) is recognized by an

NB-LRR protein, activating Effector-Triggered Immunity (ETI), an amplified version of PTI that

often exceeds a threshold for induction of hypersensitive cell death (HR). In phase 4, pathogen

isolates are selected that have lost the red effector, or perhaps gained new effectors through

horizontal gene flow (in blue). These can help pathogens to suppress ETI. In phase 5, selection

favours new plant NB-LRR alleles that can recognize one of the newly acquired effectors,

resulting again in ETI.

(B) Molecular events occurring in phases 1 to 3 of the zigzag model.

(B)

(A)

PAMPs

Am

pli

tud

e o

f d

efe

nce

Threshold for HR

Threshold for effective resistance

PTI Effectors

ETS Low

High

Avr-R

ETI IHR

Ph

ase

1

Ph

ase

3

ETI

Effectors

Avr-R

Ph

ase

5

16

that the cell wall plays in resistance comes from pathogens that use mechanical force or

release cell wall degrading enzymes to overcome this barrier (Kubicek et al., 2014, Hématy

et al., 2009). For example, Erwinia spp produce pectinases to increase accessibility for other

enzymes like cellulases and xylanases and several other hydrolases to break down the

hemicellulose chains (Toth and Birch, 2005), which disrupt host cell integrity and thus

promote rotting (Toth et al., 2003). Even though in the past these structures have been

regarded as “passive”, research has shown that they are very dynamic and intricately

connected to “active” defences (Traw and Bergelson, 2003).

Beside the physical barriers against pathogen penetration, plants constantly produce various

chemical compounds that inhibit pathogen growth (Osbourn, 1996). These constitutive

chemical barriers include compounds such as antimicrobial compounds (also referred as

phytoanticipins) (Pedras and Yaya, 2015) or secondary metabolites (such as glucosinolates,

tannins, ...) (Ahuja et al., 2012, Bednarek, 2012).

1.2.2. Inducible defences

Pathogens that overcome passive defence layers are systematically perceived. Following

perception of invading microbes, plants use a two-tiered receptor-based immune system to

prevent infection (see below). This adapted defence response is mounted locally and has

been summarized in the zigzag model proposed by Jones and Dangl (Jones and Dangl,

2006), which decrypts the co-evolutionary molecular events driving the interaction between

plants and pathogens (Figure 3).

The outcome of the interaction, which varies according to the genetic determinants of each

organism, is also presented in this model. Local perception of a microorganism can, in

addition, trigger systemic defence responses that prime the plant for resistance against a

broad spectrum of pathogens. Systemic acquired resistance (SAR) and induced systemic

resistance (ISR) are systemic resistance responses that are extremely rapid and usually

Figure 4. Schematic representation of systemically induced immune responses

(Adapted from Pieterse et al., 2009). Systemic acquired resistance (SAR) is typically activated in healthy systemic tissues of locally

infected plants. Upon pathogen infection, a mobile signal travels through the vascular system to

activate defence responses in distal tissues. Salicylic acid (SA) is an essential signal molecule for

the onset of SAR, as it is required for the activation of a large set of genes that encode

pathogenesis-related (PR) proteins with antimicrobial properties. Induced systemic resistance

(ISR) is typically activated upon colonization of plant roots by beneficial microorganisms. Like SAR,

a long-distance signal travels through the vascular system to activate systemic immunity in above-

ground plant parts. ISR is commonly regulated by jasmonic acid (JA)-and ethylene (ET)-dependent

signalling pathways and is typically not associated with the direct activation of PR genes. Instead,

ISR-expressing plants are primed for accelerated JA-and ET-dependent gene expression, which

becomes evident only after pathogen attack. Both SAR and ISR are effective against a broad

spectrum of virulent plant pathogens.

17

involve an amplification of the initial response in distal tissues (Kothari and Patel, 2004)

(Figure 4).

SAR is characterised by the increased, broad spectrum resistance against pathogens

following a primary infection (Gozzo and Faoro, 2013). Development of SAR usually involves

the establishment of a slowly expanding necrotic lesion and other localised responses to

infection, the release of a signal originating from the infection site, and the subsequent

priming of the plant against further attacks, allowing a more rapid response in the case of

subsequent infections. This response is dependent on the plant hormone salicylic acid (SA)

and is associated with the transcriptional reprogramming of a number of defence genes,

including pathogenesis-related (PR) genes, leading to accumulation of PR proteins that

contribute to resistance due to their antimicrobial properties (Muthamilarasan and Prasad,

2013). The precise nature of the signal triggering SAR is still unknown. Although SA levels

increase around necrotic lesions and remain high in plants displaying SAR, a phloem-

translocated lipid molecule, and not SA itself, has been proposed as the SAR-inducing long

distance signal (Chanda et al., 2011, Maldonado et al., 2002, Chaturvedi et al., 2008) (Figure

4).

ISR is triggered by non-pathogenic root bacteria (for example, Pseudomonas fluorescens)

and confers effective resistance against a broad range of pathogens and insect herbivores

(Pieterse et al., 2014). In contrast to SAR, ISR seems to develop independently of SA and

PR gene induction and is rather dependent on the phytohormones jasmonic acid (JA) and

ethylene (ET) (Pieterse et al., 2009) (Figure 4).

· Pathogen-Triggered Immunity (PTI)

When a pathogen manages to penetrate the plant cell wall and reaches the periplasmic

space, it comes into contact with the plasma membrane of the host cell where it is exposed

to surface receptors that are capable of perceiving a great variety of microorganisms. This

detection generally occurs through the perception of microbial molecules, conventionally

Figure 5. Plant PRRs and their signalling adapters. (A) Domain structures for receptor-like kinases (RLKs) and receptor-like proteins (RLPs). The

kinase domain is absent in RLPs. SP: signal peptide; TMD: transmembrane domain; LysM:

lysine motif; LRR: leucine-rich repeats.

(B) Bacterial elicitors flagellin (flg22) and EF-Tu (elf18) are recognised by the Arabidopsis RLKs

flagellin sensing2 (FLS2) and EF-Tu receptor (EFR), respectively. FLS2, and EFR, oligomerise

with BRI1-associated kinase1 (BAK1) in a ligand-dependent manner. Chitin binds to

homodimers of the Arabidopsis lysine motif receptor kinase (LysM-RK) chitin elicitor receptor

kinase1 (CERK1) to induce immune responses. The Arabidopsis RLK PEPR1 recognises

endogenous AtPep peptides that act as danger-associated molecular patterns (DAMPs).

(B)

(A) RLP RLKs

LRR

TMD

SP

Kinase

LysM

Key:

BACTERIA DAMPs

AtPep

peptides

PEPR1

AtProPep

proteins

FLS2

Flagelin

(flg22)

EF-Tu

(elf18)

EFR

BAK1 BAK1

FUNGI

CERK1

Chitin

18

located at the surface of the microorganism called pathogen-associated molecular patterns

(PAMPs) or microbial-associated molecular patterns (MAMPs). PAMPs are typically

conserved and indispensable molecules, such as bacterial flagellin or chitin, a substance

found in fungi cell walls and the exoskeleton of insects and nematodes. However, PAMPs

can also be intracellular molecules being secreted or released from dead bacteria, which are

perceived by the plant [e.g. elongation factor (EF)-Tu]. These molecules are recognized by

cognate plasma-membrane-bound extracellular receptor proteins called pattern recognition

receptors (PRRs), on the external face of host cells. PRRs are typically plasma membrane-

bound receptor-like kinases (RLK)- or receptors-like proteins (RLP)- type proteins (Zipfel,

2014). These receptors present an extracellular leucine-rich repeat (LRR) or lysine motif

(LysM) domain allowing the recognition of "non-self" molecules, a transmembrane domain

(TMD) and a kinase domain in the case of RLKs, involved in signal transduction (Figure 5A).

Without a kinase domain, the short intracellular domain of RLPs associates with intracellular

kinase proteins in order to transduce an appropriate signal response (Böhm et al., 2014).

The Arabidopsis RLKs flagellin sensing2 (FLS2) and EF-Tu receptor (EFR) recognize

bacterial flagellin and EF-Tu, respectively, and are the best characterized plant PRRs (Zipfel,

2014). These and other examples of PRR proteins are shown in Figure 5B. It has recently

become clear that several RLKs require other RLKs for full function (Zipfel, 2014). Binding of

flg22 or elf18 (the immunogenic peptides of flagellin or EF-Tu in Arabidopsis, respectively) to

FLS2 or EFR, respectively, induces their instant association with the co-receptor RLK BRI1-

associated kinase1 (BAK1), phosphorylation of both proteins and initiation of downstream

responses (Roux et al., 2011, Schwessinger et al., 2011, Chinchilla et al., 2007) (Figure 5B).

RLKs interact also with soluble receptor-like cytoplasmic kinases. For example, Botrytis-

induced kinase1 (BIK1) and related PBS1-like (PBL) kinases constitutively associate with

FLS2 and EFR and become quickly phosphorylated and released from the PRR complexes

upon PAMP binding (Lu et al., 2010, Zhang et al., 2010). These associations play positive

regulatory roles in immunity (Böhm et al., 2014). To penetrate into the host cell, several

19

pathogens produce a range of cutin-degrading enzymes, which are often crucial for the

successful penetration of the plant tissue, and release cell wall or cuticular fragments. Plants

are also able to detect and respond to these endogenous molecules released by pathogen

invasion, called danger-associated molecular patterns (DAMPs), that are not available for

recognition under normal conditions. Polysaccharides released from the plant cell wall (e.g.

oligogalacturonides), and some endogenous peptides are DAMPs detected as “infectious

self”. When cell wall damage is detected, the cell wall is remodelled and reinforced at the

penetration site by formation of cell wall-associated structures like papillae (Hückelhoven,

2007). This reaction prevents infection of individual cells and stunts pathogen growth. The

first plant DAMP/PRR pairs have been recently identified. The Arabidopsis RLKs PEPR1 and

PEPR2 perceive AtPep peptides (Yamaguchi et al., 2006, Yamaguchi et al., 2010) (Figure

5B). These peptides are derived from the propeptides (AtProPeps) that are encoded by a

multigenic family of seven-members whose expression is induced by wounding or PAMP

perception (Krol et al., 2010). Treatment with AtPep peptides induces defence gene

expression and overexpression of AtProPep1 leads to enhanced resistance to the fungal root

pathogen Pythium irregulare. AtPep perception is part of a PTI amplification loop and is

important for the induction of systemic immunity (Zipfel, 2013).

Stimulation of PRRs triggers a set of complex signalling pathways leading to the

development of the first line of active defence responses formerly called basal or horizontal

immunity. This response is known as PAMP, DAMP or MAMP-triggered immunity (PTI, DTI

or MTI) and is sufficient to prevent the colonization of the microorganism in the plant (Beck et

al., 2012) (Figure 3). Functional PRRs and co-regulators are crucial for the success of PTI,

as mutant plants with a defective recognition system show increased susceptibility to

pathogens (Miya et al., 2007, Zipfel et al., 2006).

After pathogen detection, activation of PRRs induces a number of defence mechanisms such

as the establishment of structural and/or chemical barriers. Indeed, antimicrobial compounds

can be synthesized de novo in response to microbial attack. Such compounds are known as

20

phytoalexins. For example, when Pseudomonas syringae enters the plant via stomata,

recognition of flg22 by FLS2 stimulates production of reactive oxygen species (ROS), which

have direct antimicrobial properties but also serve as signalling molecules to activate further

immune outputs (O'Brien et al., 2012), cell walls are reinforced by callose, lignin and suberin

deposition for extra protection (Senthil-Kumar and Mysore, 2013) and production and

secretion of molecules (such as camalexin) and defence-related proteins/peptides (such as

PR1) is induced (Bigeard et al., 2015, Bednarek, 2012, Ahuja et al., 2012, Melotto et al.,

2008). This recognition also leads to closure of stomata to limit bacterial entry (Sawinski et

al., 2013). Moreover, certain proline-rich proteins of the cell wall become oxidatively cross-

linked after pathogen attack in an H2O2-mediated reaction. This process strengthens the cell

wall in the vicinity of the infection site, increasing resistance to microbial penetration

(Morimoto et al., 1999). In response to cell wall degrading enzymes secreted by

phytopathogenic microorganisms, plants have evolved a diverse battery of defence

responses including protein inhibitors of these enzymes. These include inhibitors of pectin

degrading enzymes such as polygalacturonases, pectinmethyl esterases and pectin lyases,

and hemicellulose degrading enzymes such as endoxylanases and xyloglucan

endoglucanases (Juge, 2006).

All these defensive reactions, together with passive defences, build up a so-called basal

resistance that is regarded as non-specific as it is activated regardless of the

microorganism encountered. Ultimately, basal defence generally contributes to halt infection

before the microbe gains a hold in the plant allowing resistance to a variety of

phytopathogenic organisms. However, during evolution, some pathogens developed various

strategies to counter this first line of plant defence and acquired the ability to induce the

development of symptoms leading to disease and even the death of the plant.

· Effector-Triggered Susceptibility (ETS)

Successful pathogens such as bacteria, fungi, oomycete, and nematodes have evolved

strategies to circumvent PTI responses and are able to promote pathogenesis by delivering a

Figure 6. Examples of plant targets of bacterial type III effector proteins (From

Deslandes and Rivas, 2014). At the plasma membrane (PM), activation of the receptor complexes, for example, Flagellin-

Sensitive2/BRI1-Associated Kinase1 (FLS2/BAK1) or EF-Tu receptor/BAK1 (EFR/BAK1), by recognition

of conserved bacterial pathogen-associated molecular pattern (PAMP) molecules triggers PAMP-

triggered immunity (PTI)-associated signalling. Phytopathogenic bacteria inject type III effector (T3E)

proteins into plant cells using the type III secretion system (T3SS). Following their translocation into

plant cells, T3Es may be addressed to different subcellular compartments where they may manipulate

a variety of host cellular functions. Recognition of the activity of T3Es by R proteins triggers effector-

triggered immunity (ETI) responses. AvrPto and AvrPtoB target PM-associated receptor complexes.

The cytoplasmic kinase proteins Pto and Fen act as molecular mimics of host virulence targets of

AvrPto and AvrPtoB to activate ETI. AvrPto, AvrPtoB, AvrB, AvrRpm1, AvrRpt2, and HopF2 target the

negative regulator of defence RPM1-interacting protein4 (RIN4) at the PM. AvrAC targets immune

kinases at the PM. HopZ1a, recognized by the R protein ZAR1, targets 2-hydroxyisoflavone

dehydratase (GmHID1), which is involved in isoflavone biosynthesis, and tubulin, which affects the

cellular microtubule network. Cleavage of PBS1 and additional related kinases (BIK1 and PBL1) by

AvrPphB is recognized by the R protein RPS5, triggering ETI. HopI1 and HopN1 are addressed to the

chloroplast where they respectively target the chaperone protein Hsp70, suppressing salicylic acid (SA)

accumulation, and the Photosystem II-associated protein PsbQ, diminishing reactive oxygen species

(ROS) production. HopM1 accumulates in the trans-Golgi network/early endosome (TGN/EE) where it

targets AtMIN7, a key component of vesicle trafficking, thereby suppressing cell wall-associated

defences. HopF2, HopAI1, and AvrB target mitogen-activated protein kinase (MAPK) signalling. In the

nucleus, the Transcription Activator-Like (TAL) effector AvrBs3 is able to mimic eukaryotic

transcription factors (TFs) and directly activate transcription. AvrBs3 binding to the UPA box in host

promoters induces plant cell hypertrophy, contributing to disease development. By contrast, in

resistant plants, activation of the resistance gene Bs3 leads to HR development. The effector protein

XopD targets the Arabidopsis TF AtMYB30, which a positive regulator of Arabidopsis defence

responses. This protein interaction results in repression of AtMYB30 transcriptional activation and

suppression of plant HR and defence responses. PopP2 induces nuclear relocalization of the vacuolar

protease RD19 and perception of PopP2 activity by the R protein RRS1-R activates immunity. HopU1

ribosylates GRP7/8 probably altering immunity-related RNA metabolism. Both HopA1 and AvrRps4

target the enhanced disease suceptibility1 (EDS1) immune regulator disrupting its association with

various immune surveillance proteins, including RPS4. Host targets are underlined.

21

battery of secreted molecules called effectors at the extracellular space of host plant cells

(apoplastic effectors) or inside plant cells (cytoplasmic effectors) (Win et al., 2012).

Therefore, by targeting PTI signalling (Zhang et al., 2007) or PTI receptors (Chaparro-Garcia

et al., 2015) these effectors can suppress PTI, thus resulting in effector-triggered

susceptibility (ETS) (Figure 6). During the past few years, the molecular functions of a

significant number of effectors from various phytopathogens have been discovered, revealing

an astonishing number of eukaryotic processes that are targeted by effector proteins

(Deslandes and Rivas, 2012). Apoplastic effectors are able to prevent PAMP recognition and

PRR activation, as well as chitinase or protease action (Asai and Shirasu, 2015). Pathogen

effectors delivered inside host cells suppress defence responses by targeting components of

defence signalling (details on signalling events triggered upon pathogenic infection are

provided in the next section) (Feng and Zhou, 2012). Host-translocated (cytoplasmic)

effectors are delivered into host plant cells via the type III secretion system (T3SS) (Galán et

al., 2014) or through specialized infectious structures called haustoria that are formed within

the host cell during infection (Giraldo and Valent, 2013). In plant-pathogenic bacteria, genes

encoding components of the T3SS are located in so-called hrp (HR and Pathogenicity) gene

clusters, as their mutation typically disrupts bacterial ability to cause disease on host plants

and to elicit a hypersensitive response (see section 3) on non-host plants (Tang et al., 2006).

The hrp cluster is conserved among Gram negative bacteria including Pseudomonas

syringae, Xanthomonas spp., Ralstonia solanacearum and Erwinia spp. (Galán et al., 2014).

Although the bacterial T3SS has been well studied (Galán et al., 2014), the mechanisms of

effector translocation by filamentous pathogens are still under debate (Giraldo and Valent,

2013). Once inside host cells, effectors subsequently traffic to distinct compartments

including the nucleus (Sarris et al., 2015), the plasma membrane (Wu et al., 2011), the

endoplasmic reticulum (ER) (Block et al., 2014), the tonoplast (Caillaud et al., 2012),

intracellular vesicles (Nomura et al., 2011), chloroplasts (Petre et al., 2015) or the

microtubule network (Lee et al., 2012). Figure 6 presents a general, not exhaustive, view of

the diversity of cellular processes targeted by type III effectors (T3Es) to promote disease.

Figure 7. Major families of R proteins. Representation of the location and structure of the main classes of plant disease resistance

proteins. The majority of R proteins contain tandem leucine-rich repeats (LRRs, depicted in blue),

which have a major role in recognition specificity. The most widely represented R protein family

consists of Nucleotide-Binding site–LRRs (NB-LRRs) proteins that contain a nucleotide-binding site

and a region of similarity to proteins that regulate PCD in metazoans. NB-LRR proteins are likely

localized in the cytoplasm, perhaps as peripheral membrane proteins. Some NB-LRR proteins

contain a putative coiled-coil domain (CC) at the N-terminus. Other NB-LRR proteins contain a

domain with homology to the metazoan superfamily of Toll-like innate immunity receptors (TIR).

Another class of R protein consists of an extracellular LRR (eLRR) anchored to a TM domain.

LRR

RLP (Cf-proteins)

CC

NBS

LRR

(RPM1,

RPS2) (RPS4,

RPS6)

TIR

NBS

LRR

NB-LRRs

Cytoplasm

Pla

sma

me

mb

ran

e

Nucleus

Apoplasm

22

· Effector-Triggered immunity (ETI)

Just as pathogens have evolved to disable plant defences, plants have gained the ability to

recognize and respond to these effector-mediated attacks. Indeed, plants have evolved a,

more efficient and specific, second line of active defence to recognise and protect

themselves from sneaky invaders. To detect effectors, or their interference with host

proteins, plants have developed receptors called resistance (R) proteins. The A. thaliana

Columbia (Col-0) accession presents around 150 R genes in its genome (Meyers et al.,

2003). A majority of R proteins belongs to the intracellular nucleotide-binding leucine-rich

repeat receptor (NLR or NB-LRR) protein family (Jones and Dangl, 2006) that displays

striking similarities with animal nucleotide-binding oligomerization domain-like receptors

(NLRs) or CATERPILLER proteins (Rairdan and Moffett, 2007, Inohara and Nuñez, 2003).

NLR proteins are multidomain proteins and possess a conserved architecture including a

central nucleotide binding site domain (NBS) and a C-terminal LRR domain (Takken and

Goverse, 2012) (Figure 7). Based on the presence in their N-terminal domain of either

Toll/Interleukin-1 receptor (TIR) or coiled-coil (CC) motifs, NLR proteins fall into two major

structurally distinct sub-classes named TIR-NB-LRR (TNL) or CC-NB-LRR (CNL),

respectively (Bonardi et al., 2012). Other R proteins belong to the group of eLRR

(extracellular LRR) domain proteins. This includes mainly the Receptor-Like Protein (RLP)

family characterized by an extracellular LRR domain, a TMD and a short cytoplasmic domain

(Muthamilarasan and Prasad, 2013) (Figure 7). The best characterized R proteins of this

class are Cf proteins which confer resistance of tomato to the fungal pathogen Cladosporium

fulvum (Rivas and Thomas, 2005).

This mode of recognition leads to co-evolutionary dynamics between the plant and the

pathogen that are quite different from those associated to PTI as, in contrast to PAMPs,

effectors are dispensable molecules although usually necessary for pathogenicity. This

specific recognition represents the second line of active defence known as specific

Figure 8. Model of integrated decoys in NLR protein pairs (From Cesari et al., 2014). Pathogen effectors target host proteins for manipulation in order to promote infection.

(A) Indirect recognition of effectors occurs when target proteins are guarded by host NLR

proteins,

(B) or if duplicated target genes evolve to encode decoy proteins monitored by host NLRs.

(C) Alternatively, the decoy may be integrated into the structure of the receptor component of an

NLR pair, allowing effector recognition by direct binding.

Pathogen genotype

Plant genotype Avr (Avirulent) avr (virulent)

R (Resistant) Resistance (HR) Disease

r (Susceptible) Disease Disease

Table 1. Genetic model of the gene for gene theory (From Flor, 1971). The resistance of the plant, often associated with the HR, is only established if the plant carries an

R gene corresponding to an Avr gene in the pathogen. In all other cases, disease occurs.

(A) Guard model

(B) Decoy model

(C) Integrated decoy model

23

resistance and commonly termed effector-triggered immunity (ETI) (Jones and Dangl, 2006)

(Figure 3).

Variation in host resistance is often controlled by the segregation of single R genes

(Hammond-Kosack and Jones, 1997). The genetic interaction underlying the induction of this

type of resistance is typically explained by the "gene for gene" model (Flor, 1971). This

classic concept is based on the observation that a plant carring a dominant resistance gene

(R) is resistant when it interacts with a pathogen that expresses a dominant and

complementary effector protein historically called avirulence protein (Avr). In the absence

of the R protein and/or the corresponding avirulence protein, the pathogen is not detected by

the plant, which results in the establishment of disease (Table 1) (Gassmann and

Bhattacharjee, 2012). The first biochemical interpretation of this hypothesis was a receptor–

ligand model that implies that plants activate defence mechanisms upon R-protein-mediated

recognition of pathogen-derived Avr gene products (Table 1).

More recent studies showed that direct recognition of Avr gene products by R proteins is the

exception rather than the rule and that a more prevalent mode of recognition exists that

involves indirect interaction mediated by accessory-proteins that the immune receptor

associates with and in which it recognizes effector-induced modifications. These accessory

proteins that mediate indirect recognition may either be direct virulence targets of the effector

(guard model) (Dangl and Jones, 2001, Dodds and Rathjen, 2010) (Figure 8-A) or decoy

proteins that the plant has evolved to mimic real effector targets (decoy model) (Hou et al.,

2011, van der Hoorn and Kamoun, 2008) (Figure 8-B). In some cases,a decoy protein fused

to a member of an NLR pair may act as bait to trigger defence signalling by a second NLR

member upon effector binding (integrated decoy model) (Cesari et al., 2014, Delga et al.,

2015) (Figure 8-C).

Figure 9. Major signalling mechanisms in plant defence. PAMPs perception by PRR induces rapid (seconds) immune receptor complex formation at the

plasma membrane and different auto- and transphosphorylations of the actors involved (1). BIK1

becomes quickly phosphorylated and released from the PRR complex (2). Phosphorylated Botrytis-

induced kinase1 (BIK1) has a higher binding affinity for respiratory burst oxidase homolog D

(RBOHD) and phosphorylates it (3). At the same time, BIK1 also activates Ca2+ channel(s) and

induces Ca2+ influx. A Ca2+ burst occurs (30 s to 2 min) and reaches a peak at around 4–6 min (4).

This Ca2+ influx induces opening of other membrane channels (influx of H+, efflux of K+ and Cl–),

which leads to extracellular medium alkalinization (1 min) and depolarization of the plasma

membrane (1–3 min) (5). A ROS burst then rapidly occurs (2–3 min) via RBOHD and peaks at

around 10–14 min (6). Full activation of RBOHD requires phosphorylation by BIK1 and Ca2+-

induced Calcium-dependent protein kinases (CDPKs) (6). Ca2+ also regulates RBOHD through direct

binding or modification of the protein (6). RBOHD produces O2.– in the apoplast, which is

converted into H2O2 by superoxide dismutases (SOD) (7). H2O2 can enter the cytosol and the

different organelles of the cell and is capable of inducing cytosolic Ca2+ elevations (8). 14-3-3

proteins modulate the activity of RBOHD and CDPKs (9). Effector recognition induces rapid

immune response (10). Mitogen-activated protein kinase (MAPK) modules are activated in a few

minutes leading to transcription factor (TF) activation (11). TFs participate in the regulation of

several thousand genes (12). SA, JA, and ET signalling pathways then contribute to downstream

regulation of gene expression (13). Crosstalks also occur with other phytohormones (14). This

complex signalling network finally leads to the implementation of plant-induced defences, such as

the production and secretion of antimicrobial compounds and the generation of toxic ROS (15).

Arrows denote enzymatic pathways, transport, or regulation (see text for more details). ABA:

abscisic acid; BR: brassinosteroid; CK, cytokinin; ET: ethylene; GA: gibberellic acid; JA: jasmonic

acid; P, Phosphorylation; SA: salicylic acid.

24

1.2.3. Signalling events during plant defence responses

Upon pathogen perception, the induction of defence mechanisms relies on a complex and

interconnected network of signalling events (Bigeard et al., 2015). PTI and ETI share a set of

downstream signalling components with distinct activation dynamics and amplitudes (Tsuda

and Katagiri, 2010).

Transient changes in the ion permeability of the plasma membrane appear to be a common

early element in defence signalling that stimulates ion fluxes across the plasma membrane

(Ca2+ and H+ influx, K+ and Cl– efflux) resulting in elevation of cytosolic calcium ([Ca2+]cyt)

(Reddy et al., 2011), concomitant membrane depolarization (Jeworutzki et al., 2010),

medium alkalinization and cytoplasmic acidification (Bricchi et al., 2013) (Figure 9). Stimulus-

specific responses are explained by the concept of the “Ca2+ signature” (McAinsh and

Pittman, 2009), where duration, amplitude, frequency and spatial distribution are thought to

encode stimulus-specific information that is decoded by various Ca2+-binding proteins

including calmodulins (CaMs) and CaM-Like proteins (CMLs), which regulate the production

of nitric oxide (NO) (Ma, 2011). In addition, roles of Ca2+/CaM interacting proteins such as

CaM binding protein (CBP) and CaM-binding transcription activators (CAMTAs) have been

identified in plant defence signalling cascades (Ma, 2011). Furthermore, calcium-dependent

protein kinases (CDPKs) emerged as important Ca2+ sensor proteins in transducing

differential Ca2+ signatures, triggered by PAMPs or effectors and activating complex

downstream responses (Ma, 2011) (Figure 9). For example, overexpression of Arabidopsis

AtCDPK1 confers broad-spectrum resistance to both bacteria and fungi (Coca and San

Segundo, 2010). Moreover, emerging evidence suggests that specific and overlapping

CDPKs phosphorylate distinct substrates in PTI and ETI to regulate diverse plant immune

responses (Boudsocq et al., 2010) (Figure 9).

Subsequently, ROS production often referred to as “ROS burst” is an additional early

response, starting only a few minutes after PAMP treatment and at a much slower pace

during ETI. In Arabidopsis, the plasma membrane-localized nicotinamide adenine

25

dinucleotide phosphate-oxidase (NADPH), named respiratory burst oxidase homolog D

(RBOHD), is predominantly responsible for ROS burst in response to pathogen attack

(Torres et al., 2002). RBOHD is mainly controlled by Ca2+ via direct binding to EF-hand

motifs and phosphorylation by CDPK (Dubiella et al., 2013). Recent studies have, however,

revealed a critical role for a Ca2+-independent regulation of RBOHD (Kadota et al., 2014, Li

et al., 2014). Biochemical analyses showed that RBOHD associates with the PRR complex in

vivo, and that BIK1 directly phosphorylates RBOHD upon PAMP perception (Li et al., 2014,

Kadota et al., 2014). Furthermore, abrogation of ROS accumulation, either in the rbohD

mutant or through inhibitor application, led to loss of the second peak of PAMP-induced

biphasic Ca2+ cytoplasmic changes, demonstrating a positive feedback activation of ROS on

Ca2+ signalling (Ranf et al., 2011).

14-3-3 proteins also participate in immune signal transduction. They were shown to interact

with known components of immune signal transduction, such as NtRBOHD (Elmayan et al.,

2007) or CDPKs (Camoni et al., 1998, Lachaud et al., 2013), and modulate their activity

(Figure 9). The signal is further transduced by activation of mitogen-activated protein kinase

(MAPK) proteins, typically functioning in a phosphorylation cascade that involves at least

three interlinked protein kinases (MAPKKK, MAPKK and MAPK) which are sequentially

activated by phosphorylation. Interestingly, a reduction of Ca2+ oscillations was observed

upon MAMP perception using inhibitors of serine/threonine protein kinases and MAPK

kinases, suggesting that Ca2+-PTI signalling is in part dependent on MAPK cascades (Ranf

et al., 2011, Boudsocq et al., 2010). However, although other data suggest that MAPK

activation may be independent of CDPKs (Boudsocq et al., 2010). Although several immune-

related MAPK substrates have been identified that are involved in diverse cellular functions

(Bigeard et al., 2015), almost half of bona fide immune MAPK substrates are Transcription

Factors (TFs). In this context, the identification of MYB TFs, such as MYB41 or MYB44

(Hoang et al., 2012, Nguyen et al., 2012, Persak and Pitzschke, 2013), or WRKY TFs, such

as WRKY33 or WRKY1 (Ishihama and Yoshioka, 2012), as targets of MAPK activity

Figure 10. Classic model established for the hormonal control of the plant defence

(Addapted from David De Vleesschauwer et al., 2013). Plant resistance is mainly controlled by two antagonistic hormonal pathways, those of SA and

JA/ET. They respectively promote resistance against biotrophic and necrotrophic pathogens. Auxin

induces the JA/ET pathway whereas, cytokinins and gibberellic acid (or gibberellin) induce the SA

pathway. BRs regulate plant immunity through an SA-independent pathway. ABA appears to act as

a negative regulator of defence against biotic stress, but plays a crucial role in responses to abiotic

stresses. The arrows indicate activation or positive interaction and blocked lines indicate

repression or negative interaction. Hormone abbreviations: ABA: abscisic acid; BR:

brassinosteroid; CK, cytokinin; ET: ethylene; GA: gibberellic acid; JA: jasmonic acid; SA: salicylic

acid.

26

highlights the important role of MAPKs in the transcriptional reprogramming directing the

plant defence response (Figure 9).

Transduction of this signalling cascade to the nucleus allows the activation of TFs (Tsuda

and Somssich, 2015) and results in the synthesis of molecules involved in plant immunity

such as PR proteins and antimicrobial compounds. Transcriptional regulators typicallty act

within larger networks, in which they function cooperatively or antagonistically to regulate the

expression of genes involved in feed-forward and feedback loops. The activity of these

transcriptional regulators is orchestrated by a blend of signalling hormones of which SA, JA,

and ET are particularly important (Figure 10) (Pieterse et al., 2009). Although other

phytohormones, such as abscisic acid (ABA), auxin (IAA), brassinosteroids (BR), cytokinins

(CK) and giberellins (GA) are also involved in the regulation of plant disease resistance

(Robert-Seilaniantz et al., 2011), for simplicity reasons, here we will focus on defence-related

roles of SA, JA and ET. SA acts as a signal to activate plant defence responses both locally

and systemically (An and Mou, 2011). Arabidopsis mutants with defects in SA biosynthesis

genes like isochorismate synthase1 (ics1) display reduced PR1 gene expression upon

infection and are more susceptible to certain pathogens (Wildermuth et al., 2001, Garcion et

al., 2008, Spoel et al., 2007). Along the same lines, in planta overexpression of the

Pseudomonas putida NahG gene encoding an SA hydroxylase that degrades SA, results in

increased disease susceptibility and abolished PR1 gene expression confirming a role for SA

in resistance (Delaney et al., 1994).

Numerous examples of positive and negative crosstalk between SA, JA and ET signalling

have been reported (Pieterse et al., 2009, De Vleesschauwer et al., 2013). Mutations like

ethylene insensitive 2 (ein2) and coronatine-insensitive protein 1 (coi1), which affect ET and

JA signalling pathways respectively, result in increased susceptibility to Botrytis and failure to

induce JA-responsive marker genes (Manners et al., 1998, Thomma et al., 1998, Thomma et

al., 1999). It is now evident that SA-dependent pathways play a major role in defence against

biotrophic pathogens, whereas pathogens with a necrotrophic lifestyle are commonly

Figure 11. Development of the hypersensitive response (HR) on tobacco leaf in

response to Pseudomonas syringae pv. tomato DC3000

(http://www.sidthomas.net/images/hypersensitive.jpg).

27

deterred by defences that are controlled by JA and ET (Glazebrook, 2005). Moreover, JA

also antagonizes SA-mediated responses and vice versa (El Oirdi et al., 2011). ET fine-tunes

appropriate defence responses by inhibiting JA-mediated defence suppression by SA (Leon-

Reyes et al., 2010) adding a supplementary level of control. Agonist and antagonist crosstalk

between SA, JA and ET is presented in Figure 10. For additional details about the role of

these phytohormones on plant defence regulation, see the review article at the end of this

section (Buscaill and Rivas, 2014).

The signal transduction cascade and transcriptional reprogramming triggered during ETI

typically results in the development of a rapid and localized programmed cell death (PCD)

called hypersensitive-response (HR). The HR is triggered in plant cells directly in contact

with, or in close proximity of, the invading pathogen. This localized ‘cell suicide’ leads to

survival of the plant by stopping the spread of biotrophic pathogens beyond the site of

attempted infection. The HR and, more particularly, the role of proteases in the regulation of

plant cell death-related processes are discussed in the following section.

1.3. Plant proteases: roles in life and death during plant defence signalling

PCD is a fundamental process of life. Many different forms of PCD have been described in a

remarkable variety of cell types, tissues, and organs. PCD occurs as an integral part of plant

development (dPCD), sculpting structures or deleting unwanted tissues (Van Hautegem et

al., 2015). In addition, plant reactions to biotic and abiotic environmental challenges also

invove the development of PCD (ePCD) (Lam, 2004). Indeed, as mentioned above,

incompatible interactions are frequently associated with the development of the HR

(Greenberg, 1997) (Figure 11), which results in necrotic lesions located at the points of

pathogen entry thus confining the pathogen and avoiding its spread throughout the plant (Wu

et al., 2014). This phenomenon also prevents pathogens from getting access to essential

water and nutrients, which usually stops the infection. Nevertheless, this strategy, efficient

against biotrophic and hemi-biotrophic pathogens, is not adapted to resist an attack by

28

necrotrophs (Lorang et al., 2012). In addition, virulent attacks do not trigger an HR response

in plant cells (Wu et al., 2014).

Before the HR is triggered, ROS and NO rapidly accumulate in cells and trigger a cascade of

biochemical events resulting in the localized death of host cells (Mur et al., 2008). These

events are followed by destruction of the organelles, collapse of the plasma membrane and

its separation from the cell wall, which is left deformed after the leakage of the cell contents

into the apoplast (Lam, 2004). The HR is characterized by several features in common with

animal apoptosis such as the change in nuclear morphology, chromatin condensation, DNA

fragmentation, release of cytochrome c from mitochondria and cell shrinkage (Lam, 2004,

Coll et al., 2011).

The involvement of proteases in plant defence signalling, and particularly during the HR,

was predicted on the basis of a “death pathway” conserved between plants and animals.

Although caspase-encoding genes have not been formally identified in plants (Enoksson and

Salvesen, 2010), several studies have shown that defence-associated cell death induced in

various plant species can be blocked by human caspase inhibitors, suggesting the existence

of a caspase-like activity during plant PCD (Carmona-Gutierrez et al., 2010, del Pozo and

Lam, 2003, Belenghi et al., 2003, D'Silva et al., 1998, Rozman-Pungercar et al., 2003).

Although careful interpretation of these results is needed, as these studies were based on

the use of human caspase inhibitors, a correlation between the induction of protease-

encoding genes and the establishment of plant defence responses has been demonstrated

(Hückelhoven et al., 2001, Avrova et al., 1999, Solomon et al., 1999, Huang et al., 2015,

Hoeberichts et al., 2003, Iakimova et al., 2013, Avrova et al., 2004, Pautot et al., 1993, Gu et

al., 1996, Liu et al., 2001, Kemp et al., 2005, Schiermeyer et al., 2009). A significant number

of studies have revealed that various plant proteases actively participate in the recognition of

pathogens and in the induction of effective local and systemic defence responses (van der

Hoorn and Jones, 2004, Xia, 2004). Moreover, effector proteins directly target plant

proteases to inhibit their activity (Tian et al., 2004, Tian et al., 2005, Tian and Kamoun, 2005,

Figure 12. Cleavage mechanisms of the four major catalytic classes of proteases

(From van der Hoorn, 2008).

(A) The substrate protein (green) binds via amino acid residues (R) to the substrate

binding site of the protease (gray) by interacting with substrate (S) pockets of the

enzyme. The scissile peptide bond is adjacent to a carbonyl group, which is polarized

by the enzyme by stabilizing the oxyanion hole (blue); this makes the carbonyl carbon

vulnerable for nucleophilic attack.

(B) The major differences between the catalytic classes are the nature of the nucleophile

and oxyanion stabilizer. Cysteine and serine proteases use a Cys or Ser residue as

nucleophile, activated by histidine (His) in the active site. The oxyanion hole is usually

stabilized by two residues in the backbone of the protease. Metalloproteases and

aspartic proteases use water as nucleophile, activated by electrostatic interactions

with the metal ion (Me2+) or aspartate (Asp), respectively. The oxyanion of these

proteases is stabilized by Me2+ and Asp, respectively.

(B)

(A)

Figure 13. Classification and number of the catalytic types of Arabidopsis proteases

(From van der Hoorn and Jones, 2004). A total of 488 proteases can be distinguished within the encoded Arabidopsis genome, most of

which are also represented in the MEROPS database. Proteases can be subdivided into catalytic

types on the basis of the residues used to cleave a peptide bond. The Arabidopsis genome

encodes 198 serine (S), 112 aspartic (A), 95 cysteine (C), 80 metallo (M) and 12 threonine (T)

proteases. Each protease class consists of several clans of proteases, which are identified by a

letter following the catalytic class (e.g. clan CA within class C). Members of a single clan are

believed, on the basis of their conserved tertiary structure and order and spacing of catalytic

residues, to have a common evolutionary origin. Proteases of clans starting with the letter ‘P’ can

have different catalytic residues. Each clan of proteases consists of several families which are

identified by a number following the catalytic type (e.g. family C1 within clan CA). Not all clans and

families of plant proteases are represented in Arabidopsis. The number of proteases belonging to

each family is indicated by bars, and the classification of well-studied proteases is indicated.

Table 2. Protease in plant defence (Part I).

Class Clan Family Group Name

Plant

species(a) Function in plant immunity(b) Reference

Cy

ste

ine

Pro

tea

ses

CA C1

Pa

pa

in-l

ike

pro

tea

ses

AtCathB At Positive regulator of basal resistance against

virulent Ps

McLellan et al., 2009

AtCEP1 At Positive regulator of defence response to Ec Höwing et al., 2014

C14 Nb Positive regulator in defence response to Pi Kaschani et al., 2010,

Bozkurt et al., 2011

CYP St Induced expression upon Pi infection which

correlates with resistance

Avrova et al., 1999

Mir1 Zm Positive regulator of resistance to Sf Pechan et al., 2000, Pechan

et al., 2002

NbCathB Nb Positive regulator of disease resistance induced by

Ea and Ps

Gilroy et al., 2007

NbCYP1

and

NbCYP2

Nb Positive regulators in host defence responses to Cd Hao et al., 2006

Pip1 Le Induced extracellular accumulation during infection

by Pi

Tian et al., 2007

RCR3 Le Positive regulator of resistance to Cf Krüger et al., 2002

RD21 At Positive regulator of defence response to Bc Shindo et al., 2012

StCath St Induced expression during HR response to Pi

infection

Avrova et al., 2004

CD

C14

Me

taca

spa

ses

AtMC1

and

AtMC2

At Positive and negative regulator of the HR response

to Ps, respectively

Coll et al., 2010

AtMC2-6 At Negative regulators of cell death in response to Bc,

Bt and Be

VAN Baarlen et al., 2007

AtMC4 At Negative regulator of cell death induction by Ps and

mycotoxin fumonisin B1

Watanabe and Lam, 2011

AtMC7

and

AtMYB8

At Positive regulators of cell death in response to Bc,

Bt and Be

VAN Baarlen et al., 2007

CaMC9 Ca Positive regulator of PCD and defence in response

to Xcv

Kim et al., 2013

LeMCA1 Le Induced expression during Bc infection Hoeberichts et al., 2003

NbMCA1 Nb Positive regulator of basal defence in response to

Cd

Hao et al., 2007

OsMC1-8 Os Induced or repressed expression upon Mo and Xoo

infection

Huang et al., 2015

TaMCA4 Ta Positive regulator of PCD and defence response to

Pt

Wang et al., 2012

C13

VP

Es

VPE1 At Positive regulator of defence response to Bc and Bt VAN Baarlen et al., 2007

VPE2 At Negative regulator of defence response to Bc VAN Baarlen et al., 2007

VPE3 At Negative regulator of defence response to Ha Misas-Villamil et al., 2013

VPE3 At Positive regulator of resistance and plant basal

defence to Ps, Bc and TMV

Rojo et al., 2004

VPE1a

and VPE1b

Nb Positive regulators of the HR during TMV infection Hatsugai et al., 2004

VPEs Nu Positive regulators of the cell death induced by the

fungal AAL-toxin from Aa

Mase et al., 2012

VPEs Md Induced expression during the HR-response to Ea Iakimova et al., 2013

aAa, A. alternata f.sp. lycopersici; As, Avena sativa; At, Arabidopsis; Ca, pepper; Gm, Glycin max; Le, Lycopersicon esculentum; Md,

M. domestica; Me, Manihot esculenta; Nb, N. benthamiana; Nt, tobacco; Nu, N. umbratica; Os, rice; Sl, tomato; St, Solanum

tuberosum; Ta, Wheat; Vs, Vitis species; Zm, Zea mays. bBc, B. cinerea; Bt, B. tulipae; Be, B.elliptica; Cd, C. destructivum; CEV, Citrus Exocortis Viroid; Cf, C. fulvum; Ea, E. amylovora; Ec, E.

cruciferarum; Fs, F. solani; Ha, H. arabidopsidis; Mo, M. oryzae; Pi, P. infestans; Ps, P. syringae; Pt, P. striiformis f. sp. tritici; Sf, S.

frugiperda; TMV, Tabacco Mosaic Virus; Xcv, X. campestris pv. campestris; Xoo, X. oryzae pv. oryzae.

Table 2. Protease in plant defence (Part II).

Class Clan Family Group Name

Plant

species(a) Function in plant immunity(b) Reference

Aspartic

Proteases AD A1

Pepsin-like

proteases

CDR1 At Positive regulator of resistance to Pseudomonas

strains

Xia et al., 2004

StAPs St Kill spores of Fs and Pi Mendieta et al., 2006

Me

tall

op

rote

ase

s MF M17

Acidic

leucine

amino-

peptidases

LapA Le Induced expression after Ps infection Pautot et al., 1993,

Gu et al., 1996

Le Activity increases during Ps infection Pautot et al., 2001

MA M10A

Ma

trix

me

tall

o-

pro

tein

ase

s

GmMMP2 Gm Induced expression during Ps and Phs infection Liu et al., 2001

GmMMP2

ortholog

Me Induced expression during the HR in response to

Ps

Kemp et al., 2005

NbMMP1 Nb Positive regulator of defence against Ps infection Kang et al., 2010

NtMMP1 Nt Induced expression in defence response to Ps

and At

Schiermeyer et al.,

2009

Se

rin

e p

rote

ase

s

SB S8

Su

bti

lisi

n-l

ike

pro

tea

ses

AtKTI1 At Negative regulator of PCD in response to Ps Li et al., 2008

AtSBT3.3 At Positive regulator of primed immunity against Ps Ramirez et al., 2013

AtSBT5.2 At Negative regulator of AtMYB30-mediated

defence to Ps

This work

GmPep914

and

GmPep890

Gm Activate defence-related genes Yamaguchi et al.,

2011

GmSubPep Gm Activates defence-related genes Pearce et al., 2010

P69B Le Specifically targeted by Pi effectors (EPI1 and

EPI10)

Tian et al., 2004, Tian

et al., 2005

P69B and

P69C

Le Induced expression during CEV and Ps infection Tornero et al., 1997,

Jorda et al., 1999

phytaspases Nt Present a caspase-like activity during At infection Chichkova et al., 2004

phytaspases Nt and

Os

Positive regulator of the HR-responses to TMV Chichkova et al., 2010

saspases As Secreted during PCD induced by Cv Coffeen and Wolpert,

2004

Subtilisin Vv Constitutively expressed in the resistant

genotype and induced after Pv inoculation

Figueiredo et al.,

2012, Monteiro et al.,

2013

UPI At Positive regulator of defence to Ab, Bc and Ti Laluk and Mengiste,

2011

aAs, Avena sativa; At, Arabidopsis; Gm, Glycin max; Le, Lycopersicon esculentum; Me, Manihot esculenta; Nb, N.benthamiana; Nt,

tobacco; Nu, N. umbratica; Os, rice; St, Solanum tuberosum; Vv, Vitis vinifera; Zm, Zea mays. bAb, Alternaria brassicicola; At, A. tumefaciens; Bc, B. cinerea; CEV, Citrus Exocortis Viroid; Cf, C. fulvum; Cv, C. victoriae; Fs, F.

solani; Gc, G. cichoracearum; Ha, H. arabidopsidis; Phs, Phytophtora sojae; Pv, P. viticola; TMV, Tobacco Mosaic Virus; Tn, T. ni.

Figure 14. Protein structure of proteases.

Schematic representation of aspartic endopeptidase (A), cysteine endopeptidase (B), metallo

endopeptidase (C) and serine endopeptidase (D) precursor proteins. The protease precursors

present a signal peptide (black boxes) at the N-termini. The proprotein precursors have a

cleavable prodomain (striped boxes). Plant type I metacaspases present a zinc finger motif (blue

box). After the removal of the prodomain, proprotein precursors are converted into the

respective mature enzymes (open boxes). Identities and positions of catalytic amino acid residues

are shown above each diagram. Single-letter abbreviations for the amino acid residues are as

follows: C, Cysteine; D, Aspartate; H, Histidine; N, Arginine; S, Serine. PA, Protease associated

domain.

D D

Carboxypeptidase-like (Clan SC, Family S10)

Subtilisin-like (Clan SB, Family S8)

A. thaliana (type I)

A. thaliana (type II)

Aspartic proteases ( Class A)

Cysteine proteases (Class C)

Metallo-proteases (Class M)

Serine proteases (Class S)

Metacaspases (Clan CD, Family C14)

(A)

(B)

(C)

(D)

HHH

C H N

D H N S

H C

VPEs

(Clan CD, Family C13)

Papain-likes (Clan CA, Family C1)

Pepsin-like (Clan, Family A1)

H C

S D H

Matrix

metalloproteinases (Clan MA, family M10A)

PA

29

Tian et al., 2007). In addition, a significant number of pathogen secreted proteins are active

proteases that interact with the host immune system (Nimchuk et al., 2007, Cheng et al.,

2015, Jashni et al., 2015). Together these data underline the importance of proteases and

proteolytic activity during plant-pathogen interactions.

In contrast to exopeptidases that hydrolyse terminal residues, proteases are endopeptidases

that preferentially hydrolyse internal peptide bonds in polypeptide chains (Barrett, 1994).

Depending on the amino acid involved in this reaction, proteases are classified into four

catalytic classes: cysteine proteases, serine proteases, metalloproteases, and aspartic

proteases (Barrett and Rawlings, 1995) (Figure 12). This classification also helps predict the

effect of protease inhibitors (PIs) on the members of each class of proteases. In the

MEROPS classification (http://merops.sanger.ac.uk), these catalytic classes are subdivided

into Clans, and Clans are further subdivided into Families based on their structural and

evolutionary relationships (Rawlings et al., 2014) (Figure 13).

The involvement of proteases in defence-associated cell death responses has been

extensively reported in the literature. To illustrate this involvement, and due to space

limitations, I have made a selection of particularly important examples within each of the

major four protease families. Since my PhD work has focused on the study of the

Arabidopsis subtilase AtSBT5.2, the subtilase family of serine proteases is described in more

detail. For a more complete view of proteases involved in plant defence, refer to Table 2 and

references therein.

1.3.1. Aspartic proteases

Aspartic proteases are a class of widely distributed proteases present in animals, microbes,

viruses and plants (Rawlings and Barrett, 1995, Davies, 1990). Their active site presents a

catalytic dyad of aspartate residues (D) (Figure 14), supporting a water molecule that acts as

the nucleophile during proteolysis (Figure 12). Aspartic acid proteases are the second most

abundant plant proteases, with members divided into three large families that belong to two

30

clans (AA and AD) (van der Hoorn, 2008) (Figure 13). These enzymes are produced as

preproproteases and often secreted from cells as inactive, glycosylated enzymes that

activate autocatalytically at acidic pH (Davies, 1990). Protease sequences of the A1 family

(pepsin-like protease) are known only from eukaryotes (Barrett and Rawlings, 1995) and fifty-

nine A1 proteases were identified in Arabidopsis (Beers et al., 2004). Interestingly, this class

of proteases has been identified as a player in developmental plant PCD. For example, in

barley, genes encoding aspartic proteases are specifically expressed in nucellar cells during

degeneration (Chen and Foolad, 1997). Compared to plant subtilases and cysteine

proteases, fewer aspartic proteases have been studied. Indeed, a biological role is only

known for some aspartic proteases in family A1 of clan AA (van der Hoorn, 2008, Mendieta

et al., 2006). The secreted aspartic pepsin-like protease constitutive disease resistance 1

(CDR1) represents one of the best characterised examples of the involvement of this type of

proteases in plant immunity. CDR1, which was identified during an activation tagging

experiment upon infection with Pseudomonas syringae, acts in disease resistance signalling

and its overexpression leads to constitutive disease resistance to virulent strains of P.

syringae (Xia et al., 2004). The CDR1 protein displays proteolytic activity and accumulates in

the extracellular space of plant cells during pathogen attack (Xia et al., 2004). Through its

protease activity, that is required for CDR1 function, CDR1 releases an endogenous peptide

that was proposed to act as a mobile signal to elicit systemic SA-dependent resistance

responses (Xia et al., 2004).

1.3.2. Cysteine proteases

Cysteine proteases represent a well-characterized class of proteolytic enzymes widely

distributed in living organisms that use a catalytic Cys as a nucleophile during proteolysis

(Figure 12). Plant genomes encode approximately 140 cysteine proteases that belong to five

clans (Rawlings et al., 2014). Twelve families are known in Arabidopsis (Figure 13). The

structures of proteases from different clans are distinct: clan CA contains proteases with a

papain-like fold, whereas CD proteases present a caspase-like fold (Figure 14). Plant

31

cysteine proteases are associated to biotic stress resistance during bacterial, oomycete and

fungal infections (Martínez et al., 2012, Avrova et al., 1999, Avrova et al., 2004, Kim et al.,

2013, McLellan et al., 2009, Pechan et al., 2000, Pechan et al., 2002, Bozkurt et al., 2011,

Kaschani et al., 2010, Shindo et al., 2012, VAN Baarlen et al., 2007, Hao et al., 2007,

Watanabe and Lam, 2011, Wang et al., 2012, Misas-Villamil et al., 2013, Mase et al., 2012,

Rojo et al., 2004) and to the control of PCD (Solomon et al., 1999). Studies by activity

profiling further proved the implication of cysteine proteases in plant defence (van der Hoorn

and Kaiser, 2012, Misas-Villamil et al., 2013). Among the different reported examples, some

cysteine proteases have been studied in detail and are discussed below.

Required for Cladosporium resistance3 (Rcr3), a secreted papain-like cysteine protease

from tomato, with proven proteolytic activity and essential for resistance to the fungal

pathogen Cladosporium fulvum, is one of the best characterized (Krüger et al., 2002). Rcr3 is

required for the function of the tomato Cf-2 receptor-like protein (Cladosporium fulvum

resistance-2) against the fungal pathogen Cladosporium fulvum carrying the Avr2 avirulence

gene (Krüger et al., 2002). Moreover, in the tomato genome the Rcr3 gene maps to the

Phytophthora inhibited protease 1 (PIP1) locus. PIP1 is a pathogenesis-related protein

closely related to Rcr3 and its transcript is up-regulated upon pathogen attack. As RCR3, the

PIP1 protein accumulates in the apoplast (Krüger et al., 2002, Tian et al., 2007). Microbial

effector proteins (named cystatin-like protease inhibitors) have been shown to target papain-

like Cys proteases during infection and inhibit plant defences. For example, the secreted

peptide Avr2, a cysteine-rich protein from C. fulvum, is able to physically interact with Rcr3

and inhibit its protease activity (Rooney et al., 2005). These results suggest that Rcr3 is

rather a virulence target of Avr2 that is guarded by the Cf-2 resistance protein to monitor

pathogen entry (Jones and Dangl, 2006, Rooney et al., 2005). Inhibition of Rcr3 by protease

inhibitor E-64 or the absence of Rcr3 activity in Rcr3 mutants cannot trigger the resistance

response mediated by Cf-2, suggesting that neither the product nor substrates of Rcr3, but

the Avr2-Rcr3 complex or a specific conformational change in Rcr3, is required to trigger the

32

resistance response (Rooney et al., 2005). Moreover, the affinity of the Avr2 mutants for

Rcr3 correlates with their ability to trigger a Cf-2-mediated HR (Van't Klooster et al., 2011). In

addition, similar to Avr2, EPIC1 and EPIC2B effectors from P. infestans bind and inhibit Rcr3

(Song et al., 2009).

An additional example of apoplastic papain-like cysteine proteases is CathepsinB from

Nicotiana benthamiana (NbCathB), activated upon secretion and required for the HR

induced by nonhost pathogens (Gilroy et al., 2007).

Another papain-like cysteine endopeptidase involved in pathogen defence is the Arabidopsis

cysteine endopeptidase 1 (AtCEP1) which is expressed in leaves in response to biotic stress

stimuli. atcep1 knockout mutants showed enhanced susceptibility to powdery mildew caused

by the biotrophic ascomycete Erysiphe cruciferarum. A translational fusion protein of AtCEP1

under control of the endogenous AtCEP1 promoter rescued the pathogenesis phenotype

demonstrating the function of AtCEP1 in restriction of powdery mildew (Höwing et al., 2014).

Following the discovery of caspases in animals, homologous proteases were searched in

plants. Exhaustive bioinformatic analyses showed that caspases are absent from plant

genomes, but plants contain proteases sharing sequence homology or at least structural

homology to the animal caspases. These proteins were called caspase-like proteases and

further classified within the CD clan of proteases as metacaspases (MCs) and vacuolar

processing enzymes (VPEs). Both families have been involved in plant defence

mechanisms. In Arabidopsis, AtMC1 is a positive regulator of cell death and requires

conserved caspase-like putative catalytic residues for its function whereas AtMC2 negatively

regulates cell death (Coll et al., 2010). In addition, a role of the cell-death/calcium-dependent

protease AtMC4 in releasing a mature Pep peptide from its precursor AtPROPEP1 upon cell

damage has been recently uncovered. Moreover, in vivo cleavage of AtPROPEP1 by AtMC4

is calcium-dependent and inhibited by metacaspase inhibitors (Stael S. Oral communication,

workshop on Plant Organellar Signalling, 2015, Primosten, Croatia). Finally, silencing of

33

VPEs in N. benthamiana compromises the HR triggered by Tobacco Mosaic Virus (TMV) in

plants carrying the TMV-resistance gene N (Hatsugai et al., 2004).

1.3.3. Metalloproteases

Metalloproteases contain catalytic metal ions that activate water for nucleophilic attack while

stabilizing the oxyanion hole (Figure 12). Plant genomes encode approximately 100

metalloproteases that belong to 18 families (Rawlings et al., 2014). These families are

diverse and divided in ten evolutionarily unrelated clans. Plant metalloprotease families

usually contain fewer than 20 members (Figure 13). Several metalloproteases families have

been involved in plant defence (van der Hoorn, 2008) and one of these proteases is

discussed below.

Within the metalloprotease family, matrix metalloproteinases (MPPs, Clan MA, family M10A)

are characterized by the presence of a highly conserved catalytic domain containing a zinc

binding motif (HEXXHXXGXXH) in which the two first histidine (H) residues are ligands of a

single zinc ion (van der Hoorn, 2008). Plant MMPs present a minimal domain structure, are

synthesized as prepro-enzymes and contain a signal peptide (SP) (Nagase and Woessner,

1999) (Figure 14). Apart from their involvement in development, senescence, PCD and

abiotic stresses, plant MMPs have been demonstrated to regulate host-pathogen interactions

(Marino and Funk, 2012, Pautot et al., 2001). For example, in the tobacco suspension line

BY-2, expression of NtMMP1 encoding a secreted MPP was induced after treatment with the

bacterial pathogens P. syringae and A. tumefaciens and may thus play a role in defence

against pathogens at the cell periphery (Schiermeyer et al., 2009). In addition, Kang and

collaborators proposed a positive role of NbMMP1 from N. bethamiana in defence against

compatible and incompatible bacterial infections (Kang et al., 2010).

1.3.4. Serine proteases

Serine proteases use a Ser residue at their active site as a nucleophile (Figure 12). With

more than 200 members, serine proteases belong to the largest class of proteases in plants

34

(Rawlings et al., 2014). Plant serine proteases are divided into 14 families that belong to nine

evolutionarily unrelated clans (van der Hoorn, 2008) (Figure 13). Three of them, the

chymotrypsin PA (S), subtilisin (SB) and carboxypeptidase D (SC) clans, share a common

reaction mechanism based on a well-characterized ‘‘catalytic triad’’ comprising a serine, an

aspartic acid, and a histidine residue (Schaller et al., 2012). Families S8 and S10 represent

the largest serine protease families in plants, each containing approximately 60 members

(Figure 13). Biological functions of serine proteases have been described for some

carboxypeptidases (BRS1 and SNG1/2; family S10, clan SC), subtilases (SDD1 and ALE1;

family S8, clan SB) and plastid-localized members of the S1, S26, and S14 families (DegPs,

Plsp1, and ClpPs) (van der Hoorn, 2008).

· Carboxypeptidase-like proteases

Serine carboxypeptidase protease-like proteins (SCPLs; family S10, clan SC) are widely

distributed proteases identified in higher organisms. They contain a catalytic triad in the

primary sequence order Ser, Asp, His (Figure 14). Nearly 60 SCPLs are encoded by the

Arabidopsis genome and divided into different major subfamilies (Fraser et al., 2005). Serine

carboxypeptidases act as acyltransferases in the biosynthesis of sinapoyl esters, which

provide UV-B protection (Lehfeldt et al., 2000, Shirley and Chapple, 2003, Shirley et al.,

2001). The Arabidopsis serine carboxipeptidase brassinosteroid insensitive suppressor 1

(BRS1) acts upstream brassinosteroid insensitive 1 (BRI1) in regulating BR signalling either

by activating proteins that assist in BR perception or by removing proteins that block the BR

binding site (Zhou and Li, 2005). Despite their functional role in regulating hormone

signalling, serine carboxipeptidases have not been directly involved in plant responses to

pathogen attack.

· Subtilisin-like proteases or Subtilases

Subtilisin-like proteases (subtilases, SBT, family S8, clan SB) are serine proteases

characterised by a catalytic triad containing three conserved amino acids, namely aspartate,

histidine, and serine, in their active site (Schaller et al., 2012) (Figure 14). Sequences

Figure 15. Phylogenetic tree of Arabidospsis subtilases (From Rautengarten et al.,

2005). Bootstrapped neighbour-joining tree generated from an alignment of the predicted 56 AtSBT full-

length protein sequences. The assignment of a gene to a specific subfamily was based primarily on

the position within the phylogenetic tree, as defined by the homology between the deduced full

length amino acid sequences. Different colours are used to distinguish the AtSBT1–6 subgroups.

Groups of neighbouring genes (e.g., At1g20150 and At1g20160) are distinguished by specific

colours. AtSBT5.2 (At1g20160) is indicated by a red arrow.

35

predicted to code for S8 family proteases are known in all kingdoms. According to the

MEROPS classification, eukaryotic subtilases constitute the S8 family within the SB clan of

serine proteases (Rawlings et al., 2014). Subtilases are classified into two subfamilies: true

subtilisins (S8A subfamily) and kexins (S8B subfamily). While kexins appear to be absent

from plants (Tripathi and Sowdhamini, 2006), genes predicted to encode functional subtilisins

that have been annotated in plant species are most similar to the bacterial S8A subfamily of

subtilisins (Beers et al., 2004). Subtilases are especially abundant in plants, with 63 genes

known in Oryza sativa and at least 15 in Lycopersicon esculentum genomes (Meichtry et al.,

1999, Tripathi and Sowdhamini, 2006). In Arabidopsis, the subtilase family is one of the

largest protease gene families (56 members) and has been divided into six subfamilies

(Rautengarten et al., 2005) (Figure 15).

Typical subtilases from plants and other organisms are synthesized as preproprotein

precursors, comprising a SP at the N-terminus, a cleavable prodomain, a peptidase (or

subtilisin) domain with the characteristic arrangement of catalytically important Asp, His

and Ser residues of the catalytic triad, and occasionally C-terminal extensions (Siezen and

Leunissen, 1997). Plant subtilisins are also characterized by a large insertion within the

catalytic domain between the His and Ser residues of the catalytic triad that forms an

additional protease-associated (PA) domain. The PA domain has been found to be

associated with different families of peptidases and implicated in protein/protein interactions

and substrate recognition (Mahon and Bateman, 2000, Luo and Hofmann, 2001, Ottmann et

al., 2009) (Figure 14).

Plant subtilases are usually synthesized in the form of preproprotein precursors, translocated

via an N-terminal ER-targeting SP into the endomembrane system (Schaller et al. 2012). In

addition, plant subtilases are predicted to be glycosylated in the secretory pathway and to

accumulate extracellularly (Schaller et al., 2012, Tripathi and Sowdhamini, 2006). Indeed,

proteins carrying a SP are efficiently secreted to the apoplast (Porter et al., 2015) and

typically carry complex-type N-glycans (Schähs et al., 2007). The initial steps of N-glycan

36

synthesis at the cytosolic side of the ER membrane and in the lumen of the ER are highly

conserved. In contrast, the final N–glycan processing in the Golgi apparatus is organism-

specific giving rise to a wide variety of carbohydrate structures (Lannoo and Van Damme,

2015). As implied by the pre-pro-protein structure, the maturation of the active enzyme from

its inactive precursor requires at least two processing steps. After cleavage of the SP,

subtilases are ultimately activated by cleavage of the prodomain producing the mature

active enzyme (Taylor et al., 1997). Prodomain processing in plant subtilases is an

intramolecular autocatalytic reaction that occurs late in the ER or in the early Golgi (Cedzich

et al., 2009, Chichkova et al., 2010). Cleavage of the prodomain is required for passage

through the secretory pathway (Janzik et al., 2000, Cedzich et al., 2009) and the well-

documented function of the prodomain as an auto-inhibitor in bacterial subtilisins (Baker et

al., 1993, Li et al., 1995, Takagi et al., 2001) has also been confirmed for plant subtilases

(Nakagawa et al., 2010). In some cases, further trimming has been observed at both protein

N and C-termini (Yamagata et al., 1994, Von Groll et al., 2002, Beilinson et al., 2002, Plattner

et al., 2014).

Despite their prevalence, our current understanding of the functions of plants subtilases is

still limited. There are evidences for a role in both general protein turnover (Bogacheva et al.,

1999) and highly specific regulation of plant development or responses to environmental

challenges (Figueiredo et al., 2014, Schaller et al., 2012). For example, abnormal leaf shape

(ALE1, AtSBT2.4) determines proper epidermis formation and cuticle development at the

endosperm-embryo interface during embryogenesis (Tanaka et al., 2001) and stomatal

density and distribution 1 (SDD1, AtSBT1.2) plays a role in determining stomatal density and

distribution (Berger and Altmann, 2000, Von Groll et al., 2002, Schlüter et al., 2003). Also in

Arabidopsis, expression of the S8 protease-encoding gene auxin-induced in root cultures 3

(AIR3, AtSBT5.3) is linked to lateral root emergence (Neuteboom et al., 1999) and xylem Ser

peptidase 1 (XSP1, AtSBT4.14) appears to be involved in xylem differentiation, as indicated

by its specific expression in this tissue (Zhao et al., 2000). Recently, AtSBT5.2 was identified

37

as a CO2-induced extracellular protease, named CO2 response secreted protease (CRSP)

and negatively regulating stomatal development under high CO2 conditions (Engineer et al.,

2014). Overall, the mode of action of these subtilases in the regulation of these different

developmental processes is still poorly understood.

Several reports summarized in the following have highlighted the involvement of subtilases

during the interaction of plants and microbes. The first example of a plant subtilase acting

during plant-pathogen interactions was reported in tomato, where expression of the

subtilases P69B and P69C behaved as PR genes being induced by pathogen (Citrus

Exocortis Viroid, Pseudomonas syringae) infections and SA treatment (Jordá et al., 1999,

Tornero et al., 1997). P69 genes form a distinct subgroup among the 15 genes encoding

subtilases that have been cloned from tomato (Meichtry et al., 1999). Two other subtilase-

encoding genes (P69A and D) showed constitutive expression (Jordá et al., 1999), whereas

specific developmental expression patterns were observed for the remaining two genes

(P69E and F) (Jordá et al., 2000). It was additionally proposed that these subtilases are

secreted to the plant extracellular matrix (ECM) where they accumulate (Tornero et al., 1996,

Tornero et al., 1997). Considering that ECM is where the first host–pathogen interaction,

recognition and signaling events take place (Gupta et al., 2015), the accumulation of these

subtilases in plant ECM may account for an important role during pathogenesis. More

recently, when comparing resistant and susceptible grapevine genotypes, a gene encoding a

subtilase protein was shown to be constitutively expressed in the resistant genotype, its

expression being induced after P. viticola inoculation (Figueiredo et al., 2012, Monteiro et al.,

2013). Interestingly, a defensive role for subtilases was clearly supported by the finding that

P69B is specifically targeted by virulence factors from P. infestans, the causing agent of late

blight in potato and tomato (Tian et al., 2004, Tian and Kamoun, 2005, Tian et al., 2005).

More recently, the Arabidopsis subtilase AtSBT3.3 was found to be involved in the regulation

of immune signalling (Ramírez et al., 2013). AtSBT3.3 plays a role in pathogen-mediated

induction of SA-related defence gene expression and activation of MAPK proteins. Moreover,

38

AtSBT3.3 is involved in chromatin remodelling of defence-related genes associated with the

activation of immune priming (Ramírez et al., 2013).

Upon infection of a host plant by Agrobacterium tumefaciens, the VirD2 protein is responsible

for the mobilization of the T-DNA from the Ti plasmid and becomes covalently attached to the

newly formed 5’-end of the single stranded T-DNA. After transfer into the plant cell, migration

to the nucleus is largely guided by the nuclear localization signal of VirD2 and results in the

integration of the T-DNA into the plant genome (Ziemienowicz et al., 2001). Based on the

observation that human caspase-3 is able to cleave the VirD2 protein, Chichkova and co-

workers searched for a similar proteolytic activity in plants, capable of cleaving the VirD2

protein at the caspase recognition site. Such an activity was detected in leaf extracts of

several plant species and purified from tobacco and rice. Two enzymes were identified as

subtilases, with caspase specificity distinct from that of other known caspase-like proteases,

and were named phytaspases (for plant aspartate-specific proteases) (Chichkova et al.,

2004, Chichkova et al., 2010, Chichkova et al., 2012). Cleavage by phytaspase removes

VirD2 nuclear localization signal thus preventing its nuclear accumulation and plant

transformation (Chichkova et al., 2004). Moreover, phytaspase was directly implicated in the

HR of tobacco in response to TMV by enabling a local PCD that restricted the dispersal of

viral particles through the entire plant (Chichkova et al., 2010, Chichkova et al., 2012).

Unexpectedly, phytaspase, which is synthesized constitutively and sequestered in the

apoplastic space before PCD, is re-imported into the cell during infection during PCD

(Chichkova et al., 2010).

Sensitive oat (Avena sativa) leaves treated with the victorin toxin from necrotrophic fungus

Cochliobolus victoriae show symptoms of PCD (Navarre and Wolpert, 1999). Two proteases

involved in the victorin-induced PCD signalling cascade were purified and characterized

(Coffeen and Wolpert, 2004). These proteases display caspase activity, belong to the family

of plant subtilases and were named saspases (for serine-dependent aspartate-specific

proteases) (Coffeen and Wolpert, 2004). Saspases are constitutively expressed in cells in an

39

active form and appear to function in the apoplast and being involved indirectly in victorin-

induced cleavage of Rubisco during PCD in oat (Coffeen and Wolpert, 2004).

In addition to pathogen-derived elicitors that can activate the plant innate immune response,

plant endogenous elicitors that trigger or amplify the innate immune response have also

been identified (Ryan and Pearce, 2003, Yamaguchi et al., 2011, Huffaker et al., 2006,

Huffaker and Ryan, 2007, Pearce et al., 2010). Among those, a 12-amino acid peptide

derived from the extracellular soybean subtilase GmSubPep was shown to activate

expression of defence-related genes, suggesting that, upon pathogen attack, this

endogenous peptide would be available for receptor binding and initiation of defence

signalling (Pearce et al., 2010).

Endogenous serine PIs have also been shown to be involved in plant defence responses to

pathogen attack. For example, unusual serine protease inhibitor (upi) mutant plants display

enhanced susceptibility to the necrotrophic fungi B. cinerea and Alternaria brassicicola and

reduced tolerance to feeding by the generalist insect pest Trichoplusia ni. These data

suggest that UPI is a functional PI that positively contributes to plant defence (Laluk and

Mengiste, 2011). In addition, expression of a serine protease (Kunitz trypsin) inhibitor (KTI1)

is induced by phytopathogens and fumonisin B1 treatment in Arabidopsis. KTI1 antagonizes

in plant-pathogen interaction-related PCD (Li et al., 2008). Although the mode of inhibition

strongly suggests saspases as the target proteases, the molecular details of cell death

modulation by KTI1 and the identity of its saspase target protease(s) remain to be elucidated

(Li et al., 2008). Finally, serine PIs partially inhibited the overall activation of PCD and

thereby changed the level of susceptibility of grapevine towards the oomycete P. viticola

(Gindro et al., 2012).

40

Sensing of stress signals and their transduction into appropriate responses are crucial

requirements for plant adaptation and survival. Signal transduction to the nucleus, leads to

regulation of the expression of specific genes whose products are necessary for eliciting a

specific response. The following section presents an overview of the mechanisms involved in

triggering an adapted transcriptional response in order to ensure plant disease resistance.

41

2. Transcriptional regulation of plant defence responses

In multi-cellular organisms, a tight spatio-temporal regulation of gene expression ensures

cell-to-cell communication, development, and survival in a challenging environment. In this

context, the arsenal of plant transcriptional regulators consists not only of DNA-binding TFs

that function as activators and repressors, but also of cofactors that do not physically

associate with the DNA but co-activate or co-repress transcription through interaction with

DNA-binding TFs. In addition, recent reports suggest that signal integration is dictated by TF

regulatory networks (Tsuda and Somssich, 2015). Upon receptor activation and signal

initiation, selected TFs and associated co-factors decode this information leading to adapted

transcriptional changes (Tsuda and Somssich, 2015).

Genes encoding TFs are overrepresented in plant genomes as compared to other eukaryotic

organisms. For example, in the genome of Arabidopsis thaliana, between 6% and 10% of

genes encode TFs, in contrast to only 3% in Drosophila melanogaster or 5% in humans

(Pireyre and Burow, 2015). In addition, TFs often belong to large gene families and it is worth

noting that about 45% of described plant TF families are specific to plants (Riechmann et al.,

2000). It has been proposed that this significant number of TFs present in plants may

contribute to their adaptation to rapidly changing environmental conditions (Tsuda and

Somssich, 2015).

TFs typically present a modular structure and are able to bind to cis promoter sequences

located on target genes through their DNA binding domain (DBD). This DNA binding may

have a positive (activation) or negative (repression) effect on target gene expression. Based

on structural studies and sequence comparisons of DBDs, TFs have been classified into

several families that use related structural motifs for DNA recognition (Pabo and Sauer,

1992). In plants, several families of TFs, including AP2/ERF, bHLH, bZIP, ERF, TGA, MYB,

NA, Whirly and WRKY TF families, have been shown to be involved in the regulation of plant

42

defence response against biotic stresses (Eulgem and Somssich, 2007, Dubos et al., 2010,

Seo et al., 2015).

The involvement of these TF families in plant defence has been extensively described in the

literature and recent reviews summarize this active area of research (Seo et al. 2015, Tsuda

and Somssich 2015). As in the case of proteases in the previous section, I have made a

selection of interesting examples of TFs of different families to illustrate their involvement in

the regulation of plant immune responses. Since my PhD work has focused on the study of

the Arabidopsis defence-related MYB TF MYB30 (AtMYB30), the MYB family of TFs and its

role in the regulation of the plant response to pathogens is described in more detail. Finally, a

review article highlighting the important role of the transcriptional control of plant defence

responses is presented at the end of this section.

2.1. AP2/EREBP TFs

The APETALA2/Ethylene-responsive-element-binding protein (AP2/EREBP) family

represents one of the largest plant TF families (Licausi et al., 2013) with over 140 predicted

members in Arabidopsis. All members share a common AP2/ERF domain necessary for

specific binding to DNA and can be subdivided into four subfamilies defined as (Sakuma et

al., 2002): APETALA2 (AP2), dehydration-responsive element-binding (DREB), ethylene-

responsive factors (ERF) and related to ABI3/VPI (RAV). The ERF subfamily of proteins is

unique to plants with members that participate in the regulation of genes responsive to biotic

stress upon infection, in particular related to the JA and ET signalling pathways (Yang et al.,

2015). Several Arabidopsis ERF genes respond to pathogen infection with different but

overlapping kinetics, likely helping orchestrate an adequate defence response (Oñate-

Sánchez and Singh, 2002). The ERF subfamily includes Pto-interacting4 (PTI4) that interacts

with and is phospholylated by the product of the tomato R gene Pseudomonas Tomato

Resistance (PTO). Phosphorylation of PTI4 by PTO increases the DNA-binding of the TF,

resulting in a relatively simple signal transduction pathway that leads to resistance (Gu et al.,

2000).

43

2.2. bHLH TFs

The basic-Helix-Loop-Helix (bHLH) family was estimated to comprise more than 160

members in Arabidopsis and rice (Carretero-Paulet et al., 2010). Myelocytomatosis-related

(MYC) proteins belong to a subfamily of eukaryotic bHLH TFs that have been involved in the

establishment of defence responses. MYC TFs are key transcriptional regulators of the

expression of JA-responsive genes, positively regulating wound resistance genes and acting

as negative regulators during the expression of pathogen defence genes. For example,

AtMYC2/JAI1/JIN1, along with its closely related proteins AtMYC3 and AtMYC4 are key

master regulators coordinating JA-dependent defence responses and mediating crosstalk

with other phytohormones such as SA, ABA, GA, and auxin (Kazan and Manners, 2013). For

more details on the regulation of MYC proteins see the review article at the end of this

section (Buscaill and Rivas, 2014).

2.3. bZIP TFs

The Arabidopsis genome encodes 75 distinct members of the basic domain leucine zipper

(bZIP) family. In plants, bZIP TFs are central players of the regulation of plant immunity

especially within the SA-signalling pathway conferring resistance toward biotrophic

pathogens (Gatz, 2013). Among these proteins, AtTGA2, 5 and 6 play critical roles in

establishing SAR and are also essential activators of certain ET-induced defence responses

(Zander et al., 2014).

2.4. BBX TFs

The B-box (BBX) proteins are a class of zinc-finger TFs that sometimes also feature a

CONSTANS, CO-like, and a TOC1 (CCT) domains. BBX proteins are key factors in

regulatory networks controlling growth and developmental processes as well as responses to

biotic and abiotic stresses (Gangappa and Botto, 2014). BBX proteins also participate in

wounding and defence responses (Taki et al., 2005). For example, a study showed that

BBX32 expression is increased after a short treatment with chitin, suggesting involvement of

this TF in plant defence pathways (Libault et al., 2007).

44

2.5. NAC TFs

With 135 members in rice and 117 in Arabidopsis, no apical meristem (NAM), Arabidopsis

thaliana transcription activation factor (ATAF) and cup-shaped cotyledon (CUC2) (NAC)

proteins belong to a large TF family (Nuruzzaman et al., 2013). NAC proteins, which belong

to the largest family of plant-specific TF, have been reported to be involved in plant

development and biotic and abiotic stress regulation (Puranik et al., 2012). A significant

number of reports highlight the role of NAC TFs as central regulators of the plant innate

immune system, basal defence and SAR responses (Collinge and Boller, 2001, Jensen et

al., 2007, Jensen et al., 2008, Bu et al., 2008, Nuruzzaman et al., 2013, Delessert et al.,

2005, Seo et al., 2010a). For example, the Arabidopsis NAC protein NTL6 has been shown

to positively regulate pathogen resistance against P. syringae (Seo et al., 2010a).

2.6. Whirly TFs

Members of the Whirly family of proteins are found throughout the plant kingdom and are

predicted to share the ability to bind ssDNA (single-stranded DNA) (Desveaux et al., 2005).

This TFs family has been involved in the regulation of defence gene expression. For

example, the Arabidopsis Whirly protein AtWhy1 has been shown to be required for both

basal resistance, ETI and SAR (Desveaux et al., 2005).

2.7. WRKY TFs

WRKY proteins are a plant-specific family of TFs (Rushton et al., 2010). WRKY TFs belong

to one of the largest families of transcriptional regulators in plants with 72 representatives in

Arabidopsis, and more than 100 members in rice, soybean or poplar (Bakshi and Oelmüller,

2014). Extensive research has firmly established a major role of WRKY family members in

host immunity in Arabidopsis, barley or rice (Pandey and Somssich, 2009). Specific WRKY

family members show enhanced expression and/or DNA-binding activity following induction

by a range of pathogens, defence signals and wounding (Eulgem et al., 2000). WRKY

proteins bind to the W box, which is found in the promoters of many plant defence-related

genes (Bakshi and Oelmüller, 2014). Interestingly, expression profiling revealed that

45

expression of 70% of the Arabidopsis WRKY genes is differentially regulated in response to

SA treatment or infection by various pathogens (Dong et al., 2003, Eulgem and Somssich,

2007, Ulker and Somssich, 2004). This confirms that WRKY TFs are able to regulate

defence mechanisms to different pathogens and this activity is exerted by their action as

positive (Knoth et al., 2007, Liu et al., 2006) or negative (Journot-Catalino et al., 2006, Xu et

al., 2006) regulators. Finally, it was also shown that the WRKY TFs may act on hormonal

pathways by influencing their interconnections thus enabling the activation of various

defence pathways (Chen et al., 2013). For additional details on this particular TF family and

its role on plant defence regulation, see the review article at the end of this section (Buscaill

and Rivas, 2014).

2.8. MYB TFs

The Myeloblastom (Myb) gene was first identified as the v-Myb oncogene of an avian

myeloblastosis virus (Klempnauer et al., 1982). Subsequently, members of the Myb gene

family were identified in diverse plants and animals (Lipsick, 1996). The first plant MYB gene,

C1, was isolated from Zea mays and encodes a c-MYB-like protein with structural homology

to the vertebrate cellular proto-oncogene c-MYB that is involved in anthocyanin biosynthesis

(Paz-Ares et al., 1987). Since 1987, the size of catalogue of MYB-related TFs has increased

considerably due to the big number of MYB genes identified in higher plants. Analysis of the

Arabidopsis genome identified 198 genes in the MYB superfamily (Dubos et al., 2010).

MYB superfamily members present a modular structure characterized by the presence of a

highly conserved DBD – the MYB domain – located in the N-terminal part of the protein,

which generally comprises up to four imperfect repeats called R or MYB repeats. Each

repeat consists of approximately 50–53 amino acid residues that form a Helix-Turn-Helix

(HTH) which may or may not function synergistically in their ability to bind to DNA. The

second and third helices of each repeat build a HTH structure with three regularly spaced

tryptophan (or hydrophobic) residues, forming a hydrophobic core in the 3D HTH structure

46

(Ogata et al., 1996). The third helix of each repeat is the ‘‘recognition helix’’ that is

responsible for direct contact with DNA (Jia et al., 2004).

In contrast to the MYB domain, the C-terminal region of MYB proteins, the transcriptional

activation domain (TAD) is characteristically highly variable from one MYB protein to another.

In addition, the TAD may function as an activation or a repression domain (Dubos et al.,

2010), giving rise to the wide structural and functional variability of the MYB family.

2.8.1. DNA MYB Binding Sites (MBSs)

Despite the prominent roles played by MYB TFs in the regulation of plant gene expression,

little is known about how these proteins interact with their DNA targets (Prouse and

Campbell, 2012). Six DNA sequences have been well-characterized as DNA-binding sites for

87 proteins of the MYB superfamily: CNGTT(A/G), ACC(A/T)A(A/C), TTAGGG,

AAAATATCT, GATA and TATCCA. R2R3–MYB DNA-binding data mainly result from in vitro

assays. However, in vivo assays have also been used to determine R2R3–MYB DNA-

binding specificities. R2R3–MYB family members from different species have been

previously classified into different phylogenetic clades (groups A, B, and C) based on

sequence similarities (Romero et al., 1998). These groups were then analysed for DNA-

binding specificities. It was shown that members of group A bind the so-called MBS (MYB

Binding Site) type I sequence (C(A/C/G/T)GTT(A/G)), members of group B bind equally to

MBS type I and type II (G(G/T)T(A/T)GTT(A/G)), and most members of group C bind MBS

type IIG ((C/T)ACC(A/T)A(A/C)C) (Romero et al., 1998). However, other examples show that

MYB proteins, although similar in structure and function, are able to bind to different DNA

target sites. Overall, several studies highlight the importance of conducting DNA-binding site

experiments for individual MYB proteins because it is extremely difficult to predict MBSs on

the sole basis of sequence information (Prouse and Campbell, 2012).

Figure 16. Plant MYB transcription factor classes (Adapted from Dubos et al., 2010). Plant MYB proteins are classified depending on the number of adjacent MYB repeats (R). The

primary and secondary structures of a typical R2R3-MYB are indicated. H, helix; T, turn; W,

tryptophan; X, any amino-acid residue.

R1/2 R3

R3 R2

R3 R2

R2

R1

R1/2

1R-MYB

2R-MYB

3R-MYB

4R-MYB R2 R1

R3 R2

-W-(X19)-W- …. -F/I-(X18)-W- Primary structure

Secondary structure H H H H H H T T

Figure 17. Schematic illustration of different MYB protein classes and their

functions (From Ambawat et al., 2013).

Figure 18. Schematic representation of the relationships between the different

Arabidopsis R2R3-MYB subgroups (From Dubos et al., 2010). The tree was inferred using the neighbour-joining method and 1000 bootstraps with putative

amino acid full length of 126 Arabidopsis R2R3-MYB sequences with the Clustal X2 software.

Based on the conservation of the DNA-binding domain and of amino acid motifs in the C terminal

domains, R2R3-MYB proteins have been divided into 25 subgroups. AtMYB30 is boxed in red.

47

2.8.2. Classification of MYB TFs

The highly conserved DBD in MYB proteins suggests a common mechanism for the

regulation of DNA binding, whereas the more variable TAD domain may modulate TF

activation and DBD accessibility and thereby confer specificity of DNA binding to the TF.

Animal MYB TFs present three MYB repeats (Lipsick, 1996), while in plants the number of

repeats varies from one to four allowing classification of these proteins into four classes

(Figure 16).

· 1R-MYB

This heterogeneous class designated 1R-MYB (also called ‘‘MYB-related’’) comprises TFs

that usually contain a single MYB repeat (R3-MYB) or a partial MYB repeat (R1/2-MYB)

(Rosinski and Atchley, 1998) (Figure 16). A recent report identified 68 MYB-related genes in

Arabidopsis (Du et al., 2013). 1R-MYBs have been involved in organ morphogenesis

(Stracke et al., 2001, Simon et al., 2007, Pesch and Hülskamp, 2009) or in regulation of

secondary metabolism (Dubos et al., 2010, Matsui et al., 2008). Some of these TFs have

also been described as transcriptional activators associated with the regulation of the

circadian cycle (Schaffer et al., 2001), phosphate starvation and chloroplast development

(Ambawat et al., 2013). Others are capable of controlling gene expression indirectly through

modification of histones or chromatin remodelling (Boyer et al., 2002, Clapier and Cairns,

2009, Marian and Bass, 2005) (Figure 17).

· 2R-MYB (or R2R3-MYB)

In contrast to animals, plants contain a R2R3-MYB class which is characterised by the

presence of two MYB repeats (Wilkins et al., 2009, Jiang et al., 2004). R2R3-MYB proteins

are thought to have evolved from an R1R2R3-MYB gene ancestor, by the loss of the

sequences encoding the R1 repeat and subsequent expansion of the gene family (Rosinski

and Atchley, 1998, Du et al., 2013, Stracke et al., 2001). These proteins represent the largest

MYB class present exclusively in plants with 126 members in Arabidopsis (Dubos et al.,

2010) (Figure 18). Based on the conservation of the DBD and amino acid units located in the

48

conserved C-terminal domain of the protein, MYB R2R3 TFs were classified into 25

subgroups (Dubos et al., 2010) (Figure 18). Comparative phylogenetic studies enabled the

identification of new R2R3-MYB protein subgroups in other plant species (for example, in

Poplar and Vitis) that do not have orthologs in Arabidopsis. This suggests that these proteins

may have specific roles in these species (Wilkins et al., 2009, Matus et al., 2008). R2R3-

MYBs have been involved in primary metabolism, cell fate and identity, developmental

processes as well as responses to biotic and abiotic stresses (Ambawat et al., 2013) (Figure

17).

· 3R-MYB (or R1R2R3-MYB)

3R-MYB proteins are characterized by the presence of three MYB repeats called R1, R2 and

R3 (Figure 16). Each repeat motif of 3R-MYB proteins identified in tobacco and Arabidopsis

is more closely related to the vertebrate MYB repeats than to the repeats from plant R2R3-

MYB proteins. Thus, MYB proteins with three repeats from plants and animals represent an

evolutionarily conserved group in the MYB superfamily and are collectively called three-

repeat MYB (3R MYB) proteins (Ito, 2005). 3R-MYB proteins are encoded by five genes in

Arabidopsis and are found in most eukaryotic organisms, suggesting that they represent a

class of conserved genes which derived from the R2R3-MYB (Stracke et al., 2001, Ambawat

et al., 2013). These proteins are associated with the transcriptional control of cyclins and

thus involved in the regulation of the cell cycle (Ambawat et al., 2013, Haga et al., 2007)

(Figure 17).

· 4R-MYB

4R-MYB proteins are the smallest class of MYB proteins whose members contain four

R1/R2-like repeats (Figure 16). A single 4R-MYB protein is encoded in several plant

genomes and its role remains unknown to date (Ambawat et al., 2013) (Figure 17).

49

2.8.3. Functions of MYB TFs

While vertebrates present only three MYB proteins (c-Myb, A-Myb and B-Myb) (Lipsick et al.,

2001), members of the MYB-type family of TFs in plants are over-represented. In

vertebrates, MYB-related proto-oncogenes form a small family with a central role in

controlling cellular proliferation and development. In plants; the expansion of R2R3-MYB

proteins may have occurred in response to their adaptation to a sessile lifestyle. MYB TFs

play a wide variety of physiological functions in higher plants (Dubos et al., 2010) (Figure 17),

including regulation of primary and secondary metabolism (Lepiniec et al., 2006, Gigolashvili

et al., 2007, Zhou et al., 2009), control of cell development (Kang et al., 2009) and cell cycle

(Haga et al., 2007), participation in defence and response to various abiotic stresses (Seo et

al., 2009), hormone synthesis (Seo and Park, 2010) and signal transduction (Shin et al.,

2007).

In particular, MYB TFs play important roles as regulators of defence responses against

pathogen attack in various plant species. For example, in Nicotiana tabacum, the expression

of several MYB R2R3 genes is induced in response to elicitors (Sugimoto et al., 2000). In

barley (Hordeum vulgare), HvMYB6 functions as a positive regulator of immunity responses

to B. graminis (Chang et al., 2013). In Oryza sativa, an AtMYB78-related MYB gene has

been shown to be expressed in response to fungal attack (Lee et al., 2001). In Arabidopsis,

the R2R3-MYB TF AtMYB108/BOSI1 is a positive regulator of defence, as the Arabidopsis

mutant bos1 presents increased symptoms after infection by various necrotroph or biotroph

pathogens (Mengiste et al., 2003). Moreover, the R2R3-type MYB TF gene AtMYB72 was

identified as one of the significantly induced genes in Arabidopsis roots in response to P.

fluorescens WCS417r (Verhagen et al., 2004). This root-specific TF is an early signalling

factor that functions as a node of convergence in ISR elicited by diverse beneficial microbes

(Van der Ent et al., 2008). AtMYB44 acts as an integrator of the cross-talk between SA and

JA signalling during plant defence responses (Shim and Choi, 2013). Finally, R2R3-MYB

proteins of the subgroup 1, AtMYB30 and AtMYB96, are involved in immune responses.

50

Indeed, some studies indicate that AtMYB96 acts through the ABA signalling cascade to

regulate drought stress, freezing tolerance and disease resistance (Lee et al., 2015, Seo et

al., 2009, Seo and Park, 2010). AtMYB30 is probably the Arabidopsis MYB TF that has been

the best characterized for its positive regulatory function of the establishment of the HR,

acting through the activation of very long chain fatty acid (VLCFA) synthesis, associated with

resistance to various pathogens (Lacomme and Roby, 1999, Daniel et al., 1999, Vailleau et

al., 2002). An overview of our current knowledge about AtMYB30-mediated regulation of

defence responses is presented in section 3.

51

2.9. Transcriptional control in plant defence (Review article)

During my PhD, I had the opportunity to participate in writing a review article focused on the

transcriptional control in plant immunity. In this review, the importance of (i) defence-related

hormone signalling, (ii) the role of WRKY transcription factors during the regulation of plant

responses to pathogens, (iii) nuclear functions of plant immune receptor proteins, as well as

(iv) the varied ways by which microbial effectors subvert plant transcriptional reprogramming

to promote disease are discussed in more detail.

Transcriptional control of plant defence responsesPierre Buscaill1 and Susana Rivas2

Mounting of efficient plant defence responses depends on the

ability to trigger a rapid defence reaction after recognition of the

invading microbe. Activation of plant resistance is achieved by

modulation of the activity of multiple transcriptional regulators,

both DNA-binding transcription factors and their regulatory

proteins, that are able to reprogram transcription in the plant cell

towards the activation of defence signalling. Here we provide an

overview of recent developments on the transcriptional control of

plant defence responses and discuss defence-related hormone

signalling, the role of WRKY transcription factors during the

regulation of plant responses to pathogens, nuclear functions of

plant immune receptor proteins, as well as varied ways by which

microbial effectors subvert plant transcriptional reprogramming

to promote disease.

Addresses1 INRA, Laboratoire des Interactions Plantes-Microorganismes (LIPM),

UMR441, F-31326 Castanet-Tolosan, France2CNRS, Laboratoire des Interactions Plantes-Microorganismes (LIPM),

UMR2594, F-31326 Castanet-Tolosan, France

Corresponding author: Rivas, Susana ([email protected])

Current Opinion in Plant Biology 2014, 20:35–46

This review comes from a themed issue on Biotic interactions

Edited by Makoto Hayashi and Martin Parniske

For a complete overview see the Issue and the Editorial

Available online 20th May 2014

http://dx.doi.org/10.1016/j.pbi.2014.04.004

1369-5266/# 2014 Elsevier Ltd. All rights reserved.

IntroductionBecause of the sedentary nature of plants, sensing of

stress signals and their transduction into appropriate

responses are crucial requirements for plant adaptation

and survival. The plant immune system consists of two

interconnected branches termed PAMP-triggered

immunity (PTI) and effector-triggered immunity (ETI)

[1]. PTI is initiated upon the perception of well conserved

pathogen molecular signatures, named pathogen-/

microbe-associated molecular patterns (PAMPs/

MAMPs), through pattern recognition receptors (PRRs)

at the cell surface. To counteract PTI, adapted microor-

ganisms acquired the ability to deliver effector proteins

inside host cells, resulting in enhanced virulence [2]. In

turn, some effectors are directly or indirectly recognized

by plant disease resistance (R) proteins, for the most part

members of the nucleotide-binding/leucine-rich repeat

(NLR) family of intracellular immune sensors, in a pro-

cess that activates ETI. Based on the presence of either

Toll/interleukin-1 receptor (TIR) or coiled-coil (CC)

motifs in their N-terminal domain, NLR proteins

are subdivided in two structurally distinct groups named

TNL (TIR-NB-LRR) or CNL (CC-NB-LRR),

respectively.

Activation of immune receptors is often accompanied by a

form of programmed cell death at infection sites called

hypersensitive response (HR) that limits pathogen multi-

plication, although plant resistance appears to be

uncoupled from HR cell death in some cases [3]. At the

molecular level, plant responses to infection involve the

depolarization of the plasma membrane, modification of

ion channel activity, production of reactive oxygen species

(ROS) and antimicrobial compounds, reversible protein

phosphorylation through the activation of mitogen-acti-

vated protein kinase (MAPK) cascades or calcium-depend-

ent protein kinases (CDPKs), modulation of host gene

transcription, and the deposition of lignin and callose at the

plant cell wall [4]. The onset of local responses typically

triggers systemic acquired resistance (SAR) that confers

broad-spectrum resistance to secondary infection in a

salicylic acid (SA) partially dependent manner [5]. Plant

hormones, including jasmonate (JA), SA, abscisic acid

(ABA) and ethylene (ET), are important signal molecules

for the efficient integration of biotic stimuli [6]. Typically,

SA signalling mediates resistance to biotroph and hemi-

biotroph pathogens, whereas JA- and ET-related pathways

activate resistance against necrotrophs [6].

Modulation of gene transcription is a crucial step to

mount an efficient defence response in host cells. Tran-

scriptional re-programming of the plant cell involves

major changes in gene expression to favour defence over

other cellular processes such as growth and development.

Indeed, recent reports have uncovered transcriptional

regulators that mediate the trade-off between growth

and immunity to ensure proper allocation of resources

and plant survival [7–9]. The arsenal of defence-related

transcriptional regulators consists not only of DNA-bind-

ing TFs of the AP2/ERF, NAC, MYB, MYC/bHLH,

TGA/bZIP and WRKY families but also of proteins that

interact with and regulate these TFs through varied

molecular mechanisms [10].

Here we provide an overview of the complex transcrip-

tional responses that regulate plant responses to infection.

We outline recent advances on the transcriptional regu-

lation of defence-associated hormone signalling. We also

discuss recent data on the transcriptional control of

Available online at www.sciencedirect.com

ScienceDirect

www.sciencedirect.com Current Opinion in Plant Biology 2014, 20:35–46

defence signalling by TFs of the well-characterized

WRKY family as well as on nucleocytoplasmic NLR

immune regulators and their nuclear functions during

their molecular interaction with defence-related TFs.

Finally, we discuss examples of nuclear targeted effector

proteins that evolved to subvert defence-related tran-

scriptional responses in the host in order to promote

disease.

Transcriptional control of defence-relatedhormone signallingPlant defence responses rely on a complex interplay of

hormone signalling pathways that are interconnected in

intricate networks [6].

During Arabidopsis SAR responses, increased SA levels

alter the cellular redox balance causing partial reduction

of the key transcriptional regulator NPR1 and releasing

monomeric NPR1 from cytoplasmic oligomers, through

thioredoxin-mediated reduction of intermolecular disul-

fide bridges [11��]. NPR1 monomerization facilitates its

translocation to the nucleus where NPR1 binds SA

through two key NPR1 cysteine residues via coordinated

copper, inducing a conformational change that releases

NPR1 C-terminal transactivation domain from the N-

terminal autoinhibitory domain [12��]. Nuclear NPR1

interacts with members of the TGA transcription factor

family, able to act as activators or repressors, and promotes

the transcriptional activation of the promoter of PR1 and

WRKY defence-related genes (Fig. 1a). Interestingly,

NPR1 paralogues NPR3 and NPR4 are also able to bind

SA and, as adaptors of the Cullin 3 ubiquitin E3 ligase,

mediate NPR1 proteasomal degradation, which prevents

spurious activation of NPR1 target genes in resting cells

and ensures full induction of SA responses though NPR1

recycling of NPR1 monomers in infected cells [13��]. For

a comprehensive review on the complex regulation of SA-

mediated responses, we refer the reader to [14].

In Arabidopsis, the bioactive hormone JA–isoleucine (JA–

Ile) is perceived by a receptor complex comprising the F-

box component of the SCF (Skp1/Cullin/F-box)-type E3

ubiquitin ligase COI1 and a member of the JASMO-

NATE ZIM DOMAIN (JAZ) protein family [15�,16��].

Beyond their co-receptor function, JAZ proteins are also

repressors of TFs regulating JA responses, including the

central regulator of the bHLH class MYC2, and recruit

the general co-repressors TOPLESS (TPL) and TPR

(TOPLESS-related protein) through the adaptor protein

NINJA [17�]. Following JA-Ile recognition, JAZ are ubi-

quitinated and targeted for proteasomal degradation,

which releases TFs and activates transcription of JA-

responsive genes (Fig. 1b). Additional bHLH TFs, such

as MYC3 and MYC4 show redundant functions with

MYC2 [18] whereas bHLH003, bHLH012 and

bHLH017 have been recently uncovered as transcrip-

tional repressors, also able to interact with JAZ proteins

[19], suggesting an intricate competition between activa-

tors and repressors that determines the output of JA-

dependent transcriptional responses.

In Arabidopsis, the gaseous hormone ET is perceived by

five endoplasmic reticulum (ER) located proteins: ETR1,

ERS1, ETR2, ERS2 and EIN4, all possessing an active

kinase domain related to histidine kinases. Upon ET

perception, ET receptors relieve the suppression of

downstream signalling through release and deactivation

of the negative regulator CTR1, a Raf-like kinase that is

no longer able to phosphorylate the C-terminal domain of

the central positive regulator EIN2 [20,21��,22��].

Unphosphorylated EIN2 undergoes proteolytic cleavage

and its C-terminal fragment is translocated to the nucleus

where it participates to inhibition of the proteasomal

degradation of the TFs EIN3/EIL1, thereby triggering

ET-dependent signalling (Fig. 1c) [20,21��,22��]. These

findings fill the gap between ET perception at the ER

and transcriptional activation of ET signalling in the

nucleus [23].

WRKY transcription factorsWRKY TFs (74 members in Arabidopsis) have been

extensively characterized and shown to play both positive

and negative roles during the regulation of plant defence

responses [24]. Regulation of WRKY activity by MAPK

proteins is particularly well documented, with MAPKs

playing a role not only in substrate phosphorylation, but

also in the sequestration and release of TFs, which allows

access to target promoters. In other cases, activation of

WRKY TFs by MAPKs may be caused by phosphoryla-

tion-induced structural changes [25].

In Arabidopsis, WRKY33 is involved in resistance to

necrotrophic fungi [26,27�]. Phosphorylation of WRKY33

by functionally redundant MPK3/MPK6 is required for

the transcriptional activation of camalexin biosynthetic

genes and camalexin production in response to Botrytis

cinerea [27�]. Since phosphorylation of WRKY33 does not

affect its DNA-binding, MPK3 and MPK6 are thought to

promote WRKY33 transactivation activity. Moreover,

WRKY33 directly interacts with W-boxes in its own

promoter, suggesting a potential positive feedback regu-

latory loop and activation of WRKY33 transcriptionally

and post-translationally [27�]. In addition, SIGMA FAC-

TOR BINDING PROTEIN1 (SIB1) and SIB2 interact

with WRKY33 via their VQ motif that is also required to

stimulate WRKY33 DNA-binding (but not trans-

activation) activity [28�]. Similar to WRKY33, expression

of both SIB1 and SIB2 is induced by infection with B.

cinerea and SIB1 and SIB2 act as positive regulators of

WRKY33-mediated resistance to necrotrophic fungi

[28�]. In another study, WRKY33, and the closely related

WRKY25, were shown to interact with MAP KINASE

SUBSTRATE1 (MKS1), an additional VQ motif-contain-

ing protein [29]. In resting cells, MPK4 exists in nuclear

36 Biotic interactions

Current Opinion in Plant Biology 2014, 20:35–46 www.sciencedirect.com

complexes with WRKY33 and MKS1 although no direct

interaction between MPK4 and WRKY33 has been

demonstrated [30]. Upon bacterial infection, activated

MPK4 phosphorylates MKS1, which releases MKS1-

associated WRKY33 that then targets the expression of

camalexin biosynthesis genes [30]. In summary, different

VQ motif-containing proteins appear to be able to interact

with and activate multiple WRKY proteins and, in the

case of WRKY33, the dynamic nature of these molecular

interactions suggests distinct roles and regulations during

the plant response to either bacterial or fungal pathogens

(Fig. 2b).

Transcriptional control of plant immunity Buscaill and Rivas 37

Figure 1

Nucleus

(a) (c)

(b)

Cytoplasm

EIN3

ET

EIN3

EIN3

COI1

JAZ

JA

TRX

JAZ

JA

P

ET

JA

PR1/WRKY

TGA

MYC2/3/4

MYC2/3/4

PR1/WRKY

Redox

change

(rele

ased

)

(dephosphorylation)

SA

TGA

NPR3/4

-SA +SA

ET

-ET

+ET

-JA +JA

UbbUbb

UbbbUb

NPR1NPR1

NPR1NPR1

NPR1

NPR1NPR1

NPR1NPR1

NPR1

NPR1

NPR1

ETR1

ETR1

EIN2

EIN2

NPR1

NPR3/4NPR1

TA

EIN33333

(dephosphorylation)

EIN2

PRRRR111Ubb

UbbUbbb

Ub PRR11Ubb

UbbUbbb

Ub

SA

NPR1

N33

NPR

1

TPLTPL

NIN

JA

NIN

JA

CTR

1

CTR

1

EIN2

UbUb

UbUb

UbUb

UbUb

Current Opinion in Plant Biology

Simplified view of the transcriptional regulation of some defence related hormone pathways. (a) SA signalling depends on the key transcriptional

regulator NPR1 that in resting cells is sequestered in the cytoplasm as an oligomer. In these conditions, a small amount of NPR1 monomers is

translocated to the nucleus where the protein is ubiquitinated and degraded by the proteasome via the NPR1 paralogues NPR3 and NPR4 that act as

adaptors of the Cullin 3 ubiquitin E3 ligase [13��]. Increased SA levels induce changes in the cellular redox balance releasing monomeric NPR1 through

thioredoxin-mediated reduction of disulfide bonds among NPR1 molecules. Large amounts of momomeric NPR1 are then translocated to the nucleus

where NPR1 directly binds SA, which induces a conformational change that releases NPR1 C-terminal transactivation (TA) domain [12��]. Moreover,

NPR1 interacts with members of the TGA transcription factor family and promotes the transcriptional activation of defence-related genes, including

PR1 and WRKY genes [14]. (b) In non-induced cells, JAZ proteins act as transcriptional repressors through their interaction with the transcriptional

activators MYC2/3/4 [18]. The NINJA adaptor protein acts as a scaffold and mediates the interaction of JAZ proteins with the co-repressor TPL,

preventing untimely activation of JA-related gene expression [17�]. In challenged cells, JA is perceived by a COI1-JAZ receptor complex [16��].

Subsequent ubiquitination and proteasomal degradation of the JAZ transcriptional repressors releases MYC TFs and results in transcriptional

activation of JA-responsive genes [16��]. (c) In the absence of ET, the Raf-like kinase CTR1, that is able to interact with the ET receptor ETR1 in ER

membranes, phosphorylates the C-terminal domain of the key regulator EIN2, preventing its proteolytic cleavage and nuclear translocation

[20,21��,22��]. In this context, the transcriptional activator EIN3 is degraded by the proteasome preventing activation of ET signalling [23]. In the

presence of ET, the hormone binds to the ETR1 receptor, resulting in realease and deactivation of the Raf-like kinase CTR1, that is no longer able to

phosphorylate the EIN2 C-terminal domain [20,21��,22��]. Unphosphorylated EIN2 undergoes proteolytic cleavage and its C-terminal fragment is

translocated to the nucleus where it participates to inhibition of the proteasomal degradation of the EIN3, thereby triggering ET-dependent signalling

(Fig. 1C) [20,21��,22��,23].

www.sciencedirect.com Current Opinion in Plant Biology 2014, 20:35–46

A direct interaction between MAPK and WRKY proteins

has also been reported in Nicotiana benthamiana.

NbWRKY8 (closest homolog of Arabidopsis WRKY33)

is a substrate of three pathogen-responsive MAPKs,

SIPK, WIPK, and NTF4 [31]. Phosphorylation of

NbWRKY8 enhances its DNA-binding and trans-

activation activities, increasing expression of downstream

defence-related genes. Silencing of NbWRKY8 increased

plant disease susceptibility to Phytophthora infestans and

Colletotrichum orbiculare, indicating that NbWRKY8

regulates broad-spectrum disease resistance through acti-

vation of defence gene expression [31�].

A recent report showed that WRKY TFs are additionally

able to act synergistically with CPK proteins during ETI

signalling in Arabidopsis [32��]. WRKY46 was identified as

38 Biotic interactions

Figure 2

(a)

defence

defence

resistance

MIEL1

(relocalized)

coor

dina

tion

Pathogen

perception

Pathogen

perception

Pathogen

perception

Defence gene Defence gene

Defence gene

SA

Defence gene

Pathogen perception

MYB30 MYB30

MAPK4

MAPK4

MAPK3/6

VQ

VQ

VQ

SIB1/2

SIB1/2

MKS1MKS1

P

P

P

P

WRKY33

WRKY33

WRKY33

WRKY33

WRKY33

WRKY1/2

WRKY1/2

WRKY45

WRKY45

Pb1

WRKY1

MYB6

MYB6 MYB6

cell

death

fungal

growth

MLA10

MLA10

MLA10

UbUb

UbUb

VLCFA

Cytoplasm

Nucleus

camalexin

camalexin

WRKY33

AtsPLA2α

(b)

(c)

(d)

Current Opinion in Plant Biology

Simplified models for the action of selected nuclear TFs during the regulation of plant immune responses. (a) MYB30 positively regulates Arabidopsis

defence responses to bacteria [63,64]. Interaction with AtsPLA2-a, which relocalizes to the nucleus from Golgi vesicles, and the RING-type E3 ligase

MIEL1, which leads to MYB30 ubiquitination and proteasomal degradation, both result in reduced MYB30-mediated transcriptional activation of

VLCFA-related genes and suppressed defence [66�,67]. (b) In resting Arabidopsis cells, WRKY33 is inactive in a complex with the VQ-containing motif

protein MKS1 and MPK4 [30]. Following pathogen perception, MAPK4 phosphorylates MKS1, releasing MKS1-associated WRKY33 that is then able to

activate transcription of camalexin biosynthetic genes [30]. WRKY33 phosphorylation by MPK3/MPK6 is required for the transcriptional activation of

camalexin biosynthetic genes and WRKY33 itself, which results in a positive feedback regulatory loop [27�]. WRKY33 interaction with VQ-containing

motif proteins SIB1/2 stimulates its DNA-binding activity and thus promotes defence [28�]. (c) In resting barley cells (or during a compatible

interaction), WRKY1/2 repressors suppress defence gene expression [40��] and sequester HvMYB6 activator from activating defence [41�]. Following

recognition of the AVR10 effector, the R protein MLA10 is activated and interacts with WRKY1/2, which releases HvMYB6 from suppression byWRKY1

[40��,41�]. Released HvMYB6 may then activate defence gene expression both directly and through interaction with MLA6 [41�]. Coordination of MLA

molecular activities in the cytoplasm (for cell death activation) and in the nucleus (for restriction of fungal growth) is required for efficient activation of

defence signalling [42]. (d) In resting rice cells, WRKY45 accumulation is downregulated through proteasomal degradation [37]. Following pathogen

perception, interaction with the R protein Pb1 protects WRKY45 from proteasomal degradation, resulting in transcriptional activation of SA-mediated

signalling and resistance [36�].

Current Opinion in Plant Biology 2014, 20:35–46 www.sciencedirect.com

an ETI marker, its expression being strongly induced by

activation of the NLR immune receptors RPS2 and

RPM1. Activation of closely related CPK4,5,6 and 11

results in specific phosphorylation of a subgroup of

TFs, WRKY8, 28 and 48, and activation of WRKY46

expression, and these effects are dependent on

CPK4,5,6 and 11 kinase activity [32��]. Importantly,

phosphorylation of WRKY proteins by CPKs promotes

TF binding to target promoters, which suggests syner-

gistic roles of specific CPK and WRKY proteins to activate

WRKY46 during ETI signalling. In agreement with these

findings, cpk5,6, wrky8 and wrky48 were compromised in

defence gene activation and ETI-mediated disease resist-

ance [32��], although WRKY8 and 48 had been previously

described as negative regulators of basal resistance

[33,34].

Arabidopsis WRKY8 has also been shown to function in

the long-distance movement of crucifer-infecting Tobacco

Mosaic Virus (TMVcg) by modulating both ABA and ET

signalling [35]. Following TMVcg infection, WRKY8

respectively induces and represses the expression of

genes mediating ABA and ET signalling by directly

binding to their respective promoters. Since ABA and

ET respectively hamper and induce accumulation of

TMVcg, WRKY8 has been proposed to promote plant

resistance by mediating the crosstalk between these

signalling pathways [35].

Nuclear functions of plant immune receptorsin interaction with TFsIncreasing evidence points to the crucial role of nuclear

functions of plant immune receptor proteins that display

nucleocytoplasmic partitioning, which is essential for

proper initiation of host defences. In this section we

discuss recent reports highlighting the importance of

the interaction between immune sensors and TFs during

the regulation of defence signalling in different plant

species. These findings provide a direct connection be-

tween pathogen perception and defence-related tran-

scriptional reprogramming although the details of this

molecular process are not yet fully understood.

In rice (Oryza sativa), the CNL Panicle blast1 (Pb1)

confers durable broad-spectrum resistance to Magnaporthe

oryzae. A recent report showed that the nuclear interaction

between Pb1 and OsWRKY45 is essential for blast resist-

ance [36�]. OsWRKY45, whose transcriptional activity is

regulated by the 26S proteasome, was previously

described as being required for SA-mediated rice resist-

ance to Magnaporthe [37]. Interaction with Pb1 protects

OsWRKY45 from proteasomal degradation thus resulting

in plant resistance [36�] (Fig. 2d).

In N. benthamiana, the immune receptor N resides in the

cytoplasm and the nucleus of non-infected cells. After

TMV inoculation, and in the presence of the viral helicase

p50-U1, cytoplasmic N either enters the nucleus or sends

a signal that activates the N nuclear pool, resulting in the

activation of a successful defence response [38��]. A

recent report showed that, nuclear N interacts with the

SPL6 TF to promote defence gene activation and resist-

ance to TMV in a p50-U1-dependent manner [39�].

Interestingly, the SPL6 Arabidopsis ortholog is also

required for RPS4-mediated resistance to Pseudomonas

syringae pv. tomato expressing the effector AvrRps4 (Pst-

avrRps4), suggesting a novel and conserved function for

SPL6 TFs in activation of defence-related gene expres-

sion across several plant species.

In barley (Hordeum vulgare), the CLR immune receptor

MLA10 interacts with the transcriptional repressors

HvWRKY1 and HvWRKY2 in the nucleus to induce

resistance against the powdery mildew fungus Blumeria

graminis expressing AVR10 [40��]. Since this protein

interaction depends on recognition of AVR10 by

MLA10, HvWRKY1 and HvWRKY2 likely suppress

defence activation in the absence of the pathogen and,

upon infection, binding of MLA10 de-represses transcrip-

tion and triggers defence. A recent report identified the

MYB TF HvMYB6 as an additional MLA10-interacting

TF that positively regulates resistance to B. graminis

[41�]. Notably, both MLA10-HvMYB6 and N-SPL6 mol-

ecular interactions are only detected after activation,

suggesting that elicitor-triggered conformational changes,

oligomerization or subcellular relocalization of the R

proteins are likely required for R protein-TF interactions

[41�]. HvMYB6 DNA-binding is antagonized by its direct

association with the HvWRKY1 repressor. Activated

MLA10 releases the HvMYB6 activator from HvWRKY1

repression, thereby stimulating HvMYB6-dependent

gene expression and plant resistance [41�]. A structure-

function analysis of MLA10 has shown that cell death and

disease resistance signalling triggered by MLA10 depend

on its balanced nucleocytoplasmic activities [42]. Indeed,

MLA10 activity in cell death signalling appears to be

suppressed in the nucleus but enhanced in the cytoplasm

whereas nuclear localized MLA10 is essential and suffi-

cient to mediate disease resistance against B. graminis [42]

(Fig. 2c). In Arabidopsis, mutation of WRKY18 and

WRKY40 (homologs of barley HvWRKY1 and HvWRKY2)

results in massive defence-related transcriptional repro-

gramming and SA/EDS1-dependent resistance to the

powdery mildew fungus Golovinomyces orontii, suggesting

that WRKY18 and WRKY40 facilitate Arabidopsis infec-

tion by powdery mildew [43,44].

Effector proteins modulating the hosttranscriptional response to pathogensA significant number of microbial effector proteins is

targeted to the host cell nucleus where they are able to

manipulate host transcription or directly subvert essential

host components to promote virulence by using a striking

variety of molecular strategies [45].

Transcriptional control of plant immunity Buscaill and Rivas 39

www.sciencedirect.com Current Opinion in Plant Biology 2014, 20:35–46

The TNL RPS4 confers resistance to Pst-avrRps4 and

RPS4-mediated defence responses require nucleocyto-

plasmic activities of the central regulator EDS1 [46�].

EDS1 represents a convergence point for different

immune receptors as it is additionally able to interact

with (i) RPS6, a TNL involved in ETI against the

Pseudomonas effector HopA1 [47]; (ii) SNC1, a TNL that

contributes to RPS4 defence responses through its inter-

action with transcriptional co-repressor proteins [48]; and

(iii) SRFR1, a negative defence regulator [47,49]. EDS1

association with RPS4, RPS6 and SRFR1 at endomem-

branes is disrupted by P. syringae effectors AvrRps4 and

HopA1, in a process that is considered as a first step

towards defence activation [50��,51��]. These findings

support the idea that TNL immune sensors evolved to

guard and co-opt EDS1 for execution of ETI. Moreover,

detection of RPS4-EDS1 and AvrRps4-EDS1 complexes

in the soluble fraction of resistance-activated Arabidopsis

tissues suggests the release of an activated RPS4-EDS1

complex that shuttles between the cytoplasm and the

nucleus to trigger effective defence signalling [50��].

Further characterization of the subcellular localization

of these complexes revealed distinct, but coordinated,

cell compartment-specific RPS4-EDS1 defence

branches. Indeed, as described for MLA10 [42],

AvrRps4/EDS1/RPS4-associated nuclear processes are

involved in restriction of bacterial growth whereas cyto-

plasmic AvrRps4/EDS1/RPS4 activities triggered host

cell death [50��]. Amplification of cell death and systemic

defence-related transcriptional activation require nucleo-

cytoplasmic molecular activities, suggesting that coordi-

nation between cell compartments is required for effec-

tive innate immunity [50��] (Fig. 3a). Intriguingly,

AvrRps4 has been recently located to chloroplasts

suggesting that, while AvrRps4-induced ETI requires

nucleocytoplasmic coordination, AvrRps4 virulence tar-

get(s) may reside in chloroplasts, where this effector may

suppress PTI signalling [52]. Finally, WRKY18 and

WRKY40, which act as negative regulators of basal

defence responses to G. orontii, positively regulate

RPS4-mediated resistance, indicating distinct roles for

these two TFs during basal defence and ETI signalling

[44].

Xanthomonas and Ralstonia transcription activator-like

(TAL) effectors act as transcriptional activators in the

plant cell nucleus and provide a fascinating example of

manipulation of the eukaryotic transcriptional machinery

by directly promoting specific host gene reprogramming

for the benefit of the pathogen (for recent reviews, see

[53,54]). The molecular basis of the specificity of DNA-

binding by TAL effectors was uncovered through exper-

imental, computational and crystallographic studies

showing that polymorphic repeats in their DNA-binding

domain correspond one-to-one with given nucleotides in

the DNA target sequence [55��,56��,57��,58��]. This

specificity of DNA-binding may be used for a wide range

of genome engineering applications as well as for cloning

of TAL effector-specific plant resistance genes [59,60�].

Although the host targets of TAL effectors remain poorly

characterized, several TAL proteins appear to induce

expression of so-called SWEET genes encoding sugar

transporters that have been defined as susceptibility

genes since their expression favours infection allegedly

through stimulation of sugar efflux to feed bacteria

[61�,62].

One of the best characterized MYB TFs directing

defence-related transcriptional responses is MYB30 that

positively regulates Arabidopsis defence responses by

enhancing the synthesis of very long chain fatty acids

(VLCFAs) [63,64]. Recent studies have uncovered a tight

regulation of MYB30 activity through protein–protein

interactions and post-translational modifications [65].

Indeed, MYB30 nuclear interaction with the secreted

phospholipase AtsPLA2-a or the RING-type E3 ubiquitin

ligase MIEL1 are involved in attenuation of MYB30-

mediated transcriptional responses [66�,67] (Fig. 2a).

Moreover, the modular effector XopD from the strain

B100 of Xanthomonas campestris pv. campestris was shown

to target Arabidopsis MYB30 in the nucleus resulting in

suppression of plant resistance and defence-associated

cell death responses through inhibition of the transcrip-

tional activation of MYB30 VLCFA-related target genes

[68�] (Fig. 3b). In tomato, XopD from X. euvesicatoria

(Xcv) is able to target, deSUMOylate and destabilize the

ET-responsive TF SlERF4, leading to repressed SlERF4

transcription and thus reduced SlERF4-mediated ET

production, which is required for anti-Xcv defence and

symptom development [69�] (Fig. 3c). Another effector

able to target a host TF is HopD1 from Pseudomonas

syringae, which acts as a virulence factor in Arabidopsis.

HopD1 interacts with the TF NTL9, a positive regulator

of plant defence, in the ER resulting in the suppression of

ETI but not PTI responses [70].

Remarkably, some Pseudomonas strains have evolved the

ability to disrupt plant hormone homeostasis by produ-

cing the toxin coronatine (COR), a mimic of bioactive JA-

Ile, that contributes to disease by promoting opening of

stomata, and thus bacterial entry, as well as bacterial

growth through repression of SA signalling, and thus

activation of the antagonistic JA pathway [6]. A recent

report has identified the effector HopX1 from a Pseudo-

monas strain unable to produce COR that interacts with

and promotes the degradation of JAZ proteins through its

cysteine protease activity and independently of COI1

[71��]. Ectopic expression of HopX1 in Arabidopsis acti-

vates and represses JA- and SA-dependent gene expres-

sion, respectively [71��]. Moreover, HopX1 contributes to

bacterial virulence by mimicking COR-induced suscepti-

bility and this is dependent on the catalytic activity of the

effector [71��]. In a similar vein, the Pseudomonas effector

HopZ1a interacts with and acetylates JAZ transcriptional

40 Biotic interactions

Current Opinion in Plant Biology 2014, 20:35–46 www.sciencedirect.com

repressors, leading to their degradation through an

unknown COI1-dependent mechanism [72��]. Similar

to HopX1, HopZ1a is able to partially rescue the viru-

lence defect of Pseudomonas strains unable to produce

COR [72��].

Additional illustration of the varied strategies displayed

by nuclear effectors to block defence is provided

by HopU1, a mono-ADP-ribosyltransferase from

Pseudomonas that targets several RNA-binding proteins

including GRP7 [73�,74]. GRP7 is able to bind transcripts

of the PRRs FLS2 and EFR and HopU1 hampers this

interaction, resulting in reduced FLS2 protein accumu-

lation and suppression of PTI signalling [75�] (Fig. 3d).

The broadly conserved Mediator complex (>25 proteins)

is an essential transcriptional regulator that mediates the

interaction between TFs and RNA polymerase II [76].

Transcriptional control of plant immunity Buscaill and Rivas 41

Figure 3

RPS6

SNC1

Med19

HaRxL44

(Hpa )

Nucleus

SlERF4

immunityET

4 -SUMO

SA JA/ETdefence to

biotrophs

GRP7

HopU1

(Pst )

FLS2 mRNA

FLS2

SlERF4RPS4

cell death

RPS4

bacterial

growth

Defence gene

SlERF4

EDS1EDS1

EDS1

SRFR1

RPS6

HopA1

(Pst)

RPS4EDS1

SRFR1

AvrRps4

(Pst)

(a)

Cytoplasm

MYB30

VLCFA defence

XopD

(Xcc )

XopD

(Xcv )

FL

PTI

AvrRps4

(Pst)

AvrRps4

(Pst)

(b)

(e)

(d)

(c)

Current Opinion in Plant Biology

Schematic representation of the mode of action of certain microbial effector proteins inside the plant cell nucleus. (a) In resting cells, the repressor SRFR1

interacts at endomembranes with EDS1 and inactive RPS4 and RPS6 R proteins [51��,84]. The interaction of effectors AvrRps4 and HopA1 with EDS1

results in perturbation of receptor complexes residing at endomembranes [51��]. Nucleocytoplasmic coordination of molecular activities of a released

RPS4-EDS1-AvrRPS4 signalling complex is required for the activation of cytoplasmic cell death and transcriptional activation of defence genes, whereas a

nuclear RPS4-EDS1-AvrRPS4 complex restricts bacterial growth [50��]. The R protein SNC1 contributes to RPS4 defence responses through its

interaction with transcriptional co-repressor proteins [48]. Pst: Pseudomonas syringae pv. tomato. (b) XopD from Xcc targets the Arabidopsis MYB TF

MYB30 and suppresses MYB30-mediated transcriptional activation of VLCFA-related genes thereby blocking defence responses [68�]. Xcc:

Xanthomonas campestris pv. campestris. (c) XopD from Xcv targets and deSUMOylates the tomato TF SlERF4, resulting in reduced transcriptional

activation of SlERF4 and ethylene signalling-related genes, which blocks anti-Xcv immunity [69�]. Xcv: Xanthomonas euvesicatoria. (d) The mono-ADP-

ribosyltransferase HopU1 from Pst targets the RNA-binding protein GRP7 resulting in suppressed GRP7 binding of FLS2 transcripts, reduced FSL2

accumulation and inhibited PTI signalling [73�,74]. (e) HaRxL44 from Hpa interacts with the Med19 subunit of the mediator complex, which activates and

suppresses JA/ET and SA signalling respectively suppressing defence against biotrophs. Hpa: Hyaloperonospora arabidopsidis [77��].

www.sciencedirect.com Current Opinion in Plant Biology 2014, 20:35–46

The nuclear-localized effector HaRxL44 from Hyaloper-

onospora arabidopsidis (Hpa) interacts with the Mediator

complex subunit 19a (Med19a), which is a positive reg-

ulator of resistance to Hpa [77��]. This interaction results

in Med19a proteasomal degradation and increased and

decreased JA/ET and SA signalling, respectively. In

agreement with this finding, Hpa abolishes expression

of the SA marker gene PR1 specifically in haustoria-

containing plant cells [77��]. HaRxL44-mediated degra-

dation of Med19a thus enhances susceptibility to Hpa

through attenuation of SA signalling (Fig. 3e).

ConclusionsBeyond the obvious contribution of TFs to modulation of

gene expression, recent groundbreaking studies have

shown that intracellular immune sensors display crucial

nuclear activities for efficient defence signalling, thereby

providing a direct link between pathogen perception and

defence-related gene expression. However, not every

NLR appears to function in the nucleus. For example,

the CNL RPM1 is located and functions at the plasma

membrane to initiate defence signalling [78]. Since

RPM1 was not found to translocate to the nucleus, the

question of how signals generated at the plasma mem-

brane are transduced to defence-related transcriptional

changes remains to be answered. Further work is clearly

necessary to shed light on the commonalities and specifi-

cities of the transcriptional responses downstream the

activation of nuclear and plasma membrane-located

immune receptors.

Although studies on RPS4 and MLA R proteins suggest

that programmed cell death may be uncoupled from

mechanisms that restrict bacterial growth, this is not a

universal principle for NLR defence signalling as exem-

plified by the nucleocytoplasmic CNL Rx that is acti-

vated in the cytoplasm by the Potato Virus X (PVX) coat

protein [79��]. Retention of Rx in the cytoplasm, which is

mediated by its interaction with the regulator of nucleo-

cytoplasmic traffic RanGAP2, is an essential requirement

to trigger cell death and resistance to PVX [79��,80�].

Overall, these studies suggest that coordination between

the cytoplasmic and nuclear compartments is a key fea-

ture directing effective defence outputs at least in a

significant number of cases. Nevertheless, the molecular

mechanisms that govern and are directed by nucleocyto-

plasmic partitioning of NLR immune receptors remain

elusive and clearly warrant further investigation. An

exciting possibility that needs to be explored in the future

is that nuclear-localized R proteins may directly promote

transcriptional regulation of target genes, as already

described for mammalian NLR immune regulators

[81]. For example, RPS4 and the TNL WRKY

domain-containing RRS1 cooperate genetically and con-

fer resistance to bacterial and fungal pathogens [82]. It is

thus tempting to hypothesize that effector-mediated acti-

vation of RPS4/RRS1 may promote transcriptional

regulation of target genes through RRS1 WRKY domain,

perhaps following a yet to be demonstrated RPS4–RRS1

physical interaction [83]. Finally, in order to obtain an

integrated view of the transcriptional changes associated

to plant defence responses, knowledge on mechanisms

used by plants to mount appropriate defence outputs

need to be brought together with new findings on the

molecular strategies used by microbes to subvert defence

signalling in host cells.

Within the current heyday of high-throughput genomic

and transcriptomic methodologies, a comprehensive over-

view of the global transcriptional changes in cells under-

going pathogen attack is starting to emerge. These global

studies (i) suggest a significant overlap in transcriptional

changes during PTI and ETI responses, (ii) highlight the

complexity of the molecular mechanisms involved in

defence-related transcriptional reprogramming of plant

cells, and (iii) represent an extremely valuable source of

information for further analyses aimed at understanding

the transcriptional control of plant immunity. Indeed, in

depth analysis of these global datasets should shed light

on the large-scale transcriptomic changes undergone by

plant cells during their interaction with microbes. Finally,

large-scale proteomic approaches represent an ideal

complement of global gene expression analyses and are

clearly indispensable to get a more complete picture of

the composition of protein complexes and the post-trans-

lational modifications that play a major role during the

regulation of protein activity.

AcknowledgementsWe apologize to all colleagues whose work could not be discussed becauseof space limitations. P.B. is funded by a grant from the French Ministry ofNational Education and Research/French Laboratory of Excellence project‘‘TULIP’’ (ANR-10-LABX-41; ANR-11-IDEX-0002-02). Our work issupported by the French Laboratory of Excellence project ‘‘TULIP’’(ANR-10-LABX-41; ANR-11-IDEX-0002-02).

References and recommended readingPapers of particular interest, published within the period of review,have been highlighted as:

� of special interest

�� of outstanding interest

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

2. Deslandes L, Rivas S: Catch me if you can: bacterial effectorsand plant targets. Trends Plant Sci 2012, 17:644-655.

3. Coll NS, Epple P, Dangl JL: Programmed cell death in the plantimmune system. Cell Death Differ 2011, 18:1247-1256.

4. Meng X, Zhang S: MAPK cascades in plant disease resistancesignaling. Annu Rev Phytopathol 2013, 51:245-266.

5. Spoel SH, Dong X: How do plants achieve immunity? Defencewithout specialized immune cells. Nat Rev Immunol 2012,12:89-100.

6. Pieterse CM, Van der Does D, Zamioudis C, Leon-Reyes A, VanWees SC: Hormonal modulation of plant immunity. Annu RevCell Dev Biol 2012, 28:489-521.

42 Biotic interactions

Current Opinion in Plant Biology 2014, 20:35–46 www.sciencedirect.com

7. Lozano-Duran R, Macho AP, Boutrot F, Segonzac C, Somssich IE,Zipfel C: The transcriptional regulator BZR1 mediates trade-offbetween plant innate immunity and growth. Elife 2013,2:e00983.

8. Fan M, Bai MY, Kim JG, Wang T, Oh E, Chen L, Park CH, Son SH,Kim SK, Mudgett MB et al.: The bHLH transcription factor HBI1mediates the trade-off between growth and pathogen-associated molecular pattern-triggered immunity inArabidopsis. Plant Cell 2014, 26:828-841.

9. Malinovsky FG, Batoux M, Schwessinger B, Youn JH, Stransfeld L,Win J, Kim SK, Zipfel C: Antagonistic regulation of growth andimmunity by the Arabidopsis basic helix–loop–helixtranscription factor homolog of brassinosteroid enhancedexpression2 interacting with increased leaf inclination1binding bHLH1. Plant Physiol 2014, 164:1443-1455.

10. Eulgem T: Regulation of the Arabidopsis defencetranscriptome. Trends Plant Sci 2005, 10:71-78.

11.��

Tada Y, Spoel SH, Pajerowska-Mukhtar K, Mou Z, Song J,Wang C, Zuo J, Dong X: Plant immunity requires conformationalchanges of NPR1 via S-nitrosylation and thioredoxins. Science2008, 321:952-956.

The authors show that NPR1 is sequestered in the cytoplasm as anoligomer through intermolecular disulfide bonds. Increased SA levelsinduce NPR1 monomerization through the activity of thioredoxins. Thesedata highlight the importance of pathogen-triggered redox changes forthe establishment of plant defences.

12.��

Wu Y, Zhang D, Chu JY, Boyle P, Wang Y, Brindle ID, De Luca V,Despres C: The Arabidopsis NPR1 protein is a receptor for theplant defense hormone salicylic acid. Cell Rep 2012, 1:639-647.

This work shows that the transcriptional regulator NPR1 is able to bind SAthrough two conserved cysteine residues, thereby inducing a conforma-tional change that releases the C-terminal transactivation domain inNPR1. See also Fu et al. (Ref. [13��]).

13.��

Fu ZQ, Yan S, Saleh A, Wang W, Ruble J, Oka N, Mohan R,Spoel SH, Tada Y, Zheng N et al.: NPR3 and NPR4 are receptorsfor the immune signal salicylic acid in plants. Nature 2012,486:228-232.

This work shows that NPR1 paralogues NPR3 and NPR4 are SA receptorsthat bind SA with different affinities. In addition, NPR3 and NPR4 functionas adaptors of the Cullin 3 ubiquitin E3 ligase to mediate NPR1 degrada-tion in an SA-dependent fashion. Although NPR1 did not bind SA in thisstudy, binding of SA by NPR1 was demonstrated by Wu and coworkers(Ref. [12��]).

14. Pajerowska-Mukhtar KM, Emerine DK, Mukhtar MS: Tell memore: roles of NPRs in plant immunity. Trends Plant Sci 2013,18:402-411.

15.�

Fonseca S, Chini A, Hamberg M, Adie B, Porzel A, Kramell R,Miersch O, Wasternack C, Solano R: (+)-7-Iso-jasmonoyl-L-isoleucine is the endogenous bioactive jasmonate. Nat ChemBiol 2009, 5:344-350.

This work shows that the biologically active jasmoonate hormone is theactive hormone is (+)-7-iso-JA-L-Ile, which is structurally more similar tocoronatine produced by some strains of pathogenic bacteria.

16.��

Sheard LB, Tan X, Mao H, Withers J, Ben-Nissan G, Hinds TR,Kobayashi Y, Hsu FF, Sharon M, Browse J et al.: Jasmonateperception by inositol-phosphate-potentiated COI1-JAZ co-receptor. Nature 2010, 468:400-405.

This study unraveled the mechanism of jasmonate perception by usingstructural and pharmacological approaches to show that the Arabidopsisjasmonate receptor is a complex of both the F-box protein COI1 and thetranscriptional repressor JAZ.

17.�

Pauwels L, Barbero GF, Geerinck J, Tilleman S, Grunewald W,Perez AC, Chico JM, Bossche RV, Sewell J, Gil E et al.: NINJAconnects the co-repressor TOPLESS to jasmonate signalling.Nature 2010, 464:788-791.

The autors show that JAZ proteins recruit the co-repressor TOPLESS(TPL) and related TPR proteins (TPRs) through the adaptor protein NINJA,which acts as a transcriptional repressor. Both NINJA and TPL proteinsfunction as negative regulators of JA responses.

18. Fernandez-Calvo P, Chini A, Fernandez-Barbero G, Chico JM,Gimenez-Ibanez S, Geerinck J, Eeckhout D, Schweizer F,Godoy M, Franco-Zorrilla JM et al.: The Arabidopsis bHLHtranscription factors MYC3 and MYC4 are targets of JAZ

repressors and act additively with MYC2 in the activation ofjasmonate responses. Plant Cell 2011, 23:701-715.

19. Fonseca S, Fernandez-Calvo P, Fernandez GM, Diez-Diaz M,Gimenez-Ibanez S, Lopez-Vidriero I, Godoy M, Fernandez-Barbero G, Van Leene J, De Jaeger G et al.: bHLH003, bHLH013and bHLH017 are new targets of JAZ repressors negativelyregulating JA responses. PLoS One 2014, 9:e86182.

20. Wen X, Zhang C, Ji Y, Zhao Q, He W, An F, Jiang L, Guo H:Activation of ethylene signaling is mediated by nucleartranslocation of the cleaved EIN2 carboxyl terminus. Cell Res2012, 22:1613-1616.

21.��

Qiao H, Shen Z, Huang SS, Schmitz RJ, Urich MA, Briggs SP,Ecker JR: Processing and subcellular trafficking of ER-tethered EIN2 control response to ethylene gas. Science 2012,338:390-393.

This study, together with the report by Ju et al. (Ref. [22��]), shows thatethylene perception triggers dephosphorylation and proteolytic cleavageof the key regulator EIN2, that resides at ER membranes, resulting innuclear translocation of a carboxyl-terminal EIN2 fragment. These datauncover a mechanism that links hormone perception at the ER withnuclear transcriptional regulators. See also Wen et al. (Ref. [20]).

22.��

Ju C, Yoon GM, Shemansky JM, Lin DY, Ying ZI, Chang J,Garrett WM, Kessenbrock M, Groth G, Tucker ML et al.: CTR1phosphorylates the central regulator EIN2 to control ethylenehormone signaling from the ER membrane to the nucleus inArabidopsis. Proc Natl Acad Sci U S A 2012, 109:19486-19491.

Together with the work of Qiao et al. (Ref. [21��]), this study shows thatdephosphorylation and proteolytic cleavage of EIN2 at ER membranesresults in translocation of a carboxyl-terminal EIN2 fragment to thenucleus allowing the ET signal to reach the downstream TFs. See alsoWen et al. (Ref. [20]).

23. Ji Y, Guo H: From endoplasmic reticulum (ER) to nucleus: EIN2bridges the gap in ethylene signaling. Mol Plant 2012, 6:11-14.

24. Pandey SP, Somssich IE: The role of WRKY transcriptionfactors in plant immunity. Plant Physiol 2009, 150:1648-1655.

25. Ishihama N, Yoshioka H: Post-translational regulation of WRKYtranscription factors in plant immunity. Curr Opin Plant Biol2012, 15:431-437.

26. Zheng Z, Qamar SA, Chen Z, Mengiste T: Arabidopsis WRKY33transcription factor is required for resistance to necrotrophicfungal pathogens. Plant J 2006, 48:592-605.

27.�

Mao G, Meng X, Liu Y, Zheng Z, Chen Z, Zhang S:Phosphorylation of a WRKY transcription factor by twopathogen-responsive MAPKs drives phytoalexin biosynthesisin Arabidopsis. Plant Cell 2011, 23:1639-1653.

In this work, the authors show that MPK3 and MPK6 phosphorylateWRK33, which triggers camalexin biosynthesis in response to infection byB. cinerea. WRKY33 binding to its own promoter suggests a potentialpositive feedback regulatory loop.

28.�

Lai Z, Li Y, Wang F, Cheng Y, Fan B, Yu JQ, Chen Z: Arabidopsissigma factor binding proteins are activators of the WRKY33transcription factor in plant defense. Plant Cell 2011,23:3824-3841.

This study demonstrates that VQ domain-containing SIB1 and SIB2proteins interact with WRK33 and stimulate its DNA-binding activity.Together with the work of Andreasson et al. (Ref. [29]), this work under-lines the potential role of VQ motif-containing proteins in promotingWRK33 binding to its transcriptional gene targets in response to differentpathogens.

29. Andreasson E, Jenkins T, Brodersen P, Thorgrimsen S,Petersen NH, Zhu S, Qiu JL, Micheelsen P, Rocher A, Petersen Met al.: The MAP kinase substrate MKS1 is a regulator of plantdefense responses. EMBO J 2005, 24:2579-2589.

30. Qiu JL, Fiil BK, Petersen K, Nielsen HB, Botanga CJ,Thorgrimsen S, Palma K, Suarez-Rodriguez MC, Sandbech-Clausen S, Lichota J et al.: Arabidopsis MAP kinase 4 regulatesgene expression through transcription factor release in thenucleus. EMBO J 2008, 27:2214-2221.

31.�

Ishihama N, Yamada R, Yoshioka M, Katou S, Yoshioka H:Phosphorylation of the Nicotiana benthamiana WRKY8transcription factor by MAPK functions in the defenseresponse. Plant Cell 2011, 23:1153-1170.

Transcriptional control of plant immunity Buscaill and Rivas 43

www.sciencedirect.com Current Opinion in Plant Biology 2014, 20:35–46

The authors show that in N. benthamiana NbWRKY8 is phosphorylated bythree independent MAPKs (SIPK, WIPK and NTF4) resulting in enhancedNbWRKY8 DNA-binding and transactivation activities and broad-spec-trum disease resistance.

32.��

Gao X, Chen X, Lin W, Chen S, Lu D, Niu Y, Li L, Cheng C,McCormack M, Sheen J et al.: Bifurcation of Arabidopsis NLRimmune signaling via Ca(2)(+)-dependent protein kinases.PLoS Pathog 2013, 9:e1003127.

This work shows the role of six closely related CPKs in mediating bifurcateNLR-activated immune responses. Of particular interest with regard tothe regulation of WRKY transcription factors, the authors demonstrate thephosphorylation of the subgroup of WRKY8/28/48 by closely relatedCPK4/5/6/11, which results in enhanced WRKY46 expression. The sig-nificance of this multiple level regulation is supported by the fact thatcpk5,6 wrky8 and wrk48 displayed reduced defence responses afterbacterial inoculation.

33. Chen L, Zhang L, Yu D: Wounding-induced WRKY8 is involvedin basal defense in Arabidopsis. Mol Plant Microbe Interact2010, 23:558-565.

34. Xing DH, Lai ZB, Zheng ZY, Vinod KM, Fan BF, Chen ZX: Stress-and pathogen-induced Arabidopsis WRKY48 is atranscriptional activator that represses plant basal defense.Mol Plant 2008, 1:459-470.

35. Chen L, Zhang L, Li D, Wang F, Yu D: WRKY8 transcription factorfunctions in the TMV-cg defense response by mediating bothabscisic acid and ethylene signaling in Arabidopsis. Proc NatlAcad Sci U S A 2013, 110:E1963-E1971.

36.�

Inoue H, Hayashi N, Matsushita A, Xinqiong L, Nakayama A,Sugano S, Jiang CJ, Takatsuji H: Blast resistance of CC-NB-LRRprotein Pb1 is mediated by WRKY45 through protein–proteininteraction. Proc Natl Acad Sci U S A 2013, 110:9577-9582.

This study shows that nuclear interaction between the rice R protein Pb1and the WRKY transcription factor WRKY45 protects WRKY45 fromdegradation by the 26S proteasome and is required for blast resistance.

37. Matsushita A, Inoue H, Goto S, Nakayama A, Sugano S, Hayashi N,Takatsuji H: The nuclear ubiquitin proteasome degradationaffects WRKY45 function in the rice defense program. Plant J2012, 73:302-313.

38.��

Caplan JL, Mamillapalli P, Burch-Smith TM, Czymmek K, Dinesh-Kumar SP: Chloroplastic protein NRIP1 mediates innateimmune receptor recognition of a viral effector. Cell 2008,132:449-462.

This important study showed that in the presence of its cognate elicitor,the viral helicase p50, the tobacco N immune receptor displays nucleo-cytoplasmic partitioning through the chloroplastic rhodanese sulfurtrans-ferase NRIP1. NRIP1 is able to interact with both N and p50 and isrequired for full resistance to the TMV virus.

39.�

Padmanabhan MS, Ma S, Burch-Smith TM, Czymmek K, Huijser P,Dinesh-Kumar SP: Novel positive regulatory role for the SPL6transcription factor in the N TIR-NB-LRR receptor-mediatedplant innate immunity. PLoS Pathog 2013, 9:e1003235.

This study follows up on the work of Caplan and co-workers (Ref. [38��])and shows that nuclear N interacts with the transcription factor SQUA-MOSA PROMOTER BINDING PROTEIN (SBP)-domain (SPL6) in a pro-cess that activates defence gene expression. Interestingly, function ofSPL6 proteins appears to be conserved across different plant species.

40.��

Shen QH, Saijo Y, Mauch S, Biskup C, Bieri S, Keller B, Seki H,Ulker B, Somssich IE, Schulze-Lefert P: Nuclear activity of MLAimmune receptors links isolate-specific and basal disease-resistance responses. Science 2007, 315:1098-1103.

This seminal study showed that in barley, intracellular MLA R proteinsfunction in the nucleus to confer powdery mildew resistance. Recognitionof the fungal AVR10 effector by MLA10 induces the nuclear associationbetween MLA10 and WRKY1/2 transcription factors. MLA10 interfereswith the transcriptional repressor function of WRKY1/2, thereby de-repressing basal defence responses.

41.�

Chang C, Yu D, Jiao J, Jing S, Schulze-Lefert P, Shen QH: BarleyMLA immune receptors directly interfere with antagonisticallyacting transcription factors to initiate disease resistancesignaling. Plant Cell 2013, 25:1158-1173.

This study of the interaction between MLA and the antagonistic tran-scription factors MYB6 and WRKY1 builds up on the work of Shen andcollaborators (Ref. [40��]). The DNA-binding activity of MYB6, a positiveregulator of basal and MLA-mediated immunity responses to Blumeriagraminis, is stimulated and repressed by its interaction with MLA6 and

WRKY1, respectively. MLA10 releases MYB6 from WRKY1 repressionand thus stimulates MYB6-dependent gene expression.

42. Bai S, Liu J, Chang C, Zhang L, Maekawa T, Wang Q, Xiao W, Liu Y,Chai J, Takken FL et al.: Structure–function analysis of barleyNLR immune receptor MLA10 reveals its cell compartmentspecific activity in cell death and disease resistance. PLoSPathog 2012, 8:e1002752.

43. Pandey SP, Roccaro M, Schon M, Logemann E, Somssich IE:Transcriptional reprogramming regulated by WRKY18 andWRKY40 facilitates powdery mildew infection of Arabidopsis.Plant J 2010, 64:912-923.

44. Schon M, Toller A, Diezel C, Roth C, Westphal L, Wiermer M,Somssich IE: Analyses of wrky18 wrky40 plants reveal criticalroles of SA/EDS1 signaling and indole-glucosinolatebiosynthesis for golovinomyces orontii resistance and a loss-of resistance towards Pseudomonas syringae pv. tomatoAvrRPS4. Mol Plant Microbe Interact 2013, 26:758-767.

45. Rivas S, Genin S: A plethora of virulence strategies hiddenbehind nuclear targeting of microbial effectors. Front Plant Sci2011, 2:104.

46.�

Garcia AV, Blanvillain-Baufume S, Huibers RP, Wiermer M, Li G,Gobbato E, Rietz S, Parker JE: Balanced nuclear andcytoplasmic activities of EDS1 are required for a completeplant innate immune response. PLoS Pathog 2010, 6:e1000970.

This work shows that coordinated nuclear and cytoplasmic activities ofthe EDS1 regulator are required to mount a balanced immune response.Nuclear EDS1 activity is required for R gene-dependent transcriptionalreprogramming whereas cytoplasmic EDS1 promotes effector-triggeredprogrammed cell death. See also Heidrich et al. (Ref. [50��]).

47. Kim SH, Kwon SI, Saha D, Anyanwu NC, Gassmann W:Resistance to the Pseudomonas syringae effector HopA1 isgoverned by the TIR-NBS-LRR protein RPS6 and is enhancedby mutations in SRFR1. Plant Physiol 2009, 150:1723-1732.

48. Zhu Z, Xu F, Zhang Y, Cheng YT, Wiermer M, Li X: Arabidopsisresistance protein SNC1 activates immune responses throughassociation with a transcriptional corepressor. Proc Natl AcadSci U S A 2010, 107:13960-13965.

49. Kwon SI, Kim SH, Bhattacharjee S, Noh JJ, Gassmann W: SRFR1,a suppressor of effector-triggered immunity, encodes aconserved tetratricopeptide repeat protein with similarity totranscriptional repressors. Plant J 2009, 57:109-119.

50.��

Heidrich K, Wirthmueller L, Tasset C, Pouzet C, Deslandes L,Parker JE: Arabidopsis EDS1 connects pathogen effectorrecognition to cell compartment-specific immune responses.Science 2011, 334:1401-1404.

This work, together with the report by Bhattacharjee et al. (Ref. [51��]),identifies a molecular mechanism connecting effector recognition byimmune receptors and activation of immune signalling. Detection ofAvrRPS4-EDS1 and RPS4-EDS1 soluble and nuclear complexes sug-gests the need of coordinated nucleocytoplasmic activities for effectivelocal and systemic resistance responses.

51.��

Bhattacharjee S, Halane MK, Kim SH, Gassmann W: Pathogeneffectors target Arabidopsis EDS1 and alter its interactionswith immune regulators. Science 2011, 334:1405-1408.

This study shows that AvrRps4-/HopA1-mediated disruption of EDS1association with different R proteins at endomembranes activatesdefence, suggesting that immune receptors may guard EDS1 to triggerimmunity. See also Heidrich et al. (Ref. [50��]).

52. Li G, Froehlich JE, Elowsky C, Msanne J, Ostosh AC, Zhang C,Awada T, Alfano JR: Distinct Pseudomonas type-III effectors usea cleavable transit peptide to target chloroplasts. Plant J 2013.

53. Schornack S, Moscou MJ, Ward ER, Horvath DM: Engineeringplant disease resistance based on TAL effectors. Annu RevPhytopathol 2013, 51:383-406.

54. Doyle EL, Stoddard BL, Voytas DF, Bogdanove AJ: TAL effectors:highly adaptable phytobacterial virulence factors and readilyengineered DNA-targeting proteins. Trends Cell Biol 2013,23:390-398.

55.��

Boch J, Scholze H, Schornack S, Landgraf A, Hahn S, Kay S,Lahaye T, Nickstadt A, Bonas U: Breaking the code of DNAbinding specificity of TAL-type III effectors. Science 2009,326:1509-1512.

44 Biotic interactions

Current Opinion in Plant Biology 2014, 20:35–46 www.sciencedirect.com

By projecting the sequence of repeats with amino acid repeat-variable di-residues (RVDs) of the Xanthomonas TAL effector AvrBs3 onto its targetDNA promoter sequence, the UPA box, the authors determined the TALeffector-DNA recognition code. See also Ref. [56��].

56.��

Moscou MJ, Bogdanove AJ: A simple cipher governs DNArecognition by TAL effectors. Science 2009, 326:1501.

In this study, the TAL effector-DNA recognition code was decipheredusing a computational approach. See also Ref. [55��].

57.��

Deng D, Yan C, Pan X, Mahfouz M, Wang J, Zhu JK, Shi Y, Yan N:Structural basis for sequence-specific recognition of DNA byTAL effectors. Science 2012, 335:720-723.

This study shows crystal structures of a 11.5-repeat artificially engineeredTAL effector, dHax3, in both DNA-free and DNA-bound states. See alsoRef. [58��].

58.��

Mak AN, Bradley P, Cernadas RA, Bogdanove AJ, Stoddard BL:The crystal structure of TAL effector PthXo1 bound to its DNAtarget. Science 2012, 335:716-719.

This paper shows the crystal structure of the Xanthomonas TAL effectorPthXo1 bound to its target DNA sequence. See also Ref. [57��].

59. Strauss T, van Poecke RM, Strauss A, Romer P, Minsavage GV,Singh S, Wolf C, Kim S, Lee HA, Yeom SI et al.: RNA-seqpinpoints a Xanthomonas TAL-effector activated resistancegene in a large-crop genome. Proc Natl Acad Sci U S A 2012,109:19480-19485.

60.�

Li T, Liu B, Spalding MH, Weeks DP, Yang B: High-efficiencyTALEN-based gene editing produces disease-resistant rice.Nat Biotechnol 2012, 30:390-392.

The authors exploit TALEN-based disruption to edit a specific ricesusceptibility gene, Os11N3, and engineer resistance to bacterial blight.This study illustrates the potential of TALEN-based technologies toengineer heritable genome modifications for plant resistance.

61.�

Chen LQ, Hou BH, Lalonde S, Takanaga H, Hartung ML, Qu XQ,Guo WJ, Kim JG, Underwood W, Chaudhuri B et al.: Sugartransporters for intercellular exchange and nutrition ofpathogens. Nature 2010, 468:527-532.

This work shows that biotrophic bacteria and fungi induce the expressionof SWEET genes, suggesting that sugar efflux by SWEET transporters isprobably targeted by pathogens for nutritional gain. See also Ref. [62].

62. Antony G, Zhou J, Huang S, Li T, Liu B, White F, Yang B: Rice xa13recessive resistance to bacterial blight is defeated byinduction of the disease susceptibility gene Os-11N3. Plant Cell2010, 22:3864-3870.

63. Vailleau F, Daniel X, Tronchet M, Montillet JL, Triantaphylides C,Roby D: A R2R3-MYB gene, AtMYB30, acts as a positiveregulator of the hypersensitive cell death program in plants inresponse to pathogen attack. Proc Natl Acad Sci U S A 2002,99:10179-10184.

64. Raffaele S, Vailleau F, Leger A, Joubes J, Miersch O, Huard C,Blee E, Mongrand S, Domergue F, Roby D: A MYB transcriptionfactor regulates very-long-chain fatty acid biosynthesis foractivation of the hypersensitive cell death response inArabidopsis. Plant Cell 2008, 20:752-767.

65. Raffaele S, Rivas S: Regulate and be regulated: integration ofdefense and other signals by the AtMYB30 transcriptionfactor. Front Plant Sci 2013, 4:98.

66.�

Froidure S, Canonne J, Daniel X, Jauneau A, Briere C, Roby D,Rivas S: AtsPLA2-alpha nuclear relocalization by theArabidopsis transcription factor AtMYB30 leads to repressionof the plant defense response. Proc Natl Acad Sci U S A 2010,107:15281-15286.

This study describes the nuclear relocalization of a secreted phospho-lipase to negatively regulate defence-related transcriptional responsesmediated by the transcription factor MYB30. This work identifies a novelfunction for a phospholipase beyond the classical lipid hydrolyzingactivities.

67. Marino D, Froidure S, Canonne J, Ben Khaled S, Khafif M,Pouzet C, Jauneau A, Roby D, Rivas S: Arabidopsis ubiquitinligase MIEL1 mediates degradation of the transcription factorMYB30 weakening plant defence. Nat Commun 2013, 4:1476.

68.�

Canonne J, Marino D, Jauneau A, Pouzet C, Briere C, Roby D,Rivas S: The Xanthomonas type III effector XopD targets the

Arabidopsis transcription factor AtMYB30 to suppress plantdefence. The Plant Cell 2011, 23:3498-3511.

This work shows that, in Arabidopsis, the Xanthomonas effector XopDdirectly targets the nuclear MYB transcription factor MYB30, resulting insuppressed MYB30-mediated transcriptional activation and defenceresponses.

69.�

Kim JG, Stork W, Mudgett MB: Xanthomonas type III effectorXopD desumoylates tomato transcription factor SlERF4 tosuppress ethylene responses and promote pathogen growth.Cell Host Microbe 2013, 13:143-154.

The authors demonstrate that, in tomato, the Xanthomonas effector XopDtargets and deSUMOylates the nuclear transcription factor SlERF4 sup-pressing ethylene production, which is required for immunity againstXanthomonas and symptom development.

70. Block A, Toruno TY, Elowsky CG, Zhang C, Steinbrenner J,Beynon J, Alfano JR: The Pseudomonas syringae type IIIeffector HopD1 suppresses effector-triggered immunity,localizes to the endoplasmic reticulum, and targets theArabidopsis transcription factor NTL9. New Phytol 2013,201:1358-1370.

71.��

Gimenez-Ibanez S, Boter M, Fernandez-Barbero G, Chini A,Rathjen JP, Solano R: The bacterial effector HopX1 targets JAZtranscriptional repressors to activate jasmonate signaling andpromote infection in Arabidopsis. PLoS Biol 2014, 12:e1001792.

This study shows that the effector HopX1 interacts with and promotes thedegradation of JAZ repressors, through its cysteine protease activity.Expression of HopX1 in Arabidopsis induces the expression of JA-dependent genes and represses SA signalling. When delivered by bac-teria, HopX1 promotes plant susceptibility to a similar extent as theaddition of the JA-mimicking phytotoxin coronatine. See also the workof Jian et al. (Ref. [72��]).

72.��

Jiang S, Yao J, Ma KW, Zhou H, Song J, He SY, Ma W: Bacterialeffector activates jasmonate signaling by directly targetingJAZ transcriptional repressors. PLoS Pathog 2013, 9:e1003715.

The authors show that the effector HopZ1a interacts with and acetylatesJAZ proteins. Inoculation with Pseudomonas producing HopZ1a, but nota HopZ1a catalytic mutant, promotes degradation of JAZ repressors andactivates JA signallin. Furthermore, HopZ1a partially rescues the viru-lence defect of a Pseudomonas strain unable to produce coronatine. Thiswork, together with the work of Gimenez-Inanez et al. (Ref. [71��])suggests that the JA receptor complex is potentially a target hub forbacterial pathogens.

73.�

Fu ZQ, Guo M, Jeong BR, Tian F, Elthon TE, Cerny RL, Staiger D,Alfano JR: A type III effector ADP-ribosylates RNA-bindingproteins and quells plant immunity. Nature 2007, 447:284-288.

This study identifies the Pseudomonas effector HopU1 as a mono-ADP-ribosyltransferase able to ADP-ribosylate the Arabidopsis glycine-richRNA-binding protein GRP7 and suppress plant immunity.

74. Jeong BR, Lin Y, Joe A, Guo M, Korneli C, Yang H, Wang P, Yu M,Cerny RL, Staiger D et al.: Structure function analysis of an ADP-ribosyltransferase type III effector and its RNA-binding targetin plant immunity. J Biol Chem 2011, 286:43272-43281.

75.�

Nicaise V, Joe A, Jeong BR, Korneli C, Boutrot F, Westedt I,Staiger D, Alfano JR, Zipfel C: Pseudomonas HopU1 modulatesplant immune receptor levels by blocking the interaction oftheir mRNAs with GRP7. EMBO J 2013, 32:701-712.

This work extends the findings of Fu and co-workers (Ref. [73�]). Theauthors demonstrate that GRP7 binds to FLS2 transcripts in vivo and thatthis interaction is inhibited in the presence of HopU1, resulting in reducedFLS2 protein accumulation and suppressed PTI responses.

76. Conaway RC, Conaway JW: Origins and activity of the Mediatorcomplex. Semin Cell Dev Biol 2011, 22:729-734.

77.��

Caillaud MC, Asai S, Rallapalli G, Piquerez S, Fabro G, Jones JD: Adowny mildew effector attenuates salicylic acid-triggeredimmunity in Arabidopsis by interacting with the host mediatorcomplex. PLoS Biol 2013, 11:e1001732.

This work shows the nuclear interaction of the Hyaloperonospora arabi-dopsidis effector HaRxL44 with the subunit Med19a of the Mediatorcomplex, which results in degradation of Med19a in a proteasome-dependent manner. Med19a is a positive regulator of resistance to H.Arabidopsidis and HaRxL44 targeting of Med19a leads to enhanced JA/ET and decreased SA signalling, which increases susceptibility to bio-troph pathogens by attenuating SA-dependent gene expression.

78. Gao Z, Chung EH, Eitas TK, Dangl JL: Plant intracellular innateimmune receptor resistance to Pseudomonas syringae pv.

Transcriptional control of plant immunity Buscaill and Rivas 45

www.sciencedirect.com Current Opinion in Plant Biology 2014, 20:35–46

maculicola 1 (RPM1) is activated at, and functions on, theplasma membrane. Proc Natl Acad Sci U S A 2011,108:7619-7624.

79.��

Slootweg E, Roosien J, Spiridon LN, Petrescu AJ, Tameling W,Joosten M, Pomp R, van Schaik C, Dees R, Borst JW et al.:Nucleocytoplasmic distribution is required for activation ofresistance by the potato NB-LRR receptor Rx1 and isbalanced by its functional domains. Plant Cell 2010,22:4195-4215.

This work shows that interdomain interactions and folding states deter-mine the nucleocytoplasmic distribution of the potato R protein Rx1. Incontrast to the findings by Heidrich et al. (Ref. [50��]) and by Bai et al; (Ref[42]), the authors demonstrate that Rx1 is activated in the cytoplasm andcannot be activated in the nucleus. See also Ref. [81].

80.�

Tameling WI, Nooijen C, Ludwig N, Boter M, Slootweg E,Goverse A, Shirasu K, Joosten MH: RanGAP2 mediatesnucleocytoplasmic partitioning of the NB-LRR immunereceptor Rx in the Solanaceae, thereby dictating Rx function.Plant Cell 2010, 22:4176-4194.

This study shows that nucleocytoplasmic partitioning of Rx is regulatedby its interaction with RanGAP2, a protein that is involved in nucleocy-toplasmic trafficking of macromolecules through nuclear pores and that isrequired for extreme resistance to Potato Virus X. Together with the study

by Slootweeg and collaborators (Ref. [79��]), this study shows the crucialrole of nucleocytoplasmic distribution of Rx during the establishement ofdisease resistance.

81. Meissner TB, Li A, Biswas A, Lee KH, Liu YJ, Bayir E, Iliopoulos D,van den Elsen PJ, Kobayashi KS: NLR family member NLRC5 is atranscriptional regulator of MHC class I genes. Proc Natl AcadSci U S A 2010, 107:13794-13799.

82. Narusaka M, Shirasu K, Noutoshi Y, Kubo Y, Shiraishi T,Iwabuchi M, Narusaka Y: RRS1 and RPS4 provide a dualResistance-gene system against fungal and bacterialpathogens. Plant J 2009, 60:218-226.

83. Heidrich K, Tsuda K, Blanvillain-Baufume S, Wirthmueller L,Bautor J, Parker JE: Arabidopsis TNL-WRKY domain receptorRRS1 contributes to temperature-conditioned RPS4 auto-immunity. Front Plant Sci 2013, 4:403.

84. Kim SH, Gao F, Bhattacharjee S, Adiasor JA, Nam JC,Gassmann W: The Arabidopsis resistance-like gene SNC1 isactivated by mutations in SRFR1 and contributes to resistanceto the bacterial effector AvrRps4. PLoS Pathog 2010,6:e1001172.

46 Biotic interactions

Current Opinion in Plant Biology 2014, 20:35–46 www.sciencedirect.com

Figure 19. Schematic representation of the AtMYB30 protein. As other R2R3-MYB TFs, AtMYB30 presents an N-terminal MYB domain and a C-terminal

transactivation domain. C-terminal motifs 1, 2 and 3 were used to classify AtMYB30 in the

subgroup 1 of R2R3-MYB proteins.

TAD 2 R2 R3 1 3 C N

1 11 115 234 264 323

MYB Domain

Transcriptional

Activation Domain

(A) (B)

Figure 20. Analysis of AtMYB30 expression in Arabidopsis upon bacterial infection

(From Daniel et al., 1999). (A) RT-PCR analysis of the expression of AtMYB30 in Arabidopsis resistant (Col-0) or sensitive (Sf-

2) ecotypes, following inoculation with H2O or with the avirulent strain Xcc147, as indicated.

(B) Analysis by RT-PCR of AtMYB30 expression in Arabidopsis in response to virulent Pst DC3000

or avirulent Pst DC3000 AvrB or Pst DC3000 AvrRpm1 strains.

52

3. AtMYB30 a positive regulator of the HR in A. thaliana

3.1. Identification of AtMYB30

A differential screening was previously performed in our team in order to identify genes

potentially regulating the early events of the establishment of the HR in Arabidopsis thaliana.

This screen was based on the inoculation of Arabidopsis cell suspensions pre-treated with

cycloheximide (in order to focus on genes whose expression did not depend on de novo

protein biosynthesis) with two different strains of Xanthomonas campestris pv. campestris

(Xcc): the avirulent strain Xcc147 that leads to the development of the HR in Arabidopsis,

and the non-pathogenenic strain Xcc8B2 that is asymptomatic due to the deletion of the hrp

cluster. This differential screening allowed the identification of 27 genes early transcribed

during the establishment of the HR. These genes were named Arabidopsis thaliana

hypersensitivity-related (Athsr) and grouped in seven different gene families (Lacomme and

Roby, 1999). Sequence analysis indicated that one of these clones corresponded to a single-

copy gene that encoded the protein AtMYB30, a member of the family of R2R3 class of plant

MYB TFs (Lacomme and Roby, 1999, Daniel et al., 1999) (Figure 19). Given the potentially

interesting role of AtMYB30 during the establishment of the HR, its functional analysis was

next initiated.

3.2. Expression and function of AtMYB30

The analysis of the expression pattern of AtMYB30 was performed in Arabidopsis plants and

in cultured Arabidopsis cells infected with either virulent or avirulent strains of Xcc or Pst.

This study showed that the AtMYB30 transcript is early, transiently and specifically detected

after infection by avirulent bacteria (Figure 20). In contrast, AtMYB30 expression was not

detected in compatible interactions that lead to disease development, or in response to an

asymptomatic hrp- mutant of Xcc (Lacomme and Roby, 1999, Daniel et al., 1999).

Furthermore, in order to better understand the mode of action of AtMYB30, AtMYB30 gene

expression was monitored in Arabidopsis mutant plants affected either in the initiation or in

Figure 21. Overexpression of AtMYB30 in tobacco leads to accelerated HR in

response to inoculation with different pathogens (From Vailleau et al., 2002). (A) Phenotypes of wild-type tobacco and AtMYB30-overexpressing (AtMYB30OE) lines in response

to inoculation with R. solanacearum. Different inocula were used for the avirulent strain

(GMI1000), ranging from 5x107 cfu/ml (no. 1), 107 cfu/ml (no. 2), to 5x106 cfuml (no. 3). The

virulent strain (K60) was inoculated at 5x107 cfu/ml (no. 4), and a control inoculation was

performed with water (no. 5). Symptoms are shown 20 hours post inoculation (left) and 6

days post inoculation (right).

(B) Disease symptoms caused by Cercospora nicotianae 14 days after inoculation in the wild-type

and AtMYB30-overexpressing (AtMYB30OE) lines. The black circle indicates the inoculated

zone.

(B)

(A)

Wild type AtMYB30OE

Wild type Wild type AtMYB30OE AtMYB30OE

20 hours post inoculation 6 days post inoculation

53

the propagation of the HR (“initiation class” or “propagation class” mutants). AtMYB30 was

constitutively expressed in initiation mutants lsd4 and lsd5 (exhibiting spontaneous necrotic

lesions simulating disease resistance in the absence of the pathogen) and was expressed

only under lesion promoting conditions in lsd3 mutant plants. AtMYB30 transcripts did not

accumulate in different phx (phoenix) mutants that act as suppressors of the lsd5 mutation.

Similarly, the AtMYB30 transcript was not detectable in the propagation mutant lsd1, which is

hyper-responsive to cell death initiators and unable to limit the extent of cell death. These

results suggest a strong correlation between AtMYB30 expression and genetically controlled

cell death and indicate that AtMYB30 expression is associated with the initiation rather than

the spread of the cell death (Lacomme and Roby, 1999, Daniel et al., 1999).

A functional analysis of AtMYB30 during the HR was then conducted through the use of

transgenic tobacco and Arabidopsis plants deregulated for AtMYB30 expression. In these

plants, the coding sequence of AtMYB30 was expressed under the control of the constitutive

promoter 35S from Cauliflower Mosaic Virus either in the sense (AtMYB30OE) or the

antisense (AtMYB30AS) orientation. As compared to wild-type plants, AtMYB30OE plants

showed accelerated and stronger HR, enhanced PR1 defence gene expression and reduced

bacterial growth in response to avirulent bacterial strains (Ralstonia solanacearum GMI1000

in tobacco, and Pst DC3000 AvrRpm1 or Xcc147 in Arabidopsis). Interestingly, an HR-like

response and decreased plant susceptibility were observed when tobacco and Arabidopsis

AtMYB30OE plants were infected with virulent bacteria (Vailleau et al., 2002) (Figure 21A). In

addition, AtMYB30 overexpression in tobacco increased resistance against the fungal

pathogen Cercospora nicotianae (Figure 21B). On the other hand, plants overexpressing

AtMYB30 in an antisense orientation showed reverse phenotypes, with delayed and weaker

HR and resistance in response to avirulent pathogens, and enhanced disease symptoms and

bacterial growth rates in the context of a compatible interaction with various virulent

pathogens. In both interaction contexts, the observed phenotypes were associated with lower

PR1 marker gene expression. Together these data showed that AtMYB30 is a positive

54

regulator of the signalling pathway controlling the establishment of cell death-associated

resistance against pathogen attack (Vailleau et al., 2002).

3.3. Hormonal control of the AtMYB30-mediated HR

As mentioned in the previous section, SA plays a central role in the regulation of defence and

cell death-associated responses in Arabidopsis, particularly in response to biotrophic

pathogens (Pieterse et al., 2009). In order to elucidate whether AtMYB30 expression is

regulated by the SA pathway, AtMYB30 transcript levels were first quantified in Arabidopsis

mutants affected in both SA biosynthesis and signalling. AtMYB30 expression was found to

be reduced in SA-related mutants npr1, sid1 and sid2 as well as in plants overexpressing the

NahG gene that are unable to accumulate SA. In contrast, SA treatment induced AtMYB30

expression, confirming that AtMYB30 expression is SA-dependent. Moreover, altered

expression of AtMYB30 modulated SA levels and expression of SA-associated genes such

as ICS1 and PR1, suggesting that AtMYB30 expression may be critical to trigger the SA

defence signalling pathway. Overexpression of AtMYB30 in the NahG, sid1, sid2 or npr1

genetic backgrounds suppressed the enhanced cell death and resistance phenotypes

previously observed in AtMYB30OE plants, indicating that the phenotypes conferred by over-

expression of AtMYB30 are dependent on SA accumulation and signalling. Taken together

these data suggest that AtMYB30 is involved in an amplification loop of the signalling

cascade that modulates synthesis of the plant defence-related hormone SA, which in turn

modulates cell death (Raffaele et al., 2006).

The ability of AtMYB30 to activate JA production and JA-dependent defence responses was

also investigated. Modification of oxylipin profiles in AtMYB30 transgenic plants was also

reported (Vailleau et al., 2002). JA marker-gene analysis showed that AtMYB30 interferes

with the ability of JA to activate PDF1-2 and PR4 but not VSP1 and LOX3 marker gene

expression. Moreover, AtMYB30 overexpression in a jar1 mutant background, affected in JA

signal transduction, led to enhanced HR and resistance phenotypes indistinguishable from

those displayed by AtMYB30OE plants after inoculation by an avirulent strain of Pst. This

Figure 22. AtMYB30 modulates the expression of very long chain fatty acid

(VLCFA)-related genes after bacterial inoculation (From Raffaele et al., 2008). List of 18 genes identified as putative targets of AtMYB30. aAffymetrix Probe set number. bArabidopsis Genome Initiative number and corresponding putative function. Columns in the right

show fold changes in the mean expression levels from two independent experiments in inoculated

wild type compared with the wild type at t0 (WT1/WT0) and in the AtMYB30OE

line compared with

the wild type after Xcc147 inoculations (Ox/WT).

55

study production of JA and of 12-oxo-phytodienoic acid (OPDA), a JA precursor, was

evaluated by GC-MS and revealed that JA signal transduction does not play an essential role

in AtMYB30-mediated defence signalling (Raffaele et al., 2008).

3.4. Transcriptional targets of AtMYB30

To identify putative AtMYB30 target genes, a transcriptomic analysis with the Affymetrix chip

"Complete Genome" was performed using Arabidopsis wild-type, AtMYB30OE and

AtMYB30AS plants of the Ws-0 ecotype inoculated with the avirulent strain Xcc147. The

analysis of differentially expressed genes in the first 6 hours after inoculation led to the

identification of 18 genes as potential targets of AtMYB30 (Figure 22). Interestingly, 14 of

them presented a function related to lipid metabolism, and more particularly, in the lipid

biosynthesis pathway that leads to the production of VLCFAs (Figure 22). Within the VLCFA

pathway, genes encoding subunits of the Acyl-CoA Elongase complex were over-

represented suggesting that AtMYB30 modulates the early steps of fatty acid elongation.

VLCFAs are involved in the production of lipid second messengers, such as sphingolipids or

ceramides, or in the production of epicuticular waxes and all these molecules have been

shown to be involved in defence and cell death regulation (Raffaele et al., 2009, Berkey et

al., 2012, Markham et al., 2013).

Molecular, genetic and biochemical studies were next undertaken to validate these

transcriptomic data. Regulation of the expression of these 14 putative target genes by

AtMYB30 after bacterial inoculation was further confirmed by quantitative RT-PCR not only in

the transgenic lines but also in a T-DNA insertion knockout (atmyb30ko) mutant. Interestingly,

following inoculation, the expression of these genes was higher in AtMYB30OE and lower in

atmyb30ko and AtMYB30AS lines. These results are consistent with these genes being

transcriptional targets of AtMYB30. Further support for this idea was obtained in

transactivation assays in Nicotiana benthamiana, in which AtMYB30 expression led to

activation of the reporter genes GUS and GFP fused to the promoter of the candidate target

genes. In addition, the levels of VLCFAs from a sphingolipid-enriched fraction were affected

Figure 23. Schematic overview of metabolic pathways regulated by AtMYB30 during

the incompatible interaction between Arabidopsis and avirulent bacterial pathogens

(From Raffaele et al., 2008). Elements strongly and positively regulated by AtMYB30 are indicated in red (genes of the acyl-CoA

elongase complex, accumulation of very long fatty acids (VLCFAs), accumulation of salicylic acid

(SA), and cell death). Other elements positively but weakly regulated by AtMYB30 are indicated in

orange (upstream steps of VLCFA synthesis, wax- and cutin-related genes, and wax accumulation).

Elements shown in gray are not significantly regulated by AtMYB30, at least during the early

events of the interaction, and elements shown in blue are negatively regulated (directly or

indirectly) by AtMYB30 (expression of genes encoding fatty acid (FA) desaturases). Inoculation

could trigger the synthesis of new signalling molecules as indicated by the dotted arrow. These

molecules, together with other lipid-derived signals, such as SA, oxylipins, jasmonic acid (JA), and

phospholipids, would subsequently act for activation of the hypersensitive cell death. Synthesis of

these VLCFA-derived signals may be enhanced by AtMYB30. Upregulation of VLCFA synthesis in

non-challenged AtMYB30OE

plants allows enhanced accumulation of VLCFAs and of VLCFA

metabolizing enzymes. The cell is thus predisposed for a stronger and faster response to pathogen

attack. CW, cell wall; OPDA, 12-oxo-phytodienoic acid; PM, plasma membrane; PUFAs,

polyunsaturated fatty acids.

56

in the transgenic and mutant lines, and these alterations were enhanced in response to

pathogen challenge. Finally, loss of function of the thioesterase FATB, which impairs the

supply of fatty acids for VLCFA biosynthesis, reversed the accelerated and intensified HR

phenotype of the AtMYB30OE line. Based on these findings, a model was proposed in which

AtMYB30 modulates cell death lipid signalling by enhancing the synthesis of VLCFAs in the

ER (Raffaele et al., 2008) (Figure 23). This supports the idea that AtMYB30 target genes are

involved in VLCFA biosynthesis and/or derivation and that VLCFAs and/or some of their

derivatives are involved in establishment or control of the HR cell death program (Raffaele et

al., 2008).

Recently, our group performed oriented deep sequencing of polyA-enriched RNAs from

Arabidopsis wild-type, AtMYB30OE and atmyb30ko plants of the Col-0 ecotype inoculated with

the avirulent strain Pst DC3000 AvrRpm1. The results on this RNA sequencing experiment

confirmed the VLCFA pathway as a target of AtMYB30 activity.

3.5. Regulation of AtMYB30

During the last few years, the study of AtMYB30 regulatory mechanisms has significantly

contributed to further our understanding about the mode of action of this TF and, consistent

with its role in regulating defence-related cell death associated responses, uncovered that

the activity of AtMYB30 is tightly controlled by the plant cell through different regulatory

mechanisms that are described below (Raffaele and Rivas, 2013).

3.5.1. Post-transcriptional regulation of AtMYB30

The crucial role of RNA silencing is well established in the control of plant development as

well as in the plant response to adverse environmental conditions, including biotic stresses

(Kamthan et al., 2015). AtMYB30 transcript is well detected in young seedlings while its

expression rapidly decreases later in development to be very weak in adult four weeks-old

plants. As detailed above, in adult plants AtMYB30 expression is induced after pathogen

challenge (Daniel et al., 1999). In AtMYB30OE lines the same expression pattern was

57

observed, both during development and after bacterial inoculation, despite the fact that in

these lines AtMYB30 expression is under the control of a 35S promoter. In addition, the level

of AtMYB30 protein accumulation correlated with that of the transcript, both in wild-type and

AtMYB30OE lines. These data suggest the existence of a post-transcriptional mechanism that

tightly controls expression of AtMYB30. Indeed, work from our group suggests that RNA

silencing directly mediates downregulation of AtMYB30 expression both in young seedlings

and in adult plants. In contrast, an indirect RNA silencing mechanism appears to be

responsible for the induction of AtMYB30 expression after bacterial inoculation, possibly via

the degradation of a yet unknown negative regulator of its expression (Froidure et al.,

2010b). These results underline the role of small RNAs in the regulation of TF activity both

during plant development and in response to pathogen attack.

3.5.2. Post translational modification of AtMYB30

Our group and others have shown that AtMYB30 activity is modulated through a combination

of different PTMs, some of which are described below.

In silico AtMYB30 protein sequence analysis revealed that the C-terminal region of this TF is

particularly rich in phosphorylation sites for different protein kinases. The study of the in

vivo phosphorylation of AtMYB30 was therefore carried out in the team and demonstrated

using two-dimensional electrophoresis, that AtMYB30 is phosphorylated in planta

(unpublished data). The biological role of the phosphorylation of AtMYB30 has not yet been

established but it is probable that the activity of AtMYB30 in the plant is regulated by

phosphorylation.

A second in silico analysis of the AtMYB30 protein sequence allowed the identification of

several lysine residues with different probabilities of being targeted by SUMOylation.

Interestingly, results from mass spectrometry analyses on purified AtMYB30 protein were

consistent with the fact that some of these lysine residues may be the target of PTM by small

peptides such as SUMO or ubiquitin (unpublished data). The hypothesis of AtMYB30

58

SUMOylation was tested in the group and the results showed that AtMYB30 is SUMOylated

in vitro and in vivo. Different AtMYB30 lysine residues potentially targeted by SUMOylation

were modified by site-directed mutagenesis and the analysis of the transcriptional activity of

these AtMYB30 mutant versions was initiated. Interestingly, mutation of certain lysine

residues modulates AtMYB30 capacity to activate VLCFA-related target genes in

transactivation assays in Nicotiana benthamiana (unpublished data). Moreover, studies by

Okada and colleagues identified SUMOylation sites within AtMYB30 by reconstituting the

SUMOylation cascade in Escherichia coli (Okada et al., 2009). In addition, Zheng and co-

workers show that AtMYB30 SUMOylation by SIZ1, an Arabidosis E3-SUMO protein ligase,

leads to AtMYB30 protein stabilization and affects AtMYB30-mediated transcriptional

activation of several ABA-responsive genes (Zheng et al., 2012), underlining the importance

of AtMYB30 SUMOylation during the regulation of ABA signalling. However, a more detailed

functional analysis of the importance of SUMOylation for AtMYB30-mediated defence

remains to be performed.

Research in our group showed that AtMYB30 is additionally modified by ubiquitination

following interaction with the RING-type E3 ligase protein MYB30-interacting E3 ligase1

(MIEL1). This ubiquitination leads to AtMYB30 degradation by the proteasome. More details

about the outcome of AtMYB30 ubiquitination by MIEL1 are provided in the next section.

Finally, recent studies by Tavares and coworkers demonstrate that S-nitrosylation of the

MYB domain of AtMYB30 negatively regulates its DNA-binding activity in vitro since PTM of

AtMYB30 by nitrosylation is able to modify the secondary structure and the thermal stability

of the protein (Tavares et al., 2014).

3.5.3. Regulation of AtMYB30 activity through protein-protein interactions

XopD from strain B100 of Xcc (XopDXccB100) is a type III effector protein that presents a

modular structure and contains different domains with varied biochemical activities (Canonne

et al., 2010). XopDXccB100 is targeted to plant cell nuclei (Canonne et al., 2012, Canonne et

Figure 24. Simplified model for the simultaneous regulation of AtMYB30-mediated HR

cell death through interaction with AtsPLA2- and MIEL1 (Adapted from Raffaele and

Rivas, 2013). The action of with AtsPLA2-α and MIEL1 on AtMYB30-mediated HR development is presented in cells

challenged with bacterial inoculation (A) and peripheral cells (B). Activity of the bacterial XopD

effector is shown in blue. See the text for details.

Nucleus

VLCFA genes

ETI

HR

VLCFA genes

AtMYB30

AtsPLA2-α

MIEL1 inactive AtMYB30

AtMYB30

degradation ETI

HR

IEL1AtMYB30

relocalize Nucleus

XopD inactive

AtMYB30

(B)

(A)

Challenged cell

Peripherical cell

AtsPLA2-α

MIEL1

59

al., 2011, Kim et al., 2008) and has been proposed to interact with chromatin and/or

transcriptional units, leading to modulation of host transcription by affecting chromatin

remodeling and/or TF activity.

In agreement with the idea that plant TFs and/or regulators might be direct targets of XopD,

XopDXccB100 was shown to target AtMYB30. XopDXccB100 expression leads to accumulation of

AtMYB30 in XopDXccB100-containing nuclear foci but the physical interaction between

XopDXccB100 and AtMYB30 is independent of AtMYB30 relocalization to nuclear foci, as both

proteins are also able to interact in the nucleoplasm (Canonne et al., 2011). XopDXccB100

targeting of AtMYB30 leads to reduced activation of AtMYB30 VLCFA-related target genes

and, therefore, to suppression of plant defence responses during infection by XccB100

(Canonne et al., 2011) (Figure 24). A helix-loop-helix (HLH) domain in XopDXccB100 is

necessary and sufficient to mediate the interaction with AtMYB30 and repression of

AtMYB30 transcriptional activation and plant resistance responses. Consistently, XopD from

the 8004 strain of Xcc (XopDXcc8004), which does not present the HLH domain and localizes

homogenously within plant cell nuclei, is not able to interact with AtMYB30 and has no effect

on AtMYB30 transcriptional activation or AtMYB30-mediated defence. Since R2R3-MYBs

typically function in association with bHLH factors (Pireyre and Burow, 2015), the central role

played by the HLH domain of XopD XccB100 in repressing AtMYB30 function represents an

additional example of how microbes have evolved adapted microbial molecular strategies to

subvert resistance responses by the host. Together, these data highlight the importance of

AtMYB30-mediated activation of the VLCFA pathway for directing the plant defence

response.

In addition to the regulation of AtMYB30 via the interaction with the bacterial effector protein

XopD, our group has shown that the activity of this TF is additional controlled through its

interaction with plant proteins. Indeed, overexpression of AtMYB30 in transgenic plants does

not lead to the formation of spontaneous HR-like lesions in the absence of the pathogen,

suggesting that AtMYB30 probably acts in cooperation with additional factor(s) for initiation of

60

the HR (Vailleau et al., 2002). To search for AtMYB30 interacting proteins a Yeast Two-

Hybrid (Y2H) screen was performed using an AtMYB30 version deleted from its

transcriptional activation domain (AtMYB30ΔTAD) as bait to screen a cDNA library from

Arabidopsis plants inoculated with avirulent bacteria. Several AtMYB30-interacting partners,

identified or not in this Y2H screen, are described here after.

Expression of AtMYB96, an additional MYB TF that belongs to the S1 phylogenetic

subgroup, is rapidly induced during incompatible interactions. AtMYB96 has been shown to

transcriptionally activate VLCFA biosynthesis leading to the production of epicuticular waxes

under drought conditions (Seo et al., 2011b). Our group has shown that AtMYB96 directly

interacts with AtMYB30 in the plant cell nucleus and collaborates with AtMYB30 to positively

regulate the hypersensitive cell death (unpublished data). Interestingly, AtMYB30 and

AtMYB96 are able to regulate each other’s expression, suggesting a complex molecular

interaction between these two TFs. This regulatory feedback loop, associated to the physical

interaction and possible competition towards the same target genes leads to fine and subtle

regulation of lipid metabolism. Taken together, these results show that AtMYB30 and

AtMYB96 are main components of a transcriptional rheostat that controls the establishment

of the hypersensitive cell death pathway through the production of sphingolipid-containing

VLCFAs. This is an original finding since no physical interaction between two MYB proteins

of the R2R3-type has been previously reported.

A first candidate that was characterized following its identification in the Y2H screen is

AtsPLA2-α , a secreted Arabidopsis phospholipase that is specifically relocalized to the

nucleus in the presence of AtMYB30 (Figure 24). In addition, AtsPLA2-α and AtMYB30

physically interact in the nucleus and this protein interaction leads to repression of AtMYB30

transcriptional activity and negative regulation of plant HR and defence, revealing that

AtsPLA2-α is a negative regulator of AtMYB30-mediated defence (Froidure et al., 2010a).

Interestingly, this study suggested that AtPLA2-α contributes to restrict HR development to

the inoculated zone, thereby preventing spreading of cell death throughout the leaf (Froidure

61

et al., 2010a) (Figure 24). This work (i) represented a first identification of a secreted

phospholipase as a negative regulator of defence-associated plant responses and (ii)

underlined the importance of cellular dynamics, and particularly protein translocation to the

nucleus, for defence-associated gene regulation in plants (Rivas, 2012).

A second protein identified in the previously mentioned Y2H screen is MYB30-Interacting E3

Ligase1 (MIEL1), a RING-type E3 ubiquitin-ligase that interacts with AtMYB30 in the nucleus

(Figure 24). This protein interaction leads to ubiquitination of AtMYB30 and its degradation

by the proteasome resulting in inhibition of AtMYB30 transcriptional activity and suppression

of HR and defence responses (Figure 24) (Marino et al., 2013). In agreement with these

observations, Arabidopsis miel1 mutant plants displayed enhanced HR and resistance after

inoculation with avirulent bacteria. These phenotypes were AtMYB30-dependent and

correlated with downregulation of AtMYB30 target gene expression. Following this

observation a working model was proposed where in non-infected plants, MIEL1 attenuated

cell death and defence through degradation of AtMYB30. Following bacterial inoculation,

repression of MIEL1 expression removed this negative regulation allowing sufficient

AtMYB30 accumulation in the inoculated zone to trigger HR and restrict pathogen growth

(Marino et al., 2013). This work showed the important role played by ubiquitination to control

the HR and underlined the sophisticated fine-tuning of plant responses to pathogen attack.

62

Scientific context of the PhD project

Among the potential interactor candidates identified in the previously performed Y2H screen,

a cDNA encoding the last 103 amino acids of a serin-type endopeptidase was identified

twice. This serine protease is AtSBT5.2, an Arabidopsis protease of the subtilase family that

therefore appeared as a new putative AtMYB30-interacting candidate.

As discussed above, although the importance of plant proteases during the plant response to

pathogens has been clearly established, the involvement of subtilase proteins in plant

defence is still poorly characterized (see Introduction, section 1.3). Therefore, the

identification of AtSBT5.2 as a new interactor of the well characterised defence regulator

AtMYB30 provided us with a good opportunity to characterize its potential role as a regulator

of plant disease resistance and defence-associated cell death responses.

63

Objectives of the PhD project

My thesis project is part of a more general project in our group that aims to further our

understanding of the molecular mechanisms that regulate AtMYB30 activity during the

establishment of the HR. More particularly, during my PhD, I focused on the functional

characterization of AtSBT5.2 as a new interacting partner of AtMYB30 with the following

objectives:

· To validate the interaction between AtSBT5.2 and AtMYB30 in planta.

· To study the catalytic activity of AtSBT5.2 and its effect on AtMYB30 protein

accumulation.

· To characterize the mode of action of AtSBT5.2 on AtMYB30-mediated defence

responses.

64

Results

__________________________________________________________________________

Figure 25. Interaction between AtMYB30 and AtSBT5.2 in yeast. Yeasts are shown after growth for 5 days on low stringency (left; SD/-TL) or high stringency (right;

SD/-TLHA) media. Co-expression of AtMYB30 deleted from its C-terminal transcription activation

domain (MYB30DAD) and the isolated cDNA clone encoding the last 103 amino acids of AtSBT5.2

(SBT5.2) resulted in yeast growth on selective medium. In a control experiment, yeast cells expressing

MYB30DAD or SBT5.2 with controls provided by Clontech (T-antigen or P53, respectively) were not

able to grow on selective medium. BD, GAL4 DNA-binding domain; AD, GAL4 activation domain.

65

A protease of the subtilase family negatively regulates plant defence

through its interaction with the Arabidopsis transcription factor AtMYB30

Previous results: Identification of AtSBT5.2 as a new AtMYB30 interacting partner.

In order to search for AtMYB30-interacting partners, an AtMYB30 version deleted from its C-

terminal TAD (AtMYB30ΔAD) was previously used as bait to screen a Y2H Arabidopsis

cDNA library generated from mRNAs isolated from leaf tissue inoculated with the avirulent

bacterial strain Xcc147 (Froidure et al., 2010a). A cDNA clone encoding the last 103 amino

acids of the Arabidopsis serine protease of the subtilisin-like family AtSBT5.2 (At1g20160)

was identified (Figure 25). AtSBT5.2 belongs to subgroup V, which contains 6 members

(AtSBT5.1-6), within the classification of the 56 members of the Arabidopsis subtilase family

(Figure 15) (Schaller et al., 2012, Rautengarten et al., 2005).

Figure 26. AtSBT5.2 is alternatively spliced. (A) Two gene models (At1g20160.1 and At1g20160.2) for AtSBT5.2 as shown in the TAIR database.

Exons are shown as dark blue boxes, introns as blue lines between exons and 5’ and 3’ UTRs are

shown in light blue.

(B) RT-PCR analysis of AtSBT5.2(a) and AtSBT5.2(b) transcripts in Col-0 Arabidopsis leaves.

(C) Nucleotide sequences of 5’ ends of AtSBT5.2(a) and AtSBT5.2(b) transcripts as obtained by

5’RACE amplification. The ATG encoding the fist coding Met residue in each protein is boxed in

blue. Identical sequences are highlighted in red. From the ATG in AtSBT5.2(b) both nucleotide

sequences are also identical and, thus, not shown. Sequences used to specifically amplify

AtSBT5.2(a) or AtSBT5.2(b) are underlined.

(D) Schematic representation of AtSBT5.2(a) and AtSBT5.2(b) protein sequences. The signal peptide

(SP) and the prodomain (PD) in AtSBT5.2(a) are respectively shown as black and grey boxes.

Catalytical conserved residues Asp (D), His (H), Asn (N), and Ser (S) residues are indicated.

Putative N-glycosylation sites are represented by black dots and their amino acid position

indicated. The C-terminal region of 103 amino acids encoded by the partial cDNA clone identified

in the yeast two-hybrid screen as interacting with AtMYB30DAD is boxed in blue.

(B) (A)

(C)

(D)

AtSBT5.2(a) - 769 1 -

SP PD

D H N S

AtSBT5.2(b) - 730 1 - D H N S

_ 2

25

_ 3

63

_ 4

67

_ 5

25

_ 6

36

_

65

0

_ 6

78

_ 1

86

_ 3

24

_ 4

28

_ 4

86

_ 5

97

_

61

1

_ 6

39

At1g20160.1 (AtSBT5.2(a))

At1g20160.2 (AtSBT5.2(b))

1 kb

AtSBT5.2(a)

AtSBT5.2(b)

AtSBT5.2(a)

AtSBT5.2(b)

! ! ! ! ! ! ! !

SBT5.2(a) GAATAAGTCTTTCCAGTGATTAGGAAACTACAAAGCC ATG AAAGGCATTACATTCTTCACACCCTTTTTATCATTTCTAT 80

SBT5.2(b) ACTCATAATTCTTTTGATCTATCTATAGCTTCCAGTGTCTCTCAACCT 48

! ! ! ! ! ! ! !

SBT5.2(a) ATCTCTTATGCATCTTGTTTATGACAGAAACTGAAGCTGGGTCGAGAAATGGTGATGGGGTCTACATTGTCTACATG… 157

GTATAAATACCCTTTTCTTGAGTTGAGAAACTGAAGCTGGGTCGAGAAATGGTGATGGGGTCTACATTGTCTAC ATG … 125

1000 bp -

650 bp -

500 bp -

66

1. Characterization of AtSBT5.2

1.1. AtSBT5.2 is alternative spliced and encodes two distinct isoforms.

Two gene models (splice variants) are annotated in the Arabidosis database TAIR

(http://www.arabidopsis.org) for the gene AtSBT5.2 (At1g20160), suggesting that its

transcript is alternatively spliced (Figure 26A). These two splice variants (At1g20160.1 and

At1g20160.2) were renamed AtSBT5.2(a) and AtSBT5.2(b), respectively. The two

corresponding transcripts contain different 5’UTRs and are predicted to encode two distinct

proteins.

To first investigate the existence of the two transcripts in planta, specific primers were

designed for each 5’UTR (in sense orientation) and one common primer in the fourth exon of

the AtSBT5.2 sequence (in antisense orientation). Semi quantitative RT-PCR analysis, using

cDNA from Col-0 Arabidopsis leaves, allowed amplification of both transcripts demonstrating

that both AtSBT5.2(a) and AtSBT5.2(b) are expressed in planta (Figure 26B). 5’RACE

experiments followed by cDNA sequencing confirmed that two distinct transcripts exist in

Col-0 leaves. The sequences of the 5’ end of both AtSBT5.2(a) and AtSBT5.2(b) mRNAs are

presented in Figure 26C.

AtSBT5.2(a) corresponds to a transcript of 2402 bp, which is predicted to encode a 769

amino acid preproenzyme containing a 27 amino acid signal peptide followed by a

prodomain of 74 amino acids (amino acids 35 to 108) and a 661 amino acid mature

polypeptide with a predicted molecular mass of 69.5 kDa (AtSBT5.2(a)) (Figure 26D). In

contrast, AtSBT5.2(b) corresponds to a transcript of 2373 bp, which is predicted to encode a

protein of 730 amino acids with no SP, and lacking the first five amino acids of the prodomain

in AtSBT5.2(a) and a predicted molecular mass of 77 kDa (AtSBT5.2(b)) (Figure 26D).

Except for their N-terminal differences, the two corresponding encoded proteins are identical.

Both predicted proteins contain the three conserved amino acids of the catalytic triad

characteristic of the subtilase family. Indeed, on the basis of sequence similarities with other

Figure 27. Sequence alignment of AtSBT5.2(a) and AtSBT5.2(b) proteins. Identical amino acids are highlighted in blue. The signal peptide and prodomain in AtSBT5.2(a) are

boxed in red and yellow, respectively. Catalytical conserved residues are indicated by red dots.

Putative N-glycosylation sites (PGSs) are indicated by blue dots. The 103 C-terminal amino acids

encoded by the partial AtSBT5.2 cDNA clone identified in the yeast two-hybrid screen as interacting

with AtMYB30ΔAD is boxed in blue. A putative myristoylation domain in AtSBT5.2(b) is boxed in

green.

AtSBT5.2(a)

AtSBT5.2(b)

AtSBT5.2(a)

AtSBT5.2(b)

AtSBT5.2(a)

AtSBT5.2(b)

AtSBT5.2(a)

AtSBT5.2(b)

AtSBT5.2(a)

AtSBT5.2(b)

AtSBT5.2(a)

AtSBT5.2(b)

AtSBT5.2(a)

AtSBT5.2(b)

AtSBT5.2(a)

AtSBT5.2(b)

Signal peptide Prodomain

Figure 28. Subcellular localization studies show that AtST5.2(a) is secreted whereas

AtSBT5.2(b) is intracellular.

Confocal images of epidermal cells of N. benthamiana leaves 36 hours after Agrobacterium-

mediated transient expression of the indicated constructs. Bars, 10 µm. RFP, Red Fluorescent

Protein.

AtS

BT

5.2

(a)-

RF

P

RFP fluorescence Bright Field Merged

AtS

BT

5.2

(b)-

RF

P

Figure 29. Intercellular fluid isolation confirms that AtSBT5.2(a) is secreted whereas

AtSBT5.2(b) is intracellular. Western blot analysis of total protein extracts (TE) and intercellular fluids (IF) of N. benthamiana leaf

tissue co-expressing the intracellular protein MIEL1 (Marino et al., 2013) with AtSBT proteins. All

proteins were HA-tagged and detected using anti-HA antibodies (a-HA). Molecular mass markers in

kiloDaltons are indicated on the right.

AtS

BT

5.2

(b)-

HA

MIEL1-HA +

a-HA

. 130

. 95

. 34

. 26

. 72

. 43

. 55

MIEL1

AtS

BT

5.2

(a)-

HA

67

subtilases the amino acid residues D145, H210 and S546 in AtSBT5.2(a) and D106, H171

and S507 in AtSBT5.2(b) were identified as residues of the catalytic triad (Figure 26D). The

complete deduced amino acid sequences of AtSBT5.2(a) and AtSBT5.2(b) are presented in

Figure 27.

1.2. AtSBT5.2(a) is a secreted protein whereas AtSBT5.2(b) is intracellular.

Alternative splicing of AtSBT5.2 may have important implications on the subcellular

localization of the proteins encoded by the two transcripts. The presence of a SP and a

prodomain in AtSBT5.2(a) suggests that this protein may enter the secretory pathway and be

secreted to the extracellular space. Indeed, secretion of AtSBT5.2(a) was previously reported

(Engineer et al., 2014). In contrast, the absence of the SP and the first five amino acids of

the prodomain in AtSBT5.2(b) may prevent this protein from being secreted. In order to test

this possibility, the subcellular localization of the two proteins was first investigated using

Agrobacterium-mediated transient expression of RFP-tagged AtSBT5.2(a) and AtSBT5.2(b)

versions under the control of the constitutive 35S promoter in leaf epidermal cells of N.

benthamiana. As expected, AtSBT5.2(a) was found to be located in the extracellular space,

confirming that AtSBT5.2(a) is a secreted protein (Figure 28). In contrast, AtSBT5.2(b) was

not secreted but detected in mobile punctuated structures inside cells (Figure 28).

In order to obtain biochemical validation of the distinct subcellular localization of AtSBT5.2(a)

and AtSBT5.2(b), HA-tagged versions of AtSBT5.2(a) and AtSBT5.2(b) were transiently

expressed in N. benthamiana and intercellular fluid (IF), where the secreted proteins are

expected to be present, was isolated. In order to control the detection of intracellular proteins

in the IF, the intracellular protein MIEL1 (Marino et al., 2013) was co-expressed with

AtSBT5.2 proteins in these assays. Being intracellular, MIEL1 was systematically detected in

the total extract fraction (TE) and not in the IF, confirming that the IF fraction did not contain

intracellular proteins due to undesired cellular lysis during IF isolation (Figure 29).

Figure 31. Sequence alignment of AtSBT5.2(a) and AtSBT5.1 proteins. Identical amino acids are highlighted in blue. Signal peptides and prodomains are boxed in red and

yellow, respectively. Catalytical conserved residues are indicated by red dots. Putative N-

glycosylation sites are boxed in orange. The 103 C-terminal amino acids encoded by the partial

AtSBT5.2 cDNA clone identified in the yeast two-hybrid screen as interacting with AtMYB30DAD are

underlined.

Figure 30. Schematic representation of AtSBT5.2(a) and AtSBT5.1 protein sequences. The signal peptide (SP) and the prodomain (PD) in AtSBT5.2(a) and its closest homolog AtSBT5.1 are

respectively shown as black and grey boxes. Catalytically conserved Asp (D), His (H), Asn (N), and Ser

(S) residues are indicated. Putative N-glycosylation sites are represented by black dots, and their

amino acid positions indicated, except for shared PGSs between AtSBT5.2 and AtSBT5.1, which are

represented by red dots.

AtSBT5.2(a)

AtSBT5.1

AtSBT5.2(a)

AtSBT5.1

AtSBT5.2(a)

AtSBT5.1

AtSBT5.2(a)

AtSBT5.1

AtSBT5.2(a)

AtSBT5.1

AtSBT5.2(a)

AtSBT5.1

AtSBT5.2(a)

AtSBT5.1

AtSBT5.2(a)

AtSBT5.1

AtSBT5.2(a)

AtSBT5.1

AtSBT5.2(a)

AtSBT5.1

Signal peptide Prodomain

AtSBT5.1 - 780 1 -

SP PD

D H N S

_ 1

97

_ 2

30

_ 4

71

_ 7

76

AtSBT5.2(a) - 769 1 -

SP PD

D H N S

_ 2

25

_ 3

63

_ 4

67

_ 5

25

_ 6

36

_

65

0

_ 6

78

68

AtSBT5.2(a) was detected in the IF, whereas AtSBT5.2(b) was exclusively detected in the

TE and never in the IF fraction (Figure 29). These results demonstrate the secretion and

intracellular localization of AtSBT5.2(a) and AtSBT5.2(b), respectively.

1.3. AtSBT5.2(a), but not AtSBT5.2(b), is glycosylated in planta.

In silico analysis of AtSBT5.2 proteins indicated that both polypeptides comprise seven

Putative asparagine-linked Glycosylation Sites (PGS; N in NxS/T motifs): amino acids N225,

N363, N467, N522, N636, N650 and N678 in AtSBT5.2(a), and N186, N324, N428, N486,

N597, N611 and N639 in AtSBT5.2(b) (Figure 26 and 27). According to the Rautengarten

phylogenetic classification (Figure 15), the closest homolog of AtSBT5.2 is AtSBT5.1

(At1g20150). In contrast to AtSBT5.2, a single gene model (AtSBT5.1) is annotated for

At1g20150 in the TAIR database, suggesting that AtSBT5.1 is not alternatively spliced.

AtSBT5.1 corresponds to a transcript of 2343 bp, which is predicted to encode a 780 amino

acid preproenzyme containing a 25 amino acids SP followed by an 78-amino acid prodomain

(amino acids 28 to 106) and a 674 amino acid mature polypeptide with a predicted molecular

mass of 73.8 kDa (Figure 30). The amino acid residues Asp-146, His-215 and Ser-550 were

identified as residues of the catalytic triad in AtSBT5.1 (Figure 30). AtSBT5.1 shows four

PGSs (N197, N230, N471 and N776) two of which (N230 and N471) are conserved in

AtSBT5.2 (Figure 30). Figure 31 shows an alignment of AtSBT5.2(a) and AtSBT5.1 protein

sequences.

To first investigate whether AtSBT5.2 proteins are glycosylated in planta protein extracts

from N. benthamiana leaves transiently expressing AtSBT5.2(a) and AtSBT5.2(b) were

subjected to purification using a concanavalin A resin. Concanavalin A is a lectin

(carbohydrate-binding protein) that is often used to purify glycosylated macromolecules by

affinity chromatography. As shown in Figure 32A, AtSBT5.2(a) binds to the concanavalin A

resin from which it can be eluted by adding an excess of a-methyl-D-glycosamide and a-

Figure 32. AtSBT5.2(a), but not AtSBT5.2(b), is glycosylated in planta. The indicated HA-tagged AtSBT proteins were transiently expressed in N. benthamiana leaves.

Proteins were detected by immunoblot with anti-HA antibodies (a-HA). Molecular mass markers in

kiloDaltons are indicated on the right.

(A) HA-tagged AtSBT5.1 and AtSBT5.2(a), but not AtSBT5.2(b), can be affinity purified using a

concanavalin A resin. The presence of different proteins at the indicated steps of the affinity

purification is shown.

(B) AtSBT5.1 and AtSBT5.2(a), but not AtSBT5.2(b), are deglycosylated by PNGase F or Endo H.

Protein extracts containing the indicated HA-tagged AtSBT proteins were treated (+) or not (-)

with PNGase F or Endo H.

(C) Glycosylation of HA-tagged AtSBT5.1 and AtSBT5.2(a), but not AtSBT5.2(b), is blocked by

tunicamycin treatment N. benthamiana leaves transiently expressing the indicated proteins were

treated (+) or not (-) with tunicamycin; as indicated. Ponceau S staining confirms equal loading.

(B)

(A)

(C)

69

methyl-D-manosamide. AtSBT5.1 was also purified under the same conditions (Figure 32A),

suggesting that both AtSBT5.2(a) and AtSBT5.1 are glycosylated in planta.

In order to obtain additional proof of the in planta N-glycosylation of these subtilases, protein

extracts containing AtSBT5.2(a) and AtSBT5.1 were analysed by Western blot after

treatment with deglycosylases PNGase F or EndoH. The electrophoretic mobility shift

observed in Figure 32B confirmed that AtSBT5.2(a) and AtSBT5.1 are N-deglycosylated and

that the N-linked carbohydrate is very likely to be high manose. Finally, we investigated the

effect of tunicamycin, which inhibits N-linked glycosylation of newly synthesized

glycoproteins in the ER (Bassik and Kampmann, 2011), on glycosylation of AtSBT5.2(a) and

AtSBT5.1 proteins. The reduced electrophoretic mobility observed after tunicamycin

treatment in the case of AtSBT5.2(a) and AtSBT5.1 provided further evidence of the in planta

N-glycosylation of AtSBT5.2(a) and AtSBT5.1 (Figure 32C). It is worth noting that tunicamyn

treatment resulted in reduced accumulation of both AtSBT5.2(a) and AtSBT5.1, suggesting

that glycosylation may contribute to the stability of both subtilases.

Finally, to determine the contribution of each putative PGS to AtSBT5.2(a) and AtSBT5.1 N-

glycosylation in planta, individual PGS removal mutants, in which the N residue was replaced

by A, were transiently expressed in N. benthamiana and analysed on high resolution SDS-

PAGE gels to determine their electrophoretic mobility as compared to that of wild-type

proteins. Protein extracts containing mutant proteins were separated on 7.5% acrylamide

SDS-PAGE gels and run overnight at low voltage, with interspaced wild-type AtSBT5.2(a)

protein to facilitate detection of the small size difference expected from removing just one

PGS. This analysis revealed slightly, but significantly, reduced mobility for all PGS mutants in

AtSBT5.2(a) as compared to wild-type AtSBT5.2(a) (Figure 33A), suggesting that all PGS in

AtSBT5.2(a) are used in planta. Similarly, a reduction in the electrophoretic mobility was

observed for all AtSBT5.1 N-glycosylation mutants, except N471A, indicating that all PGS but

not N471, [which is conserved (Figure 31) and glycosylated in AtSBT5.2(a) (Figure 33A)] are

used for AtSBT5.1 glycosylation in planta (Figure 33B).

Figure 33. All PGS in AtSBT5.2(a) are used for glycosylation in planta.

Electrophoretic mobility of HA-tagged individual PGS removal AtSBT5.2(a) (A) and AtSBT5.1 (B)

mutants. N to A substitutions in AtSBT5.2(a) and AtSBT5.1 are indicated. Wild-type (WT) AtSBT5.2(a)

or AtSBT5.1 proteins were interspersed to facilitate detection of the mobility shifts. The PGS not used

for glycosylation in AtSBT5.1 is indicated in red. Western blot analyses were performed using anti-HA

antibodies. Molecular mass markers in kiloDaltons are indicated on the right.

(B)

(A)

Figure 34. AtSBT5.2(a) self cleaves in planta. Protein extracts containing the indicated HA-tagged AtSBT5.2 proteins transiently expressed in N.

benthamiana and analysed by Western blot using anti-HA antibodies (a-HA). Ponceau S staining of

ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco) for confirmation of equal loading in each

lane is also shown (bottom). Molecular mass markers in kiloDaltons are indicated on the right.

. 130

. 95

Ponceau S

AtSBT5.2(a) unprocessed

AtSBT5.2(a) processed

AtSBT5.2(b)

a-HA

Rubisco

70

Taken together, affinity purification using concanavalin A, deglycosylation using PNGase F or

Endoglycosidase H, tunicamycin treatment of N. benthamiana leaves expressing the proteins

under study as well as systematic mutagenesis of PGS demonstrate that AtSBT5.2(a) and

AtSBT5.1 are glycosylated in planta.

Despite the fact that AtSBT5.2(b) contains the seven PGSs present in AtSBT5.2(a), (i) no

AtSBT5.2(b) binding to or elution from a concanavalin A resin was observed (Figure 32A) (ii)

AtSBT5.2(b) electrophoretic mobility was not modified after treatment with PNGase F or

EndoH (Figure 32B); and (iii) no effect of tunicamycin treatment on the AtSBT5.2(b)

electrophoresis profile was detected (Figure 32C). These results, which are consistent with

the absence of a SP in AtSBT5.2(b) and our previous observation that AtSBT5.2(b) is not

secreted, strongly suggest that AtSBT5.2(b) does not enter the secretory pathway and is

therefore not N-glycosylated.

1.4. AtSBT5.2(a) is an active serine protease.

As mentioned in section 1.3.4 of the introduction, subtilases, as other proteases, are typically

able to catalyze their self-processing to render a mature active polypeptide. In addition, the

presence of the three conserved amino acids in the catalytic triad of AtSBT5.2 proteins is

consistent with a possible protease activity of these proteins. When transiently expressing

AtSBT5.2(a) in leaf epidermal cells of N. benthamiana, two protein bands were detected that

may correspond to the processed and unprocessed forms of the protease (Figure 34). A

single band was detected for AtSBT5.2(b), suggesting that this protein is either not

processed or fully processed in planta (Figure 34). In order to learn more about proteolytic

cleavage of AtSBT5.2 proteins, AtSBT5.2(a) and AtSBT5.2(b) catalytic mutant versions were

generated, in which the conserved histidine residue in the catalytic triad of the proteins was

mutated to Alanine, (AtSBT5.2(a)-H210A and AtSBT5.2(b)-H171A). Following transient

expression in N. benthamiana leaf epidermal cells, mutation of the catalytic histidine residue

did not affect migration of AtSBT5.2(b) as compared to the wild-type protein, suggesting that,

consistent with the absence of a SP in AtSBT5.2(b), this protein is not processed in planta

Figure 35. AtSBT5.2(a) is an active serine hydrolase.

Protein extracts from N. benthamiana leaves transiently expressing the indicated HA-tagged AtSBT5.2

proteins were incubated with the ActivX Desthiobiotin-FP serine hydrolase probe. Protein extracts

from untreated N. benthamiana leaves and leaves transformed with an empty vector construct were

included as controls. Western blot analyses were performed using anti-HA or streptavidin antibodies.

Ponceau S staining confirms equal loading. Molecular mass markers in kiloDaltons are indicated on

the right. This experiment was performed five times with identical results.

Figure 36. Mutation of some glycosylated residues affects the catalytic activity of

AtSBT5.2(a). N. benthamiana leaf extracts expressing either HA-tagged AtSBT5.2(a) wild type, catalytic mutant or

PGS point mutant versions were incubated with ActivX Desthiobiotin-FP serine hydrolase probe.

Western blot analyses were performed using anti-HA or streptavidin antibodies. Ponceau staining

confirms equal loading. Molecular mass markers in kiloDaltons are indicated on the right. Similar

results were obtained in three independent experiments.

71

(Figure 34). In contrast, when expressing AtSBT5.2(a)-H210A, only the slow migrating band,

that very likely corresponds to the unprocessed form of the protein, was detected thus

suggesting that AtSBT5.2(a) is able to auto-process in planta and thus be active as a

protease (Figure 34).

Being a subtilase, we may hypothesize that AtSBT5.2(b) interaction with AtMYB30 may lead

to proteolytic degradation of the TF, perhaps providing a molecular mechanism of alleviation

of its activity. The catalytic activity of HA-tagged AtSBT5.2(a) and AtSBT5.2(b) was thus

investigated using an activity-based profiling (ABP) assay with a biotin-conjugated

fluorophosphonate (FP) probe able to bind specifically and irreversibly to the catalytic site of

active serine proteases. Biotin conjugation of the FP probe allows ready detection of active

serine proteases by Western blot using HRP-labelled streptavidin. AtSBT5.2(a), AtSBT5.2(b)

and their respective catalytic mutants were expressed in N. benthamiana leaves and the

corresponding protein extracts were incubated with the FP serine hydrolase probe. A specific

labelled band was detected for AtSBT5.2(a), whereas no signal was detected for its catalytic

mutant version AtSBT5.2(a)H210A, reinforcing the idea that the detected band reflects the

catalytic activity of AtSBT5.2(a) (Figure 35). Surprisingly, the detected band presents a

molecular size that is smaller than that detected for AtSBT5.2(a) by Western blot with anti-

HA antibodies (Figure 35). This result is consistent with an active AtSBT5.2(a) version that

would be both N- and C-terminally processed and thus undetectable by Western blot

analysis, as previously reported for some subtilase proteins (Yamagata et al., 1994, Von

Groll et al., 2002, Beilinson et al., 2002, Plattner et al., 2014). In contrast, no specific labelling

was detected for AtSBT5.2(b) or its catalytic mutant under the same conditions (Figure 35),

which does not allow to conclude on the catalytic activity of AtSBT5.2(b) .

In order to investigate the effect of N-glycosylation on the catalytic activity of AtSBT5.2(a),

AtSBT5.2(a) PGS removal mutants were tested for serine protease activity using the same

ABP assay. Protein extracts of N. benthamiana leaves expressing wild-type, catalytic mutant

or PGS mutant AtSBT5.2(a) versions were incubated with the FP probe as described above.

72

As shown in Figure 36, different effects were observed for different PGS mutant versions.

For example, AtSBT5.2(a)N678A appeared to display weaker protein accumulation and

stronger serine protease activity than the wild-type protein. In contrast, AtSBT5.2(a)N525A

appeared to display unchanged protein accumulation but weaker serine protease activity

than the wild-type protein suggesting that glycosylation of these residues may play a role in

modulating the catalytic activity of AtSBT5.2(a).

Figure 37. Neither AtSBT5.2(a) nor AtSBT5.2(b) affect AtMYB30 protein accumulation

in planta.

(A) TAP-Tagged AtMYB30 was transiently co-expressed with HA-tagged AtSBT5.2(a), or AtSBT5.2(b)

and their respective catalytic mutants side by side in the same N. benthamiana leaf.

(B) TAP-Tagged AtMYB30 was transiently co-expressed with the indicated HA-tagged AtSBT5.2

proteins in N. benthamiana leaves. Proteins were detected by Western blot analysis using anti-

HA or anti-TAP antibodies. Ponceau S staining confirms equal loading. Molecular mass markers in

kiloDaltons are indicated on the right. Identical results were obtained in five independent

experiments.

AtMYB30-TAP +

AtMYB30

Ponceau S

a-HA

. 130

. 95

. 72

a-TAP

(B)

(A)

AtMYB30-TAP +

AtSBT5.2(a)HA

AtMYB30-TAP +

AtSBT5.2(a)H210A-HA

AtMYB30-TAP +

AtSBT5.2(b)HA

AtMYB30-TAP +

AtSBT5.2(b)H171A-HA

73

2. Characterization of the interaction between AtMYB30 and AtSBT5.2

2.1. Neither AtSBT5.2(a) nor AtSBT5.2(b) affect AtMYB30 protein accumulation in planta.

Since the activity-based probe assay did not allow to conclude on the catalytic activity of

AtSBT5.2(b), the in planta accumulation of AtMYB30 when expressed with the different

AtSBT5.2 proteins was next analysed. To minimize differences in protein expression, which are

inherent to transient assays, AtMYB30 was co-expressed with AtSBT5.2(a) (or AtSBT5.2(b))

and their respective catalytic mutants, side by side in the same N. benthamiana leaf (Figure

37A). To avoid ex planta protein degradation, leaves were pre-treated with the protease inhibitor

phenylmethylsulfonyl fluoride (PMSF) 30 minutes before harvesting the tissue for protein

extraction. Western blot analysis consistently showed that AtMYB30 accumulation is not

significantly altered in the presence of AtSBT5.2(a) or AtSBT5.2(b) as compared to the

expression observed in the presence of the subtilase catalytic mutant versions (Figure 37B).

These results suggest that neither AtSBT5.2(a) nor AtSBT5.2(b) are able to proteolytically

cleave AtMYB30.

2.2. AtSBT5.2(b), but not AtSBT5.2(a), colocalizes with AtMYB30 in planta.

AtMYB30 was previously localized to the nucleus of Arabidopsis and N. benthamiana cells

(Froidure et al., 2010a). In order to investigate AtMYB30 potential colocalization with

AtSBT5.2(a) and/or AtSBT5.2(b), a Green Fluorescent Protein (GFP)-tagged AtMYB30 was

co-expressed with RFP-tagged AtSBT5.2(a) or AtSBT5.2(b) (Figure 38). Confocal

microscopy analysis of N. benthamiana leaves transiently co-expressing AtMYB30-GFP and

AtSBT5.2(a)-RFP showed that these proteins do not co-localize in the plant cell (Figure 38).

Indeed, AtSBT5.2(a)-RFP was located extracellularly whereas AtMYB30-GFP was located

inside the nucleus (Figure 38). Surprisingly, when co-expressed with AtSBT5.2(b)-RFP,

AtMYB30-GFP was excluded from the nucleus and localized to the same punctuated

Figure 38. AtSBT5.2(b), but not AtSBT5.2(a), colocalises with and retains AtMYB30

outside the nucleus.

Confocal images of epidermal cells of N. benthamiana leaves 36 hours after Agrobacterium-

mediated transient co-expression of GFP-tagged AtMYB30 with either AtSBT5.2(a) (top) or

AtSBT5.2(b) (down) RFP-tagged versions. Bars, 10µm. GFP, Green Fluorescent Protein; RFP, Red

Fluorescent Protein.

AtM

YB

30

-GF

P

+ A

tSB

T5

.2(a

)-R

FP

AtM

YB

30

-GF

P

+ A

tSB

T5

.2(b

)-R

FP

GFP Fluorescence RFP Fluorescence Merged

Table 3. FRET-FLIM analysis shows that AtSBT5.2(b) physically interacts with

AtMYB30 in N. benthamiana epidermal cells.

Donor Acceptor Lifetime(a) SEM(b) N(c) E(d) p-value (e)

AtMYB30-GFP - 2.669 0.009 82 - -

AtMYB30-GFP AtSBT5.2(b)-HA 2.552 0.013 57 - -

AtMYB30-GFP AtSBT5.2(b)-RFP 2.274 0.019 54 10.86 5.85 x 10-21

AtMYB30-GFP AtSBT5.2(b)362-730-RFP 2.271 0.015 51 14.91 3.79 x 10-49

AtMYB30-GFP AtSBT5.1405-780-RFP 2.592 0.017 44 2.86 4.23 x 10-5

AtMYB123-GFP

AtMYB123-GFP AtMYB123-GFP

-

AtSBT5.2(b)362-730-RFP AtSBT5.1405-780-RFP

2.512

2.366 2.386

0.012

0.018 0.015

45

39 45

-

5.84 5.05

-

2.00 x 10-9 8.92 x 10-9

aMean lifetime in nanoseconds. bStandard error of the mean. cTotal number of measures. dPercentage of FRET efficiency (E = 1 - !DA/!D) calculated by comparing the lifetime of the donor in the

presence of the acceptor (!DA) with its lifetime in the absence of the acceptor (!D). eP value of the difference between the donor lifetimes in the presence and in the absence of the

acceptor (t-test).

Figure 40. AtSBT5.2(b) does not affect AtMYB123 nuclear localization.

Confocal images of epidermal cells of N. benthamiana leaves 36 hours after Agrobacterium-

mediated transient expression of GFP-tagged AtMYB123 co-expressed with RFP-tagged

AtSBT5.2(b). White arrows indicate cell nuclei. Bars, 10 µm. GFP, Green Fluorescent Protein; RFP,

Red Fluorescent Protein.

GFP Fluorescence RFP Fluorescence Merged

AtM

YB

12

3 -

GF

P

+ A

tSB

T5

.2(b

)-R

FP

Figure 39. AtSBT5.2(b)-mediated retention of AtMYB30 outside the nucleus is

independent of C-terminal tagging of the subtilase.

Confocal images of epidermal cells of N. benthamiana leaves 36 hours after Agrobacterium-

mediated transient expression of GFP-tagged AtMYB30 co-expressed with HA-tagged (left) or

untagged (right) versions of AtSBT5.2(b). Bars, 10 µm. GFP, Green Fluorescent Protein.

AtM

YB

30

-GF

P

+ A

tSB

T5

.2(b

)-H

A

AtM

YB

30

-GF

P

+ A

tSB

T5

.2(b

)

GFP fluorescence GFP fluorescence

74

structures where AtSBT5.2(b)-RFP was localized, suggesting a possible in planta interaction

between the two proteins outside the nucleus (Figure 38).

2.3. AtSBT5.2(b), but not (a), interacts with AtMYB30 in planta.

Subcellular co-localisation is a first prerequisite for the study of protein-protein interactions.

Since no co-localization was detected between AtMYB30 (nuclear) and AtSBT5.2(a)

(secreted), and because AtSBT5.2(b) co-localized with AtMYB30 in the small vesicle-like

structures described in the previous section, we sought out to investigate the interaction

between AtSBT5.2(b) and AtMYB30 in living plant cells.

The physical interaction between AtSBT5.2(b) and AtMYB30 was tested using a quantitative

non-invasive FLIM approach to monitor the Förster Resonance Energy Transfer (FRET)

between the GFP (donor) and RFP (acceptor) molecules fused to AtMYB30 and

AtSBT5.2(b), respectively. If these two proteins interact, the transfer of energy from the

donor to the acceptor decreases the fluorescence lifetime (average time that a molecule

remains in its excited state prior to returning to its basal state) of the donor fluorophore. The

relative difference of lifetime is a measure of FRET efficiency (E) that is considered to be

significant when higher than 7%. The average GFP lifetime in nuclei expressing AtMYB30-

GFP was 2.669 ± 0.009 ns (mean ± sem), whereas the GFP lifetime of AtMYB30-GFP in the

vesicle-like structures when co-expressed with AtSBT5.2(b)-RFP was 2.274 ± 0.019 ns

(Table 3; Figure 38). Although this reduction in GFP lifetime may result from the physical

interaction between AtMYB30 and AtSBT5.2(b), it was important to rule out that this

difference did not reflect distinct molecular environments of different subcellular

compartments in which AtMYB30 is localized (nuclear when expressed alone or in

intracellular vesicles when co-expressed with AtSBT5.2(b)). The subcellular localization of of

GFP-tagged AtMYB30 when co-expressed with non-fluorescent HA-tagged or untagged

versions of AtSBT5.2(b) was therefore investigated. As shown in Figure 39, both

Figure 41. AtMYB30 and AtMYB123 colocalize in nuclei with both AtSBT5.2362-730 and

AtSBT5.1405-780. Confocal images of epidermal cells of N. benthamiana leaves 36 hours after Agrobacterium-

mediated transient expression of GFP-tagged AtMYB30 co-expressed with the indicated truncated

versions of AtSBT proteins. White arrows indicate cell nuclei. Bars, 10 µm. GFP, Green Fluorescent

Protein; RFP, Red Fluorescent Protein.

GFP Fluorescence RFP Fluorescence Merged

AtM

YB

30

-GF

P

+ A

tSB

T5

.2(b

) 36

2-7

51R

FP

AtM

YB

30

-GF

P

+ A

tSB

T5

.14

05

-78

0R

FP

AtM

YB

12

3-G

FP

+ A

tSB

T5

.2(b

) 36

2-7

30R

FP

AtM

YB

12

3-G

FP

+ A

tSB

T5

.14

05

-78

0R

FP

75

AtSBT5.2(b)-HA and untagged AtSBT5.2(b), which are not able to act as acceptors for the

GFP fluorescence, still lead to AtMYB30 localization to mobile intracellular vesicles. As

shown in Table 3, a slight decrease in GFP-lifetime was indeed detected for AtMYB30-GFP

when co-expressed with AtSBT5.2(b)-HA (in intracellular punctuated structures) as

compared to AtMYB30-GFP expressed alone (located inside nuclei), suggesting differences

in the molecular environment between the two compartments. Importantly, a significant

reduction of GFP lifetime was observed when AtMYB30-GFP was co-expressed with RFP-

tagged AtSBT5.2(b) (2.274 ± 0.019 ns) as compared to co-expression with AtSBT5.2(b)-HA

(2.552 ± 0.013 ns), confirming the physical interaction between the two proteins in

intracellular vesicles.

Together, our data suggests that AtSBT5.2(b) may high jack AtMYB30 in the vesicle-like

structures, thus impeding its nuclear entry.

2.4. The AtSBT5.2(b)-AtMYB30 interaction is specific and mediated through AtSBT5.2(b) C-

terminal domain

The large number of subtilase proteins and MYB TFs in Arabidopsis suggests functional

redundancy, although it may also be indicative of functional diversification. Therefore, we

next addressed the question of specificity of the interaction between AtSBT5.2(b) and

AtMYB30.

In order to test whether AtSBT5.2(b) was able to induce mislocalization of other TFs in the

vesicle-like structures, AtSBT5.2(b)-RFP was co-expressed with the unrelated nuclear MYB

TF AtMYB123. Despite the formation of the vesicle-like structures containing AtSBT5.2(b)-

RFP, GFP-tagged AtMYB123 was consistently detected inside the nucleus (Figure 40),

suggesting that AtSBT5.2(b)-mediated mislocalization of AtMYB30 and, thus the interaction

between the two proteins, is specific.

>AtSBT5.2(b)(362-730)

MVKGKIVLCENVGGSYYASSARDEVKSKGGTGCVFVDDRTRAVASAYGSFPTTVIDSKEAAEIFSYL

NSTKDPVATILPTATVEKFTPAPAVAYFSSRGPSSLTRSILKPDITAPGVSILAAWTGNDSSISLEGKPA

SQYNVISGTSMAAPHVSAVASLIKSQHPTWGPSAIRSAIMTTATQTNNDKGLITTETGATATPYDS

GAGELSSTASMQPGLVYETTETDYLNFLCYYGYNVTTIKAMSKAFPENFTCPADSNLDLISTINYPSI

GISGFKGNGSKTVTRTVTNVGEDGEAVYTVSVETPPGFNIQVTPEKLQFTKDGEKLTYQVIVSATAS

LKQDVFGALTWSNAKYKVRSPIVISSESSRTN

>AtSBT5.1(405-780)

MVKGKIVVCDSDLDNQVIQWKSDEVKRLGGIGMVLVDDESMDLSFIDPSFLVTIIKPEDGIQIMSYI

NSTREPIATIMPTRSRTGHMLAPSIPSFSSRGPYLLTRSILKPDIAAPGVNILASWLVGDRNAAPEGKP

PPLFNIESGTSMSCPHVSGIAARLKSRYPSWSPAAIRSAIMTTAVQMTNTGSHITTETGEKATPYDF

GAGQVTIFGPSSPGLIYETNHMDYLNFLGYYGFTSDQIKKISNRIPQGFACPEQSNRGDISNINYPSIS

ISNFNGKESRRVSRTVTNVASRLIGDEDTVYTVSIDAPEGLLVRVIPRRLHFRKIGDKLSYQVIFSSTTTI

LKDDAFGSITWSNGMYNVRSPFVVTSKDDNDSER

Figure 42. Sequence alignment of AtSBT5.2(b)362-730 and AtSBT5.1405-780 proteins. Identical amino acids are highlighted in blue. The 103 C-terminal amino acids encoded by the partial

AtSBT5.2 cDNA clone identified in the yeast two-hybrid screen as interacting with AtMYB30DAD are

underlined.

AtSBT5.2(b)362-730

AtSBT5.1405-780

AtSBT5.2(b)362-730

AtSBT5.1405-780

AtSBT5.2(b)362-730

AtSBT5.1405-780

AtSBT5.2(b)362-730

AtSBT5.1405-780

AtSBT5.2(b)362-730

AtSBT5.1405-780

76

The identification of a partial AtSBT5.2 cDNA clone (the last 103 amino acids of AtSBT5.2) in

the previously perfomed Y2H screen (Froidure et al., 2010a) suggests that the AtMYB30-

AtSBT5.2(b) interaction is mediated by the C-terminus of AtSBT5.2(b). In order to test this

idea, a truncated AtSBT5.2(b) version containing the last 368 amino acids of the protein

(AtSBT5.2(b)362-730) fused to the RFP was generated and transiently expressed in N.

benthamiana leaves. Confocal microscopy analysis showed that AtSBT5.2(b)362-730-RFP

presents a nucleocytoplasmic localization and colocalises with AtMYB30 in the plant cell

nucleus (Figure 41). A significant reduction of the average GFP lifetime was measured in

nuclei coexpressing AtMYB30-GFP and AtSBT5.2(b)362-730-RFP (2.271 ± 0.112 ns), as

compared to nuclei expressing AtMYB30-GFP alone (2.669 ± 0.081 ns) (Figure 41; Table 3),

showing that the C-terminus of AtSBT5.2(b) is sufficient for the interaction with AtMYB30 in

the nucleus. The specificity of this interaction was highlighted by the absence of interaction

between AtMYB30-GFP and the equivalent C-terminal domain of AtSBT5.1 (AtSBT5.1405-780-

RFP), (average GFP lifetime of 2.592 ± 0.119 ns in nuclei co-expressing both proteins)

(Figure 41; Table 3). Alignement of C-terminal domains of AtSBT5.2(b)362-730 and AtSBT5.

1405-780 is presented in Figure 42. Moreover, the unrelated nuclear TF AtMYB123, whose

nuclear localization was not affected in the presence of full length AtSBT5.2(b) (Figure 40),

was not able to interact with the C-terminus of AtSBT5.2(b) (AtSBT5.2(b)362-730-RFP). Indeed,

despite their nuclear co-localization, the FRET efficiency in nuclei co-expressing both

proteins was lower than 7% (Figure 41; Table 3), further confirming the specificity of the

interaction between AtMYB30 and AtSBT5.2(b).

2.5. AtSBT5.2(b) localization in intracellular vesicles is mediated through its N-terminal domain

The nucleocytoplasmic localization of AtSBT5.2(b)362-730 suggests that the N-terminal region

of AtSBT5.2(b) is required to mediate its localization in the intracellular vesicles described

above. In order to test this idea, a truncated AtSBT5.2(b) version deleted from its first 161

amino acids was generated (AtSBT5.2(b)162-730). Transiently expressed RFP-tagged

Figure 43. AtMYB30 localisation in intracellular vesicles is AtSBT5.2(b) N-terminal

domain dependant.

Confocal images of epidermal cells of N. benthamiana leaves 36 hours after Agrobacterium-

mediated transient expression of GFP-tagged AtMYB30 co-expressed with RFP-tagged

AtSBT5.2(b)162-730 or an N-terminally tagged RFP fusion of AtSBT5.2(b). White arrows indicate cell

nuclei. Bars, 10 µm. GFP, Green Fluorescent Protein; RFP, Red Fluorescent Protein.

AtM

YB

30

-GF

P

+ A

tSB

T5

.2(b

) 16

2-7

30-R

FP

RFP fluorescence GFP fluorescence Merged

AtM

YB

30

-GF

P

+ R

FP

-AtS

BT

5.2

(b)

77

AtSBT5.2(b)162-730 was indeed localized in the cytoplasm on N. bethamiana cells (Figure 43).

Furthermore, an N-terminally tagged RFP fusion of AtSBT5.2(b) also presented a

cytoplasmic localization (Figure 43), suggesting that a free AtSBT5.2(b) N-terminus is

necessary for targeting AtSBT5.2(b) to intracellular vesicles. In addition, both AtSBT5.2(b)162-

730 and N-terminally tagged RFP-AtSBT5.2(b) did not affect AtMYB30 nuclear localization

(Figure 43), indicating that a free N-terminal domain in AtSBT5.2(b) is necessary to target

AtMYB30 to vesicles. Close inspection of AtSBT5.2(b) N-terminal region uncovered the

presence of a stretch of amino acid residues containing a putative myristoylation site (Figure

27), thus warranting ongoing future experiments in our group to test the role of the Gly2

residue in AtSBT5.2(b) myristoylation and vesicle targeting.

0

0,2

0,4

0,6

0,8

1

1,2

1,4

1,6

AtSBT5.1 AtSBT5.2 AtSBT5.3 AtSBT5.6

Re

lati

ve

ge

ne

exp

resi

on

(a

.u.)

WT (Col-0)

atsbt5.2-1

atsbt5.2-2

Figure 44. Genetic analysis of AtSBT5.2 and AtSBT5.1 Arabidopsis mutant lines.

(A) Representation of AtSBT5.2 genomic organization. Exons (E1-E9) are represented with open

arrows and introns (I1-I8) are shown as black thick lines. The insertion sites and the SALK

numbers of the T-DNA in the atsbt5.2 and atsbt5.1 mutants are indicated. Positions of the

primers used for qRT-PCR analysis are indicated by black arrows.

(B) qRT-PCR analysis of the expression of AtSBT genes in subgroup V (Rautengarten et al., 2005) in

leaves of wild-type Col-0 (WT) and atsbt5.2 mutant lines (atsbt5.2-1 and atsbt5.2-2) leaves.

Statistical significance according to a Student’s t test P value <10−7 is indicated by asterisks.

(C) qRT-PCR analysis of the expression of AtSBT5.1 in leaves of wild-type Col-0 (WT) and atsbt5.1

mutant lines (atsbt5.1-1 and atsbt5.1-2).

(A)

500bp

atsbt5.2-1

atsbt5.2-2

atsbt5.1-2

atsbt5.1-1

E1 E2 E3 E4 E5 E6 E7 E8 E9 I1 I2 I3 I5 I4 I6 I7 I8

LB T-DNA LB T-DNA

T-DNA LB T-DNA LB

SALK_121716

insertion at 52bp

SALK_017993

insertion at 447bp

SALK_132812C

insertion at 2274bp

SALK_132812C

insertion at 2247bp

AtSBT5.2

E1 E2 E3 E4 E5 E6 E7 E8 E9 I1 I2 I3 I5 I4 I6 I7 I8

AtSBT5.1 SALK_1

E6 E7I6 I7

(C) (B)

* *

0

0,01

0,02

0,03

0,04

0,05

0,06

0,07

0,08

0,09

AtSBT5.1

Re

lati

ve

ge

ne

exp

ress

ion

(a

.u.)

WT (Col-0)

atsbt5.1-1

atsbt5.1-2

78

3. Functional analysis of AtSBT5.2 in plant defence

Having validated the in planta interaction between AtMYB30 and AtSBT5.2(b), we sought out

to investigate the outcome of this protein association, and of AtSBT5.2(b)-mediated nuclear

exclusion of AtMYB30, in AtMYB30-mediated defence responses in Arabidopsis. Since

AtSBT5.2(b) induced retention of AtMYB30 outside the nucleus, it was tempting to

hypothesize that this subitilase may act as a negative regulator of AtMYB30-mediated

defence response.

3.1. AtSBT5.2 negatively regulates plant defence and HR.

To study the role of the detected interaction between AtSBT5.2 and AtMYB30 in the control

of the plant defence response, the phenotypic analysis (HR cell death and bacterial growth

rates) of plants showing deregulated AtSBT5.2 expression (knockout and overexpressing

plants) was carried out.

To investigate the function of AtSBT5.2 in the establishment of the plant response to

bacterial pathogens, we first searched for Arabidopsis atsbt5.2 null mutants in the SALK

collection (http://signal.salk.edu). Genetic characterization of four candidate lines allowed us

to identify two homozygous atsbt5.2 mutant lines (SALK_012113 and SALK_132812C), both

containing a T-DNA insertion in the last exon (Exon 9, Figure 44A). These lines were

renamed atsbt5.2-1 and atsbt5.2-2, respectively. PCR fragments obtained using a primer in

the left border of the T-DNA and an internal primer in the AtSBT5.2 gene were sequenced,

showing that atsbt5.2-1 and atsbt5.2-2 carried a T-DNA insertion at positions 2247 and 2274

bp of the predicted open reading frame (ORF), respectively (Figure 44A).

The effect of the T-DNA insertions on AtSBT5.2 transcript levels was next analysed by qRT-

PCR using AtSBT5.2 gene-specific primers and revealed high reduction (more than 97%) of

AtSBT5.2 expression in homozygous atsbt5.2 mutant plants (Figure 44B). Expression of the

Log

10C

FU

/cm

2

(B)

(A)

Figure 45. AtSBT5.2 negatively regulates HR development and plant resistance to

bacterial inoculation. (A) Quantification of cell death by measuring electrolyte leakage before (white bars) and 24 hours

after inoculation (gray bars) of the indicated lines with Pst AvrRpm1 (5 x 106 cfu/ml). Cell death

values are related to the value displayed by wild-type Col-0 plants, which is set at 100%.

Statistical differences using Multiple Factor ANOVA (P value < 10-4) are indicated by asterisks.

(B) Growth of Pst DC3000 AvrRpm1 in the indicated Arabidopsis lines 3 days after inoculation with a

bacterial suspension of 5 × 105 cfu/ml. Mean bacterial densities were calculated from 6

independent experiments with 6 individual plants (4 leaves/plant). Statistical differences

according to a Multiple Factor ANOVA test (P value < 0.01) are indicated by asterisks.

79

other members of the AtSBT5 subtilase subfamily was also analysed in atsbt5.2-1 and

atsbt5.2-2 lines. No significant effect on the expression of AtSBT5.1, 3 or 6 could be detected

in the atsbt5.2 mutant lines. It is worth noting that, despite the close proximity of the T-DNA

insertions to the AtSBT5.1 promoter region, expression of AtSBT5.1 was not affected in any

of the two atsbt5.2 mutant lines (Figure 44B) Finally, no amplification was obtained for

AtSBT5.4 and 5, indicating that consistent with a previous report (Rautengarten et al., 2005)

the expression of these two genes is extremely weak in Arabidopsis leaves (Figure 44B),

Despite the severe reduction of AtSBT5.2 expression in the mutant lines, no obvious

macroscopic phenotypes were observed in plants grown under our conditions. The

phenotype of these lines in response to bacterial inoculation was next analysed. Similar to a

previously characterized miel1 mutant line (Marino et al., 2013), which was included as a

control for enhanced cell death and resistance in these assays, atsbt5.2 mutant plants

showed clear HR cell death symptoms after inoculation with Pst DC3000 AvRpm1 as

compared to Col-0 wild-type plants. This phenotype was quantified by ion leakage

measurements in leaf disks assays. Conductivity values measured in atsbt5.2 and miel1

plants significantly higher than those displayed by Col-0 wild type plants after bacterial

inoculation (Figure 45A). In addition, this enhanced cell death phenotype correlated with

reduced in planta bacterial growth indicating that, in agreement with faster HR development,

atsbt5.2 plants showed increased resistance in response to inoculation with Pst DC3000

AvrRpm1 as compared to wild-type plants (Figure 45B). Together, these results suggest a

role for AtSBT5.2 as a negative regulator of disease resistance to avirulent bacteria.

In order to investigate whether the enhanced HR and defence responses displayed by

atsbt5.2 plants are specific, we next tested whether the closest homolog AtSBT5.2

Arabidopsis homolog (AtSBT5.1) may also play a role in the response to bacterial infection.

Two independent T-DNA lines for AtSBT5.1, SALK_017993 and SALK_121716 (renamed

atsbt5.1-1 and atsbt5.1-2, respectively) were kindly provided by Dr. Andreas Schaller (Figure

44A). Characterization of these lines showed that atsbt5.1-1 and atsbt5.1-2 contain a T-DNA

Figure 46. AtSBT5.2 is a negative regulator of AtMYB30-mediated HR cell death.

Quantification of cell death by measuring electrolyte leakage before (white bars) and 24 hours

after inoculation (gray bars) of the indicated lines with Pst AvrRpm1 (5 x 106 cfu/ml). Cell death

values are related to the value displayed by wild-type Col-0 plants, which is set at 100%. Statistical

differences using Multiple Factor ANOVA (P value < 10-4) are indicated by asterisks.

Ion Leakage myb30ko x sbt5.2ko

* *

0

20

40

60

80

100

120

140

160

WT MYB30OE sbt5.2 myb30ko x

sbt5.2ko - 1

myb30ko x

sbt5.2ko - 2

Co

nd

uct

ivit

y (

a.u

.)

0 hpi 24 hpi

80

insertion in exon 3 at position 447 and in exon 1 at position 52 of the predicted ORF,

respectively. Although qRT-PCR analysis, using AtSBT5.1 gene-specific primers, showed no

significant reduction of AtSBT5.1 expression in homozygous atsbt5.1 mutant plants (Figure

44C), these two lines were considered as functional knocked out mutants since the T-DNA

insertion leads to production an aberrant protein in atsbt5.1-1 and to the introduction of an

early STOP codon in atsbt5.1-2. Following inoculation with avirulent Pst DC3000 AvrRpm1,

atsbt5.1-1 and atsbt5.1-2 mutant plants showed conductivity and bacterial growth rates

indistinguishable from those displayed by Col-0 wild-type plants (Figure 45A and 45B),

indicating that AtSBT5.1 is not involved in immune responses to bacterial infection and thus

underlining the specific effect of AtSBT5.2 on the control of plant defence responses.

Taken together, these results indicate that AtSBT5.2 acts as a negative regulator of the HR

and defence in Arabidopsis.

3.2. AtSBT5.2 controls the HR via AtMYB30.

In order to determine whether the enhanced resistance phenotype displayed by atsbt5.2

plants is dependent on AtMYB30, we generated two independent atsbt5.2/atmyb30 double

mutant lines and tested their response to bacterial infection. As observed before, and similar

to AtMYB30OE plants, atsbt5.2 displayed an enhanced HR as compared to Col-0 wild type

plants (Figure 46). Importantly, the increased HR displayed by atsbt5.2 mutant plants was

suppressed in the atmyb30 mutant background (atsbt5.2/atmyb30; Figure 46), suggesting

that negative regulation of defence-related cell death responses by AtSBT5.2 is mediated via

its effect on AtMYB30.

Figure 48. AtSBT5.2(b), but not AtSBT5.2(a), negatively regulates defence-related

HR cell death.

Quantification of cell death by measuring electrolyte leakage before (white bars) and 24 hours

after inoculation (gray bars) of the indicated lines with Pst AvrRpm1 (5 x 106 cfu/ml). Cell death

values are related to the value displayed by wild-type Col-0 plants, which is set at 100%. Statistical

differences using Multiple Factor ANOVA (P value < 10-4) are indicated by asterisks.

Figure 47. Characterization of AtSBT5.2(a) and AtSBT5.2(b) overexpressing

Arapidopsis lines.

(A) Relative expression of AtSBT5.2 in Arabidopsis leaves in wild-type Col-0, 35S:AtSBT5.2(a)-HA

(AtSBT5.2(a)OE) and 35S:AtSBT5.2(b)-HA (AtSBT5.2(b)OE) lines was determined by qRT-PCR.

The expression values were normalized by using SAND family gene as internal standards,

and related to the value of Col-0, which is set at 1. Mean and SEM values were calculated

from four individual plants per line.

(B) Western blot analysis showing protein accumulation of AtSBT5.2(a)-HA or AtSBT5.2(b) in

Arabidopsis transgenic and Col-0 wild-type plants. Ponceau S staining confirms equal

loading. Molecular mass markers in kiloDaltons are indicated on the right.

(A) (B)

*

*

* *

81

3.3. AtSBT5.2(b), but not AtSBT5.2(a), negatively regulates defence-associated cell death

responses.

In atsbt5.2 mutant plants, expression of both AtSBT5.2(a) and AtSBT5.2(b) is affected. In

order to obtain additional confirmation of the negative role specifically played by AtSBT5.2(b)

on AtMYB30-mediated defence, atsbt5.2 mutant plants were independently transformed with

HA-tagged versions of either AtSBT5.2(a) or AtSBT5.2(b) expressed under the control of the

35S promoter. Expression of AtSBT5.2 gene and AtSBT5.2 protein accumulation were

monitored by qRT-PCR and Western Blot analysis, respectively, in two T4 independent

homozygous lines for each construct and overexpression of the desired transcript and

protein were confirmed in the different lines (Figure 47A and 47B). Importantly, the increased

HR phenotype displayed by atsbt5.2 was specifically suppressed in atsbt5.2 plants

overexpressing AtSBT5.2(b), but not AtSBT5.2(a) (Figure 48). These results demonstrate

that, in agreement with the specific AtSBT5.2(b)-mediated retention of and interaction with

AtMYB30 in perinuclear structures, negative regulation of AtMYB30-mediated defence-

related cell death is specifically controlled by AtSBT5.2(b). Overall, this work identifies

AtSBT5.2(b) as a new negative regulator or AtMYB30-mediated defence responses.

3.4. Analysis of AtSBT5.2 expression after bacterial inoculation.

To gain further knowledge on the mode of action of AtSBT5.2 during the plant interaction

with HR-inducing bacteria, we analysed the expression profile of AtSBT5.2(a) and

AtSBT5.2(b) at different times after inoculation of wild-type Col-0 Arabidopsis plants with

avirulent Pst DC3000 AvrRpm1. As previously described, AtMYB30 expression was clearly

induced 4 hpi (Figure 49A) (Marino et al., 2013). Surprisingly, expression of AtSBT5.2(a)

appeared to be rapidly downregulated after inoculation (Figure 49B). This result suggests a

potential role for AtSBt5.2(a) during the plant response to bacteria, although independently of

AtMYB30. In contrast, AtSBT5.2(b) expression appears to follow the expression profile of

AtMYB30 after bacterial inoculation (Figure 49C). Indeed, although AtSBT5.2(b) expression

0

0,05

0,1

0,15

0,2

0,25

0,3

0 1 2 3 4 Re

lati

ve

ge

ne

exp

ress

ion

(a

.u.)

hpi

AtSBT5.2(b)

AtSBT5.1

0

0,5

1

1,5

2

2,5

3

0 1 2 3 4 Re

lati

ve

ge

ne

exp

ress

ion

(a

.u.)

hpi

AtMYB30

0

1

2

3

4

5

6

7

8

9

0 1 2 3 4 Re

lati

ve

ge

ne

exp

ress

ion

(a

.u.)

hpi

AtSBT5.2(b)

AtMYB30

0

0,1

0,2

0,3

0,4

0,5

0,6

0,7

0,8

0 1 2 3 4 Re

lati

ve

ge

ne

exp

ress

ion

(a

.u.)

hpi

AtSBT5.2(a)

Figure 49. AtSBT5.2(a), AtSBT5.2(b), AtSBT5.1 and AtMYB30 expression profile

during avirulent HR-inducing bacteria infection.

Expression analysis of AtMYB30 (A), AtSBT5.2(a) (B), AtSBT5.2(b) and AtSBT5.1 (C) and AtMYB30

and AtSBT5.2(b) (D) in wild-type Col-0 Arabidopsis lines after inoculation with Pst AvrRpm1 (5x107

cfu/ml). In (D), expression values were related to the value of each gene at time 0, which is set at

1. In all cases, expression values of the individual genes were normalized using SAND family as

internal standard. Mean and SEM values were calculated from 4 experiments with 4 replicates. 0,

1, 2 and 4 indicate hours post-inoculation.

(A) (B)

(C) (D)

82

appears to be weaker than that of AtMYB30, when gene expression values were related to

the expression value of each gene at time 0, both AtSBT5.2(b) and AtMYB30 display

overlapping expression profiles and identical induction rates after bacterial inoculation

(Figure 49C). Finally, in agreement with the fact that the closest AtSBT5.2 homolog,

AtSBT5.1, does not seem to regulate HR and bacterial growth in Arabidopsis (Figure 45), no

change in AtSBT5.1 expression was detected in these assays (Figure 49C), underlining the

specificity of AtSBT5.2(b)-mediated regulation of the HR.

Our qRT-PCR analysis indicated that expression of both AtSBT5.2(a) and AtSBT5.2(b) is

very weak in Arabidopsis, in particular in the case of AtSBT5.2(b). In order to investigate the

relative expression levels of both isoforms and the potential changes of these relative levels

after bacterial inoculation, we looked at available RNA sequencing data recently obtained in

our group using Arabidopsis plants inoculated with avirulent Pst DC3000 AvrRpm1. To be

able to distinguish between the two isoforms, we searched for differences in the number of

reads located at the 5’UTR of each isoform before and at different times after inoculation.

Unfortunately, the extremely low number of reads obtained at the 5’UTR (that is, the only

region that allows differentiating both isoforms) confirmed the very low expression of

AtSBT5.2 transcripts but unabled us to conclude on this matter.

83

Discussion

__________________________________________________________________________

84

My PhD work has encovered a new regulator of AtMYB30 activity that contributes to the

attenuation of MYB30-mediated defence-associated hypersensitive cell death. This new

regulator is AtSBT5.2, a protease of the subtilase family. During my PhD, I have shown that

the gene encoding AtSBT5.2 is alternatively spliced, giving rise to two distinct isofoms: one

encoding a canonical subtilase protein localized in the apoplastic space and a second

transcript encoding a protein located in intracellular mobile vesicles. Importantly, regulation of

AtMYB30 is specifically mediated by the interaction of AtMYB30 with this latter isofom,

whereas the secreted AtSBT5.2 version does not seem to play a role in regulating AtMYB30

activity.

1. AS, an emerging regulatory mechanism of plant defence

Removal of introns through splicing of pre-mRNAs is a key step in eukaryotic gene

expression (Han et al., 2011). Alternative splicing (AS) describes the processing of a single

pre-mRNA to produce multiple transcript isoforms (Nilsen and Graveley, 2010). AS has

important consequences for the cell, both at the RNA and protein levels. First, AS can

regulate transcript levels by the introduction of premature termination codons, which commit

the transcript isoform to degradation by the nonsense-mediated decay (NMD) pathway

(McGlincy and Smith, 2008, Nicholson and Mühlemann, 2010). In Arabidopsis, at least 13%

of genes undergo AS-NMD (Kalyna et al., 2012). The second main consequence of AS is the

production of transcript isoforms giving rise to proteins that differ in their sequence and

domain arrangement and thus may widely differ in subcellular localization, stability, or

function (Syed et al., 2012). Proteins or polypeptides that are truncated as a consequence of

AS can act as dominant-negative inhibitors of the authentic proteins (e.g., through

unproductive interaction with dimerization partners or nucleic acids) and have been

designated micropeptides or small interfering peptides (Seo et al., 2011a). Moreover,

variation affecting splicing/AS outcomes can provide flexibility in the transcriptome and

proteome to contribute to the ability of plants to adapt to their environment (Kazan, 2003).

85

AS is involved in most plant processes and is particularly prevalent in plants exposed to

environmental stress. A growing body of evidence suggests that AS in involved in the

regulation of cell fate, circadian clock, plant defence, and tolerance/sensitivity to abiotic

stress, thus uncovering a fundamental role of AS in plant growth, development, and

responses to external cues (Yang et al., 2014, Staiger and Brown, 2013). As mentioned

above, AS has been connected to NMD, which is also involved in plant disease resistance

(Rayson et al., 2012, Gloggnitzer et al., 2014). In addition, a genome-wide transcriptome

analysis (RNA-Seq) of Arabidopsis plants infected with Pseudomonas syringae indicated a

surprisingly large number of AS events. Indeed, more than 44% of multi-exon genes showed

evidence for novel AS, demonstrating that the Arabidopsis transcriptome annotation is still

highly incomplete, and that AS events are more abundant than expected (Howard et al.,

2013).

Although we are still far from understanding the functional implications of this transcriptome

complexity, a growing number of reports indicate at least some functions for AS during the

interaction of plants with pathogens. Several plant disease resistance (R) genes undergo AS,

and several R genes require alternatively spliced transcripts to produce R proteins that can

specifically recognize pathogen invasion (Gassmann, 2008). How AS of R genes functions in

disease resistance remains mostly unknown, but it has been suggested that the truncated

proteins may promote R gene function by alleviating self-inhibition of the intact protein

(Zhang and Gassmann, 2003). Alternatively, the truncated proteins may interfere with

downstream signalling. The N gene, encoding the TIR-NB-LRR R protein N that confers

resistance to TMV in tobacco, is alternatively spliced (Erickson et al., 1999). After TMV

infection, AS induces the production of a shorter N variant that lacks 13 out of 14 of the LRR

repeat domains found in the longer transcript and that is required for resistance to TMV

(Dinesh-Kumar and Baker, 2000). Similarly, alternative splicing of the Arabidopsis R genes

RPS4 (Zhang and Gassmann, 2007), RPS6 (Marquez et al., 2012) or SNC1 (Xu et al., 2012)

is critically important for defence against P. syringae. In a recent study, the splicing factors

86

SUppressor of ABI3-5 (SUA) and Required for SNC4-1D 2 (RSN2) were identified as

regulators of AS events in two Arabidopsis RLKs: Suppressor of NPR1-1 Constitutive4

(SNC4) and Chitin Elicitor Receptor Kinase1 (CERK1). In sua and rsn4 Arabidopsis mutants,

SNC4 splicing is altered and the amount of SNC4 transcripts is reduced (Zhang et al., 2014).

SUA and RSN2 are also required for the proper splicing of CERK1, which encodes a

receptor for chitin (Figure 5). In sua and rsn4 mutants, chitin-mediated ROS production is

reduced and correlates with enhanced growth of P. syringae, suggesting that AS plays

important roles during plant immunity (Zhang et al., 2014).

Our discovery that AS of AtSBT5.2 results in the production of the AtSBT5.2(b), which

specifically regulates AtMYB30-mediated hypersensitive cell death, contributes to further our

understanding of the various roles of AS during the regulation of plant immunity.

87

2. Regulation of AtSBT5.2 function through AS

2.1. AS affects the subcellular localization of resulting AtSBT5.2 protein variants

As mentioned earlier, alternative splicing may result in different subcellular localization of the

proteins encoded by distinct splice variants. For example, tissue-specific AS of the auxin

biosynthesis gene YUCCA4 generates two distinct YUCCA4 splice variants. One isoform is

restricted to flowers and contains a predicted C-terminal hydrophobic transmembrane

domain (TMD) that determines anchoring of the resulting protein to the cytosolic face of the

ER membrane, whereas the other isoform is present in all tissues and distributed throughout

the cytosol (Kriechbaumer et al., 2012). In addition, the subcellular localisation of the tomato

protein phosphatase 5 (LePP5) isoforms is determined by AS. AS of PP5 results in a longer

transcript containing an additional exon encoding two putative TMDs. Subcellular localization

studies indicated that the short isoform is localized in both the nucleus and the cytoplasm,

whereas the longer isoform is targeted to the ER and the nuclear envelope (de la Fuente van

Bentem et al., 2003). In Arabidopsis, Inositol-Requiring Enzyme1a (IRE1A) and IRE1B

catalyse the unconventional splicing of the TF bZIP60 in response to stress, resulting in a

frameshift that replaces the C-terminal region of bZIP60, including the TMD, by a shorter

region without TMD. Since the C-terminal TMD in bZIP60 anchors the protein to the ER,

removal of this region allows the functional form of bZIP60 to enter the nucleus, where it

activates transcription of ER stress-inducible genes (Nagashima et al., 2011, Deng et al.,

2011).

Similar to the previously described examples, we have shown that AS of AtSBT5.2 results in

two isoforms with distinct subcellular localizations. To our knowledge, our work represents

the first described example of AS affectting the localization of a protein of the subilase family,

whose members are typically secreted to the apoplastic space. According to the TAIR

database, out of the Arabidopsis 56 subtilase-encoding genes, only five (namely, AtSBT2.2,

AtSBT3.6, AtSBT4.11, AtSBT4.12 and AtSBT5.2) are predicted to be alternatively spliced.

AtSBT5.2 is the only spliced subtilase of the six members of clade V (Figure 15). The

AtSBT2.2(a)

AtSBT2.2(b)

AtSBT2.2(a)

AtSBT2.2(b)

AtSBT2.2(a)

AtSBT2.2(b)

AtSBT2.2(a)

AtSBT2.2(b)

AtSBT2.2(a)

AtSBT2.2(b)

AtSBT2.2(a)

AtSBT2.2(b)

AtSBT2.2(a)

AtSBT2.2(b)

AtSBT2.2(a)

AtSBT2.2(b)

AtSBT2.2(a)

AtSBT2.2(b)

AtSBT3.6(a)

AtSBT3.6(b)

AtSBT3.6(c)

AtSBT3.6(a)

AtSBT3.6(b)

AtSBT3.6(c)

AtSBT3.6(a)

AtSBT3.6(b)

AtSBT3.6(c)

AtSBT3.6(a)

AtSBT3.6(b)

AtSBT3.6(c)

AtSBT3.6(a)

AtSBT3.6(b)

AtSBT3.6(c)

AtSBT3.6(a)

AtSBT3.6(b)

AtSBT3.6(c)

AtSBT3.6(a)

AtSBT3.6(b)

AtSBT3.6(c)

AtSBT3.6(a)

AtSBT3.6(b)

AtSBT3.6(c)

(B)

(A) Signal peptide Prodomain

AtSBT4.11(a)

AtSBT4.11(b)

AtSBT4.11(a)

AtSBT4.11(b)

AtSBT4.11(a)

AtSBT4.11(b)

AtSBT4.11(a)

AtSBT4.11(b)

AtSBT4.11(a)

AtSBT4.11(b)

AtSBT4.11(a)

AtSBT4.11(b)

AtSBT4.11(a)

AtSBT4.11(b)

AtSBT4.11(a)

AtSBT4.11(b)

AtSBT4.12(a)

AtSBT4.12(b)

AtSBT4.12(c)

AtSBT4.12(a)

AtSBT4.12(b)

AtSBT4.12(c)

AtSBT4.12(a)

AtSBT4.12(b)

AtSBT4.12(c)

AtSBT4.12(a)

AtSBT4.12(b)

AtSBT4.12(c)

AtSBT4.12(a)

AtSBT4.12(b)

AtSBT4.12(c)

AtSBT4.12(a)

AtSBT4.12(b)

AtSBT4.12(c)

AtSBT4.12(a)

AtSBT4.12(b)

AtSBT4.12(c)

AtSBT4.12(a)

AtSBT4.12(b)

AtSBT4.12(c)

(D)

(C)

Figure 50. Predicted effects of AS on the proteins encoded by AtSBT2.2, AtSBT3.6,

AtSBT4.11 and AtSBT4.12 splice variants. Sequence alignment of predicted proteins encoded by AtSBT2.2 (A), AtSBT3.6 (B), AtSBT4.11 (C) and

AtSBT4.12 (D) splice variants. Identical amino acids are highlighted in blue. The signal peptide and

prodomain are boxed in red and orange, respectively. Catalytical conserved residues are indicated by

red dots. Putative N-glycosylation sites are indicated by blue dots. The amino acids of the predicted

transmembrane domain (TMD) in AtSBT3.6(a) are underlined.

88

predicted effects of AS on the proteins encoded by AtSBT2.2, AtSBT3.6, AtSBT4.11,

AtSBT4.12 splice variants are shown in Figure 50. AS of AtSBT2.2 leads to removal of amino

acids 452 to 475 in isoform (b), although the functional implications of this deletion are

difficult to predict at this stage. In the case of AtSBT3.6, the (a) isoform presents a predicted

SP and a TMD, and in silico analysis suggests that the protein would be located in the ER.

This ER localization could be altered in the two other AtSBT3.6 splice variants, which present

neither the SP nor the TMD, are rather predicted to be cytoplasmic and are differentiated by

an N-terminal region of 72 amino acids that is absent in the (b) isoform. As for AtSBT4.11

and AtSBT4.12, all isoforms are expected to be secreted although deletions of small regions

in the proteins are predicted. Intriguingly, AS of both AtSBT4.11 and AtSBT4.12 is expected

to lead to a deletion at the same position in the prodomain of both proteins although whether

and how these deletions may affect processing of the prodomain is just a speculation.

Indeed, since the function of these subtiliases has not been yet investigated, it is not possible

to speculate about the functional outcome of these predicted changes at present.

2.1.1. AtSBT5.2(b) localizes to endosomes

Defence-related proteases are found at different subcellular locations, as presented in the

Introduction. For example, VPEs have been shown to be located in the vacuole (Hatsugai et

al., 2004, Kuroyanagi et al., 2005), while subtilases, including P69B and AtSBT3.3, are

secreted to the apoplast (Tian et al., 2004, Tian et al., 2005, Tornero et al., 1997, Ramírez et

al., 2013). Similarly, in healthy tobacco tissues, mature phytaspase was located in the

apoplast thanks to a SP that directs the protein to the secretory pathway (Chichkova et al.,

2010). Unexpectedly, triggering of PCD by biotic or abiotic stresses resulted in relocalization

of phytaspase from the apoplast to the cytoplasm (Chichkova et al., 2010). Since inhibition of

protein synthesis did not affect this process, phytaspase relocalization likely reflects its

physical redistribution from the apoplast to the cytosol, rather than the arrest of phytaspase

secretion. Other proteases have been found to be located in the endomembrane system

during infection. For example, before inoculation, a functional GFP fusion of AtCEP1, a

89

papain-like cysteine protease involved in restriction of powdery mildew was undetectable. In

contrast, after treatment with Erysiphe cruciferarum, AtCEP1 localized to the ER of cells that

were successfully penetrated by the fungus and accumulated especially around established

haustoria (Höwing et al., 2014). An additional papain-like cystein protease from Arabidopsis,

Responsive to Desiccation21 (RD21), is characterized by the presence of a C-terminal

granulin domain and accumulates in vesicles that originate from the ER (ER bodies)

(Koizumi et al., 1993, Yamada et al., 2001, Gu et al., 2012). Importantly, RD21 is a central

component of the plant immune response against fungal and oomycete pathogens (Bozkurt

et al., 2011, Shindo et al., 2012).

Subtilases are expected to be secreted proteins. Consistent with harboring all canonical

features of a subtilase protein, including the N-terminal SP and prodomain, AtSBT5.2(a) was

previously shown to accumulate extracellularly (Engineer et al., 2014). In this work, through

isolation of intercellular (apoplastic) fluids and subcellular localization studies, we confirmed

that AtSBT5.2(a) enters the secretory pathway and is secreted to the extracellular space. In

contrast, AtSBT5.2(b) is an intracellular protein located in small vesicle-like structures. To

determine the nature of these intracellular structures, we conducted a series of colocalization

experiments with diverse subcellular markers. AtSBT5.2(b)-containing vesicles did not co-

localise with the Golgi cisternae marker GmMan49 (Nelson et al., 2007) or the ER marker

HDEL (Gomord et al., 1997) (data not shown). Given the mobile character and varied sizes

of the vesicle-like structures, additional experiments have recenltly been performed in our

group to investigate components of the endocytic cycle. Using protoplast from stable

transgenic Arabidopsis lines expressing a fluorescently tagged version of the endosomal

marker Ara7, a clear co-localization of Ara7 and AtSBT5.2(b) was shown. As a member of

the RAB GTPase family, Ara7 is a key regulator of endosomal trafficking, endocytosis, and

vacuolar transport (Kotzer et al., 2004, Richter et al., 2009) and has been found at both early

endosome (EE) and late endosome (LE) compartments, thus representing a suitable marker

for both endosomal populations (Ueda et al., 2004). Co-localization with Ara7 thus suggests

90

that AtSBT5.2(b) localizes to endosomes. In order to further validate this finding, co-

localization experiments with markers of additional endosomal compartments [including

the vacuolar-type H+-ATPase subunit VHA-a1 (Dettmer et al., 2006) and the Soluble N-

ethylmaleimide sensitive factor Attachment protein REceptor SYntaxin of Plant 61 (SNARE

SYP61) (Robert et al., 2008) for TGN/EE, as well as the RAB GTPase Ara6 (Ueda et al.,

2004) and the SNARE SYP21 (Sanderfoot et al., 2001) for LEs] are in progress. In these

experiments, we will thus test whether AtSBT5.2(b) is specifically localized to EEs, LEs or

both. In addition, endosomal compartments will be labelled with the lipophilic endocytic

tracer FM4-64 and co-localization of the signal with AtSBT5.2(b) will be investigated. Finally,

plant tissue expressing AtSBT5.2(b) will be treated with different inhibitors of vesicular

trafficking, such as wortmannin [(Wm), a phosphatidylphosphate-3-kinase inhibitor that

interferes with vesicle formation from the plasma membrane and the maturation of LEs and

multivesicular bodies (MVBs), resulting in their enlargement (Tse et al., 2004, Wang et al.,

2009)], brefeldin A [(BFA) an inhibitor of endosomal recycling to the plasma membrane,

acting by targeting the ADP ribosylation factor-GTP exchange factor GNOM and leading to

the accumulation of so-called BFA bodies (Geldner et al., 2003, Robinson et al., 2008)] or

concanamycin A [(Conc A), another well established membrane trafficking inhibitor that is

known to block vacuolar transport (Dettmer et al., 2006) by targeting vacuolar ATPase

activity at the TGN/EE, which is required for MVB/LE biogenesis (Scheuring et al., 2011)].

Together, these ongoing experiments will allow confirming and further characterizing

AtSBT5.2(b) localization to endosomes.

Subcellular localisation studies of N-terminally tagged or truncated AtSBT5.2(b) versions

showed that AtSBT5.2(b) localization in intracellular vesicles is mediated through its N-

terminal domain and, most likely, through a putative myristoylation site that is found

specifically in AtSBT5.2(b) after AS. N-myristoylation is the irreversible attachment of a

saturated 14-carbon myristate moiety to a protein, specifically on the α -amine group of an N-

terminal glycine (Boyle and Martin, 2015), which can reversibly direct both protein–

91

membrane and protein-protein interactions (Maurer-Stroh and Eisenhaber, 2004). This

putatively myristoylated Gly residue [Gly2 in AtSBT5.2(b)] is part of the prodomain in

AtSBT5.2(a) [Gly41 in AtSBT5.2(a)] and thus removed during AtSBT5.2(a) processing within

the secretory pathway (Figure 27). Endosomal localization of AtSBT5.2(b) is reminiscent of

that of Ara6, another member of the RAB GTPase family that is tightly associated with LEs

(Ueda et al., 2001). Interestingly, N-terminal myristoylation and palmitoylation of Ara6 appear

to be used as a substitute membrane anchoring system, essential for localization of Ara6 to

endosomes.

To investigate the role of Glyc2 in AtSBT5.2(b) myristolylation and localization to vesicles, a

mutant AtSBT5.2(b)G2A version in which the Gly2 residue has been substituted by an Ala

has been recently generated. If, as expected, this mutation abolishes AtSBT5.2(b)

endosomal localization and AtMYB30 nuclear exclusion, atsbt5.2 mutant plants will be

transformed with AtSBT5.2(b)G2A and inoculated with bacteria in order to obtain further

proof that myristoylation-mediated AtSBT5.2(b) location to endosomes is required to regulate

Arabidopsis HR through AtMYB30.

2.1.2. Endosomes as important sites for regulation of defence signalling

Our work suggests that endosomes are the subcellular sites of AtMYB30-AtSBT5.2(b)

interaction and support the emerging idea that endomembranes play a primary role in the

regulation of plant defence through endocytic trafficking. Endosomal trafficking pathways are

central regulators of plasma membrane protein homeostasis and also control developmental

processes and multiple signalling pathways, including those involved in plant disease

resistance (Reyes et al., 2011). Consistent with this idea, Lu and co-workers showed that

development of the plant-haustorium interface in compatible interactions with filamentous

pathogens Phytophthora infestans and Hyaloperonospora arabidopsidis is accompanied by

secretory vesicles and endosomal compartments surrounding haustoria, which suggests a

role for vesicle trafficking in the pathogen-controlled biogenesis of the extrahaustorial

membrane (Lu et al., 2012). This idea was further supported by enhanced susceptibility of

92

plants impaired in endosome-mediated trafficking regulators, emphasizing the importance of

endocytic processes in establishing susceptibility or resistance (Lu et al., 2012).

The fact that prominent immune-related cargos of plant trafficking pathways are PRRs, that

must be present at the plasma membrane to sense microbes in the apoplast, further

underlines the prominent role of endomembrane systems in determining plant resistance

(Beck et al., 2012, Ben Khaled et al., 2015). Indeed, upon flg22 stimulation, FLS2 is

ubiquitinated, internalized, and accumulates in late endosomes, prior to its degradation

(Chinchilla et al., 2006, Robatzek et al., 2006, Lu et al., 2011, Salomon and Robatzek, 2006).

Critically, endocytosis of FLS2 has been reported to be required for efficient PTI (Robatzek et

al., 2006), establishing a precedent for signalling from endosomes in plants. A prominent role

of both secretory and endocytic trafficking in defence against different pathogens is

additionally suggested by the transcriptional (Wang et al., 2014, Livaja et al., 2008, Navarro

et al., 2004) and posttranslational (Heese et al., 2005, Wang et al., 2014) changes of many

trafficking regulators upon triggering with MAMPs and apoplastic effectors. Moreover, ligand-

induced endocytosis is conserved across PRR families and different plant species (Robatzek

et al., 2006, Sharfman et al., 2011) and endosomal trafficking and localization of PRRs are

highly regulated processes that can be reprogrammed by pathogens (Chaparro-Garcia et al.,

2015, Du et al., 2015, Nomura et al., 2011, Göhre et al., 2008).

Beyond the presence of PRRs in the endocytic pathway, a number of NB-LRR resistance

proteins have been shown to be constitutively localized to endomembranes, although the

mechanistic basis for these localization in terms of effector recognition and HR signalling is

unknown (Takemoto et al., 2012). For example, the flax L6 protein conferring resistance to

the flax rust fungus Melampsora lini presents a predicted N-terminal signal anchor domain

that targets the protein to the endomembrane system (Takemoto et al., 2012). In addition,

the potato (Solanum tuberosum) resistance protein R3a relocates from the cytoplasm to

endosomal compartments only when coexpressed with the recognized Phytophthora

infestans effector form AVR3aKI and not with the unrecognized form AVR3aEM (Engelhardt et

93

al., 2012). Similarly, in the absence of R3a, AVR3aKI is cytoplasmic and does not associate

with vesicles (Bos et al., 2010, Gilroy et al., 2011). Engelhardt and co-workers showed

specific relocalization of AVR3aKI, but not AVR3aEM, to late endosomes, in the presence of

untagged (active) R3a and prior to the development of HR symptoms (Engelhardt et al.,

2012). Critically, treatment with inhibitors of the endocytic pathway (BFA and Wm) attenuated

accumulation of R3a at late endosomes and HR initiation. This work indicates that HR

signalling is not exclusively triggered within the nucleus and establishes vesicles as reported

sites from which signalling can be initiated to activate an efficient immune response

(Engelhardt et al., 2012).

2.2. AS affects the glycosylation status of resulting AtSBT5.2 protein variants

Proteins are synthesized in cytosolic ribosomes and delivered to their sites of function

through encoded or post-translationally appended signals. Most proteins that contain a SP

are committed to the secretory pathway (Porter et al., 2015). Membrane-bound and soluble

proteins of the secretory pathway are commonly glycosylated in the ER, where

carbohydrates are added primarily to newly synthesized, unfolded proteins (Xu and Ng,

2015). Within the secretory pathway of eukaryotic cells, two types of sugars, asparagines-

linked (N-linked) and serine/threonine-linked (O-linked) can be attached to proteins. N-linked

sugars are added in the ER as nascent proteins emerge from the translocon channel if the

proteins contain an N-X-S/T consensus sequence (X being any amino acid except proline)

(Hebert et al., 2005, Costantini and Snapp, 2013). It has been estimated that the majority of

secretory proteins are N-glycosylated (Apweiler et al., 1999). Interestingly, N-linked

glycosylation was recently shown to play a critical role in extracellular secretion (Kim et al.,

2015).

In this study, we demonstrate that, consistent with its entering the secretory pathway,

AtSBT5.2(a) is glycosylated. We also observed a reduced accumulation of AtSBT5.2(a)

protein after deglycosylation (Figure 32), suggesting that glycosylation may play a crucial role

in protein stabilization or protection from degradation, as previously described for other

94

proteins in plant (Kadek et al., 2014), yeast and animal cells (Wang et al., 1996, Sarkar and

Wintrode, 2011). Moreover, in our assays both processed and non-processed forms of

AtSBT5.2(a) appear to be deglycosylated (Figure 32), suggesting that cleavage of the

prodomain occurs after glycosylation. Finally, as previously described for the tomato

subtilase SlSBT3, the unprocessed AtSBT5.2(a)H210A mutant version was also sentitive to

deglycosylation by Endo H and PNGase F treatment (data not shown), further suggesting

that prodomain processing occurs late in the ER (Cedzich et al., 2009). In contrast to the

results observed for AtSBT5.2(a), AtSBT5.2(b) is not glycosylated (Figure 32), consistent

with the lack of an N-terminal SP and its not entering the secretory pathway.

2.3. AS appears to affect the catalytic activity of resulting AtSBT5.2 protein variants

As already mentioned, prodomain processing in plant subtilases is an intramolecular

autocatalytic reaction that occurs late in the ER and results in the formation of the active

mature form of the protease (Cedzich et al., 2009, Chichkova et al., 2010). For example,

cleavage of the prodomain of the tomato subtilase SlSBT3 prodomain occurred by

autoprocessing to render the active subtilase enzyme (Cedzich et al., 2009). Similarly,

AtSBT5.2(a) enters the secretory pathway and self cleaves in planta (Figure 34).

Interestingly, while two bands [likely corresponding to the processed and non-processed

forms of AtSBT5.2(a)] were detected after AtSBT5.2(a) transient expression in N.

benthamiana leaves, only the processed form of AtSBT5.2(a) is detected in Arabidopsis

protoplasts (data not shown), suggesting a faster or a more efficient processing in

Arabidopsis cells.

Processing-incompetent subtilase mutants were shown to accumulate intracellularly as

inactive proenzymes indicating that cleavage of the prodomain is, at least in same cases,

required for full passage through the secretory pathway (Cedzich et al., 2009, Janzik et al.,

2000). For example, processing of the prodomain of tomato SlSBT3 in the ER is a

prerequisite for targeting to the secretory pathway and secretion of the protein (Cedzich et

95

al., 2009). In contrast, our AtSBT5.2(a)H210A catalytic mutant was detected in apoplastic

fluids (data not shown), indicating that mutation of the catalytic site of AtSBT5.2(a) does not

affect secretion of the protein to the extracellular space.

The function of the prodomain as an auto-inhibitor of protease activity has been

demonstrated for plant proteases (Gu et al., 2012, Taylor et al., 1995). For example, the

prodomains of plant papain and papaya proteinase IV were reported to inhibit the activity of

their cognate proteases (Taylor et al., 1995, Groves et al., 1998). As other proteases,

subtilases typically comprise a prodomain, that assist in the folding of the protease (Li et al.,

1995, Baker et al., 1993, Takagi et al., 2001) and serves as an intramolecular inhibitor of

enzymatic activity (Ohta et al., 1991, Huang et al., 1997, Li et al., 1995). For instance, the

prodomain of cucumisin, a melon (Cucumis melo) subtilase protein, was produced in

recombinant form and shown to act as a tight-binding competitive inhibitor of mature

cucumisin (Nakagawa et al., 2010). Along the same lines, cucumisin activity was also

strongly inhibited by the prodomains of two other plant subtilases [Arabidopsis ARA12

(AtSBT1.7) (Rautengarten et al., 2008) and rice RSP1 (Yamagata et al., 2000, Nakagawa et

al., 2010)]. In contrast to AtSBT5.2(a), and in agreement with the absence of a SP and the

first five amino acids of its prodomain, AtSBT5.2(b) does not enter the secretory pathway and

is not processed in planta. Since AtSBT5.2(b) prodomain is not cleaved, it may fold onto the

catalytic domain of AtSBT5.2(b) and inhibit its catalytic activity as a subtilase.

Indeed, although both AtSBT5.2(a) and AtSBT5.2(b) present a conserved serine protease

catalytic triad, we were only able to detect a protease activity (that depends on the integrity of

its catalytic His residue) in the case of AtSBT5.2(a), both by its ability to self-process (Figure

34) and in ABP assays (Figure 35). Since no protease activity could be detected for

AtSBT5.2(b) in ABP assays, our group is currently investigating the protease activity of

AtSBT5.2(a) and AtSBT5.2(b) in additional assays, using the generic protease substrate

casein fluorescein isothiocyanate (FITC) to obtain further proof of the detected differences in

activity between the two proteins. The absence of detected protease activity in AtSBT5.2(b)

96

can be explained by the presence of the autoinhibitory prodomain in AtSBT5.2(b). Our

results are in agreement with previous reports that described AtSBT5.2 as an active Ser

hydrolase. Indeed, AtSBT5.2 was previously shown to be able to cleave synthetic Epidermal

Patterning Factor 2 (EPF2) peptides in vitro (Engineer et al., 2014). EPF2 encodes an

extracellular pro-peptide ligand that belongs to a family of 11EPF peptide proteins, which are

predicted to be converted to active peptide ligands upon cleavage. In addition, an ABP study

using FP-based probes to display the activities of serine hydrolases was previously

performed in Arabidopsis. Mass spectrometry analysis revealed over 50 serine hydrolases,

including the six subtilases AtSBT1.4, AtSBT1.7, AtSBT1.8, AtSBT3.13, AtSBT5.2 and

AtSBT6.2 (Kaschani et al., 2009). Suprisingly, in our ABP assays, the molecular size of the

band corresponding to active AtSBT5.2(a) was smaller than that detected using antibodies

against the HA tag positioned at the C-terminus of the protein (Figure 35). This result would

be consistent with additional C-terminal processing of the protein to render the active

subtilase, as previously shown for a subtilase from barley (Plattner et al., 2014). It is worth

however noting that no band of small size (that could correspond to the C-terminal processed

AtSBT5.2(a) domain) was detected in our assays, neither when AtSBT5.2(a) was C-

terminally tagged with a small epitope (HA tag), as in Figure 35, nor when a bigger epitope

(GFP tag) was used for C-terminal tagging (data not shown). This suggests that, if a C-

terminal AtSBT5.2(a) fragment is generated through processing, it may be rapidly degraded.

To further investigate this putative C-terminal processing of AtSBT5.2(a), it would be

interesting to generate an antibody against the protease. In addition, AtSBT5.2(a) could be

purified using the biotin-conjugated FP probe and analysed by mass spectrometry in order

the determine the N- and C-terminal sequences of the fully processed, active AtSBT5.2(a)

protein.

In additional experiments, we showed that AtMYB30 accumulation in planta is not affected by

the presence of AtSBT5.2(a) or AtSB5.2(b), indicating that these proteins are not able to

proteolytically cleave the TF. In the case of AtSBT5.2(a), this can be explained by the lack of

97

co-localization of both proteins in planta (AtSBT5.2(a) being secreted and AtMYB30 being

located in the nucleus). Despite their subcellular co-localization and physical interaction, lack

of modification of AtMYB30 accumulation in the presence of AtSBT5.2(b) is consistent with

AtSBT5.2(b) being inactive as a protease and suggests alternative modes of action of

AtSBT5.2(b) on AtMYB30 activity, likely related to the AtSBT5.2(b)-mediated AtMYB30

nuclear exclusion discussed below. Finally, the interaction between AtMYB30 and both

AtSBT5.2(b) wild-type and AtSBT5.2(b)H171A catalytic mutant versions has been recently

confirmed in our team by FRET-FLIM assays in endosomes of Arabidopsis protoplasts (data

not shown), thus confirming that AtSBT5.2(b) catalytic activity is not involved in AtMYB30

retention outside the nucleus.

98

3. The apoplast as a privileged site for anti-microbial defence

The plant apoplast has been described as a protease-rich environment in which proteases

are central components of the plant defence response (Xia et al., 2004, Krüger et al., 2002).

In particular, apoplastic subtilases have been proposed to be involved in pathogen

recognition (Jordá et al., 1999, Tornero et al., 1997) although functional proof of this

hypothesis has only been obtained recently (Ramírez et al., 2013). Similarly to the tomato

P69C, the secreted Arabidopsis AtSBT3.3 protein was linked to pathogen recognition and

activation of signalling processes. Indeed, AtSBT3.3 plays a role in pathogen-mediated

induction of SA-related defence gene expression and MAPK activation. The expression of

AtSBT3.3 is rapidly induced during PTI activation, preceeding the induction of SA-responsive

genes and responding very rapidly to PAMP-triggered ROS production. Moreover, AtSBT3.3

is involved in chromatin remodelling of defence-related genes associated with the activation

of immune priming (Ramírez et al., 2013). Although no AtSBT3.3 substrate was identified, it

was hypothesized that AtSBT3.3 may cleave a protein, likely functioning as a receptor

located in the plasma membrane. After proteolytic cleavage of its extracellular domain, the

receptor could become activated and initiate downstream immune signalling, as described in

animals (Ramírez et al., 2013, Ossovskaya and Bunnett, 2004).

AtSBT5.2 was previously described as a negative regulator of stomatal density under high

CO2 conditions through cleavage of the extracellular pro-peptide ligand EPF2 (Engineer et

al., 2014). Plants adapt to the continuing rise in atmospheric CO2 concentration by reducing

their stomatal density (that is, the number of stomata per unit of epidermal surface area).

Indeed, elevated CO2 induced upregulation of EPF2 and AtSBT5.2 transcripts (Engineer et

al., 2014). Interestingly, Engeneer and co-workers proposed that, similar to ERECTA, the

wide expression pattern of AtSBT5.2 indicates that the AtSBT5.2 protein could have

additional roles in plant growth and development.

As mentioned in the Introduction, pathogen entry into host tissue is a critical first step to

cause infection. For foliar bacterial plant pathogens, natural surface openings, such as

99

stomata, are important entry sites. However, diverse MAMPs are capable of inducing

stomatal closure. In agreement, mutant fls2 plants are impaired in stomatal closure in

response to flg22 and show increased susceptibility to Pto DC3000 when sprayed onto the

leaf surface but not when infiltrated into leaves (Chinchilla et al., 2006, Gómez-Gómez and

Boller, 2000, Zipfel et al., 2004, Zeng and He, 2010). In addition, Pst DC3000 produces the

diffusible virulence factor coronatine (COR) to promote stomatal re-opening (Melotto et al.,

2006), providing evidence that pathogen adjusts the plant physiological state associated

with their infection cycle, keeping stomata open at the beginning of the infection phase.

Interestingly, increasing atmospheric CO2 concentrations reduce plant stomatal opening thus

enhancing tomato defence against P. syringae (Li et al., 2015). AtSBT5.2(a) expression

appears to be rapidly downregulated after inoculation with Pst AvrRpm1 (Figure 49B) but the

significance of this finding remains to be investigated. Although transformation of atsbt5.2

mutant plants with a construct overexpressing AtSBT5.2(a) did not restore AtMYB30-

mediated HR responses (Figure 48), considering the described role of AtSBT5.2(a) in

regulation of stomatal density, we cannot rule out AtSBT5.2(a) to be involved in stomatal

regulation of pathogen entry during PTI. In order to investigate the potential role of

AtSBT5.2(a) in the regulation of PTI responses, expression of AtSBT5.2(a) should be

additionally monitored after elicitor treatment or after inoculation with virulent Pst DC3000 or

T3S-deficient Pst DC3000 (hrcU-) strains. Furthermore, it would also be interesting to

measure stomatal density (and closure) as well as bacterial growth rates, in atsbt5.2 mutant

plants and in plants overexpressing AtSBT5.2(a) in an atsbt5.2 mutant background, after

inoculation with these latter strains.

100

4. Nuclear exclusion through interaction with AtSBT5.2(b): a new

regulatory mechanism of AtMYB30 activity

During the last few years, different regulatory mechanisms of AtMYB30-mediated HR have

been uncovered in our group (Raffaele and Rivas, 2013). Indeed, spatio-temporal control of

AtMYB30 activity through the action of the secreted phospholipase AtsPLA2-α [that is

specifically relocalized to the nucleus in the presence of AtMYB30 (Froidure et al., 2010a)]

and the RING-type E3 ligase MIEL1 [that ubiquitinates AtMYB30 and leads to its

proteasomal degradation (Marino et al., 2013)] has been revealed. During my PhD work, I

have shown that AtMYB30 is excluded from the nucleus when co-expressed with

AtSBT5.2(b), suggesting that AtSBT5.2(b) may act as a negative regulator of AtMYB30-

mediated HR responses. This idea has been confirmed by the finding that atsbt5.2 mutant

plants display enhanced HR and that this phenotype is lost in an atsbt5.2/atmyb30 double

mutant background (Figure 46). Based on our results, this negative regulation appears to be

mediated by the capacity of AtSBT5.2(b) to retain AtMYB30 in intracellular vesicles outside

the nucleus. This would prevent AtMYB30 from activating its target genes, resulting in

attenuation of the HR.

Upon stimulation by internal and external signals, expression of numerous genes encoding

TFs is induced, and the newly synthesized TFs are typically transported into the nucleus.

Studies have shown that TF activity may be regulated at various levels after gene

transcription, both post-transcriptionally and post-translationally (See Introduction, section

3.5). Controlled nuclear localization is a well recognized mechanism regulating the activities

of TFs and co-regulatory proteins in eukaryotes (Zhang et al., 2001, Chariot et al., 1999). For

example, nuclear import of “dormant” TFs plays an important role in regulation of gene

expression. A small group of NAC and basic leucine zipper TFs was shown to be stored as

“dormant” forms in association with cellular membranes, including plasma membranes,

nuclear membranes, and ER membranes (Chen et al., 2008, Seo et al., 2008). Upon

exposure to environmental stresses, several membrane-bound TFs have been shown to be

101

proteolytically activated by either ubiquitin-mediated proteasome activities or by specific

membrane-bound proteases (Hoppe et al., 2001). Subtilases are known to be involved in the

release of “dormant” membrane-bound TFs. For example, AtSBT6.1 is able to cleave the

ER-located TF AtbZIP17 that, once released, moves to the nucleus to activate transcription

of salt stress genes (Liu et al., 2007). In this context, controlled proteolytic activation of

membrane-bound TFs has been proposed as a way of triggering quick transcriptional

responses that are necessary to ensure plant survival under stressful conditions (Kim et al.,

2006, Seo et al., 2010b).

Interestingly, nuclear exclusion by localization to small vesicle-like structures has been

reported as a negative regulatory mechanism of TF activity.The Arabidopsis small interfering

protein, MIni zinc Finger1 (MIF1) presents a ZF motif but lacks the homeodomain motifs

necessary for TF activity (Hu et al., 2008). MIF1 and its functional homologues physically

interact with a group of Zinc finger HomeoDomain (ZHD) TFs, including ZHD5, that regulate

flower architecture and leaf development. MIF1 blocks DNA-binding and transcriptional

activation of ZHD5 homodimers by competitively forming MIF1-ZHD5 heterodimers. Notably,

MIF1 interferes with the nuclear localization or promotes the nuclear exclusion of ZHD5 by

relocalizing the TF to small vesicle-like structuresin which MIF1 is localized (Hong et al.,

2011).

Considering the overlapping expression profiles between AtMYB30 and AtSBT5.2(b) after

inoculation with avirulent bacteria (Figure 49), it is tempting to speculate that AtMYB30

nuclear exclusion after bacterial inoculation may provide a molecular mechanism to

attenuate the HR. In order to further investigate AtSBT5.2(b)-mediated AtMYB30 nuclear

exclusion, Arabidopsis lines expressing a GFP-tagged AtMYB30 version have been recently

generated and transformed with a dexamethasone (dex)-inducible version of AtSBT5.2(b).

These lines will be used in future experiments to follow the dynamics of AtSBT5.2(b) and

AtMYB30 subcellular distribution before and after bacterial inoculation. In addition,

expression of AtSBT5.2(b) in the area immediately surrounding the inoculated (HR

102

developping) zone should be monitored in the future. Indeed, expression of the AtMYB30

negative regulator AtsPLA2-a peaks 6 hpi in peripheral but not in challenged cells,

suggesting that AtPLA2-a may contribute to restrict the development of the HR to the

inoculated zone, thereby preventing spreading of cell death throughout the leaf (Froidure et

al., 2010a). These experiments should contribute to further our understanding of the mode of

action of AtSBT5.2(b) in the regulation of AtMYB30-mediated HR.

103

Materials and Methods

__________________________________________________________________________

Name Sequence (5' - 3') Polarity Comment

At2g28390-S 5ɂ AACTCTATGCAGCATTTGATCCACT (+) SAND expression

At2g28390-AS 5ɂ TGATTGCATATCTTTATCGCCATC (-) SAND expression

AtMYB30 qRT-PCR-S 5' TCAAGAGTGATGATGGGAAGGAGT (+) AtMYB30 expression

AtMYB30 qRT-PCR-AS 5' GTCCACCAGAATCCTCAAACA (-) AtMYB30 expression

AtSBT5.1 N197A-S 5' GAGCAAGGTATTACGCTTCGTCATTCTTCTTA (+) PGS mutant

AtSBT5.1 N197A-AS 5' TAAGAAGAATGACGAAGCGTAATACCTTGCTC (-) PGS mutant

AtSBT5.1 N230A-S 5' GGCAAATAATAGCAGCTGCGTCCTACTATG (+) PGS mutant

AtSBT5.1 N230A-AS 5' CATAGTAGGACGCAGCTGCTATTATTTGCC (-) PGS mutant

AtSBT5.1 N471A-S 5' ATCATGTCCTACATAGCTTCAACAAGAGAACC (+) PGS mutant

AtSBT5.1 N471A-AS 5' GGTTCTCTTGTTGAAGCTATGTAGGACATGAT (-) PGS mutant

AtSBT5.1 N776A-S 5' CAGCAAAGACGACGCCGATAGCGAACGTTA (+) PGS mutant

AtSBT5.1 N776A-AS 5' TAACGTTCGCTATCGGCGTCGTCTTTGCTG (-) PGS mutant

AtSBT5.1 qRT-PCR-S 5ɂ CAGATACCCTTCATGGAGTCCT (+) AtSBT5.1 expression

AtSBT5.1 qRT-PCR-AS 5ɂ CCTGTGTTGGTCATTTGCAC (-) AtSBT5.1 expression

AtSBT5.2 1482-AS 5’ AAAATATGCAACAGCAGGGG (-) Sequencing

AtSBT5.2 261-AS 5’ GGCCTCTTCTGCTGTCAAAC (-) Sequencing

AtSBT5.2 969-AS 5’ AGCGCTCTTAGCAGACTTGC (-) Sequencing

AtSBT5.2 His to Ala-S 5' AGGGATGTCATCGGTGCCGGTTCTCATGTGTC (+) Catalytic mutant

AtSBT5.2 His to Ala-AS 5' GACACATGAGAACCGGCACCGATGACATCCCT (-) Catalytic mutant

AtSBT5.2 N225A-S 5' TCTGCCGTGGAGGCTGCTTCCTACTAT (+) PGS mutant

AtSBT5.2 N225A-AS 5' ATAGTAGGAAGCAGCCTCCACGGCAGA (-) PGS mutant

AtSBT5.2 N363A-S 5' GGTATACACTTTTCAGCCGTTAGTAAATCTCCT (+) PGS mutant

AtSBT5.2 N363A-AS 5' AGGAGATTTACTAACGGCTGAAAAGTGTATACC (-) PGS mutant

AtSBT5.2 N467A-S 5' CTTCTCCTACCTCGCCTCAACCAAAGATCC (+) PGS mutant

AtSBT5.2 N467A-AS 5' GGATCTTTGGTTGAGGCGAGGTAGGAGAAG (-) PGS mutant

AtSBT5.2 N525A-S 5' TGCATGGACTGGAGCCGACTCAAGCATTTC (+) PGS mutant

AtSBT5.2 N525A-AS 5' GAAATGCTTGAGTCGGCTCCAGTCCATGCA (-) PGS mutant

AtSBT5.2 N636A-S 5' TGTTACTATGGATATGCCGTAACCACAATAAAG (+) PGS mutant

AtSBT5.2 N636A-AS 5' CTTTATTGTGGTTACGGCATATCCATAGTAACA (-) PGS mutant

AtSBT5.2 N650A-S 5' AAGCTTTTCCAGAGGCTTTTACTTGCCCTG (+) PGS mutant

AtSBT5.2 N650A-AS 5' CAGGGCAAGTAAAAGCCTCTGGAAAAGCTT (-) PGS mutant

AtSBT5.2 N678A-S 5' CTGGATTCAAAGGAGCTGGTAGCAAGACAG (+) PGS mutant

AtSBT5.2 N678A-AS 5' CTGTCTTGCTACCAGCTCCTTTGAATCCAG (-) PGS mutant

AtSBT5.2(a) qRT-PCR-S 5' GCCATGAAAGGCATTACATTCT (+) AtSBT5.2(a) expression

AtSBT5.2(b) qRT-PCR-S 5' GATCTATCTATAGCTTCCAGTG (+) AtSBT5.2(b) expression

AtSBT5.2(a) and (b) qRT-PCR-AS 5' GAAGCTGATCCCATGTAGACAA (-) AtSBT5.2(a) and AtSBT5.2(b) expression

AtSBT5.2 qRT-PCR-S 5ɂ CCTCACAAGAAGCATTCTCAAAC (+) AtSBT5.2(a) and AtSBT5.2(b) expression

AtSBT5.2 qRT-PCR-AS 5ɂ CCTGATATGACGTTATACTGAGAAGC (-) AtSBT5.2(a) and AtSBT5.2(b) expression

AtSBT5.2(a) S 5' GAATAAGTCTTTCCAGTGATTAG (+) used to specifically amplify AtSBT5.2(a)

AtSBT5.2(b) S 5' GATCTATCTATAGCTTCCAGTG (+) used to specifically amplify AtSBT5.2(b)

AtSBT5.2 563AS 5' CCAATGATCTTTCTGTTACAGT (+) used to amplify AtSBT5.2(a) and AtSBT5.2(b)

AtSBT5.3 qRT-PCR-S 5' CAAGATATATCAGCCAAACCAAGAA (-) AtSBT5.3 expression

AtSBT5.3 qRT-PCR-AS 5' CCATTACAGGCGCTGGTT (+) AtSBT5.3 expression

AtSBT5.6 qRT-PCR-S 5' CGTCGGTGTTCTCGACAGT (-) AtSBT5.6 expression

AtSBT5.6 qRT-PCR-AS 5' GGCAGATTCCTTTCCATGATT (+) AtSBT5.6 expression

attB1-AtSBT5.1 5' ggggacaagtttgtacaaaaaagcaggcttaATGATGAGATGCCTCACTATC (+) GW cloning

attB1-AtSBT5.1 (405-780) 5' ggggacaagtttgtacaaaaaagcaggcttaATGGTAAAAGGGAAGATTGTGT (+) GW cloning (Truncated version)

attB1-AtSBT5.2(a) 5’ ggggacaagtttgtacaaaaaagcaggcttaATGAAAGGCATTACATTCTTCA (+) GW cloning

attB1-AtSBT5.2(b) 5' ggggacaagtttgtacaaaaaagcaggcttaATGGGATCAGCTTCCTCTGC (+) GW cloning

attB1-AtSBT5.2(b) (362-730) 5' ggggacaagtttgtacaaaaaagcaggcttaATGGTAAAAGGGAAGATTGTGT (+) GW cloning (Truncated version)

attB1-AtSBT5.2(b) (162-730) 5' ggggacaagtttgtacaaaaaagcaggcttaATGTACTATACCACAAGGGATG (+) GW cloning (Truncated version)

attB2-AtSBT5.1 A 5' ggggaccactttgtacaagaaagctgggtcTTAACGTTCGCTATCGTTGTC (-) GW cloning

attB2-AtSBT5.1 B 5' ggggaccactttgtacaagaaagctgggtcACGTTCGCTATCGTTGTCGT (-) GW cloning

attB2-AtSBT5.2(a) and (b) A 5’ ggggaccactttgtacaagaaagctgggtcGTTTGTGCGGCTACTCTCG (-) GW cloning

attB2-AtSBT5.2(a) and (b) B 5’ ggggaccactttgtacaagaaagctgggtcTCAGTTTGTGCGGCTACTCT (-) GW cloning

T DNA-LB 5’ CCCTTTAGGGTTCCGATTTAGTGCT (+/-) T-DNA lines

Supplemental Table 1. Oligonucleotide primers used in this study.

104

Yeast assays

The Y2H screen and methods used for identifi cation of AtSBT5.2 were previously described

(Froidure et al., 2010a). Briefl y, an Arabidopsis thaliana Gal4 yeast two-hybrid cDNA prey

library (MatchMaker; Clontech) was generated from mRNA isolated from leaves of 4-week-

old plants (Ws-4 ecotype), syringe-infi ltrated with the Xanthomonas campestris pv.

campestris 147 strain (Raffaele et al., 2008). An AtMYB30 version deleted from its C-terminal

activation domain (amino acids 1–234) was used as bait for screening 2 x 106 independent

transformants exhibiting His auxotrophy on selective plates.

Plasmid Constructions

Plasmids used in this study were constructed by Gateway technology (GW; Invitrogen)

following the instructions of the manufacturer. All primer sequences are listed in

Supplementary Table 1. PCR products flanked by the attB sites were recombined into the

pDONR207 vector (Invitrogen) via a BP reaction to create the corresponding entry clones

with attL sites (pENTR). Inserts cloned into the entry clones were subsequently recombined

into the destination vectors via an LR reaction to create the expression constructs.

For yeast assays, GAL4-BD-AtMYB30ΔAD, and GAL4-BD-AtMYB123ΔAD fusions were

previously described (Froidure et al., 2010a). AD-AtSBT5.2 and AD-AtSBT5.1 constructs

were generated from recombination of the corresponding entry constructs with the pGAD-

AD-GW vector (Froidure et al., 2010a).

AtSBT5.2(a), AtSBT5.2(b) and AtSBT5.1 coding sequences were amplified, using primers

attB1-AtSBT5.2(a) (5’ ggggacaagtttgtacaaaaaagcaggcttaATGAAAGGCATTACATTCTTCA)

and attB2-AtSBT5.2(a) (5’ ggggaccactttgtacaagaaagctgggtcGTTTGTGCGGCTACTCTCG),

attB1-AtSBT5.2(b) (5' ggggacaagtttgtacaaaaaagcaggcttaATGGGATCAGCTTCCTCTGC)

and attB2-AtSBT5.2(b) (5’ ggggaccactttgtacaagaaagctgggtcGTTTGTGCGGCTACTCTCG),

105

and attB1-AtSBT5.1 (5’ ggggacaagtttgtacaaaaaagcaggcttaATGATGAGATGCCTCACTATC)

and attB2-AtSBT5.1 (5’ ggggaccactttgtacaagaaagctgggtcTTAACGTTCGCTATCGTTGTC),

respectively, from first-strand cDNAs synthesized from 1.5 µg of total RNA (Col-0; 1 week-old

seedlings for AtSBT5.2 and flowers for AtSBT5.1) using oligo (dT) primer and Transcriptor

Reverse Transcriptase (Roche Diagnostics, Meylan, France).

Point mutations were generated using the QuikChange mutagenesis kit (Stratagene) using

the pENTR-AtSBT5.2(a), pENTR-AtSBT5.2(b) or pENTR-AtSBT5.1 as templates and

following the manufacturer’s instructions. Primers used for mutagenesis are shown in

Supplementary Table 1.

TAP- , HA-, eGFP- and RFP-tagged constructs were generated by recombination of the

corresponding entry vectors with pBin19-35S-GW-3xFLAG, pBin19-35S-GW-TAP, pBin19-

35S-GW-3xHA, pB7FWG2-35S-GW-eGFP or pB7WGF2-35S-eGFP-GW and pB7RWG2-

35S-GW-RFP destination vectors, respectively.

P35S:AtMYB30-TAP construct was previously described (Froidure et al., 2010a).

Bacterial Strains

Escherichia coli DH5 alpha (Novagen) was grown at 37°C on Luria broth medium containing

the required antibiotics.

Agrobacterium tumefaciens C58C1 cells were transformed with pB7GW-derived constructs

using a standard electroporation method and grown on low-salt LB agar medium containing

rifampicin (50 μ g/mL), tetracycline (10 μ g/mL) and spectinomycin (50 μ g/mL) at 28°C. A.

tumefaciens C58C1 cells transformed with pBin19 derived constructs were selected on

rifampicin (50 μ g/mL), tetracycline (10 μ g/mL) and kanamycin (50 μ g/mL) at 28°C.

106

Pseudomonas syringae pv. tomato AvrRpm1 (Pst AvrRpm1) strain was grown on King’s B

medium containing rifampicin (50 μ g/mL) and tetracycline (10 μ g/mL) at 28 °C.

Plant materials and inoculation methods

All Arabidopsis lines used in this study were in the Columbia background. As a wild-type

control, we used Col-0 (Nottingham Arabidopsis Stock Centre [NASC] accession number

N1093). Plants were grown in Jiffy pots under controlled conditions. Briefly, seeds were

germinated on Murashige and Skoog (MS) medium and plants were grown in Jiffy pots in a

growth chamber at 22°C, with a 9 hour light period and a light intensity of 190 µmol.m-2.s-1.

For transient expression of proteins in N. benthamiana, overnight bacterial cultures of

Agrobacterium tumefaciens strain C58C1 expressing the protein of interest were harvested

by centrifugation. Cells were resuspended in induction buffer (10 mM MgCl2, 10 mM MES,

pH 5.6, and 150 mM acetosyringone) to an OD600 of 0.5. After 1 h at 22°C, cells were

infiltrated into leaves of 4-week-old N. benthamiana plants. Two days after A. tumefaciens

infiltration, leaf discs used for experiments were harvested and processed, or frozen

immediately in liquid nitrogen and stored at -80°C.

Arabidopsis 4-week-old plants were kept at high humidity 12 h before inoculation and

injected with a bacterial suspension of Pst AvrRpm1 at the indicated bacterial densities using

a blunt syringe on the abaxial side of the leaves. For determination of in planta bacterial

growth, leaves samples were harvested 0 and 3 days post-inoculation and ground on sterile

water. A 1:1000 dilution for each sample was plated on King’s B medium and incubated at

28°C for 2 days. Data were submitted to a statistical analysis. The effect of the genotype was

tested by using Student test (t- test; P < 0.05).

Protein extraction and Western blot analysis

107

N. benthamiana leaf discs were ground in liquid nitrogen and resuspended in 2 volumes of

extraction buffer [50 mM Tris-HCl (pH 7.4), 150 mM NaCl, 10% glycerol (v/v), 1 mM DTT, 1

mM PMSF, 1% plant protease inhibitor cocktail (Sigma)] and centrifuged at 10,000g for 10

min at 4°C. Protein concentration in the supernatant was determined with the Bradford

protein assay kit (Bio-Rad), using BSA as a standard. Fifty micrograms of total protein were

separated on a 7.5% polyacrylamide gel (Mini-PROTEAN® TGX™ Precast Protein Gels,

BioRad) according to the manufacturer’s instructions and transferred onto nitrocellulose

membranes (Trans-Blot® Turbo™ RTA Midi Nitrocellulose Transfer Kit, BioRad) by semi-dry

blotting systems (Trans-Blot® Turbo™ Transfer System; Bio-Rad).

Antibodies used for western blotting were rabbit anti-PAP soluble complex-HRP (Sigma,

1:2000), ratmonoclonal anti-HA-HRP (3F10, Roche, 1:5000), anti-FLAG-HRP (M2, Sigma,

1:5000), mouse monoclonal anti-GFP IgG 1 K (clones 7.1 and 13.1; Roche) and goat anti-

mouse IgG-HRP (Santa Cruz, 1:10000). Proteins were visualized using the Immobilon kit

(Millipore) under standard conditions.

Isolation of intercellular (apoplastic) fluid

N. benthamiana leaves transiently expressing the proteins of interest were harvested 48

hours after agroinfiltration and infiltrated with water. Intercellular fluids (IF) were isolated by

centrifugation at 3,000g as previously described (De Wit and Spikman, 1982).

Concanavalin A Purification

Proteins were extracted in 50 mM Tris-HCl, pH 7.5, 150 mM NaCl, 10% [v/v] glycerol, 1 mM

PMSF, and 1% plant protease inhibitor cocktail (Sigma-Aldrich) and centrifuged at 14,000g

for 10 min at 4°C. The supernatant was equilibrated in concanavalin A buffer (0.2 M Tris-HCl

pH 7.5, 1 M NaCl, 200 mM MgCl2, 200 mM CaCl2) and applied to concanavalin A-agarose

108

resin from Canavalia ensiformis (Sigma-Aldrich) pre-equilibrated in concanavalin A buffer.

After three steps of washing with concanavalin A buffer, glycosylated proteins were eluted in

concanavalin A buffer supplemented with 0.75 M a-methyl-D-glycosamide and 0.75 M a-

methyl-D-manosamide. The presence of HA-tagged AtSBT5.2 in the eluted proteins was

confirmed by Western blot using anti HA antibodies.

Deglycosylation experiments

Proteins were extracted in 50 mM Tris-HCl, pH 7.5, 150 mM NaCl, 10% [v/v] glycerol, 1 mM

PMSF, and 1% plant protease inhibitor cocktail (Sigma-Aldrich) and centrifuged at 14,000g

for 10 min at 4°C. Proteins in the supernatant were denatured and then incubated with

PNGase F or Endo H (New England Biolabs) following the instructions of the manufacturer.

Deglycosylation reactions were performed for 30 minutes and stopped by adding SDS-PAGE

loading buffer and boiling. Proteins were detected by Western blot using anti-HA antibodies.

For tunicamycin treatment, 24 hours after Agrobacterium-mediated transient expression, leaf

discs of N. benthamiana leaves were incubated in a solution of 10 µM tunicamycin for 20

hours at room temperature and later frozen in liquid nitrogen before processing. Proteins

were detected by Western blot using anti-HA antibodies.

Specific detection of Active Serine Hydrolases

N. benthamiana leaf tissue expressing the indicated proteins were ground in 25 mM Tris-HCl

(pH 7.5), 1 % NP-40, 150 mM NaCl, 5 % glycerol. 50 µl protein extracts (1.5 mg/ml) were

incubated with 1 µl of ActivX Desthiobiotin-FP Serine Hydrolase Probe (Thermo Scientific).

The final probe concentration in the reaction was 2 µM. After 1 hour of incubation at room

temperature, the labelling reaction was stopped by adding 50 µl of 2x gel loading buffer and

109

boiling. Labelled proteins were visualised by Western blot using HRP-labelled streptavidin

(1:10,000).

Fluorescence Microscopy

The eGFP and RFP fluorescence in N. benthamiana leaves was analyzed with a confocal

laser scanning microscope (TCS SP2-AOBS; Leica) by using a ×63 water immersion

objective lens (numerical aperture 0.9; PL APO). eGFP fluorescence was excited with the

488 nm ray line of the argon laser and recorded in one of the confocal channels in the 500-

550nm emission range. RFP fluorescence was excited with the 561 nm line ray of the argon

laser and detected in the range between 570 and 630 nm. Images were acquired in the

sequential mode (20 z plains per stack of images; 0.5 μm per z plain) by using Leica LCS

software (version 2.61). Image overlays have been realized on LAS AF Leica Software.

FRET-FLIM Measurements

Fluorescence lifetime of the donor (eGFP) was experimentally measured in the presence and

absence of the acceptor (RFP or Sytox Orange). FRET effi ciency (E) was calculated by

comparing the lifetime of the donor in the presence (tDA) or absence (tD) of the acceptor: t =

1 - (tDA)/( tD). Statistical comparisons between control (donor) and assay (donor + acceptor)

lifetime values were performed by Student t-test. For each experiment, eight leaf discs

removed from four A. tumefaciens infiltrated leaves were used to collect data. Fluorescence

lifetime measurements were performed using a FLIM system coupled to a streak camera.

The light source is a mode-locked Ti:sapphire laser (Tsunami, model 3941, Spectra-Physics,

USA) pumped by a 10W diode laser (Millennia Pro, Spectra-Physics) and delivering ultrafast

femtosecond pulses of light with a fundamental frequency of 80MHz. A pulse picker (model

3980, Spectra-Physics) is used to reduce the repetition rate to 2MHz to satisfy the

110

requirements of the triggering unit (working at 2MHz). The experiments were carried out at λ

= 860 nm (multiphoton excitation mode). All images were acquired with a x 63 oil immersion

lens (Plan Apo 1.4 numerical aperture, IR) mounted on an inverted microscope (Eclipse

TE2000E, Nikon, Japan) coupled to the FLIM system. The fl uorescence emission was

directed back out into the detection unit through a short pass fi lter (λ <750 nm) and a band

pass filter (515/30 nm). The detector was composed of a streak camera (Streakscope

C4334, Hamamatsu Photonics, Japan) coupled to a fast and high sensitivity CCD camera

(model C8800-53C, Hamamatsu). For each subcellular domain, average fluorescence decay

profiles were plotted and lifetimes were estimated by fitting data with bi-exponential function

using a non-linear least-squares estimation procedure.

Molecular analysis of Arabidopsis T-DNA mutant lines

The AtSBT5.2 (atsbt5.2-1: SALK_012113 and atsbt5.2-2: SALK_132812C) T-DNA insertion

lines were derived from the SALK collection (http://signal.salk.edu). The AtSBT5.1 (atsbt5.1-

1: SALK_017993 and atsbt5.1-2: SALK_121716) T-DNA insertion lines were kindly provided

by Prof. Dr. Andreas Schaller (University of Hohenheim, Germany).

The position of the T-DNA insertion was confirmed by sequencing PCR fragments obtained

from genomic DNA, using a T-DNA left border (T-DNA-LB) as well as AtSBT5.2 or AtSBT5.1

gene-specific primers. F2 homozygous plants for the T-DNA insertion were selected by PCR.

Quantification of cell death using electrolyte leakage

Four leaf discs (6 mm diameter) were harvested 24 hpi, washed and incubated at room

temperature in 5 ml of distilled water before measuring conductivity. Four independent

experiments were performed with three plants (four leaves per plant).

111

RNA extraction and quantitative Real-Time-PCR (qRT-PCR) analysis

Arabidopsis Col-0 4-week-old plants were inoculated with a bacterial suspension of

Pseudomonas syringae pv. tomato AvrRpm1 (5 x 107 cfu/ml). Leaf samples were harvested

at the indicated time points inside the infiltrated zone and ground in liquid nitrogen. Total

RNA was isolated by using the Nucleospin RNA plant kit (Macherey-Nagel) according to the

manufacturer’s recommendations. Reverse transcription was performed by using 1.5 μ g of

total RNA. Quantitative RT-PCR (qRT-PCR) was performed on a Light Cycler 480 machine

(Roche Diagnostics, Meylan, France), using Roche reagents.

Primers used for qRT-PCR analysis in Arabidopsis were the following: for AtMYB30,

AtMYB30 qRT-PCR-S (5′ tcaagagtgatgatgggaaggagt) and AtMYB30 qRT-PCR-AS (5′

gtccaccagaatcctcaaaca); for AtSBT5.2(a), AtSBT5.2(a)-S (5' gccatgaaaggcattacattct) and

AtSBT5.2(a and b)-AS (5' gaagctgatcccatgtagacaa); for AtSBT5.2(b), AtSBT5.2(b)-S (5'

gatctatctatagcttccagtg ) and AtSBT5.2(a and b)-AS (5' gaagctgatcccatgtagacaa); for internal

controls SAND family, At2g28390-S (5′ aactctatgcagcatttgatccact) and At2g28390-AS (5′

tgattgcatatctttatcgccatc).

Relative expression was calculated as the ΔCp between each gene and the internal control

SAND family gene (At2g28390). Average ΔCp was calculated from three independent

experiments with three replicates and related to the value of each gene at time 0, which is

set at 1.

5’ RACE assays

5’ ends of AtSBT5.2 mRNA were determined using the GeneRacerTM RACE Ready kit

(Invitrogen, France) according to manufacturers’ instructions and using RNA from Col-0

leaves. Products from consecutive PCRs using AtSBT5.2 1611-AS, AtSBT5.2 1482-AS and

112

AtSBT5.2 969-AS as gene specific primers were cloned in pGEM-T Easy vector (Promega

Corporation) and sequenced.

113

Other results

__________________________________________________________________________

114

During my thesis, I had the opportunity to participate in a collaborative project among

different European laboratories that led to a publication of a research article entitled “A

Conserved Core of Programmed Cell Death Indicator Genes Discriminates Developmentally

and Environmentally Induced Programmed Cell Death in Plants” in the journal Plant

Physiology (Olvera-Carrillo et al., 2015).

In this study, a set of publicly available genome-wide transcriptome data that were

associated with different forms of cell death in Arabidopsis thaliana were exploited, with the

aim of comparatively characterizing distinct plant PCD types. This meta-analysis allowed to

identify largely non-overlapping sets of differentially regulated genes in differentiation-

induced/developmental (dPCD) and environmental (ePCD) situations known to provoke

PCD. This observation suggested that dPCD and ePCD processes are regulated in a largely

independent manner. In order to confirm these results experimentally, I contributed to this

work by analyzing the expression profile of canonical dPCD marker genes during the HR

induced in Arabidopsis plants inoculated with the bacterial strain Pseudomonas syringae

DC3000 (AvrRpm1). Interestingly, expression of none of the nine dPCD selected marker

genes was upregulated during HR PCD.

Our study indicates that the transcriptional signatures of dPCD are largely distinct from those

associated with ePCD. Moreover, different cases of dPCD share a set of cell death-

associated genes. Most of these genes are evolutionary conserved within the green plant

lineage, arguing for an evolutionary-conserved core machinery of developmental PCD.

Supplemental data: http://www.plantphysiol.org/content/169/4/2684/suppl/DC1

A Conserved Core of Programmed Cell Death IndicatorGenes Discriminates Developmentally andEnvironmentally Induced Programmed CellDeath in Plants1[OPEN]

Yadira Olvera-Carrillo2, Michiel Van Bel, Tom Van Hautegem, Matyáš Fendrych3, Marlies Huysmans,Maria Simaskova, Matthias van Durme, Pierre Buscaill, Susana Rivas, Nuria S. Coll, Frederik Coppens,Steven Maere, and Moritz K. Nowack*

Department of Plant Systems Biology, Vlaams Instituut voor Biotechnologie, and Department of PlantBiotechnology and Bioinformatics, Ghent University, 9052 Ghent, Belgium (Y.O.-C., M.V.B., T.V.H., M.F., M.H.,M.S., M.v.D., F.C., S.M., M.K.N.); Institut National de la Recherche Agronomique, Laboratoire des InteractionsPlantes-Microorganismes, Unité Mixte de Recherche 441, and Centre National de la Recherche Scientifique,Laboratoire des Interactions Plantes-Microorganismes, Unité Mixte de Recherche 2594, F–31326 Castanet-Tolosan, France (P.B., S.R.); and Center for Research in Agricultural Genomics, Bellaterra-Cerdanyola del Valles,08193 Barcelona, Spain (N.S.C.)

ORCID IDs: 0000-0001-6161-7053 (Y.O.-C.); 0000-0003-0792-8736 (M.H.); 0000-0001-6565-5145 (F.C.); 0000-0001-8918-7577 (M.K.N.).

A plethora of diverse programmed cell death (PCD) processes has been described in living organisms. In animals and plants,different forms of PCD play crucial roles in development, immunity, and responses to the environment. While the molecularcontrol of some animal PCD forms such as apoptosis is known in great detail, we still know comparatively little about theregulation of the diverse types of plant PCD. In part, this deficiency in molecular understanding is caused by the lack ofreliable reporters to detect PCD processes. Here, we addressed this issue by using a combination of bioinformatics approachesto identify commonly regulated genes during diverse plant PCD processes in Arabidopsis (Arabidopsis thaliana). Our results indicatethat the transcriptional signatures of developmentally controlled cell death are largely distinct from the ones associated withenvironmentally induced cell death. Moreover, different cases of developmental PCD share a set of cell death-associated genes.Most of these genes are evolutionary conserved within the green plant lineage, arguing for an evolutionary conserved coremachinery of developmental PCD. Based on this information, we established an array of specific promoter-reporter lines fordevelopmental PCD in Arabidopsis. These PCD indicators represent a powerful resource that can be used in addition toestablished morphological and biochemical methods to detect and analyze PCD processes in vivo and in planta.

Programmed cell death (PCD) is a fundamentalprocess of life. Already present in clonal colonies ofprokaryotes (Bayles, 2014), PCD has evolved to becomean essential mechanism in multicellular eukaryotes(Wang and Bayles, 2013). Many different forms of PCDhave been recognized, but a unifying definition char-acterizes PCD as genetically encoded, actively con-trolled cellular suicide.

In animals and plants, PCD is involved in many as-pects of development, sculpting structures or deleting un-wanted tissues (Fuchs and Steller, 2011; Van Hautegemet al., 2015). Over the last two decades, intensive in-vestigations have revealed mechanisms controllingdifferent forms of animal PCD; the most prominentamong them is apoptotic PCD (Green, 2011). In com-parison, there is still little knowledge of the molecularnetworks controlling PCD in plants, despite its abun-dance and its importance for plant life: plant PCD oc-curs as an integral part of development (dPCD) as wellas of the plant’s reactions to biotic and abiotic environ-mental challenges (ePCD; Lam, 2004). Concerning dPCD,

1 This work was supported by the Consejo Nacional de Ciencia yTecnología (postdoctoral fellowship registration nos. 186253 and203288 to Y.O.-C.) and the French Laboratory of Excellence (projectTULIP ANR–10–LABX–41 and ANR–11–IDEX–0002–02).

2 Present address: TOKU-E N.V., Poortakkerstraat 21–001, 9051 Sint-Denijs-Westrem, Belgium.

3 Present address: Institute of Science and Technology Austria, AmCampus 1, A–3400 Klosterneuburg, Austria.

* Address correspondence to [email protected] author responsible for distribution of materials integral to the

findings presented in this article in accordance with the policy de-scribed in the Instructions for Authors (www.plantphysiol.org) is:Moritz K. Nowack ([email protected]).

Y.O.-C. and M.K.N. conceived and coordinated the study; M.V.B.,F.C., and S.M. designed, and M.V.B. and S.M. performed the bioinfor-matics analyses; P.B., S.R., andN.S.C. designed, performed, and analyzedthe biotic stress experiments; Y.O.-C. performed and analyzed the wetlabexperiments and designed the figures together with M.F. and M.V.D.;T.V.H., M.H., M.S., and M.V.D. participated in the microscopy work;M.H. performed the terminal deoxynucleotidyl transferase dUTPnick-end labeling assays; Y.O.-C., S.M., and M.K.N. wrote the article.

[OPEN] Articles can be viewed without a subscription.www.plantphysiol.org/cgi/doi/10.1104/pp.15.00769

2684 Plant Physiology!, December 2015, Vol. 169, pp. 2684–2699, www.plantphysiol.org " 2015 American Society of Plant Biologists. All Rights Reserved.www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from

Copyright © 2015 American Society of Plant Biologists. All rights reserved.www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from

Copyright © 2015 American Society of Plant Biologists. All rights reserved.www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from

Copyright © 2015 American Society of Plant Biologists. All rights reserved.www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from

Copyright © 2015 American Society of Plant Biologists. All rights reserved.www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from

Copyright © 2015 American Society of Plant Biologists. All rights reserved.www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from

Copyright © 2015 American Society of Plant Biologists. All rights reserved.www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from

Copyright © 2015 American Society of Plant Biologists. All rights reserved.www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from

Copyright © 2015 American Society of Plant Biologists. All rights reserved.www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from

Copyright © 2015 American Society of Plant Biologists. All rights reserved.www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from

Copyright © 2015 American Society of Plant Biologists. All rights reserved.www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from

Copyright © 2015 American Society of Plant Biologists. All rights reserved.www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from

Copyright © 2015 American Society of Plant Biologists. All rights reserved.www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from

Copyright © 2015 American Society of Plant Biologists. All rights reserved.www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from

Copyright © 2015 American Society of Plant Biologists. All rights reserved.www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from

Copyright © 2015 American Society of Plant Biologists. All rights reserved.www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from

Copyright © 2015 American Society of Plant Biologists. All rights reserved.www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from

Copyright © 2015 American Society of Plant Biologists. All rights reserved.www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from

Copyright © 2015 American Society of Plant Biologists. All rights reserved.www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from

Copyright © 2015 American Society of Plant Biologists. All rights reserved.www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from

Copyright © 2015 American Society of Plant Biologists. All rights reserved.www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from

Copyright © 2015 American Society of Plant Biologists. All rights reserved.www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from

Copyright © 2015 American Society of Plant Biologists. All rights reserved.www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from

Copyright © 2015 American Society of Plant Biologists. All rights reserved.www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from

Copyright © 2015 American Society of Plant Biologists. All rights reserved.www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from

Copyright © 2015 American Society of Plant Biologists. All rights reserved.www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from

Copyright © 2015 American Society of Plant Biologists. All rights reserved.www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from

Copyright © 2015 American Society of Plant Biologists. All rights reserved.www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from

Copyright © 2015 American Society of Plant Biologists. All rights reserved.

a distinction can be made between (1) differentiation-induced PCD that occurs as final differentiation step inspecific cell types, for instance, in xylem tracheary ele-ments, the root cap, or the anther tapetum layer (Plackettet al., 2011; Bollhöner et al., 2012; Fendrych et al., 2014),and (2) age-induced PCD as the last step of organ se-nescence that occurs in all tissues of an organ or even theentire plant at the end of its life cycle (Thomas, 2013).Regarding ePCD, one of themost studied PCDprocessesoccurs during the hypersensitive response (HR), a local-ized plant response upon pathogen recognition (Collet al., 2011; Wu et al., 2014). Also abiotic stresses such asheat, UV radiation, or salt stress can lead to cell deathdisplaying certain hallmarks of PCD (Chen et al., 2009;Qiet al., 2011; Nawkar et al., 2013; Petrov et al., 2015).It is still unclear whether different PCD types in plants

share common regulatory mechanisms or if they arecontrolled by distinct pathways. Due to the scarcity ofmolecular information, most comparative analyses havebeen based on morphological and biochemical charac-teristics. Vacuolar cell death, defined by accumulation ofautophagosomes, vacuolar collapse, and corpse degra-dation, has been opposed to necrotic cell death, withswelling of mitochondria, protoplast shrinkage and un-processed cell corpses (vanDoorn et al., 2011). However,some types of PCD, including HR cell death, pollen self-incompatibility, or endosperm cell death, do not fall intoeither of these proposed classes (van Doorn et al., 2011).Here, we exploited publicly available genome-wide

transcriptome data that were associated with differentforms of cell death in the model plant Arabidopsis(Arabidopsis thaliana), with the aim to comparativelycharacterize plant PCD types. We identified distinct setsof differentially regulated genes in several develop-mental and environmental situations known to provokeplant cell death, suggesting that dPCD and ePCD pro-cesses are characterized by separate regulatory path-ways. Focusing on dPCD,we identified a conserved coreof transcriptionally controlled dPCD-associated genes.Based on this information, we created and analyzed anarray of promoter-reporter lines that are expressed incells preparing for different types of dPCD. The pre-sented data will be a powerful tool to complementmorphological analysis when attempting PCD discov-ery, recognition, and analysis of dPCD types in plants.

RESULTS

Meta-Analysis of Available ATH1 Data Sets RevealsDistinct Gene Expression Patterns Characterizing dPCDand ePCD

To get a viewon similarities anddifferences in the geneexpression profiles of different PCD types,we carried outa meta-analysis of Arabidopsis Affymetrix GeneChipGenome Array (ATH1) data sets. Based on their accom-panying experimental descriptions, we selected a total of59 ATH1 data sets associated with a range of generallyaccepted or hypothetical PCD contexts. For simplicity,we will refer to all of these contexts as PCD, though for

some of them, the actively controlled nature of the celldeath has not been unambiguously shown. From thiscompendium, we extracted 82 conditions, contrastingdifferent cell death situations with their correspondingnon-PCD controls (Table I; Supplemental Tables S1 andS2). We assigned these contrasts to nine categoriesbased on their experimental context. The dPCD categorydifferentiation-induced cell death contains experimentsdescribing specific cell types undergoing cell death as partof their differentiation program, while the senescence-induced cell death category comprises data sets pro-duced fromentire organs during late stages of senescence.In the ePCD categories, data sets produced from patho-gen assays (biotic stress), from plants experiencing oxi-dative stress, and from plants exposed to UV irradiation,genotoxic compounds, high or low temperatures, andosmotic and salt stresses were included. Finally, data setsfrom hormone treatments leading to cell death completethe list of putative PCD categories (Fig. 1).

To define the relatedness of the ATH1 data sets, in-dependent of predefined PCD categories, we performeda hierarchical clustering analysis (HCA) based on theexpression profiles of all genes that are differentiallyexpressed in at least one condition. Although the overallsimilarity of the entire compendium is low, it was foundto contain several functionally coherent clusters (Fig. 1).At a Pearson’s correlation distance threshold of 0.4, threeclusters of more than five conditions could be defined.The biotic stress clustermainly contains pathogen-relateddata sets but also contains some senescence, UV stress,and oxidative stress conditions. The osmotic stress clustercontains salt stress and osmotic stress conditions, and athird cluster indicates the tight relationship of most ge-notoxic stress conditions. At a more relaxed correlationdistance threshold, a fourth sizeable cluster emerges. Thiscluster, although containing more diverse expressionpatterns than the other three, is also functionally coher-ent, encompassing all differentiation-induced dPCDconditions alongwith two senescence-related conditions(Fig. 1, developmental cluster).

We compared the gene expression profiles of theconditions that fell in these four clusters and identifiedcommonly regulated genes within the clusters. In thedevelopmental cluster, we found SERINE CARBOXY-PEPTIDASE-LIKE48 (SCPL48), the aspartic proteasePASPA3, BIFUNCTIONAL NUCLEASE1 (BFN1), RIBO-NUCLEASE3 (RNS3), CALCIUM-DEPENDENT NU-CLEASE1 (CAN1), and a DOMAIN OF UNKNOWNFUNCTION679 MEMBRANE PROTEIN2 (DMP2) ofunknown function up-regulated in at least 10 out of 12conditions (Supplemental Table S3, developmental clus-ter). Additionally, 19 genes were found to be commonlyup-regulated in at least eight out of 12 conditions, in-cluding genes of families previously implicated in dif-ferent PCD processes, e.g., VACUOLAR PROCESSINGENZYMES (Hara-Nishimura and Hatsugai, 2011). Thebiotic cluster exhibits up-regulation of genes involved insalicylic acid (SA) and Ca2+ signaling: the SA-inducedgenes ENHANCED DISEASE SUSCEPTIBILITY5,PHYTOALEXIN DEFICIENT3, andWRKY DNA-BINDING

Plant Physiol. Vol. 169, 2015 2685

Transcriptome Meta-Analysis of Plant Cell Death

www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from Copyright © 2015 American Society of Plant Biologists. All rights reserved.

PROTEIN75; the calcium-binding protein-encoding geneIQ-MOTIF PROTEIN1; and AUTOINHIBITED CA2+

ATPASE12 (Supplemental Table S3, biotic cluster). Theseresults reflect the importance of calcium andSA signalingin the HR (Ma and Berkowitz, 2007;Mur et al., 2008). Theintegration of biotic stress conditions as well as senes-cence conditions in the biotic cluster suggests the acti-vation of conserved processes during biotic stress andsenescence conditions. In the osmotic cluster, 12 geneswere up-regulated in all 14 conditions of mannitol, salt,and cold stress treatments (Supplemental Table S3, os-motic cluster), including SENESCENCE-ASSOCIATEDGENE113 and several LATE EMBRYOGENESIS ABUN-DANT genes, which are known to be involved in cellularprotection and stress tolerance (Olvera-Carrillo et al.,2010; Candat et al., 2014). The genotoxic cluster com-prised DNA repair genes such as BREAST CANCERSUSCEPTIBILITY1, RAD51 (At5g20850), and two of itsparalogs, RAD17 and RAD21 (Trapp et al., 2011). Fur-thermore, nucleotide metabolism genes such as TSOMEANING UGLY IN CHINESE2 (TSO2, AT3G27060)

and THYMIDINE KINASE1A (Roa et al., 2009) and sev-eral plant-specific SIAMESE (SIM)/SIAMESE-RELATED(SMR) CYCLIN-DEPENDENT KINASE (CDK) inhibi-tors (Yi et al., 2014) were commonly up-regulated in thiscluster (Supplemental Table S3, genotoxic cluster).

In contrast to the observed correlation within each ofthe four clusters, there was little similarity between thegene expression profiles across the clusters. These re-sults indicate that distinct gene expression patternscharacterize different forms of PCD, in particulardifferentiation-induced dPCD and ePCD types. How-ever, which of the differentially expressed genes areeffectively involved in PCD regulation and which onesare elicited as part of processes other than PCD remainsto be investigated case by case.

Most dPCD-Regulated Genes Are Not Up-Regulated inePCD Situations

To test the hypothesis of distinct gene regulationoccurring in differentiation-induced dPCD and ePCD

Figure 1. PCD-associatedATH1 transcriptomedata sets group indistinct clusters.HCA showing theclustering of 82putativedPCDandePCDconditions based on the log-fold expression values of differentially regulated genes and indicating their affiliation to different putative PCDcategories (arrow). Four clusters are highlighted indicating the relatedness of data sets falling in the developmental, the biotic stress, the osmoticstress, and the genotoxic stress clusters. The color coding from blue to yellow indicates an increase in the Pearson’s correlation distance.

2686 Plant Physiol. Vol. 169, 2015

Olvera-Carrillo et al.

www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from Copyright © 2015 American Society of Plant Biologists. All rights reserved.

conditions, we applied two-dimensional clustering tothe expression profiles of a curated gene set, which con-tains the genes that are most commonly up-regulated inthe four clusters described above, as identified usingcustom R scripts (see “Materials and Methods” andSupplemental Table S3). The resulting gene clades mir-ror the original four clusters, and it again appears thereis little common regulation of these genes across clusters(Fig. 2).The genotoxic cluster appears most distinct; only few

genotoxic marker genes were up-regulated in the otherconditions. One example is the CDK-inhibitor SMR5(At1g07500) that is up-regulated in several salt andosmotic stress conditions (Fig. 2). Gene expressionprofiles in the osmotic stress and biotic stress clustershave larger overlaps; many biotic marker genes are up-regulated duringmannitol and salt treatments, and viceversa some osmotic marker genes are up-regulated as aconsequence of inoculation with the necrotrophicpathogen Botrytis cinerea (Fig. 2). The up-regulation ofdevelopmental marker genes is largely confined to thedifferentiation-induced dPCD data sets. Some genes,however, are also up-regulated in osmotic and saltstress conditions, suggesting a certain degree of com-mon gene regulation (see the lower tier of develop-mental marker genes in Fig. 2). Interestingly, conditionsof organ senescence lead to up-regulation of severalbiotic, osmotic, and developmental marker genes (Fig.2, arrow), suggesting that plant senescence activates acombination of pathways. Most developmental markergenes, however, are almost exclusively up-regulated indifferentiation-induced dPCD situations, suggestingthat the transcriptional regulation differs substantiallybetween these and ePCD contexts.

Supervised Classification of PCD Samples Based on TheirGene Expression Profiles Is Possible for Some PCD TypesBut Not for Others

Prompted by the results of the unsupervised clus-tering approaches in distinguishing PCD types (Figs.1 and 2), we attempted to classify the different putativePCD categories (Fig. 1) using supervised classificationalgorithms, based on their ATH1 expression profilesand the putative PCD class labels assigned to themfrom the experimental descriptions (see “Materials andMethods”). The aim of building such classifiers is toassess the feasibility of predicting the PCD type of anunlabeled experimental sample based on its gene ex-pression profile.We first built Support Vector Machine (SVM; Cortes

and Vapnik, 1995) and Random Forest (RF; Breiman,2001) classifiers distinguishing ePCD- from dPCD-related conditions, based on the expression profiles ofall genes. A moderate classification performance wasobtained on the full data set of ePCD and dPCD condi-tions (Supplemental Table S4). The performance in-creased markedly when excluding minority subclasses,i.e., senescence (for dPCD) and/or temperature stress,

UV stress, oxidative stress, and hormone treatments (forePCD). Using the curated set of putative PCD indicatorsfor the four major PCD clusters described above (Fig. 2;Supplemental Table S3 instead of all genes as classifica-tion features did not generally lead to improved classi-fication performance (Supplemental Table S4). Theseresults indicate that a clear molecular distinction ofePCD versus dPCD is hampered by expression similar-ities between certain subtypes of ePCD and dPCD. Inparticular, the expression profile similarities betweensenescence-induced dPCD and various ePCD conditions(see Fig. 2) appear to have a negative impact on thedPCD/ePCD classification performance (SupplementalTable S4). To investigate which PCD subtypes suffer themost from expression similarities with other subtypes,we attempted to classify particular PCD subtypesagainst all other types (Supplemental Table S4).Whereasthe maximum classification performance is high fordifferentiation-induced dPCD, genotoxic cell death, andosmotic cell death, the performance values for other PCDsubtypes are moderate to low, reflecting a lack of ade-quately distinctive expression signatures to separatethese poorly performing PCD subtypes from some of theother PCD types grouped together as the alternative la-bel set. Taken together, with the ATH1 data sets that arepublically available at this point, unconditionally dis-tinctive sets of marker genes are hard to find for manyPCD subtypes, even when using supervised classifica-tion strategies.

Identification of Unique dPCD Indicator Genes

The information that genes are predominantly up-regulated in differentiation-induced dPCD types openedthe possibility of testing some of these genes for theiraptitude as dPCD reporter genes. We took three com-plementary approaches to identify individual genes thatcould potentially be used as dPCD markers.

First, we compiled a list of genes that are significantlyup-regulated at least 2-fold in at least 60% of the dPCDdata sets in the ATH1 compendium described above.SCPL48 and PASPA3 show the highest frequency of up-regulation in all dPCD data sets (89% and 84%, re-spectively). TELOMERIC DNA-BINDING PROTEIN1(At5g13820) is up-regulated in 79%of all dPCD contrasts,and BFN1 is up-regulated in 74% of the contrasts. Ad-ditional commonly up-regulated genes inmore than 60%of all 19 dPCD contrasts include RNS3, CAN1, THIO-REDOXIN H-TYPE5, and three genes of unknownfunction (Supplemental Table S3, dPCD contrasts).

Second, we used the Genevestigator Condition Searchand Similarity Search tools (Hruz et al., 2008) to identifygenes that are commonly coregulated with BFN1,PASPA3, METACASPASE9 (MC9), and CYSTEINE EN-DOPEPTIDASE (CEP1), four genes that have been as-sociated with or functionally implicated in dPCD inseveral Arabidopsis cell types (Farage-Barhom et al.,2008; Helm et al., 2008; Ohashi-Ito et al., 2010; Bollhöneret al., 2013; Fendrych et al., 2014; Zhang et al., 2014).

Plant Physiol. Vol. 169, 2015 2687

Transcriptome Meta-Analysis of Plant Cell Death

www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from Copyright © 2015 American Society of Plant Biologists. All rights reserved.

Figure 2. Commonly up-regulated geneswithin PCD clusters are largely distinct between clusters. Two-dimensional clustering ofthe gene-conditionmatrix plotting the expression profiles of themost commonly regulated genes of the four clusters highlighted inFigure 1 over all conditions. The separate blocks of dark blue fields indicate that the regulation of commonly expressed markergenes is largely distinct for each cluster. The arrow indicates a cluster of senescence-related data sets that show up-regulation ofbiotic, osmotic, and developmental marker genes.

2688 Plant Physiol. Vol. 169, 2015www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from

Copyright © 2015 American Society of Plant Biologists. All rights reserved.

Seven genes were found to be commonly coregulatedwith these four target genes (Table II; Supplemental TableS5). Reiterating the analysis with these seven genes, weobtained a list of 154 genes that are coregulated with atleast two out of the seven genes (Supplemental Table S5).Of these genes, four are coregulated with all target genes:BFN1 andMC9, as well as RNS3 and DMP4 (At4g18425,a paralog of DMP2). An unknown gene that we dubbedEXITUS1 (EXI1; At2g14095) and the transcription factorANAC083 (for NO APICAL MERISTEM; ARABIDOPSISTRANSCRIPTION ACTIVATION FACTOR; CUP-SHAPED COTYLEDON [NAC]DOMAIN CONTAININGPROTEIN83) are coregulated with at least six of theseven target genes.Third, we constructed a list of genes potentially in-

volved in dPCD by comparing the gene expression pro-files of two root tissues that are known to execute dPCDas afinal differentiation step, the root cap (Fendrych et al.,2014), and the xylem tracheary elements (Bollhöner et al.,2012) with expression profiles of other tissues. Usingthe Visual Lateral Root Transcriptome Compendium(VLRTC; Parizot et al., 2010) based on a gene expressionatlas of the Arabidopsis root (Brady et al., 2007), wefound 95 genes commonly up-regulatedmore than 2-foldin xylem and lateral root cap (LRC) compared with roottissues not undergoing PCD (Supplemental Table S6).Eight of these genes are among the 154 genes identifiedbyGenevestigator as coexpressedwith at least two out ofseven genes in the target gene set, significantlymore thanexpected by chance (P = 1.5267e-06, hypergeometric test).Next to BFN1,MC9, PASPA3, SCPL48, and RNS3, a fattyacid desaturase family gene (At1g06090), the tran-scription factor ANAC046 (At3g04060), and SCPL20 arecommonly up-regulated, suggesting that these genes

might be involved in dPCD processes in the xylem andthe LRC.

Although the data sets used in the ATH1 meta-analysis and VLRTC approaches overlap to some ex-tentwith each other (the root cap data sets inVLRTCandthe meta-analysis are the same) and with the Geneves-tigator data, the different screening methodologies usedled to the identification of candidate reporter gene setsthat are only partially overlapping. By virtue of beingcommonly up-regulated in different differentiation-induced dPCD contexts, these genes can be consideredpotential dPCD reporters. To test the aptitude of thesegenes in this respect, we picked a set of 10 genes forin-depth characterization of their expression patterns:CEP1, PASPA3, BFN1, MC9, ANAC046, CAN1, RNS3,SCPL48, EXI1, and DMP4.

dPCD Reporters Are a Powerful Resource to DetectPutative dPCD Processes in Planta

The putative 59-regulatory regions (promoters) of theeight candidate dPCD reporter genes were cloned andfused to a Gal4 DNA binding domain fused to thetranscriptional activator domain of the herpes simplexvirus VP16 protein (GAL4-VP16) transcriptional activa-tor, combined with a GAL4-activated upstream activa-tion sequence (UAS) driving a nuclear-localized histone2A-GFP (..H2A-GFP) reporter gene. These lines can beused in a versatile manner: as marker lines to detect andanalyze PCD processes in planta, as driver lines to con-trol the transcription of transgenes in a PCD-specificspatial and temporal pattern, and as tools to sort GFP-tagged protoplasts or nuclei for tissue-specific -omicsanalyses.

Table I. Overview of the number of conditions profiled per PCD subcategory in the ATH1 compendium

PCD Category PCD Subcategory Tissue, Organ, and Stress Type No. of Conditions

Tracheary elements 4LRC 1

Differentiation-induced Endosperm 3Seed coat 2

Developmental (dPCD) Leaves 4Senescence-induced Petals 1

Sepals 1Siliques 1Mutant seedlings 2

Biotic stress-induced Fungal elicitor 12Bacterial elicitor 3Viral protein 1

Environmental (ePCD) Abiotic stress-induced Oxidative stress 11UV stress 5Genotoxic stress 8Heat stress 2Cold stress 3Osmotic stress 6Salt stress 7

Hormone treatment Ethylene 3SA 2

Total 82

Plant Physiol. Vol. 169, 2015 2689

Transcriptome Meta-Analysis of Plant Cell Death

www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from Copyright © 2015 American Society of Plant Biologists. All rights reserved.

In a first round, over a dozen independent lines perpromoter-reporter construct were investigated in T2,and lines with a single transfer DNA insertion locus anda representative GFP expression pattern were selectedfor in-depth analysis in T3. As the pEXI1 ..H2A-GFP,the pANAC046..H2A-GFP and the pCAN1..H2A-GFPconstructs conferred weak or inconsistent GFP signals,these lines were not included for further analysis. FromCEP1, PASPA3, MC9, and BFN1, which have been pre-viously reported as PCD-associated (Farage-Barhomet al., 2008; Helm et al., 2008; Bollhöner et al., 2013;Fendrych et al., 2014), we chose to display the ex-pression patterns conferred by pCEP1 and pPASPA3 asa reference for the expression pattern of the remaininggenes.

We focused our expression analysis on Arabidopsistissues or cell types known to undergo differentiation-induced dPCD: the tapetum layer in the developinganther, the protoxylem cells in the growing root, and thecells of the LRC. Cells are also dying in the central en-dosperm and in senescing petals, butmuch less is knownabout the nature of the cell death in these tissues (forreview, see Van Hautegem et al., 2015). We exploited atonoplast integrity marker (ToIM; Fendrych et al., 2014)to investigate vacuolar collapse, a hallmark of vacuolarPCD (van Doorn et al., 2011). In all tissues or cell types,the ToIM expression controlled by the pPASPA3 pro-moter shows that vacuolar collapse precedes cell deathinPASPA3-expressing cells (Fig. 3). In petals and the rootcap, we additionally performed whole-mount terminaldeoxynucleotidyl transferase dUTP nick-end labeling(TUNEL) assays, indicating that DNA fragmentationoccurs in these tissues in the stages investigated for theexpression pattern of the promoter-reporter constructs(Fig. 3).

The promoters of RNS3, PASPA3, and DMP4 con-ferred largely similar expression patterns in the degen-erating endosperm from torpedo stage onwards, in theanther tapetum layer before tapetum cell death, in dif-ferentiating LRC cells and tracheary elements, and insenescing petals (Fig. 4; Supplemental Figs. S1–S3). Inaccordance with the ATH1-derived expression data, thepRNS3 promoter conferred the strongest GFP expres-sion, while pDMP4..H2A-GFP produced weaker GFPsignals. Note that very high expression levels led to afailure to contain the H2A-GFP protein in the nucleus.Similar to pPASPA3, pRNS3 is activated many hoursbefore PCD in the LRC, leading to a broader expression

pattern compared with the one conferred by pDMP4,which only activated H2A-GFP expression shortly be-fore PCD, leading to a narrower expression pattern in theLRC (Fig. 4). In developing petals of pRNS3, pPASPA3,and pDMP4 reporter lines, expressionwas first restrictedto the tracheary elements, while expression spreadthroughout the entire organ during petal senescence(Supplemental Fig. S1). During anther development,pPASPA3 activation was confined to the differentiatingtapetum layer, while both pRNS3 and pDMP4 were ac-tive in the outer layers of the anthers in later stages offlower development (Supplemental Fig. S2). In devel-oping seeds, both pDMP4 and pPASPA3 exclusivelyconferred expression in differentiating endosperm fromthe torpedo stage onwards, while pRNS3 ..H2A-GFPsignals were also detected in the differentiating seed coatof later seed stages (Supplemental Fig. S3).

Compared with pRNS3, pPASPA3, and pDMP4, thepSCPL48 promoter conferred a broader spatial andtemporal expression pattern; it was, for instance, al-ready expressed in petals at anthesis (Supplemental Fig.S1) and in the entire LRC, and not only confined totracheary elements, but also expressed in their neigh-boring cells (Fig. 4). In developing anthers, pSCPL48activity was not confined to the tapetum but spread tothe outer anther layers (Supplemental Fig. S3). Duringseed development, pSCPL48 was not activated in theendosperm but strongly up-regulated in the differentlayers of the differentiating seed coat (SupplementalFig. S3).

Finally, the pCEP1 ..H2A-GFP expression patternwas confined to the dying LRC cells close to the root tip(Fig. 4), in accordance with earlier reports (Helm et al.,2008). Additionally, pCEP1 ..H2A-GFP conveys astrong expression in epidermal cells in the root hairzone, though these are not known to undergo cell death(data not shown). Interestingly, the close CEP1 homo-log CEP2 is highly expressed in LRC cells in the roottransition zone (Hierl et al., 2014) and might take overCEP1 functions here. In the developing seed, pCEP1 isactive in the embryonic suspensor during early embryodevelopment (data not shown) and is present in laterstages both in the seed coat and the differentiating en-dosperm (Supplemental Fig. S3).

In summary, most promoter-reporter lines are spe-cifically expressed in differentiating cells known toundergo dPCD or are associated with cellular degra-dation events that are thus far not well defined. Not all

Table II. Commonly coexpressed genes of BFN1, MC9, PASPA3, and CEP1

Coexpression scores as calculated by Genevestigator. AGI, Arabidopsis Genome Initiative gene code.

AGI Gene Name Coexpression Score BFN1 Coexpression Score MC9 Coexpression Score PASPA3 Coexpression Score CEP1

AT5G04200 MC9 0.7951 1 0.6195 0.6754AT4G18550 DSEL 0.6771 0.649 0.7146 0.6622AT4G18425 DMP4 0.8552 0.8989 0.7395 0.7205AT4G04460 PASPA3 0.6586 0.668 1 0.6105AT1G11190 BFN1 1 0.882 0.7808 0.7112AT2G14095 EXI1 0.837 0.7802 0.7406 0.6818AT1G26820 RNS3 0.7672 0.9078 0.6587 0.6031

2690 Plant Physiol. Vol. 169, 2015

Olvera-Carrillo et al.

www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from Copyright © 2015 American Society of Plant Biologists. All rights reserved.

Figure 3. Cell death processes occur in different developmental contexts. AToIM combines expression of a freeGFPaccumulating inthe cytoplasm and the nucleus (green) and of a vacuolar-localized tagRFP (a monomeric derivative of a red fluorescent protein fromEntacmaea quadricolor ; red). Vacuolar rupture is indicated by the loss of compartmentalization and the merging of the two fluo-rescent signals. Note that in some dPCD cases, cytoplasmic acidification dampens the GFP signal, making the tagRFP signal moreprominent. The ToIM is expressed under the control of the pPASPA3 promoter. A, Time lapse imaging of dPCD in a protoxylemelement. The arrowheads indicate the cytoplasm around the cell’s nucleus, which is invaded by tagRFP upon vacuolar rupture(asterisk). B, Time lapse imaging of dPCD in a root cap cell. The arrowheads indicate the cell with intact vacuole, while the asteriskmarks the cell once vacuolar rupture has occurred. C, Time lapse imaging of dPCD in petal cells at the base of a petal. The ar-rowheads indicate the cell with intact vacuole, while the asterisk marks the cell once vacuolar rupture has occurred. D to F,Vibratome sections through developing anthers around the time point of tapetum dPCD. D shows a locule lined by pPASPA3::ToIM-expressing, viable tapetum cells. E shows a locule in which dPCD is ongoing; the arrowheads point at partly degenerated cells.F shows a locule after tapetum dPCD in which degraded remains of tapetum cells line the inside of the locule. G and H, Vibratomesections through a seed in the walking stick state of embryo (em) development. H is a detail of G. Arrowheads point at ToIM-expressing but intact endospermcells, while the asterisks indicate cells in the process of degeneration. I to K, TUNEL of whole-mountpetals and root tips. 49,6-diamidino-2-phenylindole (DAPI) staining is shown in red, and TUNEL signal is shown in green. I, Ar-rowheads indicate dying or dead TUNEL-positive root cap cells. J, Arrowheads indicate two fields of TUNEL-positive petal cells. K,TUNEL-positive control treated with DNase to induce tissue-wide DNA fragmentation, showing the overlap of TUNEL and DAPIsignals. In A to C, time is indicated in minutes. Bars = 50 mm (A–C, G, and I–K) and 20 mm (D–F and H).

Plant Physiol. Vol. 169, 2015 2691

Transcriptome Meta-Analysis of Plant Cell Death

www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from Copyright © 2015 American Society of Plant Biologists. All rights reserved.

T promoter-reporter lines are present in every dPCDprocess, and not all gene expression patterns are re-stricted to cells preparing for dPCD, but the combina-tion of these marker lines provides a powerful tool toidentify and analyze putative dPCD processes in a de-velopmental context in vivo and in planta.

dPCD Indicators Are Not Up-Regulated by Biotic andAbiotic Stresses Causing Cell Death

Our meta-analysis indicated that largely nonoverlap-ping sets of genes are up-regulated in differentiation-induced dPCD and various abiotic stress-induced ePCDtypes (Fig. 2), indicating that distinct transcriptionalprograms are activated in these plant cell death types. Tofurther test this hypothesis experimentally, we analyzeddPCD marker expression in roots of Arabidopsis seed-lings upon a variety of abiotic stresses. Propidium iodide(PI), which only enters cells with compromised plasmamembrane integrity (Truernit and Haseloff, 2008), wasused to highlight dead and dying cells. We investi-gated three marker constructs, pSCPL48, pRNS3, andpPASPA3, which showed a specific expression pattern inthe LRC of the control root tips. Upon treatments withhydroxyurea, bleomycin, UV-B irradiation, hydrogenperoxide, and NaCl, increasing numbers of PI-positivecells indicated the occurrence of cell death during thedifferent stress treatments (Fig. 5, arrowheads). Whilegenotoxic andUV stress led to localized cell death of rootmeristem cells, oxidative and salt stress produced morewidespread cell death. In all cases, cell death was neitherpreceded by ectopic dPCD marker expression at earlytime points nor accompanied by dPCD marker expres-sion at late time points. However, whether the observedcell death is a result of PCD programs activated by thestress treatments or is caused by direct cellular damageis difficult to ascertain. We performed whole-mountTUNEL and found that hydrogen peroxide treatmentleads to TUNEL-positive root cells (Supplemental Fig.S4). All other stress treatments did not lead to clearlyTUNEL-positive cells, apart from the dying root cap cellsthat are TUNEL positive due to stress-independentdPCD (Supplemental Fig. S4). These results confirmthat the stresses used to produce the ATH1 data setsmeta-analyzed in our study were sufficient to cause celldeath, but they leave open whether this cell death is anactive PCD or a passive, unregulated form of cell death.Although abiotic stresses have been shown to provokecell death displaying hallmarks of PCD (Chen et al.,2009; Qi et al., 2011; Nawkar et al., 2013; Petrov et al.,2015), detailed case-by-case investigations have to showif genuine actively controlled, genetically encoded pro-grams are responsible for these types of cell death.

Although there appears to be an overlap between thegenes up-regulated during abiotic stress-induced celldeath and pathogen-induced ePCD, our meta-analysissuggested that pathogen-related ePCD and differentiation-induced dPCD are regulated largely independently(Fig. 2). To confirm these results experimentally, we

Figure 4. Selected promoter-reporter lines highlighting cells preparingfor dPCD. PASPA3, RNS3, SCPL48, CEP1, and DMP4 expression pat-terns in developing seeds, developing anthers, the root cap, the xylem,and senescing petals (columns from left to right). pPASPA3..H2A-GFPis expressed in the embryo-surrounding region of the endosperm fromtorpedo stage onwards in the tapetum layer of the anther, in the LRCand the xylem, and in mature petals nearing floral organ senescence.pRNS3 ..H2A-GFP shows a very similar pattern. The SCPL48 pro-moter confers a broader spatial and temporal expression pattern and isnot only restricted to cells preparing for dPCD. pCEP1 ..H2A-GFPshows GFP expression in the endosperm and seed coat of developingseeds, the tapetum and its surrounding anther tissues, cells from thelowest tier of the LRC, differentiating xylem vessels, and the agingpetals. pDMP4 ..H2A-GFP is again more similar in expression topPASPA3 and pRNS3. TE, Tracheary elements. Bar = 50 mm.

2692 Plant Physiol. Vol. 169, 2015

Olvera-Carrillo et al.

www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from Copyright © 2015 American Society of Plant Biologists. All rights reserved.

performed a quantitative reverse transcription (qRT)-PCRexperiment of plants inoculated with an HR-inducingPseudomonas syringae strain. In contrast to HR markergenes, none of the canonical dPCD marker gene tran-scripts were significantly up-regulated during HR (Fig. 6;Supplemental Fig. S5). To test if only individual cells

express dPCD reporter genes, whichmight not register ona tissue-wide scale of RNA quantification, we also inves-tigated promoter-reporter lines but did not find any GFPsignals in or around HR lesions (data not shown). Theseresults confirm that dPCD marker genes are not tran-scriptionally regulated during HR-related ePCD.

Core dPCD Marker Genes Are Evolutionary Conserved inLand Plants

The phenomenon of developmentally regulated PCDis most likely evolutionary ancient and occurs also insimple land plants, for instance the moss Physcomitrellapatens (Xu et al., 2014). To assess the degree towhich themolecular regulation of dPCD might be evolutionaryconserved, we investigated the conservation of thedPCD indicator genes identified in Arabidopsis withinthe plant kingdom as well as between plants and ver-tebrates. According to the plant comparative genomicsplatform PLAZA (Proost et al., 2015), RNS3, BFN1,PASPA3, MC9, and SCPL48 are widely conserved inthe green plant lineage, while BFN1 appears to be re-stricted to the land plant lineage (Table III). Using thecomparative online tool Phytozome (Goodstein et al.,2012), we identified putative homologs of these dPCDmarkers in different angiosperm lineages, as well as inthe basal angiosperm Amborella trichopoda and in thelower land plants P. patens and Selaginella moellendorffii.In the green alga Chlamydomonas reinhardtii, we identi-fied protein sequences related to SCPL48, PASPA3, andRNS3, sequences with limited blast length for MC9 butno clear homolog for BFN1 (Supplemental Table S7).Outside the plant kingdom, the HomoloGene algo-rithm (Sayers et al., 2012) indicated conservation ofRNS3, SCPL48, and PASPA3 in all eukaryotes, whileBFN1 and MC9 appeared not to be conserved betweenplants and vertebrates (Table III). Interestingly, a pu-tative RNS3 homolog, the RNase T2, has recently beenimplicated in the control of melanocyte apoptosis viathe tumor necrosis factor receptor-associated factor2pathway in vitiligo patients (Wang et al., 2014). Fur-thermore, the putative PASPA3 homolog Cathepsin Dfunctions as a proapoptotic gene targeting Bid afterrelease from the lysosome (Appelqvist et al., 2012;Repnik et al., 2014).

These results suggest a high degree of conservation ofcore dPCDmarker geneswithin the green plant lineage.Whether the proapoptotic roles of PASPA3- and RNS3-related enzymes in mammals is due to functional con-servation or due to convergent evolution is difficult todetermine. Nevertheless, it is tempting to speculate thatsimilar mechanisms are functional in both animal andplant PCD types.

DISCUSSION

To date, despite the undisputed importance of thediverse forms of plant PCD for development and forenvironmental interactions (Wu et al., 2014; Petrov

Figure 5. Abiotic stress treatments cause cell death without the up-regulation of dPCD reporters. Abiotic stress treatments applied to 5-d-old seedlings from dPCDmarkers SCPL48, RNS3, and PASPA3. Pictureswere taken after the indicated time points and treatments at the root tipto show the expression around the LRC and were stained with PI tohighlight the cell walls and cells with compromised plasma membraneintegrity indicative of cell death (arrowheads). BM, Bleomycin; HU,hydroxyurea.

Plant Physiol. Vol. 169, 2015 2693

Transcriptome Meta-Analysis of Plant Cell Death

www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from Copyright © 2015 American Society of Plant Biologists. All rights reserved.

et al., 2015; Van Hautegem et al., 2015), still only little isknown about the molecular regulation of these pro-cesses. During the plant life cycle, PCD is induced atnumerous occasions, but it is unclear whether there arecommon mechanisms involved in controlling differentPCD types. Attempts to characterize and relate differ-ent plant PCD types have been made, based chiefly onmorphological and ultrastructural features of dyingcells (van Doorn, 2011; van Doorn et al., 2011). Here, weexplored the possibility to characterize different typesof plant PCD using molecular information. As tran-scriptional regulation has been implicated in plantPCD control (Van Hautegem et al., 2015), and as so faronly scarce proteomic data in plant PCD contexts exist,this transcriptome meta-analysis is a first step into asystematic molecular characterization of plant PCDprocesses.

One aim of our study was to investigate whetherexisting transcriptome data might be useful for a mo-lecular categorization of plant PCD types. By compar-ing transcriptome profiles of different developmentalstages and environmental stresses leading to cell death,we expected to find similarities and differences thatwould allow relating different PCD types based on thedegree of common gene regulation. Such informationcould be used to complement PCD characterizationbased on morphological and biochemical hallmarks(vanDoorn, 2011; van Doorn et al., 2011). Our approachof exploiting publicly available ATH1 data sets bymeans of several bioinformatics approaches was suc-cessful in identifying unique dPCD indicator genes.Promoter-reporter constructs of these genes markedcells preparing for cell death in well-defined PCD set-tings, e.g., the xylem (Bollhöner et al., 2012), the root cap

(Fendrych et al., 2014), or the tapetum (Plackett et al.,2011), but also highlighted cell types in which so faronly scarce genetic evidence exists for the occurrence ofPCD, e.g., the seed coat (Haughn andChaudhury, 2005)or the endosperm (Waters et al., 2013) in developingseeds. These results suggest that a conserved core ofPCD-associated genes is commonly regulated in di-verse dPCD contexts, and our findings will give im-pulses to investigate developmentally regulated PCDprocesses in more detail.

Among the genes that we found to be transcriptionallyregulated during differentiation-induced dPCD wereseveral genes encoding nucleases, including CAN1,BFN1, and RNS3. BFN1 is a well-known leaf senescencereporter, which has also been shown to function inchromatin breakdown during root cap PCD in Arabi-dopsis and tracheary element PCD in Zinnia elegans (Itoand Fukuda, 2002; Fendrych et al., 2014). CAN1 is astaphylococcal-like plasma membrane-bound nucleasewhose expression has been associated with PCD eventsbefore (Le!sniewicz et al., 2012), but its exact role re-mains unclear. RNS3 belongs to the evolutionary con-served family of T2 endoribonucleases that cleavesingle-stranded RNA. T2 endoribonucleases have beensuggested to perform a variety of functions, includingscavenging of nucleic acids, degradation of self-RNA,modulating host immune responses, and serving as cel-lular cytotoxins (Luhtala and Parker, 2010). In plants, T2ribonucleases are induced during phosphate starvationand have been hypothesized to function in providingphosphates from nucleic acids (Taylor et al., 1993; Bariolaet al., 1994). Our results showRNS3 up-regulation duringleaf and floral organ senescence, correlating senescence-induced cell death with differentiation-induced dPCD.

Figure 6. dPCD marker genes are not up-regulated during HR PCD. qRT-PCR of Col-0 wild-type plants inoculated with anavirulent HR-inducing P. syringae strain in a time course experiment after infection. Relative expression of the indicated genesboth in the inoculated area and in noninoculated tissue was determined by qRT-PCR at the indicated time points. PATHO-GENESIS RELATED1 (PR1), MC1, and MYB DOMAIN PROTEIN30 (MYB30) were used as HR marker genes. Expression valueswere normalized using the SAND family gene as internal standard. Ratios of the expression values for each gene in the inoculatedzone with respect to the noninoculated area are presented for each time point. Mean and SE of the mean values were calculatedfrom three independent experimentswith three replicates. Statistical significance according to a Student’s t test P value of 0.005 isindicated by asterisks. hpi, Hours after inoculation; a.u., arbitrary units.

2694 Plant Physiol. Vol. 169, 2015

Olvera-Carrillo et al.

www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from Copyright © 2015 American Society of Plant Biologists. All rights reserved.

Next to nucleases, also protein hydrolases includingPASPA3, MC9, SCPL48, and CEP1 are among the coreof dPCD-associated proteins. For most of these prote-ases, little data exist regarding their actual function andsubstrates. The degradome of MC9 has been investi-gated in detail, thoughmost substrates identified rathersuggested functions other than PCD (Tsiatsiani et al.,2013). Nevertheless,MC9 has been shown to be a part ofa proteolytic cascade effecting postmortem cell clear-ance of tracheary elements in Arabidopsis (Bollhöneret al., 2013). Recently, MC9 activity was found to be im-portant to mediate oxidative stress-dependent cell deathvia the cleavage of GRIM REAPER (Wrzaczek et al.,2015).Next to hydrolytic enzymes, genes encoding several

proteins of unknown functions such as the plasmamembrane-localized DMP4 were up-regulated duringseveral dPCD processes. DMPs represent a uniquefamily of plant-specific plasma membrane proteins ofunknown function that have been recently identified ina screen for senescence-associated genes in Arabidopsis(Kasaras and Kunze, 2010). Of the 10 Arabidopsis DMPparalogs, DMP4 is coregulated with the core of dPCDmarker genes and up-regulated in several dPCD con-ditions. Additionally, expression of DMP4 has beendescribed in abscission zones of floral organs (Kasarasand Kunze, 2010), although involvement of PCD in thisabscission process has not been investigated. The mo-lecular function of DMP proteins still needs to be de-termined, but misexpression of Arabidopsis DMP1 ledto an aberrant endoplasmic reticulum and in some casesto the death of transfected cells (Kasaras et al., 2012).Despite the fact that different PCD types appear to

exhibit different gene expression profiles, an adequatesupervised classification of PCD types based on theavailable transcriptome data proved to be possible onlyfor some PCD subtypes, possibly due to the nature andquantity of the available transcriptome data sets. Mostdata sets analyzed were not explicitly designed to

characterize gene expression changes associated withcell death processes but rather to identify regulators ofprocesses that precede or even might counteract celldeath. What is more, deducing from the experimentalmetadata which particular PCD subtype, if any, isrepresented in a given data set is not always straight-forward.

To reliably classify different types of plant PCD, amore thorough understanding of their molecular reg-ulation will be necessary. A means to this end will bethe generation of specific transcriptome profiles ofprecisely described PCD systems. For differentiation-induced dPCD, this is a challenging task, as only sin-gle cells, or small groups of cells, are undergoing celldeath at a time. Techniques of isolating these cells ortheir nuclei for transcriptome analysis by fluorescent-associated cell sorting or isolation of nuclei tagged inspecific cell types (Deal and Henikoff, 2011) will be in-strumental to obtain meaningful data sets. The dPCDpromoter-reporter constructs presented in this studywill facilitate these approaches. At least for closelyrelated PCD types, for instance, different forms ofdifferentiation-induced dPCD, such a comparative ap-proach will become valuable to reveal unique PCDmarkers and putative core PCD regulators. This ap-proach, accompanied by thorough morphological,molecular genetics and cell biological analyses, willopen the way to a more comprehensive understandingof PCD as a fundamental cellular process in plants.

CONCLUSION

Despite the progress achieved over the last decade bya relatively small research community dedicated toplant PCD, the molecular regulation of PCD largelyremains a terra incognita. To fill the white spots on themap, and to relate the findings made in different plantPCD systems, we need to understand more of the

Table III. Evolutionary conservation of putative core dPCD markers

% ID reflects the sequence similarity between the Arabidopsis gene and the best blast hit in vertebrates. All genes belong to PLAZA 3.0 orthologousgene families that encompass the Magnoliophyta (RNS3, BFN1, and MC9) or Viridiplantae (SCPL48 and PASPA3) clade (not shown). AGI, Arabi-dopsis Genome Initiative gene code; HOM, homologous; Nuclease PA3-like, a predicted protease (GenBank accession number XP_005974529.1);CPVL, carboxypeptidase, vitellogenic-like.

AGI (Name)PLAZA 3.0 Homologous

Gene Family

PLAZA 3.0 Plant

Clade of HOM FamilyHomoloGene Blast Hits in Vertebrates % ID Role in Vertebrates

AT1G26820 (RNS3) HOM03D000496 Viridiplantae 31190, conservedin Eukaryota

RNase T2 30 Potentially skincell apoptosis(Wang et al.,2014)

AT1G11190 (BFN1) HOM03D001490 Embryophyta No information Nuclease PA3-like 27AT5G04200 (MC9) HOM03D001276 Viridiplantae No information No clear hits 0AT3G45010 (SCPL48) HOM03D000050 Viridiplantae 137548, conserved

in EukaryotaSerine CPVL 30

AT1G62290 (PASPA3) HOM03D000729 Viridiplantae 124002, conservedin Eukaryota

Cathepsin D 50 Proapoptoticgene (Appelqvistet al., 2012;Repnik et al.,2014)

Plant Physiol. Vol. 169, 2015 2695

Transcriptome Meta-Analysis of Plant Cell Death

www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from Copyright © 2015 American Society of Plant Biologists. All rights reserved.

molecular principles that govern plant cell death. Here,we made a step toward a comprehensive understand-ing of plant PCD by integrating genome-wide tran-scriptome profiles of different established as well as lesswell-known cell death systems. With the recognizedexpression patterns and dPCD reporter lines, we cannow progress to a more specific mode of analysis. Ofcourse, transcriptional regulation is only a fraction ofthe molecular control that leads to an ordered andtimely termination of vital processes of a cell under-going PCD. A challenge for the next decade will be todefine themolecular modes of function of putative PCDregulators and the posttranslational modifications thatlead to the rapid execution of cell death observed inmany systems.

MATERIALS AND METHODS

Meta-Analysis Data Retrieval

We retrieved Affymetrix ATH1 CEL files for the various transcriptome data

sets from ArrayExpress (http://www.ebi.ac.uk/arrayexpress/), Gene Ex-

pression Omnibus (http://www.ncbi.nlm.nih.gov/geo/), Riken Expression

Array Database (http://read.gsc.riken.go.jp/), or other third-party data pro-

viders. Multiple rounds of preanalysis processing steps (data curation and fil-

tering to remove conditions with little or no differential gene expression) were

performed to retain an optimal selection of expression data sets.

Meta-Analysis Detection of Differential Expression

Themicroarray datawere preprocessedwith the RobustMultiarrayAverage

procedure, as implemented in BioConductor (Irizarry et al., 2003; Gentleman

et al., 2004). An up-to-date Chip Defnition File based on the latest version of the

Arabidopsis genome annotation by The Arabidopsis Information Resource was

retrieved from BrainArray (http://brainarray.mbni.med.umich.edu) to define

probe-gene relations. A filtering of differentially expressed genes was per-

formed using the R/Bioconductor software package Limma (Ritchie et al., 2015)

to retain only those genes with an adjusted P # 0.05 and absolute log2 fold

change . 1.

HCA

The expression profiles of theATH1 genes showingdifferential expression in

at least one PCD condition were hierarchically clustered with the Orange

Canvas software (http://orange.biolab.si/) using Pearson’s correlation dis-

tance as the distance measure and the average linkage clustering option. To

identify the most commonly up-regulated genes in particular PCD clusters, we

used an R script that, given a cluster of interest, ranks genes according to the

number of conditions in the cluster in which they are significantly up-regulated

(P , 0.05) at least 2-fold. For each cluster, the resulting ranked gene list was

truncated at a specific number of observed up-regulations to obtain lists for all

clusters of 25 to 30 genes each (Supplemental Table S3). A similar analysis was

done to identify genes that are commonly up-regulated across all conditions

labeled as dPCD. A gene was considered to be commonly up-regulated in

dPCD when it was designated as significantly up-regulated (P , 0.05) at least

2-fold in 60% of the dPCD-labeled conditions.

Supervised Classification Analyses

SVM (Cortes and Vapnik, 1995) and RF (Breiman, 2001) analyses were

performed using the Orange toolbox (Demšar et al., 2013) by writing Python

scripts accessing the Orange API. In each analysis, an automated exhaustive

search of the algorithm parameter space was performed to optimize the pa-

rameter settings. These settings are reported per analysis in Supplemental Table

S4. Comparison of the classification performance across analyses and algo-

rithms was done by means of the Matthews Correlation Coefficient (MCC) as

reported after 5- or 10-fold cross validation (5-fold cross validation was used

when the number of contrasts in one of the classes was ,10). The MCC is a

balanced measure of binary classification performance that is particularly

useful if the classes are of different sizes. MCC scores range from 1 for perfect

classifiers to –1 when there is a total disagreement between the predicted and

observed class labels, with a score of 0 indicating that the classifier does not

perform better than random.

For the analyses on balanced dPCD- and ePCD-labeled data, 19 (or 10) ePCD

experiments were randomly sampled without replacement out of the relevant

ePCD subset and added to the 19 (or 10) dPCD experiments, after which SVM

and RF classifiers were learned. This random selection was performed

100 times, and the average MCC score is reported in Supplemental Table S4.

Genevestigator Coexpression Tool Search

The query genes were screened with the Conditions Search and Similarity

Search tools of Genevestigator. To find the relevant conditions that induce the

expression of the query gene, all ATH1microarrayswere given as input into the

Conditions search (Perturbation tool) and filtered by selecting microarrays

showing a log-fold change of the query gene greater than or equal to 2 and

a P value greater than or equal to 0.01. The resulting microarrays were saved

in a new list and fed in the Coexpression tool to find the top 200 positively

correlated genes in the Perturbation option. To identify commonly coregulated

genes between different genes, coregulated genes with a Genevestigator

score (Pearson’s correlation coefficient) greater than 0.6 were selected for

each gene. The resulting gene lists were fed into the online Venn Diagram tool

program provided by the VIB-Ghent University Bioinformatics and Systems

Biology laboratory at http://bioinformatics.psb.ugent.be/cgi-bin/liste/

Venn/calculate_venn.htpl to identify commonly regulated genes.

Identification of Genes Coregulated in Maturing Xylemand LRC

The VLRTC method (Parizot et al., 2010) was used to reanalyze the data

from Brady et al. (2007) as described in Fendrych et al. (2014). The candidate

genes were first thresholded for their expression in the LRC and in thematuring

tracheary elements as follows:

TRUE  if !

EXPLRC$ 2 averageEXPrest

"

AND!

EXPXM$ 2 averageEXPrest

"

And these genes were further ranked according to:

rank ¼ ðaveragefEXPLRC;EXPXMgÞ=ðMAXfEXPrestgÞ

MAX refers to the maximum expression value; EXP to normalized expression

values, rest to {Stele (wol), Stele(J2501),Protophloem(S32), Phloem+Companion

Cells(APL), PhloemCompanionCells (SUC2),DevellopingXylem(S4), Pericycle

(J2661), Pericycle Phloem Pole (S17), Pericycle Xylem Pole (JO121), Primordia

(rm1000), Ground Tissues (J0571), Endodermis (scr5), CORTEX, Epidermis

Atrichoblast (gl2), and Epidermis Trichoblast (COBL9)}; and XM to xylem

maturing.

Plant Material and Growth Conditions

For the root imaging, seedlingsweregrownvertically 5dafter sowingonone-

half-strength Murashige and Skoog (MS) plates (2.15 g L–1 MS salts [Caisson

Labs], 0.1 g L–1 MES [Sigma], pH 5.8 [KOH], and 0.8% [w/v] agar [Lab M]) in a

16-h-light/ 8-h-dark photoperiod at 21°C with 70% humidity. For the imaging

of anthers, petals, and developing seeds, 5-week-old plants were grown in jiffy

pots in a 16-h-light/8-h-dark photoperiod at 21°C and kept under optimal

irrigation and nutrient supply conditions throughout the plant life cycle.

Stress Treatments

Three biological replicates of 5-d-old seedlings from each of the marker lines

analyzed were transferred from one-half-strength MS plates to one-half-

strength MS plates containing 0, 140, and 250 mM NaCl (VWR), 5 and 20 mM

hydrogen peroxide (Merck), 5 mM Hydroxyurea (Sigma), and 0.6 ug mL–1

Bleomycin (Duchefa) for the indicated times before confocal imaging. For UV

stress, the seedlings were UV-B treated for 15 min, 30 min, 45 min and 1 h with

UV-B 313 EL lamps (Q-Lab) at an intensity of 1 W m–2 measured with the

Spectrasense 2+ meter coupled to the compatible UV-B sensor (Skye

2696 Plant Physiol. Vol. 169, 2015

Olvera-Carrillo et al.

www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from Copyright © 2015 American Society of Plant Biologists. All rights reserved.

Instruments). The UV-B lamps were in an incubator where the conditions were

18°C and 67% relative humidity.

Cloning and Transgenic Lines Preparation

The proCEP1 and proPASPA3were obtained as Gateway cloning-compatible

amplicons from the systematic analysis of Arabidopsis promoters collection

(Benhamed et al., 2008) and were recombined into the pDONRP4P1r vector

(Invitrogen). The proCEP1 spans 1,626 bp, and the proPASPA3 spans 1,997 bp

upstream of the respective start codon. The proSCPL48, proRNS3, and proDMP-4

were isolated from Arabidopsis (Arabidopsis thaliana) ecotype Columbia (Col-0)

genomic DNAusing gene-specific primers (Supplemental Table S8) and adding

BamHI and XhoI restriction sites to clone directionally into pENTRL4-R1,

a Gateway-compatible entry vector containing a cassette with a multiple

cloning site (https://gateway.psb.ugent.be/). The proSCPL48 spans 2,054 bp

(including the first 24 bp after the start codon), proRNS3 spans 1,440 bp, and

proDMP4 spans 1,352 bp upstream of the respective start codon. Sequence

information about these genes can be found in The Arabidopsis Information

Resource under the following accession numbers: BFN1 (Atg11190), CEP1

(At5g50260), PASPA3 (At4g04460), SCPL48 (At3g45010), RNS3 (At1g26820),

andDMP-4 (At4g18425). The promoters were assembled in amultisite Gateway

reaction using LR clonase II+ (Invitrogen) with the GAL4 coding sequence and

the destination vector pB9-H2A-UAS-7m24GW to create activator lines. These

lines can be used for transactivation, and at the same time, the nuclei of the cells

where the promoter is expressed are marked with GFP. This vector contains a

HISTONE 2A-6 (H2A) coding sequence (At5g59870) fused to eGFP and driven

by the repetitive UAS promoter. This vector is part of a transactivation driver

line-effector line set as described (Karimi et al., 2005).

The expression clones obtained were transformed into Agrobacterium tume-

faciens C58C1 (pMP90)-competent cells using electroporation, and these bac-

teria were used for a modified floral dip method to stably transform

Arabidopsis Col-0 plants. One milliliter of Yeast Extract Broth-grown culture

was incubated 6 h at 28°C, and 10 mL of Yeast Extract Broth was added and

grown overnight at 28°C. Plants were dipped with the overnight culture,

adding 40 mL of floral dip medium (10% [w/v] Suc and 0.05% [v/v] Silwet

L-77). All analyses were performed with T3 homozygous plants with a single-

locus insertion determined by segregation analysis.

Confocal Imaging and Image Processing

Confocal images were acquired using a Zeiss 710 CLSM microscope.

Objectives used were Plan-Apochromat 203/0.8 Dry (most images) and EC

Plan-Neofluar 103/0.30 Dry. GFP was excited with the 488-nm laser line of

the argon laser, and the emission was detected between 495 and 545 nm.

Propidium iodide (PI, Sigma) was excited by 561 nm and detected between

580 and 680 nm. PI was dissolved in one-tenth-strength MS (0.43 g L–1 MS

salts and 4 mg mL–1 PI).

Siliques and anthers fromdifferent developmental stageswerefixed for 2 h at

room temperature in a 3.7% (w/v) paraformaldehyde solution dissolved in

50 mM PIPES, 5 mM EGTA, and 1 mM MgSO4 buffer, embedded in 5% (w/v)

agarose blocks, and sectioned using a vibratome (Campden Instruments). The

samples from developing seeds were dissected in a binocular microscope to

remove the valves before fixation. The samples from senescing petals were

mounted in the glass slides using one-tenth-strengthMS and 0.01% (v/v) Triton

X-100.

Image processing was done using Fiji (Schindelin et al., 2012). Some panels

were assembled using the stiching plugin.

TUNEL Assay

For the TUNEL, seedlings were fixed for 1 h in 4% (v/v) paraformaldehyde in

phosphate-buffered saline (PBS), pH 7.4, under vacuum at room temperature.

After fixation, seedlings were washed five times in PBS and permeabilized for

2min on ice in a 0.1% (w/v) sodium citrate solutionwith 0.1% (v/v) Triton X-100.

Afterward, seedlings were washed five times in PBS. For the positive control,

fixed andpermeabilizedwild-type seedlingswere treatedwith DNaseI for 15min

at room temperature and washed three times with PBS. For the TUNEL reaction,

label solution and enzyme solution were mixed according to the manufacturer’s

manual (In Situ Cell Death Detection Kit, Fluorescein, Roche Applied Science),

and 50 mL was added to a 1.5-mL microcentrifuge tube together with the seed-

lings. For the negative control, only label solution was used. All samples were

incubated at 37°C in the dark for 1 h. Afterward, the seedlingswerewashed three

times with PBS and mounted with an antifading agent (citifluor, Citifluor Ltd.)

containing 1 mg mL–1 DAPI. The same procedure was used for petals. Stress

treatments of seedlings were the same as described before.

Pathogen Assays, RNA Extraction, and qRT-PCR Analysis

Arabidopsis Col-0 4-week-old plants were inoculated with a bacterial

suspension of Pseudomonas syringae pv tomato AvrRpm1 (5 3 107 colony

forming units mL–1). Leaf samples were harvested at the indicated time points

both inside the infiltrated zone and in noninoculated areas and ground in

liquid nitrogen. Total RNA was isolated using the Nucleospin RNA plant kit

(Macherey-Nagel) according to the manufacturer’s recommendations. Re-

verse transcription was performed using 1.5 mg of total RNA. Real-time

quantitative PCR was performed on a Light Cycler 480 II machine (Roche

Diagnostics) using Roche reagents. Primers used for qRT-PCR are shown in

Supplemental Table S8. Relative expression was calculated as the crossing

point difference between each gene and the internal control SAND family

gene (At2g28390). Average crossing point difference was calculated from

three independent experiments with three replicates and related to the value

of each gene at time 0, which is set at 1.

Supplemental Data

The following supplemental materials are available.

Supplemental Figure S1. Developmental series for petal senescence.

Supplemental Figure S2. Developmental series for tapetum differentia-

tion.

Supplemental Figure S3. Developmental series for seed development.

Supplemental Figure S4. Whole-mount TUNEL of 5- to 6-day-old root tip

after different abiotic stresses provoking cell death.

Supplemental Figure S5. dPCD marker genes are not transcriptionally

regulated during HR-related ePCD.

Supplemental Table S1. Detailed overview of the ATH1 microarray ex-

periments used for the meta-analysis.

Supplemental Table S2. Overview of the number of up- and down-

regulated genes per condition in the experiments used in the meta-analysis.

Supplemental Table S3. Genes commonly regulated in different PCD clusters.

Supplemental Table S4. Performance results of SVM and RF classification

of dPCD versus ePCD instances based on the expression profiles of

various gene (feature) sets in various experiment subsets.

Supplemental Table S5. Commonly coregulated genes ofMC9, RNS3, BFN1,

ARABIDOPSIS THALIANA DAD1-LIKE SEEDING ESTABLISHMENT-

RELATED LIPASE (DSEL), EXI1, PASPA3, and DMP4.

Supplemental Table S6. Ninety-five commonly regulated genes between

the LRC and differentiating tracheary elements, of which eight genes are

common with the 154 coregulated dPCD genes (Supplemental Table S5).

Supplemental Table S7. Phytozome blast search for putative homologs of

the Arabidopsis dPCD marker genes MC9, BFN1, PASPA3, RNS3, and

SCPL48.

Supplemental Table S8. Primers used for promoter cloning and qRT-PCR.

ACKNOWLEDGMENTS

We thank allmembers of the PCD research team at the Vlaams Instituut voor

Biotechnologie-Plant Systems Biology Department department for critical read-

ing of the article, Annick Bleys for help in revising the cited references, other

members of the Vlaams Instituut voor Biotechnologie-Plant Systems Biology

Department for sharing fields of expertise, Dr. Marc Heijde and Dr. Toon Cools

for the genotoxic stress experiments, and Dr. Pavel Kerchev for the oxidative

stress experiments.

Received May 26, 2015; accepted September 30, 2015; published October 5,

2015.

Plant Physiol. Vol. 169, 2015 2697

Transcriptome Meta-Analysis of Plant Cell Death

www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from Copyright © 2015 American Society of Plant Biologists. All rights reserved.

LITERATURE CITED

Appelqvist H, Johansson AC, Linderoth E, Johansson U, Antonsson B,Steinfeld R, Kågedal K, Öllinger K (2012) Lysosome-mediated apo-

ptosis is associated with cathepsin D-specific processing of Bid at Phe24,

Trp48, and Phe183. Ann Clin Lab Sci 42: 231–242Bariola PA, Howard CJ, Taylor CB, Verburg MT, Jaglan VD, Green PJ

(1994) The Arabidopsis ribonuclease gene RNS1 is tightly controlled in

response to phosphate limitation. Plant J 6: 673–685Bayles KW (2014) Bacterial programmed cell death: making sense of a

paradox. Nat Rev Microbiol 12: 63–69Benhamed M, Martin-Magniette ML, Taconnat L, Bitton F, Servet C, De

Clercq R, De Meyer B, Buysschaert C, Rombauts S, Villarroel R, et al(2008) Genome-scale Arabidopsis promoter array identifies targets of

the histone acetyltransferase GCN5. Plant J 56: 493–504Bollhöner B, Prestele J, Tuominen H (2012) Xylem cell death: emerging

understanding of regulation and function. J Exp Bot 63: 1081–1094Bollhöner B, Zhang B, Stael S, Denancé N, Overmyer K, Goffner D, Van

Breusegem F, Tuominen H (2013) Post mortem function of AtMC9 in

xylem vessel elements. New Phytol 200: 498–510Brady SM, Orlando DA, Lee JY, Wang JY, Koch J, Dinneny JR, Mace D,

Ohler U, Benfey PN (2007) A high-resolution root spatiotemporal map

reveals dominant expression patterns. Science 318: 801–806Breiman L (2001) Random forests. Mach Learn 45: 5–32Candat A, Paszkiewicz G, NeveuM, Gautier R, Logan DC, Avelange-Macherel

MH, Macherel D (2014) The ubiquitous distribution of late embryogenesis

abundant proteins across cell compartments in Arabidopsis offers tailored

protection against abiotic stress. Plant Cell 26: 3148–3166Chen X, Wang Y, Li JY, Jiang A, Cheng Y, Zhang W (2009) Mitochondrial

proteome during salt stress-induced programmed cell death in rice.

Plant Physiol Biochem 47: 407–415Coll NS, Epple P, Dangl JL (2011) Programmed cell death in the plant

immune system. Cell Death Differ 18: 1247–1256Cortes C, Vapnik V (1995) Support-vector networks. Mach Learn 20: 273–

297

Deal RB, Henikoff S (2011) The INTACT method for cell type-specific gene

expression and chromatin profiling in Arabidopsis thaliana. Nat Protoc 6:56–68

Demšar J, Curk T, Erjavec A, Gorup C, Ho"cevar T, Milutinovi"c M, Mo"zinaM, Polajnar M, Toplak M, Stari"c A, et al (2013) Orange: Data mining

toolbox in Python. J Mach Learn Res 14: 2349–2353Farage-Barhom S, Burd S, Sonego L, Perl-Treves R, Lers A (2008) Ex-

pression analysis of the BFN1 nuclease gene promoter during senes-

cence, abscission, and programmed cell death-related processes. J Exp

Bot 59: 3247–3258Fendrych M, Van Hautegem T, Van Durme M, Olvera-Carrillo Y,

Huysmans M, Karimi M, Lippens S, Guérin CJ, Krebs M, SchumacherK, et al (2014) Programmed cell death controlled by ANAC033/

SOMBRERO determines root cap organ size in Arabidopsis. Curr Biol

24: 931–940Fuchs Y, Steller H (2011) Programmed cell death in animal development

and disease. Cell 147: 742–758Gentleman RC, Carey VJ, Bates DM, Bolstad B, Dettling M, Dudoit S,

Ellis B, Gautier L, Ge Y, Gentry J, et al (2004) Bioconductor: open

software development for computational biology and bioinformatics.

Genome Biol 5: R80Goodstein DM, Shu S, Howson R, Neupane R, Hayes RD, Fazo J, Mitros

T, Dirks W, Hellsten U, Putnam N, et al (2012) Phytozome: a com-

parative platform for green plant genomics. Nucleic Acids Res 40:D1178–D1186

Green DR (2011) Means to an End. Apoptosis and Other Cell Death

Mechanisms. Cold Spring Harbor Laboratory Press, Cold Spring Har-

bor, NY

Hara-Nishimura I, Hatsugai N (2011) The role of vacuole in plant cell

death. Cell Death Differ 18: 1298–1304Haughn G, Chaudhury A (2005) Genetic analysis of seed coat development

in Arabidopsis. Trends Plant Sci 10: 472–477Helm M, Schmid M, Hierl G, Terneus K, Tan L, Lottspeich F, Kieliszewski

MJ, Gietl C (2008) KDEL-tailed cysteine endopeptidases involved in

programmed cell death, intercalation of new cells, and dismantling of

extensin scaffolds. Am J Bot 95: 1049–1062Hierl G, Höwing T, Isono E, Lottspeich F, Gietl C (2014) Ex vivo processing

for maturation of Arabidopsis KDEL-tailed cysteine endopeptidase 2

(AtCEP2) pro-enzyme and its storage in endoplasmic reticulum derived

organelles. Plant Mol Biol 84: 605–620Hruz T, Laule O, Szabo G, Wessendorp F, Bleuler S, Oertle L, Widmayer

P, Gruissem W, Zimmermann P (2008) Genevestigator V3: a reference

expression database for the meta-analysis of transcriptomes. Adv Bio-

informatics 2008: 420747Irizarry RA, Hobbs B, Collin F, Beazer-Barclay YD, Antonellis KJ, Scherf

U, Speed TP (2003) Exploration, normalization, and summaries of high

density oligonucleotide array probe level data. Biostatistics 4: 249–264Ito J, Fukuda H (2002) ZEN1 is a key enzyme in the degradation of nuclear

DNA during programmed cell death of tracheary elements. Plant Cell

14: 3201–3211Karimi M, De Meyer B, Hilson P (2005) Modular cloning in plant cells.

Trends Plant Sci 10: 103–105Kasaras A, Kunze R (2010) Expression, localisation and phylogeny of a

novel family of plant-specific membrane proteins. Plant Biol 12: 140–152Kasaras A, Melzer M, Kunze R (2012) Arabidopsis senescence-associated

protein DMP1 is involved in membrane remodeling of the ER and ton-

oplast. BMC Plant Biol 12: 54Lam E (2004) Controlled cell death, plant survival and development. Nat

Rev Mol Cell Biol 5: 305–315Le!sniewicz K, Poręba E, Smolarkiewicz M, Wolff N, Stanisławski S,

Wojtaszek P (2012) Plant plasma membrane-bound staphylococcal-like

DNases as a novel class of eukaryotic nucleases. BMC Plant Biol 12: 195Luhtala N, Parker R (2010) T2 family ribonucleases: ancient enzymes with

diverse roles. Trends Biochem Sci 35: 253–259Ma W, Berkowitz GA (2007) The grateful dead: calcium and cell death in

plant innate immunity. Cell Microbiol 9: 2571–2585Mur LAJ, Kenton P, Lloyd AJ, Ougham H, Prats E (2008) The hypersen-

sitive response: The centenary is upon us but how much do we know?

J Exp Bot 59: 501–520Nawkar GM, Maibam P, Park JH, Sahi VP, Lee SY, Kang CH (2013) UV-

Induced cell death in plants. Int J Mol Sci 14: 1608–1628Ohashi-Ito K, Oda Y, Fukuda H (2010) Arabidopsis VASCULAR-RELATED

NAC-DOMAIN6 directly regulates the genes that govern programmed

cell death and secondary wall formation during xylem differentiation.

Plant Cell 22: 3461–3473Olvera-Carrillo Y, Campos F, Reyes JL, Garciarrubio A, Covarrubias AA

(2010) Functional analysis of the group 4 late embryogenesis abundant

proteins reveals their relevance in the adaptive response during water

deficit in Arabidopsis. Plant Physiol 154: 373–390Parizot B, De Rybel B, Beeckman T (2010) VisuaLRTC: a new view on

lateral root initiation by combining specific transcriptome data sets.

Plant Physiol 153: 34–40Petrov V, Hille J, Mueller-Roeber B, Gechev TS (2015) ROS-mediated

abiotic stress-induced programmed cell death in plants. Front Plant

Sci 6: 69Plackett ARG, Thomas SG, Wilson ZA, Hedden P (2011) Gibberellin control of

stamen development: a fertile field. Trends Plant Sci 16: 568–578Proost S, Van Bel M, Vaneechoutte D, Van de Peer Y, Inzé D, Mueller-

Roeber B, Vandepoele K (2015) PLAZA 3.0: an access point for plant

comparative genomics. Nucleic Acids Res 43: D974–D981

Qi Y, Wang H, Zou Y, Liu C, Liu Y, Wang Y, Zhang W (2011) Over-

expression of mitochondrial heat shock protein 70 suppresses pro-

grammed cell death in rice. FEBS Lett 585: 231–239Repnik U, Hafner "Cesen M, Turk B (2014) Lysosomal membrane per-

meabilization in cell death: concepts and challenges. Mitochondrion 19: 49–57Ritchie ME, Phipson B, Wu D, Hu Y, Law CW, Shi W, Smyth GK (2015)

limma powers differential expression analyses for RNA-sequencing and

microarray studies. Nucleic Acids Res 43: e47Roa H, Lang J, Culligan KM, Keller M, Holec S, Cognat V, Montané MH,

Houlné G, Chabouté ME (2009) Ribonucleotide reductase regulation in

response to genotoxic stress in Arabidopsis. Plant Physiol 151: 461–471Sayers EW, Barrett T, Benson DA, Bolton E, Bryant SH, Canese K,

Chetvernin V, Church DM, Dicuccio M, Federhen S, et al (2012) Data-

base resources of the National Center for Biotechnology Information.

Nucleic Acids Res 40: D13–D25

Schindelin J, Arganda-Carreras I, Frise E, Kaynig V, Longair M, Pietzsch T,Preibisch S, Rueden C, Saalfeld S, Schmid B, et al (2012) Fiji: an open-

source platform for biological-image analysis. Nat Methods 9: 676–682Taylor CB, Bariola PA, delCardayré SB, Raines RT, Green PJ (1993) RNS2:

a senescence-associated RNase of Arabidopsis that diverged from the

S-RNases before speciation. Proc Natl Acad Sci USA 90: 5118–5122

2698 Plant Physiol. Vol. 169, 2015

Olvera-Carrillo et al.

www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from Copyright © 2015 American Society of Plant Biologists. All rights reserved.

Thomas H (2013) Senescence, ageing and death of the whole plant. New

Phytol 197: 696–711Trapp O, Seeliger K, Puchta H (2011) Homologs of breast cancer genes in

plants. Front Plant Sci 2: 19Truernit E, Haseloff J (2008) A simple way to identify non-viable cells within

living plant tissue using confocal microscopy. Plant Methods 4: 15Tsiatsiani L, Timmerman E, De Bock PJ, Vercammen D, Stael S, van de

Cotte B, Staes A, Goethals M, Beunens T, Van Damme P, et al (2013)The Arabidopsis metacaspase9 degradome. Plant Cell 25: 2831–2847

van Doorn WG (2011) Classes of programmed cell death in plants, com-

pared to those in animals. J Exp Bot 62: 4749–4761van Doorn WG, Beers EP, Dangl JL, Franklin-Tong VE, Gallois P, Hara-

Nishimura I, Jones AM, Kawai-Yamada M, Lam E, Mundy J, et al (2011)Morphological classification of plant cell deaths. Cell Death Differ 18: 1241–1246

Van Hautegem T, Waters AJ, Goodrich J, Nowack MK (2015) Only in

dying, life: programmed cell death during plant development. Trends

Plant Sci 20: 102–113Wang J, Bayles KW (2013) Programmed cell death in plants: lessons from

bacteria? Trends Plant Sci 18: 133–139Wang Q, Jiang M, Wu J, Ma Y, Li T, Chen Q, Zhang X, Xiang L (2014)

Stress-induced RNASET2 overexpression mediates melanocyte apopto-

sis via the TRAF2 pathway in vitro. Cell Death Dis 5: e1022

Waters A, Creff A, Goodrich J, Ingram G (2013) “What we’ve got here is

failure to communicate”: zou mutants and endosperm cell death in seed

development. Plant Signal Behav 8: e24368Wrzaczek M, Vainonen JP, Stael S, Tsiatsiani L, Help-Rinta-Rahko H,

Gauthier A, Kaufholdt D, Bollhöner B, Lamminmäki A, Staes A, et al(2015) GRIM REAPER peptide binds to receptor kinase PRK5 to trigger

cell death in Arabidopsis. EMBO J 34: 55–66Wu L, Chen H, Curtis C, Fu ZQ (2014) Go in for the kill. Virulence 5: 710–

721

Xu B, Ohtani M, Yamaguchi M, Toyooka K, Wakazaki M, Sato M, KuboM, Nakano Y, Sano R, Hiwatashi Y, et al (2014) Contribution of NAC

transcription factors to plant adaptation to land. Science 343: 1505–1508

Yi D, Alvim Kamei CL, Cools T, Vanderauwera S, Takahashi N,Okushima Y, Eekhout T, Yoshiyama KO, Larkin J, Van den Daele H,et al (2014) The Arabidopsis SIAMESE-RELATED cyclin-dependent ki-

nase inhibitors SMR5 and SMR7 regulate the DNA damage checkpoint

in response to reactive oxygen species. Plant Cell 26: 296–309Zhang D, Liu D, Lv X, Wang Y, Xun Z, Liu Z, Li F, Lu H (2014) The

cysteine protease CEP1, a key executor involved in tapetal programmed

cell death, regulates pollen development in Arabidopsis. Plant Cell 26:2939–2961

Plant Physiol. Vol. 169, 2015 2699

Transcriptome Meta-Analysis of Plant Cell Death

www.plant.org on December 11, 2015 - Published by www.plantphysiol.orgDownloaded from Copyright © 2015 American Society of Plant Biologists. All rights reserved.

115

References

__________________________________________________________________________

116

Ahuja, I., Kissen, R. and Bones, A. M. (2012) 'Phytoalexins in defense against pathogens', Trends Plant Sci, 17(2), pp. 73-90.

Ambawat, S., Sharma, P., Yadav, N. R. and Yadav, R. C. (2013) 'MYB transcription factor genes as regulators for plant responses: an overview', Physiol Mol Biol Plants, 19(3), pp. 307-21.

An, C. and Mou, Z. (2011) 'Salicylic acid and its function in plant immunity', J Integr Plant Biol, 53(6), pp. 412-28.

Apweiler, R., Hermjakob, H. and Sharon, N. (1999) 'On the frequency of protein glycosylation, as deduced from analysis of the SWISS-PROT database', Biochim Biophys Acta, 1473(1), pp. 4-8.

Asai, S. and Shirasu, K. (2015) 'Plant cells under siege: plant immune system versus pathogen effectors', Curr Opin Plant Biol, 28, pp. 1-8.

Assmann, S. M. and Shimazaki, K. (1999) 'The multisensory guard cell. Stomatal responses to blue light and abscisic acid', Plant Physiol, 119(3), pp. 809-16.

Avrova, A. O., Stewart, H. E., De Jong, W. D., Heilbronn, J., Lyon, G. D. and Birch, P. R. (1999) 'A cysteine protease gene is expressed early in resistant potato interactions with Phytophthora infestans', Mol Plant Microbe Interact, 12(12), pp. 1114-9.

Avrova, A. O., Taleb, N., Rokka, V. M., Heilbronn, J., Campbell, E., Hein, I., Gilroy, E. M., Cardle, L., Bradshaw, J. E., Stewart, H. E., Fakim, Y. J., Loake, G. and Birch, P. R. (2004) 'Potato oxysterol binding protein and cathepsin B are rapidly up-regulated in independent defence pathways that distinguish R gene-mediated and field resistances to Phytophthora infestans', Mol Plant Pathol, 5(1), pp. 45-56.

Baker, D., Shiau, A. K. and Agard, D. A. (1993) 'The role of pro regions in protein folding', Curr Opin Cell Biol, 5(6), pp. 966-70.

Bakshi, M. and Oelmüller, R. (2014) 'WRKY transcription factors: Jack of many trades in plants', Plant Signal Behav, 9(2), pp. e27700.

Barrett, A. J. (1994) 'Classification of peptidases', Methods Enzymol, 244, pp. 1-15. Barrett, A. J. and Rawlings, N. D. (1995) 'Families and clans of serine peptidases', Arch

Biochem Biophys, 318(2), pp. 247-50. Bassik, M. C. and Kampmann, M. (2011) 'Knocking out the door to tunicamycin entry', Proc

Natl Acad Sci U S A, 108(29), pp. 11731-2. Beck, M., Heard, W., Mbengue, M. and Robatzek, S. (2012) 'The INs and OUTs of pattern

recognition receptors at the cell surface', Curr Opin Plant Biol, 15(4), pp. 367-74. Bednarek, P. (2012) 'Sulfur-containing secondary metabolites from Arabidopsis thaliana and

other Brassicaceae with function in plant immunity', Chembiochem, 13(13), pp. 1846-59.

Beers, E. P., Jones, A. M. and Dickerman, A. W. (2004) 'The S8 serine, C1A cysteine and A1 aspartic protease families in Arabidopsis', Phytochemistry, 65(1), pp. 43-58.

Beilinson, V., Moskalenko, O. V., Livingstone, D. S., Reverdatto, S. V., Jung, R. and Nielsen, N. C. (2002) 'Two subtilisin-like proteases from soybean', Physiol Plant, 115(4), pp. 585-597.

Belenghi, B., Acconcia, F., Trovato, M., Perazzolli, M., Bocedi, A., Polticelli, F., Ascenzi, P. and Delledonne, M. (2003) 'AtCYS1, a cystatin from Arabidopsis thaliana, suppresses hypersensitive cell death', Eur J Biochem, 270(12), pp. 2593-604.

Ben Khaled, S., Postma, J. and Robatzek, S. (2015) 'A moving view: subcellular trafficking processes in pattern recognition receptor-triggered plant immunity', Annu Rev Phytopathol, 53, pp. 379-402.

Berger, D. and Altmann, T. (2000) 'A subtilisin-like serine protease involved in the regulation of stomatal density and distribution in Arabidopsis thaliana', Genes Dev, 14(9), pp. 1119-31.

Berkey, R., Bendigeri, D. and Xiao, S. (2012) 'Sphingolipids and plant defense/disease: the "death" connection and beyond', Front Plant Sci, 3, pp. 68.

Bigeard, J., Colcombet, J. and Hirt, H. (2015) 'Signaling mechanisms in pattern-triggered immunity (PTI)', Mol Plant, 8(4), pp. 521-39.

117

Block, A., Toruño, T. Y., Elowsky, C. G., Zhang, C., Steinbrenner, J., Beynon, J. and Alfano, J. R. (2014) 'The Pseudomonas syringae type III effector HopD1 suppresses effector-triggered immunity, localizes to the endoplasmic reticulum, and targets the Arabidopsis transcription factor NTL9', New Phytol, 201(4), pp. 1358-70.

Bogacheva, A. M., Rudenskaya, G. N., Preusser, A., Tchikileva, I. O., Dunaevsky, Y. E., Golovkin, B. N. and Stepanov, V. M. (1999) 'A new subtilisin-like proteinase from roots of the dandelion Taraxacum officinale Webb S. L', Biochemistry (Mosc), 64(9), pp. 1030-7.

Bonardi, V., Cherkis, K., Nishimura, M. T. and Dangl, J. L. (2012) 'A new eye on NLR proteins: focused on clarity or diffused by complexity?', Curr Opin Immunol, 24(1), pp. 41-50.

Bos, J. I., Armstrong, M. R., Gilroy, E. M., Boevink, P. C., Hein, I., Taylor, R. M., Zhendong, T., Engelhardt, S., Vetukuri, R. R., Harrower, B., Dixelius, C., Bryan, G., Sadanandom, A., Whisson, S. C., Kamoun, S. and Birch, P. R. (2010) 'Phytophthora infestans effector AVR3a is essential for virulence and manipulates plant immunity by stabilizing host E3 ligase CMPG1', Proc Natl Acad Sci U S A, 107(21), pp. 9909-14.

Boudsocq, M., Willmann, M. R., McCormack, M., Lee, H., Shan, L., He, P., Bush, J., Cheng, S. H. and Sheen, J. (2010) 'Differential innate immune signalling via Ca(2+) sensor protein kinases', Nature, 464(7287), pp. 418-22.

Boyer, L. A., Langer, M. R., Crowley, K. A., Tan, S., Denu, J. M. and Peterson, C. L. (2002) 'Essential role for the SANT domain in the functioning of multiple chromatin remodeling enzymes', Mol Cell, 10(4), pp. 935-42.

Boyle, P. C. and Martin, G. B. (2015) 'Greasy tactics in the plant-pathogen molecular arms race', J Exp Bot, 66(6), pp. 1607-16.

Bozkurt, T. O., Schornack, S., Win, J., Shindo, T., Ilyas, M., Oliva, R., Cano, L. M., Jones, A. M., Huitema, E., van der Hoorn, R. A. and Kamoun, S. (2011) 'Phytophthora infestans effector AVRblb2 prevents secretion of a plant immune protease at the haustorial interface', Proc Natl Acad Sci U S A, 108(51), pp. 20832-7.

Bricchi, I., Occhipinti, A., Bertea, C. M., Zebelo, S. A., Brillada, C., Verrillo, F., De Castro, C., Molinaro, A., Faulkner, C., Maule, A. J. and Maffei, M. E. (2013) 'Separation of early and late responses to herbivory in Arabidopsis by changing plasmodesmal function', Plant J, 73(1), pp. 14-25.

Brown, J. K. (2003) 'A cost of disease resistance: paradigm or peculiarity?', Trends Genet, 19(12), pp. 667-71.

Bu, Q., Jiang, H., Li, C. B., Zhai, Q., Zhang, J., Wu, X., Sun, J., Xie, Q. and Li, C. (2008) 'Role of the Arabidopsis thaliana NAC transcription factors ANAC019 and ANAC055 in regulating jasmonic acid-signaled defense responses', Cell Res, 18(7), pp. 756-67.

Burdon, J. J. (1987) Diseases and Plant Population Biology. Cambridge University Press. Buscaill, P. and Rivas, S. (2014) 'Transcriptional control of plant defence responses', Curr

Opin Plant Biol, 20, pp. 35-46. Böhm, H., Albert, I., Fan, L., Reinhard, A. and Nürnberger, T. (2014) 'Immune receptor

complexes at the plant cell surface', Curr Opin Plant Biol, 20, pp. 47-54. Caillaud, M. C., Piquerez, S. J., Fabro, G., Steinbrenner, J., Ishaque, N., Beynon, J. and

Jones, J. D. (2012) 'Subcellular localization of the Hpa RxLR effector repertoire identifies a tonoplast-associated protein HaRxL17 that confers enhanced plant susceptibility', Plant J, 69(2), pp. 252-65.

Camoni, L., Harper, J. F. and Palmgren, M. G. (1998) '14-3-3 proteins activate a plant calcium-dependent protein kinase (CDPK)', FEBS Lett, 430(3), pp. 381-4.

Canonne, J., Marino, D., Jauneau, A., Pouzet, C., Brière, C., Roby, D. and Rivas, S. (2011) 'The Xanthomonas type III effector XopD targets the Arabidopsis transcription factor MYB30 to suppress plant defense', Plant Cell, 23(9), pp. 3498-511.

Canonne, J., Marino, D., Noël, L. D., Arechaga, I., Pichereaux, C., Rossignol, M., Roby, D. and Rivas, S. (2010) 'Detection and functional characterization of a 215 amino acid N-terminal extension in the Xanthomonas type III effector XopD', PLoS One, 5(12), pp. e15773.

118

Canonne, J., Pichereaux, C., Marino, D., Roby, D., Rossignol, M. and Rivas, S. (2012) 'Identification of the protein sequence of the type III effector XopD from the B100 strain of Xanthomonas campestris pv campestris', Plant Signal Behav, 7(2), pp. 184-7.

Cardinale, B. J., Matulich, K. L., Hooper, D. U., Byrnes, J. E., Duffy, E., Gamfeldt, L., Balvanera, P., O'Connor, M. I. and Gonzalez, A. (2011) 'The functional role of producer diversity in ecosystems', Am J Bot, 98(3), pp. 572-92.

Carmona-Gutierrez, D., Fröhlich, K. U., Kroemer, G. and Madeo, F. (2010) 'Metacaspases are caspases. Doubt no more', Cell Death Differ, 17(3), pp. 377-8.

Carretero-Paulet, L., Galstyan, A., Roig-Villanova, I., Martínez-García, J. F., Bilbao-Castro, J. R. and Robertson, D. L. (2010) 'Genome-wide classification and evolutionary analysis of the bHLH family of transcription factors in Arabidopsis, poplar, rice, moss, and algae', Plant Physiol, 153(3), pp. 1398-412.

Cedzich, A., Huttenlocher, F., Kuhn, B. M., Pfannstiel, J., Gabler, L., Stintzi, A. and Schaller, A. (2009) 'The protease-associated domain and C-terminal extension are required for zymogen processing, sorting within the secretory pathway, and activity of tomato subtilase 3 (SlSBT3)', J Biol Chem, 284(21), pp. 14068-78.

Cesari, S., Bernoux, M., Moncuquet, P., Kroj, T. and Dodds, P. N. (2014) 'A novel conserved mechanism for plant NLR protein pairs: the "integrated decoy" hypothesis', Front Plant Sci, 5, pp. 606.

Chanda, B., Xia, Y., Mandal, M. K., Yu, K., Sekine, K. T., Gao, Q. M., Selote, D., Hu, Y., Stromberg, A., Navarre, D., Kachroo, A. and Kachroo, P. (2011) 'Glycerol-3-phosphate is a critical mobile inducer of systemic immunity in plants', Nat Genet, 43(5), pp. 421-7.

Chang, C., Yu, D., Jiao, J., Jing, S., Schulze-Lefert, P. and Shen, Q. H. (2013) 'Barley MLA immune receptors directly interfere with antagonistically acting transcription factors to initiate disease resistance signaling', Plant Cell, 25(3), pp. 1158-73.

Chaparro-Garcia, A., Schwizer, S., Sklenar, J., Yoshida, K., Petre, B., Bos, J. I., Schornack, S., Jones, A. M., Bozkurt, T. O. and Kamoun, S. (2015) 'Phytophthora infestans RXLR-WY Effector AVR3a Associates with Dynamin-Related Protein 2 Required for Endocytosis of the Plant Pattern Recognition Receptor FLS2', PLoS One, 10(9), pp. e0137071.

Chariot, A., Gielen, J., Merville, M. P. and Bours, V. (1999) 'The homeodomain-containing proteins: an update on their interacting partners', Biochem Pharmacol, 58(12), pp. 1851-7.

Chaturvedi, R., Krothapalli, K., Makandar, R., Nandi, A., Sparks, A. A., Roth, M. R., Welti, R. and Shah, J. (2008) 'Plastid omega3-fatty acid desaturase-dependent accumulation of a systemic acquired resistance inducing activity in petiole exudates of Arabidopsis thaliana is independent of jasmonic acid', Plant J, 54(1), pp. 106-17.

Chen, F. and Foolad, M. R. (1997) 'Molecular organization of a gene in barley which encodes a protein similar to aspartic protease and its specific expression in nucellar cells during degeneration', Plant Mol Biol, 35(6), pp. 821-31.

Chen, L., Zhang, L., Li, D., Wang, F. and Yu, D. (2013) 'WRKY8 transcription factor functions in the TMV-cg defense response by mediating both abscisic acid and ethylene signaling in Arabidopsis', Proc Natl Acad Sci U S A, 110(21), pp. E1963-71.

Chen, Y. N., Slabaugh, E. and Brandizzi, F. (2008) 'Membrane-tethered transcription factors in Arabidopsis thaliana: novel regulators in stress response and development', Curr Opin Plant Biol, 11(6), pp. 695-701.

Cheng, Z., Li, J. F., Niu, Y., Zhang, X. C., Woody, O. Z., Xiong, Y., Djonović , S., Millet, Y., Bush, J., McConkey, B. J., Sheen, J. and Ausubel, F. M. (2015) 'Pathogen-secreted proteases activate a novel plant immune pathway', Nature, 521(7551), pp. 213-6.

Chichkova, N. V., Kim, S. H., Titova, E. S., Kalkum, M., Morozov, V. S., Rubtsov, Y. P., Kalinina, N. O., Taliansky, M. E. and Vartapetian, A. B. (2004) 'A plant caspase-like protease activated during the hypersensitive response', Plant Cell, 16(1), pp. 157-71.

119

Chichkova, N. V., Shaw, J., Galiullina, R. A., Drury, G. E., Tuzhikov, A. I., Kim, S. H., Kalkum, M., Hong, T. B., Gorshkova, E. N., Torrance, L., Vartapetian, A. B. and Taliansky, M. (2010) 'Phytaspase, a relocalisable cell death promoting plant protease with caspase specificity', EMBO J, 29(6), pp. 1149-61.

Chichkova, N. V., Tuzhikov, A. I., Taliansky, M. and Vartapetian, A. B. (2012) 'Plant phytaspases and animal caspases: structurally unrelated death proteases with a common role and specificity', Physiologia Plantarum, 145(1), pp. 77-84.

Chinchilla, D., Bauer, Z., Regenass, M., Boller, T. and Felix, G. (2006) 'The Arabidopsis receptor kinase FLS2 binds flg22 and determines the specificity of flagellin perception', Plant Cell, 18(2), pp. 465-76.

Chinchilla, D., Boller, T. and Robatzek, S. (2007) 'Flagellin signalling in plant immunity', Adv Exp Med Biol, 598, pp. 358-71.

Clapier, C. R. and Cairns, B. R. (2009) 'The biology of chromatin remodeling complexes', Annu Rev Biochem, 78, pp. 273-304.

Coca, M. and San Segundo, B. (2010) 'AtCPK1 calcium-dependent protein kinase mediates pathogen resistance in Arabidopsis', Plant J, 63(3), pp. 526-40.

Coffeen, W. C. and Wolpert, T. J. (2004) 'Purification and characterization of serine proteases that exhibit caspase-like activity and are associated with programmed cell death in Avena sativa', Plant Cell, 16(4), pp. 857-73.

Coll, N. S., Epple, P. and Dangl, J. L. (2011) 'Programmed cell death in the plant immune system', Cell Death Differ, 18(8), pp. 1247-56.

Coll, N. S., Vercammen, D., Smidler, A., Clover, C., Van Breusegem, F., Dangl, J. L. and Epple, P. (2010) 'Arabidopsis type I metacaspases control cell death', Science, 330(6009), pp. 1393-7.

Collinge, M. and Boller, T. (2001) 'Differential induction of two potato genes, Stprx2 and StNAC, in response to infection by Phytophthora infestans and to wounding', Plant Mol Biol, 46(5), pp. 521-9.

Costantini, L. M. and Snapp, E. L. (2013) 'Fluorescent proteins in cellular organelles: serious pitfalls and some solutions', DNA Cell Biol, 32(11), pp. 622-7.

D'Silva, I., Poirier, G. G. and Heath, M. C. (1998) 'Activation of cysteine proteases in cowpea plants during the hypersensitive response--a form of programmed cell death', Exp Cell Res, 245(2), pp. 389-99.

Dangl, J. L., Horvath, D. M. and Staskawicz, B. J. (2013) 'Pivoting the plant immune system from dissection to deployment', Science, 341(6147), pp. 746-51.

Dangl, J. L. and Jones, J. D. (2001) 'Plant pathogens and integrated defence responses to infection', Nature, 411(6839), pp. 826-33.

Daniel, X., Lacomme, C., Morel, J. B. and Roby, D. (1999) 'A novel myb oncogene homologue in Arabidopsis thaliana related to hypersensitive cell death', Plant J, 20(1), pp. 57-66.

Davies, D. R. (1990) 'The structure and function of the aspartic proteinases', Annu Rev Biophys Biophys Chem, 19, pp. 189-215.

de la Fuente van Bentem, S., Vossen, J. H., Vermeer, J. E., de Vroomen, M. J., Gadella, T. W., Haring, M. A. and Cornelissen, B. J. (2003) 'The subcellular localization of plant protein phosphatase 5 isoforms is determined by alternative splicing', Plant Physiol, 133(2), pp. 702-12.

De Vleesschauwer, D., Gheysen, G. and Höfte, M. (2013) 'Hormone defense networking in rice: tales from a different world', Trends Plant Sci, 18(10), pp. 555-65.

De Wit, P. and Spikman, G. (1982) 'Evidence for the occurrence of race and cultivar-specific elicitors of necrosis in intercellular fluids of compatible interactions of Cladosporium fulvum and tomato', Physiological Plant Pathology, 21(1), pp. 1-11.

del Pozo, O. and Lam, E. (2003) 'Expression of the baculovirus p35 protein in tobacco affects cell death progression and compromises N gene-mediated disease resistance response to Tobacco mosaic virus', Mol Plant Microbe Interact, 16(6), pp. 485-94.

120

Delaney, T. P., Uknes, S., Vernooij, B., Friedrich, L., Weymann, K., Negrotto, D., Gaffney, T., Gut-Rella, M., Kessmann, H., Ward, E. and Ryals, J. (1994) 'A central role of salicylic Acid in plant disease resistance', Science, 266(5188), pp. 1247-50.

Delessert, C., Kazan, K., Wilson, I. W., Van Der Straeten, D., Manners, J., Dennis, E. S. and Dolferus, R. (2005) 'The transcription factor ATAF2 represses the expression of pathogenesis-related genes in Arabidopsis', Plant J, 43(5), pp. 745-57.

Delga, A., Le Roux, C. and Deslandes, L. (2015) 'Plant immune receptor decoy: pathogens in their own trap', Oncotarget, 6(18), pp. 15748-9.

Deng, Y., Humbert, S., Liu, J. X., Srivastava, R., Rothstein, S. J. and Howell, S. H. (2011) 'Heat induces the splicing by IRE1 of a mRNA encoding a transcription factor involved in the unfolded protein response in Arabidopsis', Proc Natl Acad Sci U S A, 108(17), pp. 7247-52.

Deslandes, L. and Rivas, S. (2012) 'Catch me if you can: bacterial effectors and plant targets', Trends Plant Sci, 17(11), pp. 644-55.

Desveaux, D., Maréchal, A. and Brisson, N. (2005) 'Whirly transcription factors: defense gene regulation and beyond', Trends Plant Sci, 10(2), pp. 95-102.

Dettmer, J., Hong-Hermesdorf, A., Stierhof, Y. D. and Schumacher, K. (2006) 'Vacuolar H+-ATPase activity is required for endocytic and secretory trafficking in Arabidopsis', Plant Cell, 18(3), pp. 715-30.

Dinesh-Kumar, S. P. and Baker, B. J. (2000) 'Alternatively spliced N resistance gene transcripts: their possible role in tobacco mosaic virus resistance', Proc Natl Acad Sci U S A, 97(4), pp. 1908-13.

Dodds, P. N. and Rathjen, J. P. (2010) 'Plant immunity: towards an integrated view of plant-pathogen interactions', Nat Rev Genet, 11(8), pp. 539-48.

Dong, J., Chen, C. and Chen, Z. (2003) 'Expression profiles of the Arabidopsis WRKY gene superfamily during plant defense response', Plant Mol Biol, 51(1), pp. 21-37.

Du, H., Wang, Y. B., Xie, Y., Liang, Z., Jiang, S. J., Zhang, S. S., Huang, Y. B. and Tang, Y. X. (2013) 'Genome-wide identification and evolutionary and expression analyses of MYB-related genes in land plants', DNA Res, 20(5), pp. 437-48.

Du, Y., Mpina, M. H., Birch, P. R., Bouwmeester, K. and Govers, F. (2015) 'Phytophthora infestans RXLR Effector AVR1 Interacts with Exocyst Component Sec5 to Manipulate Plant Immunity', Plant Physiol, 169(3), pp. 1975-90.

Dubiella, U., Seybold, H., Durian, G., Komander, E., Lassig, R., Witte, C. P., Schulze, W. X. and Romeis, T. (2013) 'Calcium-dependent protein kinase/NADPH oxidase activation circuit is required for rapid defense signal propagation', Proc Natl Acad Sci U S A, 110(21), pp. 8744-9.

Dubos, C., Stracke, R., Grotewold, E., Weisshaar, B., Martin, C. and Lepiniec, L. (2010) 'MYB transcription factors in Arabidopsis', Trends Plant Sci, 15(10), pp. 573-81.

El Oirdi, M., El Rahman, T. A., Rigano, L., El Hadrami, A., Rodriguez, M. C., Daayf, F., Vojnov, A. and Bouarab, K. (2011) 'Botrytis cinerea manipulates the antagonistic effects between immune pathways to promote disease development in tomato', Plant Cell, 23(6), pp. 2405-21.

Elmayan, T., Fromentin, J., Riondet, C., Alcaraz, G., Blein, J. P. and Simon-Plas, F. (2007) 'Regulation of reactive oxygen species production by a 14-3-3 protein in elicited tobacco cells', Plant Cell Environ, 30(6), pp. 722-32.

Engelhardt, S., Boevink, P. C., Armstrong, M. R., Ramos, M. B., Hein, I. and Birch, P. R. (2012) 'Relocalization of late blight resistance protein R3a to endosomal compartments is associated with effector recognition and required for the immune response', Plant Cell, 24(12), pp. 5142-58.

Engineer, C. B., Ghassemian, M., Anderson, J. C., Peck, S. C., Hu, H. and Schroeder, J. I. (2014) 'Carbonic anhydrases, EPF2 and a novel protease mediate CO2 control of stomatal development', Nature, 513(7517), pp. 246-50.

Enoksson, M. and Salvesen, G. S. (2010) 'Metacaspases are not caspases--always doubt', Cell Death Differ, 17(8), pp. 1221.

121

Erickson, F. L., Holzberg, S., Calderon-Urrea, A., Handley, V., Axtell, M., Corr, C. and Baker, B. (1999) 'The helicase domain of the TMV replicase proteins induces the N-mediated defence response in tobacco', Plant J, 18(1), pp. 67-75.

Eulgem, T., Rushton, P. J., Robatzek, S. and Somssich, I. E. (2000) 'The WRKY superfamily of plant transcription factors', Trends Plant Sci, 5(5), pp. 199-206.

Eulgem, T. and Somssich, I. E. (2007) 'Networks of WRKY transcription factors in defense signaling', Curr Opin Plant Biol, 10(4), pp. 366-71.

Fan, M., Bai, M. Y., Kim, J. G., Wang, T., Oh, E., Chen, L., Park, C. H., Son, S. H., Kim, S. K., Mudgett, M. B. and Wang, Z. Y. (2014) 'The bHLH transcription factor HBI1 mediates the trade-off between growth and pathogen-associated molecular pattern-triggered immunity in Arabidopsis', Plant Cell, 26(2), pp. 828-41.

Feng, F. and Zhou, J. M. (2012) 'Plant-bacterial pathogen interactions mediated by type III effectors', Curr Opin Plant Biol, 15(4), pp. 469-76.

Figueiredo, A., Monteiro, F., Fortes, A. M., Bonow-Rex, M., Zyprian, E., Sousa, L. and Pais, M. S. (2012) 'Cultivar-specific kinetics of gene induction during downy mildew early infection in grapevine', Funct Integr Genomics, 12(2), pp. 379-86.

Figueiredo, A., Monteiro, F. and Sebastiana, M. (2014) 'Subtilisin-like proteases in plant-pathogen recognition and immune priming: a perspective', Front Plant Sci, 5, pp. 739.

Flor, H. H. (1971) 'Current Status of the Gene-For-Gene Concept', Annual Review of Phytopathology, 9(1), pp. 275-296.

Fraser, C. M., Rider, L. W. and Chapple, C. (2005) 'An expression and bioinformatics analysis of the Arabidopsis serine carboxypeptidase-like gene family', Plant Physiol, 138(2), pp. 1136-48.

Froidure, S., Canonne, J., Daniel, X., Jauneau, A., Brière, C., Roby, D. and Rivas, S. (2010a) 'AtsPLA2-alpha nuclear relocalization by the Arabidopsis transcription factor AtMYB30 leads to repression of the plant defense response', Proc Natl Acad Sci U S A, 107(34), pp. 15281-6.

Froidure, S., Roby, D. and Rivas, S. (2010b) 'Expression of the Arabidopsis transcription factor AtMYB30 is post-transcriptionally regulated', Plant Physiol Biochem, 48(8), pp. 735-9.

Galán, J. E., Lara-Tejero, M., Marlovits, T. C. and Wagner, S. (2014) 'Bacterial type III secretion systems: specialized nanomachines for protein delivery into target cells', Annu Rev Microbiol, 68, pp. 415-38.

Gangappa, S. N. and Botto, J. F. (2014) 'The BBX family of plant transcription factors', Trends Plant Sci, 19(7), pp. 460-70.

Garcion, C., Lohmann, A., Lamodière, E., Catinot, J., Buchala, A., Doermann, P. and Métraux, J. P. (2008) 'Characterization and biological function of the ISOCHORISMATE SYNTHASE2 gene of Arabidopsis', Plant Physiol, 147(3), pp. 1279-87.

Gassmann, W. (2008) 'Alternative splicing in plant defense', Curr Top Microbiol Immunol, 326, pp. 219-33.

Gassmann, W. and Bhattacharjee, S. (2012) 'Effector-triggered immunity signaling: from gene-for-gene pathways to protein-protein interaction networks', Mol Plant Microbe Interact, 25(7), pp. 862-8.

Gatz, C. (2013) 'From pioneers to team players: TGA transcription factors provide a molecular link between different stress pathways', Mol Plant Microbe Interact, 26(2), pp. 151-9.

Geldner, N., Anders, N., Wolters, H., Keicher, J., Kornberger, W., Muller, P., Delbarre, A., Ueda, T., Nakano, A. and Jürgens, G. (2003) 'The Arabidopsis GNOM ARF-GEF mediates endosomal recycling, auxin transport, and auxin-dependent plant growth', Cell, 112(2), pp. 219-30.

Gigolashvili, T., Berger, B., Mock, H. P., Müller, C., Weisshaar, B. and Flügge, U. I. (2007) 'The transcription factor HIG1/MYB51 regulates indolic glucosinolate biosynthesis in Arabidopsis thaliana', Plant J, 50(5), pp. 886-901.

122

Gilroy, E. M., Hein, I., van der Hoorn, R., Boevink, P. C., Venter, E., McLellan, H., Kaffarnik, F., Hrubikova, K., Shaw, J., Holeva, M., López, E. C., Borras-Hidalgo, O., Pritchard, L., Loake, G. J., Lacomme, C. and Birch, P. R. (2007) 'Involvement of cathepsin B in the plant disease resistance hypersensitive response', Plant J, 52(1), pp. 1-13.

Gilroy, E. M., Taylor, R. M., Hein, I., Boevink, P., Sadanandom, A. and Birch, P. R. (2011) 'CMPG1-dependent cell death follows perception of diverse pathogen elicitors at the host plasma membrane and is suppressed by Phytophthora infestans RXLR effector AVR3a', New Phytol, 190(3), pp. 653-66.

Gindro, K., Berger, V., Godard, S., Voinesco, F., Schnee, S., Viret, O. and Alonso-Villaverde, V. (2012) 'Protease inhibitors decrease the resistance of Vitaceae to Plasmopara viticola', Plant Physiol Biochem, 60, pp. 74-80.

Giraldo, M. C. and Valent, B. (2013) 'Filamentous plant pathogen effectors in action', Nat Rev Microbiol, 11(11), pp. 800-14.

Glazebrook, J. (2005) 'Contrasting mechanisms of defense against biotrophic and necrotrophic pathogens', Annu Rev Phytopathol, 43, pp. 205-27.

Gloggnitzer, J., Akimcheva, S., Srinivasan, A., Kusenda, B., Riehs, N., Stampfl, H., Bautor, J., Dekrout, B., Jonak, C., Jiménez-Gómez, J. M., Parker, J. E. and Riha, K. (2014) 'Nonsense-mediated mRNA decay modulates immune receptor levels to regulate plant antibacterial defense', Cell Host Microbe, 16(3), pp. 376-90.

Gomord, V., Denmat, L. A., Fitchette-Lainé, A. C., Satiat-Jeunemaitre, B., Hawes, C. and Faye, L. (1997) 'The C-terminal HDEL sequence is sufficient for retention of secretory proteins in the endoplasmic reticulum (ER) but promotes vacuolar targeting of proteins that escape the ER', Plant J, 11(2), pp. 313-25.

Gopal, M., Gupta, A. and Thomas, G. V. (2013) 'Bespoke microbiome therapy to manage plant diseases', Front Microbiol, 4, pp. 355.

Gozzo, F. and Faoro, F. (2013) 'Systemic acquired resistance (50 years after discovery): moving from the lab to the field', J Agric Food Chem, 61(51), pp. 12473-91.

Greenberg, J. T. (1997) 'PROGRAMMED CELL DEATH IN PLANT-PATHOGEN INTERACTIONS', Annu Rev Plant Physiol Plant Mol Biol, 48, pp. 525-545.

Groves, M. R., Coulombe, R., Jenkins, J. and Cygler, M. (1998) 'Structural basis for specificity of papain-like cysteine protease proregions toward their cognate enzymes', Proteins, 32(4), pp. 504-14.

Gu, C., Shabab, M., Strasser, R., Wolters, P. J., Shindo, T., Niemer, M., Kaschani, F., Mach, L. and van der Hoorn, R. A. (2012) 'Post-translational regulation and trafficking of the granulin-containing protease RD21 of Arabidopsis thaliana', PLoS One, 7(3), pp. e32422.

Gu, G., Cevallos-Cevallos, J. M. and van Bruggen, A. H. (2013) 'Ingress of Salmonella enterica Typhimurium into tomato leaves through hydathodes', PLoS One, 8(1), pp. e53470.

Gu, Y. Q., Chao, W. S. and Walling, L. L. (1996) 'Localization and post-translational processing of the wound-induced leucine aminopeptidase proteins of tomato', J Biol Chem, 271(42), pp. 25880-7.

Gu, Y. Q., Yang, C., Thara, V. K., Zhou, J. and Martin, G. B. (2000) 'Pti4 is induced by ethylene and salicylic acid, and its product is phosphorylated by the Pto kinase', Plant Cell, 12(5), pp. 771-86.

Gudesblat, G. E., Torres, P. S. and Vojnov, A. A. (2009) 'Stomata and pathogens: Warfare at the gates', Plant Signal Behav, 4(12), pp. 1114-6.

Gupta, R., Lee, S. E., Agrawal, G. K., Rakwal, R., Park, S., Wang, Y. and Kim, S. T. (2015) 'Understanding the plant-pathogen interactions in the context of proteomics-generated apoplastic proteins inventory', Front Plant Sci, 6, pp. 352.

Gómez-Gómez, L. and Boller, T. (2000) 'FLS2: an LRR receptor-like kinase involved in the perception of the bacterial elicitor flagellin in Arabidopsis', Mol Cell, 5(6), pp. 1003-11.

Göhre, V., Spallek, T., Häweker, H., Mersmann, S., Mentzel, T., Boller, T., de Torres, M., Mansfield, J. W. and Robatzek, S. (2008) 'Plant pattern-recognition receptor FLS2 is

123

directed for degradation by the bacterial ubiquitin ligase AvrPtoB', Curr Biol, 18(23), pp. 1824-32.

Haga, N., Kato, K., Murase, M., Araki, S., Kubo, M., Demura, T., Suzuki, K., Müller, I., Voss, U., Jürgens, G. and Ito, M. (2007) 'R1R2R3-Myb proteins positively regulate cytokinesis through activation of KNOLLE transcription in Arabidopsis thaliana', Development, 134(6), pp. 1101-10.

Hammond-Kosack, K. E. and Jones, J. D. (1997) 'PLANT DISEASE RESISTANCE GENES', Annu Rev Plant Physiol Plant Mol Biol, 48, pp. 575-607.

Han, J., Xiong, J., Wang, D. and Fu, X. D. (2011) 'Pre-mRNA splicing: where and when in the nucleus', Trends Cell Biol, 21(6), pp. 336-43.

Hao, L., Goodwin, P. H. and Hsiang, T. (2007) 'Expression of a metacaspase gene of Nicotiana benthamiana after inoculation with Colletotrichum destructivum or Pseudomonas syringae pv. tomato, and the effect of silencing the gene on the host response', Plant Cell Rep, 26(10), pp. 1879-88.

Hatsugai, N., Kuroyanagi, M., Yamada, K., Meshi, T., Tsuda, S., Kondo, M., Nishimura, M. and Hara-Nishimura, I. (2004) 'A plant vacuolar protease, VPE, mediates virus-induced hypersensitive cell death', Science, 305(5685), pp. 855-8.

Hebert, D. N., Garman, S. C. and Molinari, M. (2005) 'The glycan code of the endoplasmic reticulum: asparagine-linked carbohydrates as protein maturation and quality-control tags', Trends Cell Biol, 15(7), pp. 364-70.

Heese, A., Ludwig, A. A. and Jones, J. D. (2005) 'Rapid phosphorylation of a syntaxin during the Avr9/Cf-9-race-specific signaling pathway', Plant Physiol, 138(4), pp. 2406-16.

Hoang, M. H., Nguyen, X. C., Lee, K., Kwon, Y. S., Pham, H. T., Park, H. C., Yun, D. J., Lim, C. O. and Chung, W. S. (2012) 'Phosphorylation by AtMPK6 is required for the biological function of AtMYB41 in Arabidopsis', Biochem Biophys Res Commun, 422(1), pp. 181-6.

Hoeberichts, F. A., ten Have, A. and Woltering, E. J. (2003) 'A tomato metacaspase gene is upregulated during programmed cell death in Botrytis cinerea-infected leaves', Planta, 217(3), pp. 517-22.

Hong, S. Y., Kim, O. K., Kim, S. G., Yang, M. S. and Park, C. M. (2011) 'Nuclear import and DNA binding of the ZHD5 transcription factor is modulated by a competitive peptide inhibitor in Arabidopsis', J Biol Chem, 286(2), pp. 1659-68.

Hoppe, T., Rape, M. and Jentsch, S. (2001) 'Membrane-bound transcription factors: regulated release by RIP or RUP', Curr Opin Cell Biol, 13(3), pp. 344-8.

Hou, S., Yang, Y., Wu, D. and Zhang, C. (2011) 'Plant immunity: evolutionary insights from PBS1, Pto, and RIN4', Plant Signal Behav, 6(6), pp. 794-9.

Howard, B. E., Hu, Q., Babaoglu, A. C., Chandra, M., Borghi, M., Tan, X., He, L., Winter-Sederoff, H., Gassmann, W., Veronese, P. and Heber, S. (2013) 'High-throughput RNA sequencing of pseudomonas-infected Arabidopsis reveals hidden transcriptome complexity and novel splice variants', PLoS One, 8(10), pp. e74183.

Hu, W., dePamphilis, C. W. and Ma, H. (2008) 'Phylogenetic analysis of the plant-specific zinc finger-homeobox and mini zinc finger gene families', J Integr Plant Biol, 50(8), pp. 1031-45.

Huang, H. W., Chen, W. C., Wu, C. Y., Yu, H. C., Lin, W. Y., Chen, S. T. and Wang, K. T. (1997) 'Kinetic studies of the inhibitory effects of propeptides subtilisin BPN' and Carlsberg to bacterial serine proteases', Protein Eng, 10(10), pp. 1227-33.

Huang, L., Zhang, H., Hong, Y., Liu, S., Li, D. and Song, F. (2015) 'Stress-Responsive Expression, Subcellular Localization and Protein-Protein Interactions of the Rice Metacaspase Family', Int J Mol Sci, 16(7), pp. 16216-41.

Huffaker, A., Pearce, G. and Ryan, C. A. (2006) 'An endogenous peptide signal in Arabidopsis activates components of the innate immune response', Proc Natl Acad Sci U S A, 103(26), pp. 10098-103.

Huffaker, A. and Ryan, C. A. (2007) 'Endogenous peptide defense signals in Arabidopsis differentially amplify signaling for the innate immune response', Proc Natl Acad Sci U S A, 104(25), pp. 10732-6.

124

Hématy, K., Cherk, C. and Somerville, S. (2009) 'Host-pathogen warfare at the plant cell wall', Curr Opin Plant Biol, 12(4), pp. 406-13.

Höwing, T., Huesmann, C., Hoefle, C., Nagel, M. K., Isono, E., Hückelhoven, R. and Gietl, C. (2014) 'Endoplasmic reticulum KDEL-tailed cysteine endopeptidase 1 of Arabidopsis (AtCEP1) is involved in pathogen defense', Front Plant Sci, 5, pp. 58.

Hückelhoven, R. (2007) 'Cell wall-associated mechanisms of disease resistance and susceptibility', Annu Rev Phytopathol, 45, pp. 101-27.

Hückelhoven, R., Dechert, C., Trujillo, M. and Kogel, K. H. (2001) 'Differential expression of putative cell death regulator genes in near-isogenic, resistant and susceptible barley lines during interaction with the powdery mildew fungus', Plant Mol Biol, 47(6), pp. 739-48.

Iakimova, E. T., Sobiczewski, P., Michalczuk, L., Wę grzynowicz-Lesiak, E., Mikiciń ski, A. and Woltering, E. J. (2013) 'Morphological and biochemical characterization of Erwinia amylovora-induced hypersensitive cell death in apple leaves', Plant Physiol Biochem, 63, pp. 292-305.

Inohara, N. and Nuñez, G. (2003) 'NODs: intracellular proteins involved in inflammation and apoptosis', Nat Rev Immunol, 3(5), pp. 371-82.

Ishihama, N. and Yoshioka, H. (2012) 'Post-translational regulation of WRKY transcription factors in plant immunity', Curr Opin Plant Biol, 15(4), pp. 431-7.

Ito, M. (2005) 'Conservation and diversification of three-repeat Myb transcription factors in plants', J Plant Res, 118(1), pp. 61-9.

Janzik, I., Macheroux, P., Amrhein, N. and Schaller, A. (2000) 'LeSBT1, a subtilase from tomato plants. Overexpression in insect cells, purification, and characterization', J Biol Chem, 275(7), pp. 5193-9.

Jashni, M. K., Mehrabi, R., Collemare, J., Mesarich, C. H. and de Wit, P. J. (2015) 'The battle in the apoplast: further insights into the roles of proteases and their inhibitors in plant-pathogen interactions', Front Plant Sci, 6, pp. 584.

Jensen, M. K., Hagedorn, P. H., de Torres-Zabala, M., Grant, M. R., Rung, J. H., Collinge, D. B. and Lyngkjaer, M. F. (2008) 'Transcriptional regulation by an NAC (NAM-ATAF1,2-CUC2) transcription factor attenuates ABA signalling for efficient basal defence towards Blumeria graminis f. sp. hordei in Arabidopsis', Plant J, 56(6), pp. 867-80.

Jensen, M. K., Rung, J. H., Gregersen, P. L., Gjetting, T., Fuglsang, A. T., Hansen, M., Joehnk, N., Lyngkjaer, M. F. and Collinge, D. B. (2007) 'The HvNAC6 transcription factor: a positive regulator of penetration resistance in barley and Arabidopsis', Plant Mol Biol, 65(1-2), pp. 137-50.

Jeworutzki, E., Roelfsema, M. R., Anschütz, U., Krol, E., Elzenga, J. T., Felix, G., Boller, T., Hedrich, R. and Becker, D. (2010) 'Early signaling through the Arabidopsis pattern recognition receptors FLS2 and EFR involves Ca-associated opening of plasma membrane anion channels', Plant J, 62(3), pp. 367-78.

Jia, L., Clegg, M. T. and Jiang, T. (2004) 'Evolutionary dynamics of the DNA-binding domains in putative R2R3-MYB genes identified from rice subspecies indica and japonica genomes', Plant Physiol, 134(2), pp. 575-85.

Jiang, C., Gu, J., Chopra, S., Gu, X. and Peterson, T. (2004) 'Ordered origin of the typical two- and three-repeat Myb genes', Gene, 326, pp. 13-22.

Jones, J. D. and Dangl, J. L. (2006) 'The plant immune system', Nature, 444(7117), pp. 323-9.

Jordá, L., Coego, A., Conejero, V. and Vera, P. (1999) 'A genomic cluster containing four differentially regulated subtilisin-like processing protease genes is in tomato plants', J Biol Chem, 274(4), pp. 2360-5.

Jordá, L., Conejero, V. and Vera, P. (2000) 'Characterization of P69E and P69F, two differentially regulated genes encoding new members of the subtilisin-like proteinase family from tomato plants', Plant Physiol, 122(1), pp. 67-74.

Journot-Catalino, N., Somssich, I. E., Roby, D. and Kroj, T. (2006) 'The transcription factors WRKY11 and WRKY17 act as negative regulators of basal resistance in Arabidopsis thaliana', Plant Cell, 18(11), pp. 3289-302.

125

Juge, N. (2006) 'Plant protein inhibitors of cell wall degrading enzymes', Trends Plant Sci, 11(7), pp. 359-67.

Kadek, A., Tretyachenko, V., Mrazek, H., Ivanova, L., Halada, P., Rey, M., Schriemer, D. C. and Man, P. (2014) 'Expression and characterization of plant aspartic protease nepenthesin-1 from Nepenthes gracilis', Protein Expr Purif, 95, pp. 121-8.

Kadota, Y., Sklenar, J., Derbyshire, P., Stransfeld, L., Asai, S., Ntoukakis, V., Jones, J. D., Shirasu, K., Menke, F., Jones, A. and Zipfel, C. (2014) 'Direct regulation of the NADPH oxidase RBOHD by the PRR-associated kinase BIK1 during plant immunity', Mol Cell, 54(1), pp. 43-55.

Kalyna, M., Simpson, C. G., Syed, N. H., Lewandowska, D., Marquez, Y., Kusenda, B., Marshall, J., Fuller, J., Cardle, L., McNicol, J., Dinh, H. Q., Barta, A. and Brown, J. W. (2012) 'Alternative splicing and nonsense-mediated decay modulate expression of important regulatory genes in Arabidopsis', Nucleic Acids Res, 40(6), pp. 2454-69.

Kamthan, A., Chaudhuri, A., Kamthan, M. and Datta, A. (2015) 'Small RNAs in plants: recent development and application for crop improvement', Front Plant Sci, 6, pp. 208.

Kang, S., Oh, S., Kim, J., Choi, D. and Baek, K. (2010) 'NMMP1, a Matrix Metalloprotease in Nicotiana benthamiana Has a Role in Protection against Bacterial Infection', Plant Pathol. J., 26(4), pp. 402-408.

Kang, Y. H., Kirik, V., Hulskamp, M., Nam, K. H., Hagely, K., Lee, M. M. and Schiefelbein, J. (2009) 'The MYB23 gene provides a positive feedback loop for cell fate specification in the Arabidopsis root epidermis', Plant Cell, 21(4), pp. 1080-94.

Kaschani, F., Gu, C., Niessen, S., Hoover, H., Cravatt, B. F. and van der Hoorn, R. A. (2009) 'Diversity of serine hydrolase activities of unchallenged and botrytis-infected Arabidopsis thaliana', Mol Cell Proteomics, 8(5), pp. 1082-93.

Kaschani, F., Shabab, M., Bozkurt, T., Shindo, T., Schornack, S., Gu, C., Ilyas, M., Win, J., Kamoun, S. and van der Hoorn, R. A. (2010) 'An effector-targeted protease contributes to defense against Phytophthora infestans and is under diversifying selection in natural hosts', Plant Physiol, 154(4), pp. 1794-804.

Kazan, K. (2003) 'Alternative splicing and proteome diversity in plants: the tip of the iceberg has just emerged', Trends Plant Sci, 8(10), pp. 468-71.

Kazan, K. and Manners, J. M. (2013) 'MYC2: the master in action', Mol Plant, 6(3), pp. 686-703.

Kemp, B. P., Beeching, J. R. and Cooper, R. M. (2005) 'cDNA-AFLP reveals genes differentially expressed during the hypersensitive response of cassava', Mol Plant Pathol, 6(2), pp. 113-23.

Kim, J. G., Taylor, K. W., Hotson, A., Keegan, M., Schmelz, E. A. and Mudgett, M. B. (2008) 'XopD SUMO protease affects host transcription, promotes pathogen growth, and delays symptom development in xanthomonas-infected tomato leaves', Plant Cell, 20(7), pp. 1915-29.

Kim, S. M., Bae, C., Oh, S. K. and Choi, D. (2013) 'A pepper (Capsicum annuum L.) metacaspase 9 (Camc9) plays a role in pathogen-induced cell death in plants', Mol Plant Pathol, 14(6), pp. 557-66.

Kim, Y. H., Kwak, M. S., Park, J. B., Lee, S. A., Choi, J. E., Cho, H. S. and Shin, J. S. (2015) 'N-linked glycosylation plays a critical role for the secretion of HMGB1', J Cell Sci.

Kim, Y. S., Kim, S. G., Park, J. E., Park, H. Y., Lim, M. H., Chua, N. H. and Park, C. M. (2006) 'A membrane-bound NAC transcription factor regulates cell division in Arabidopsis', Plant Cell, 18(11), pp. 3132-44.

Klempnauer, K. H., Gonda, T. J. and Bishop, J. M. (1982) 'Nucleotide sequence of the retroviral leukemia gene v-myb and its cellular progenitor c-myb: the architecture of a transduced oncogene', Cell, 31(2 Pt 1), pp. 453-63.

Knoth, C., Ringler, J., Dangl, J. L. and Eulgem, T. (2007) 'Arabidopsis WRKY70 is required for full RPP4-mediated disease resistance and basal defense against Hyaloperonospora parasitica', Mol Plant Microbe Interact, 20(2), pp. 120-8.

126

Koizumi, M., Yamaguchi-Shinozaki, K., Tsuji, H. and Shinozaki, K. (1993) 'Structure and expression of two genes that encode distinct drought-inducible cysteine proteinases in Arabidopsis thaliana', Gene, 129(2), pp. 175-82.

Kothari, I. L. and Patel, M. (2004) 'Plant immunization', Indian J Exp Biol, 42(3), pp. 244-52. Kotzer, A. M., Brandizzi, F., Neumann, U., Paris, N., Moore, I. and Hawes, C. (2004)

'AtRabF2b (Ara7) acts on the vacuolar trafficking pathway in tobacco leaf epidermal cells', J Cell Sci, 117(Pt 26), pp. 6377-89.

Kriechbaumer, V., Wang, P., Hawes, C. and Abell, B. M. (2012) 'Alternative splicing of the auxin biosynthesis gene YUCCA4 determines its subcellular compartmentation', Plant J, 70(2), pp. 292-302.

Krol, E., Mentzel, T., Chinchilla, D., Boller, T., Felix, G., Kemmerling, B., Postel, S., Arents, M., Jeworutzki, E., Al-Rasheid, K. A., Becker, D. and Hedrich, R. (2010) 'Perception of the Arabidopsis danger signal peptide 1 involves the pattern recognition receptor AtPEPR1 and its close homologue AtPEPR2', J Biol Chem, 285(18), pp. 13471-9.

Krüger, J., Thomas, C. M., Golstein, C., Dixon, M. S., Smoker, M., Tang, S., Mulder, L. and Jones, J. D. (2002) 'A tomato cysteine protease required for Cf-2-dependent disease resistance and suppression of autonecrosis', Science, 296(5568), pp. 744-7.

Kubicek, C. P., Starr, T. L. and Glass, N. L. (2014) 'Plant cell wall-degrading enzymes and their secretion in plant-pathogenic fungi', Annu Rev Phytopathol, 52, pp. 427-51.

Kuroyanagi, M., Yamada, K., Hatsugai, N., Kondo, M., Nishimura, M. and Hara-Nishimura, I. (2005) 'Vacuolar processing enzyme is essential for mycotoxin-induced cell death in Arabidopsis thaliana', J Biol Chem, 280(38), pp. 32914-20.

Lachaud, C., Prigent, E., Thuleau, P., Grat, S., Da Silva, D., Brière, C., Mazars, C. and Cotelle, V. (2013) '14-3-3-regulated Ca(2+)-dependent protein kinase CPK3 is required for sphingolipid-induced cell death in Arabidopsis', Cell Death Differ, 20(2), pp. 209-17.

Lacomme, C. and Roby, D. (1999) 'Identification of new early markers of the hypersensitive response in Arabidopsis thaliana(1)', FEBS Lett, 459(2), pp. 149-53.

Laluk, K. and Mengiste, T. (2011) 'The Arabidopsis extracellular UNUSUAL SERINE PROTEASE INHIBITOR functions in resistance to necrotrophic fungi and insect herbivory', Plant J, 68(3), pp. 480-94.

Lam, E. (2004) 'Controlled cell death, plant survival and development', Nat Rev Mol Cell Biol, 5(4), pp. 305-15.

Lannoo, N. and Van Damme, E. J. (2015) 'Review/N-glycans: The making of a varied toolbox', Plant Sci, 239, pp. 67-83.

Lee, A. H., Hurley, B., Felsensteiner, C., Yea, C., Ckurshumova, W., Bartetzko, V., Wang, P. W., Quach, V., Lewis, J. D., Liu, Y. C., Börnke, F., Angers, S., Wilde, A., Guttman, D. S. and Desveaux, D. (2012) 'A bacterial acetyltransferase destroys plant microtubule networks and blocks secretion', PLoS Pathog, 8(2), pp. e1002523.

Lee, K., Lee, H. G., Yoon, S., Kim, H. U. and Seo, P. J. (2015) 'The Arabidopsis MYB96 Transcription Factor Is a Positive Regulator of ABSCISIC ACID-INSENSITIVE4 in the Control of Seed Germination', Plant Physiol, 168(2), pp. 677-89.

Lee, M. W., Qi, M. and Yang, Y. (2001) 'A novel jasmonic acid-inducible rice myb gene associates with fungal infection and host cell death', Mol Plant Microbe Interact, 14(4), pp. 527-35.

Lehfeldt, C., Shirley, A. M., Meyer, K., Ruegger, M. O., Cusumano, J. C., Viitanen, P. V., Strack, D. and Chapple, C. (2000) 'Cloning of the SNG1 gene of Arabidopsis reveals a role for a serine carboxypeptidase-like protein as an acyltransferase in secondary metabolism', Plant Cell, 12(8), pp. 1295-306.

Leon-Reyes, A., Du, Y., Koornneef, A., Proietti, S., Körbes, A. P., Memelink, J., Pieterse, C. M. and Ritsema, T. (2010) 'Ethylene signaling renders the jasmonate response of Arabidopsis insensitive to future suppression by salicylic Acid', Mol Plant Microbe Interact, 23(2), pp. 187-97.

127

Lepiniec, L., Debeaujon, I., Routaboul, J. M., Baudry, A., Pourcel, L., Nesi, N. and Caboche, M. (2006) 'Genetics and biochemistry of seed flavonoids', Annu Rev Plant Biol, 57, pp. 405-30.

Li, J., Brader, G. and Palva, E. T. (2008) 'Kunitz trypsin inhibitor: an antagonist of cell death triggered by phytopathogens and fumonisin b1 in Arabidopsis', Mol Plant, 1(3), pp. 482-95.

Li, L., Li, M., Yu, L., Zhou, Z., Liang, X., Liu, Z., Cai, G., Gao, L., Zhang, X., Wang, Y., Chen, S. and Zhou, J. M. (2014) 'The FLS2-associated kinase BIK1 directly phosphorylates the NADPH oxidase RbohD to control plant immunity', Cell Host Microbe, 15(3), pp. 329-38.

Li, X., Sun, Z., Shao, S., Zhang, S., Ahammed, G. J., Zhang, G., Jiang, Y., Zhou, J., Xia, X., Zhou, Y., Yu, J. and Shi, K. (2015) 'Tomato-Pseudomonas syringae interactions under elevated CO concentration: the role of stomata', J Exp Bot, 66(1), pp. 307-16.

Li, Y., Hu, Z., Jordan, F. and Inouye, M. (1995) 'Functional analysis of the propeptide of subtilisin E as an intramolecular chaperone for protein folding. Refolding and inhibitory abilities of propeptide mutants', J Biol Chem, 270(42), pp. 25127-32.

Libault, M., Wan, J., Czechowski, T., Udvardi, M. and Stacey, G. (2007) 'Identification of 118 Arabidopsis transcription factor and 30 ubiquitin-ligase genes responding to chitin, a plant-defense elicitor', Mol Plant Microbe Interact, 20(8), pp. 900-11.

Licausi, F., Ohme-Takagi, M. and Perata, P. (2013) 'APETALA2/Ethylene Responsive Factor (AP2/ERF) transcription factors: mediators of stress responses and developmental programs', New Phytol, 199(3), pp. 639-49.

Lipsick, J. S. (1996) 'One billion years of Myb', Oncogene, 13(2), pp. 223-35. Lipsick, J. S., Manak, J., Mitiku, N., Chen, C. K., Fogarty, P. and Guthrie, E. (2001)

'Functional evolution of the Myb oncogene family', Blood Cells Mol Dis, 27(2), pp. 456-8.

Liu, J. X., Srivastava, R., Che, P. and Howell, S. H. (2007) 'Salt stress responses in Arabidopsis utilize a signal transduction pathway related to endoplasmic reticulum stress signaling', Plant J, 51(5), pp. 897-909.

Liu, X., Wang, X., Pang, Y., Liang, J., Liu, S., Sun, X. and Tang, K. (2006) '[Molecular cloning and characterization of a novel WRKY gene from Brassica chinensis]', Mol Biol (Mosk), 40(5), pp. 816-24.

Liu, Y., Dammann, C. and Bhattacharyya, M. K. (2001) 'The matrix metalloproteinase gene GmMMP2 is activated in response to pathogenic infections in soybean', Plant Physiol, 127(4), pp. 1788-97.

Livaja, M., Zeidler, D., von Rad, U. and Durner, J. (2008) 'Transcriptional responses of Arabidopsis thaliana to the bacteria-derived PAMPs harpin and lipopolysaccharide', Immunobiology, 213(3-4), pp. 161-71.

Loqué, D., Scheller, H. V. and Pauly, M. (2015) 'Engineering of plant cell walls for enhanced biofuel production', Curr Opin Plant Biol, 25, pp. 151-61.

Lorang, J., Kidarsa, T., Bradford, C. S., Gilbert, B., Curtis, M., Tzeng, S. C., Maier, C. S. and Wolpert, T. J. (2012) 'Tricking the guard: exploiting plant defense for disease susceptibility', Science, 338(6107), pp. 659-62.

Lorrain, S., Vailleau, F., Balagué, C. and Roby, D. (2003) 'Lesion mimic mutants: keys for deciphering cell death and defense pathways in plants?', Trends Plant Sci, 8(6), pp. 263-71.

Lozano-Durán, R., Macho, A. P., Boutrot, F., Segonzac, C., Somssich, I. E. and Zipfel, C. (2013) 'The transcriptional regulator BZR1 mediates trade-off between plant innate immunity and growth', Elife, 2, pp. e00983.

Lu, D., Lin, W., Gao, X., Wu, S., Cheng, C., Avila, J., Heese, A., Devarenne, T. P., He, P. and Shan, L. (2011) 'Direct ubiquitination of pattern recognition receptor FLS2 attenuates plant innate immunity', Science, 332(6036), pp. 1439-42.

Lu, D., Wu, S., Gao, X., Zhang, Y., Shan, L. and He, P. (2010) 'A receptor-like cytoplasmic kinase, BIK1, associates with a flagellin receptor complex to initiate plant innate immunity', Proc Natl Acad Sci U S A, 107(1), pp. 496-501.

128

Lu, Y. J., Schornack, S., Spallek, T., Geldner, N., Chory, J., Schellmann, S., Schumacher, K., Kamoun, S. and Robatzek, S. (2012) 'Patterns of plant subcellular responses to successful oomycete infections reveal differences in host cell reprogramming and endocytic trafficking', Cell Microbiol, 14(5), pp. 682-97.

Luo, X. and Hofmann, K. (2001) 'The protease-associated domain: a homology domain associated with multiple classes of proteases', Trends Biochem Sci, 26(3), pp. 147-8.

Ma, W. (2011) 'Roles of Ca2+ and cyclic nucleotide gated channel in plant innate immunity', Plant Sci, 181(4), pp. 342-6.

Mahon, P. and Bateman, A. (2000) 'The PA domain: a protease-associated domain', Protein Sci, 9(10), pp. 1930-4.

Maldonado, A. M., Doerner, P., Dixon, R. A., Lamb, C. J. and Cameron, R. K. (2002) 'A putative lipid transfer protein involved in systemic resistance signalling in Arabidopsis', Nature, 419(6905), pp. 399-403.

Malinovsky, F. G., Batoux, M., Schwessinger, B., Youn, J. H., Stransfeld, L., Win, J., Kim, S. K. and Zipfel, C. (2014) 'Antagonistic regulation of growth and immunity by the Arabidopsis basic helix-loop-helix transcription factor homolog of brassinosteroid enhanced expression2 interacting with increased leaf inclination1 binding bHLH1', Plant Physiol, 164(3), pp. 1443-55.

Manners, J. M., Penninckx, I. A., Vermaere, K., Kazan, K., Brown, R. L., Morgan, A., Maclean, D. J., Curtis, M. D., Cammue, B. P. and Broekaert, W. F. (1998) 'The promoter of the plant defensin gene PDF1.2 from Arabidopsis is systemically activated by fungal pathogens and responds to methyl jasmonate but not to salicylic acid', Plant Mol Biol, 38(6), pp. 1071-80.

Marian, C. O. and Bass, H. W. (2005) 'The Terminal acidic SANT 1 (Tacs1) gene of maize is expressed in tissues containing meristems and encodes an acidic SANT domain similar to some chromatin-remodeling complex proteins', Biochim Biophys Acta, 1727(2), pp. 81-6.

Marino, D., Froidure, S., Canonne, J., Ben Khaled, S., Khafif, M., Pouzet, C., Jauneau, A., Roby, D. and Rivas, S. (2013) 'Arabidopsis ubiquitin ligase MIEL1 mediates degradation of the transcription factor MYB30 weakening plant defence', Nat Commun, 4, pp. 1476.

Marino, G. and Funk, C. (2012) 'Matrix metalloproteinases in plants: a brief overview', Physiol Plant, 145(1), pp. 196-202.

Markham, J. E., Lynch, D. V., Napier, J. A., Dunn, T. M. and Cahoon, E. B. (2013) 'Plant sphingolipids: function follows form', Curr Opin Plant Biol, 16(3), pp. 350-7.

Marquez, Y., Brown, J. W., Simpson, C., Barta, A. and Kalyna, M. (2012) 'Transcriptome survey reveals increased complexity of the alternative splicing landscape in Arabidopsis', Genome Res, 22(6), pp. 1184-95.

Martínez, M., Cambra, I., González-Melendi, P., Santamaría, M. E. and Díaz, I. (2012) 'C1A cysteine-proteases and their inhibitors in plants', Physiol Plant, 145(1), pp. 85-94.

Mase, K., Mizuno, T., Ishihama, N., Fujii, T., Mori, H., Kodama, M. and Yoshioka, H. (2012) 'Ethylene signaling pathway and MAPK cascades are required for AAL toxin-induced programmed cell death', Mol Plant Microbe Interact, 25(8), pp. 1015-25.

Matsui, K., Umemura, Y. and Ohme-Takagi, M. (2008) 'AtMYBL2, a protein with a single MYB domain, acts as a negative regulator of anthocyanin biosynthesis in Arabidopsis', Plant J, 55(6), pp. 954-67.

Matus, J. T., Aquea, F. and Arce-Johnson, P. (2008) 'Analysis of the grape MYB R2R3 subfamily reveals expanded wine quality-related clades and conserved gene structure organization across Vitis and Arabidopsis genomes', BMC Plant Biol, 8, pp. 83.

Maurer-Stroh, S. and Eisenhaber, F. (2004) 'Myristoylation of viral and bacterial proteins', Trends Microbiol, 12(4), pp. 178-85.

McAinsh, M. R. and Pittman, J. K. (2009) 'Shaping the calcium signature', New Phytol, 181(2), pp. 275-94.

129

McGlincy, N. J. and Smith, C. W. (2008) 'Alternative splicing resulting in nonsense-mediated mRNA decay: what is the meaning of nonsense?', Trends Biochem Sci, 33(8), pp. 385-93.

McLellan, H., Gilroy, E. M., Yun, B. W., Birch, P. R. and Loake, G. J. (2009) 'Functional redundancy in the Arabidopsis Cathepsin B gene family contributes to basal defence, the hypersensitive response and senescence', New Phytol, 183(2), pp. 408-18.

Meichtry, J., Amrhein, N. and Schaller, A. (1999) 'Characterization of the subtilase gene family in tomato (Lycopersicon esculentum Mill.)', Plant Mol Biol, 39(4), pp. 749-60.

Melotto, M., Underwood, W. and He, S. Y. (2008) 'Role of stomata in plant innate immunity and foliar bacterial diseases', Annu Rev Phytopathol, 46, pp. 101-22.

Melotto, M., Underwood, W., Koczan, J., Nomura, K. and He, S. Y. (2006) 'Plant stomata function in innate immunity against bacterial invasion', Cell, 126(5), pp. 969-80.

Mendieta, J. R., Pagano, M. R., Muñoz, F. F., Daleo, G. R. and Guevara, M. G. (2006) 'Antimicrobial activity of potato aspartic proteases (StAPs) involves membrane permeabilization', Microbiology, 152(Pt 7), pp. 2039-47.

Mengiste, T., Chen, X., Salmeron, J. and Dietrich, R. (2003) 'The BOTRYTIS SUSCEPTIBLE1 gene encodes an R2R3MYB transcription factor protein that is required for biotic and abiotic stress responses in Arabidopsis', Plant Cell, 15(11), pp. 2551-65.

Meyers, B. C., Kozik, A., Griego, A., Kuang, H. and Michelmore, R. W. (2003) 'Genome-wide analysis of NBS-LRR-encoding genes in Arabidopsis', Plant Cell, 15(4), pp. 809-34.

Misas-Villamil, J. C., Toenges, G., Kolodziejek, I., Sadaghiani, A. M., Kaschani, F., Colby, T., Bogyo, M. and van der Hoorn, R. A. (2013) 'Activity profiling of vacuolar processing enzymes reveals a role for VPE during oomycete infection', Plant J, 73(4), pp. 689-700.

Miya, A., Albert, P., Shinya, T., Desaki, Y., Ichimura, K., Shirasu, K., Narusaka, Y., Kawakami, N., Kaku, H. and Shibuya, N. (2007) 'CERK1, a LysM receptor kinase, is essential for chitin elicitor signaling in Arabidopsis', Proc Natl Acad Sci U S A, 104(49), pp. 19613-8.

Monteiro, F., Sebastiana, M., Pais, M. S. and Figueiredo, A. (2013) 'Reference gene selection and validation for the early responses to downy mildew infection in susceptible and resistant Vitis vinifera cultivars', PLoS One, 8(9), pp. e72998.

Morimoto, S., Tateishi, N., Inuyama, M., Taura, F., Tanaka, H. and Shoyama, Y. (1999) 'Identification and molecular characterization of novel peroxidase with structural protein-like properties', J Biol Chem, 274(37), pp. 26192-8.

Mur, L. A., Laarhoven, L. J., Harren, F. J., Hall, M. A. and Smith, A. R. (2008) 'Nitric oxide interacts with salicylate to regulate biphasic ethylene production during the hypersensitive response', Plant Physiol, 148(3), pp. 1537-46.

Muthamilarasan, M. and Prasad, M. (2013) 'Plant innate immunity: an updated insight into defense mechanism', J Biosci, 38(2), pp. 433-49.

Nagase, H. and Woessner, J. F. (1999) 'Matrix metalloproteinases', J Biol Chem, 274(31), pp. 21491-4.

Nagashima, Y., Mishiba, K., Suzuki, E., Shimada, Y., Iwata, Y. and Koizumi, N. (2011) 'Arabidopsis IRE1 catalyses unconventional splicing of bZIP60 mRNA to produce the active transcription factor', Sci Rep, 1, pp. 29.

Nakagawa, M., Ueyama, M., Tsuruta, H., Uno, T., Kanamaru, K., Mikami, B. and Yamagata, H. (2010) 'Functional analysis of the cucumisin propeptide as a potent inhibitor of its mature enzyme', J Biol Chem, 285(39), pp. 29797-807.

Navarre, D. A. and Wolpert, T. J. (1999) 'Victorin induction of an apoptotic/senescence-like response in oats', Plant Cell, 11(2), pp. 237-49.

Navarro, L., Zipfel, C., Rowland, O., Keller, I., Robatzek, S., Boller, T. and Jones, J. D. (2004) 'The transcriptional innate immune response to flg22. Interplay and overlap with Avr gene-dependent defense responses and bacterial pathogenesis', Plant Physiol, 135(2), pp. 1113-28.

130

Nelson, B. K., Cai, X. and Nebenführ, A. (2007) 'A multicolored set of in vivo organelle markers for co-localization studies in Arabidopsis and other plants', Plant J, 51(6), pp. 1126-36.

Neuteboom, L. W., Veth-Tello, L. M., Clijdesdale, O. R., Hooykaas, P. J. and van der Zaal, B. J. (1999) 'A novel subtilisin-like protease gene from Arabidopsis thaliana is expressed at sites of lateral root emergence', DNA Res, 6(1), pp. 13-9.

Nguyen, X. C., Hoang, M. H., Kim, H. S., Lee, K., Liu, X. M., Kim, S. H., Bahk, S., Park, H. C. and Chung, W. S. (2012) 'Phosphorylation of the transcriptional regulator MYB44 by mitogen activated protein kinase regulates Arabidopsis seed germination', Biochem Biophys Res Commun, 423(4), pp. 703-8.

Nicholson, P. and Mühlemann, O. (2010) 'Cutting the nonsense: the degradation of PTC-containing mRNAs', Biochem Soc Trans, 38(6), pp. 1615-20.

Nilsen, T. W. and Graveley, B. R. (2010) 'Expansion of the eukaryotic proteome by alternative splicing', Nature, 463(7280), pp. 457-63.

Nimchuk, Z. L., Fisher, E. J., Desveaux, D., Chang, J. H. and Dangl, J. L. (2007) 'The HopX (AvrPphE) family of Pseudomonas syringae type III effectors require a catalytic triad and a novel N-terminal domain for function', Mol Plant Microbe Interact, 20(4), pp. 346-57.

Nomura, K., Mecey, C., Lee, Y. N., Imboden, L. A., Chang, J. H. and He, S. Y. (2011) 'Effector-triggered immunity blocks pathogen degradation of an immunity-associated vesicle traffic regulator in Arabidopsis', Proc Natl Acad Sci U S A, 108(26), pp. 10774-9.

Nuruzzaman, M., Sharoni, A. M. and Kikuchi, S. (2013) 'Roles of NAC transcription factors in the regulation of biotic and abiotic stress responses in plants', Front Microbiol, 4, pp. 248.

Nürnberger, T., Brunner, F., Kemmerling, B. and Piater, L. (2004) 'Innate immunity in plants and animals: striking similarities and obvious differences', Immunol Rev, 198, pp. 249-66.

O'Brien, J. A., Daudi, A., Butt, V. S. and Bolwell, G. P. (2012) 'Reactive oxygen species and their role in plant defence and cell wall metabolism', Planta, 236(3), pp. 765-79.

Ogata, K., Kanei-Ishii, C., Sasaki, M., Hatanaka, H., Nagadoi, A., Enari, M., Nakamura, H., Nishimura, Y., Ishii, S. and Sarai, A. (1996) 'The cavity in the hydrophobic core of Myb DNA-binding domain is reserved for DNA recognition and trans-activation', Nat Struct Biol, 3(2), pp. 178-87.

Ohta, Y., Hojo, H., Aimoto, S., Kobayashi, T., Zhu, X., Jordan, F. and Inouye, M. (1991) 'Pro-peptide as an intramolecular chaperone: renaturation of denatured subtilisin E with a synthetic pro-peptide [corrected]', Mol Microbiol, 5(6), pp. 1507-10.

Okada, S., Nagabuchi, M., Takamura, Y., Nakagawa, T., Shinmyozu, K., Nakayama, J. and Tanaka, K. (2009) 'Reconstitution of Arabidopsis thaliana SUMO pathways in E. coli: functional evaluation of SUMO machinery proteins and mapping of SUMOylation sites by mass spectrometry', Plant Cell Physiol, 50(6), pp. 1049-61.

Olvera-Carrillo, Y., Van Bel, M., Van Hautegem, T., Fendrych, M., Huysmans, M., Simaskova, M., van Durme, M., Buscaill, P., Rivas, S., S Coll, N., Coppens, F., Maere, S. and Nowack, M. K. (2015) 'A Conserved Core of Programmed Cell Death Indicator Genes Discriminates Developmentally and Environmentally Induced Programmed Cell Death in Plants', Plant Physiol, 169(4), pp. 2684-99.

Osbourn, A. E. (1996) 'Preformed Antimicrobial Compounds and Plant Defense against Fungal Attack', Plant Cell, 8(10), pp. 1821-1831.

Ossovskaya, V. S. and Bunnett, N. W. (2004) 'Protease-activated receptors: contribution to physiology and disease', Physiol Rev, 84(2), pp. 579-621.

Ottmann, C., Rose, R., Huttenlocher, F., Cedzich, A., Hauske, P., Kaiser, M., Huber, R. and Schaller, A. (2009) 'Structural basis for Ca2+-independence and activation by homodimerization of tomato subtilase 3', Proc Natl Acad Sci U S A, 106(40), pp. 17223-8.

131

Oñate-Sánchez, L. and Singh, K. B. (2002) 'Identification of Arabidopsis ethylene-responsive element binding factors with distinct induction kinetics after pathogen infection', Plant Physiol, 128(4), pp. 1313-22.

Pabo, C. O. and Sauer, R. T. (1992) 'Transcription factors: structural families and principles of DNA recognition', Annu Rev Biochem, 61, pp. 1053-95.

Pandey, S. P. and Somssich, I. E. (2009) 'The role of WRKY transcription factors in plant immunity', Plant Physiol, 150(4), pp. 1648-55.

Pautot, V., Holzer, F. M., Chaufaux, J. and Walling, L. L. (2001) 'The induction of tomato leucine aminopeptidase genes (LapA) after Pseudomonas syringae pv. tomato infection is primarily a wound response triggered by coronatine', Mol Plant Microbe Interact, 14(2), pp. 214-24.

Pautot, V., Holzer, F. M., Reisch, B. and Walling, L. L. (1993) 'Leucine aminopeptidase: an inducible component of the defense response in Lycopersicon esculentum (tomato)', Proc Natl Acad Sci U S A, 90(21), pp. 9906-10.

Paz-Ares, J., Ghosal, D., Wienand, U., Peterson, P. A. and Saedler, H. (1987) 'The regulatory c1 locus of Zea mays encodes a protein with homology to myb proto-oncogene products and with structural similarities to transcriptional activators', EMBO J, 6(12), pp. 3553-8.

Pearce, G., Yamaguchi, Y., Barona, G. and Ryan, C. A. (2010) 'A subtilisin-like protein from soybean contains an embedded, cryptic signal that activates defense-related genes', Proc Natl Acad Sci U S A, 107(33), pp. 14921-5.

Pechan, T., Cohen, A., Williams, W. P. and Luthe, D. S. (2002) 'Insect feeding mobilizes a unique plant defense protease that disrupts the peritrophic matrix of caterpillars', Proc Natl Acad Sci U S A, 99(20), pp. 13319-23.

Pechan, T., Ye, L., Chang, Y., Mitra, A., Lin, L., Davis, F. M., Williams, W. P. and Luthe, D. S. (2000) 'A unique 33-kD cysteine proteinase accumulates in response to larval feeding in maize genotypes resistant to fall armyworm and other Lepidoptera', Plant Cell, 12(7), pp. 1031-40.

Pedras, M. S. and Yaya, E. E. (2015) 'Plant chemical defenses: are all constitutive antimicrobial metabolites phytoanticipins?', Nat Prod Commun, 10(1), pp. 209-18.

Persak, H. and Pitzschke, A. (2013) 'Tight interconnection and multi-level control of Arabidopsis MYB44 in MAPK cascade signalling', PLoS One, 8(2), pp. e57547.

Pesch, M. and Hülskamp, M. (2009) 'One, two, three...models for trichome patterning in Arabidopsis?', Curr Opin Plant Biol, 12(5), pp. 587-92.

Petre, B., Lorrain, C., Saunders, D. G., Win, J., Sklenar, J., Duplessis, S. and Kamoun, S. (2015) 'Rust fungal effectors mimic host transit peptides to translocate into chloroplasts', Cell Microbiol.

Pieterse, C. M., Leon-Reyes, A., Van der Ent, S. and Van Wees, S. C. (2009) 'Networking by small-molecule hormones in plant immunity', Nat Chem Biol, 5(5), pp. 308-16.

Pieterse, C. M., Zamioudis, C., Berendsen, R. L., Weller, D. M., Van Wees, S. C. and Bakker, P. A. (2014) 'Induced systemic resistance by beneficial microbes', Annu Rev Phytopathol, 52, pp. 347-75.

Pillitteri, L. J. and Dong, J. (2013) 'Stomatal development in Arabidopsis', Arabidopsis Book, 11, pp. e0162.

Pireyre, M. and Burow, M. (2015) 'Regulation of MYB and bHLH transcription factors: a glance at the protein level', Mol Plant, 8(3), pp. 378-88.

Plattner, S., Gruber, C., Altmann, F. and Bohlmann, H. (2014) 'Self-processing of a barley subtilase expressed in E. coli', Protein Expr Purif, 101, pp. 76-83.

Porter, B. W., Yuen, C. Y. and Christopher, D. A. (2015) 'Dual protein trafficking to secretory and non-secretory cell compartments: clear or double vision?', Plant Sci, 234, pp. 174-9.

Prouse, M. B. and Campbell, M. M. (2012) 'The interaction between MYB proteins and their target DNA binding sites', Biochim Biophys Acta, 1819(1), pp. 67-77.

Puranik, S., Sahu, P. P., Srivastava, P. S. and Prasad, M. (2012) 'NAC proteins: regulation and role in stress tolerance', Trends Plant Sci, 17(6), pp. 369-81.

132

Raffaele, S., Leger, A. and Roby, D. (2009) 'Very long chain fatty acid and lipid signaling in the response of plants to pathogens', Plant Signal Behav, 4(2), pp. 94-9.

Raffaele, S. and Rivas, S. (2013) 'Regulate and be regulated: integration of defense and other signals by the AtMYB30 transcription factor', Front Plant Sci, 4, pp. 98.

Raffaele, S., Rivas, S. and Roby, D. (2006) 'An essential role for salicylic acid in AtMYB30-mediated control of the hypersensitive cell death program in Arabidopsis', FEBS Lett, 580(14), pp. 3498-504.

Raffaele, S., Vailleau, F., Léger, A., Joubès, J., Miersch, O., Huard, C., Blée, E., Mongrand, S., Domergue, F. and Roby, D. (2008) 'A MYB transcription factor regulates very-long-chain fatty acid biosynthesis for activation of the hypersensitive cell death response in Arabidopsis', Plant Cell, 20(3), pp. 752-67.

Rairdan, G. and Moffett, P. (2007) 'Brothers in arms? Common and contrasting themes in pathogen perception by plant NB-LRR and animal NACHT-LRR proteins', Microbes Infect, 9(5), pp. 677-86.

Ramírez, V., López, A., Mauch-Mani, B., Gil, M. J. and Vera, P. (2013) 'An extracellular subtilase switch for immune priming in Arabidopsis', PLoS Pathog, 9(6), pp. e1003445.

Ranf, S., Eschen-Lippold, L., Pecher, P., Lee, J. and Scheel, D. (2011) 'Interplay between calcium signalling and early signalling elements during defence responses to microbe- or damage-associated molecular patterns', Plant J, 68(1), pp. 100-13.

Rautengarten, C., Steinhauser, D., Büssis, D., Stintzi, A., Schaller, A., Kopka, J. and Altmann, T. (2005) 'Inferring hypotheses on functional relationships of genes: Analysis of the Arabidopsis thaliana subtilase gene family', PLoS Comput Biol, 1(4), pp. e40.

Rautengarten, C., Usadel, B., Neumetzler, L., Hartmann, J., Büssis, D. and Altmann, T. (2008) 'A subtilisin-like serine protease essential for mucilage release from Arabidopsis seed coats', Plant J, 54(3), pp. 466-80.

Rawlings, N. D. and Barrett, A. J. (1995) 'Families of aspartic peptidases, and those of unknown catalytic mechanism', Methods Enzymol, 248, pp. 105-20.

Rawlings, N. D., Waller, M., Barrett, A. J. and Bateman, A. (2014) 'MEROPS: the database of proteolytic enzymes, their substrates and inhibitors', Nucleic Acids Res, 42(Database issue), pp. D503-9.

Rayson, S., Arciga-Reyes, L., Wootton, L., De Torres Zabala, M., Truman, W., Graham, N., Grant, M. and Davies, B. (2012) 'A role for nonsense-mediated mRNA decay in plants: pathogen responses are induced in Arabidopsis thaliana NMD mutants', PLoS One, 7(2), pp. e31917.

Reddy, A. S., Ali, G. S., Celesnik, H. and Day, I. S. (2011) 'Coping with stresses: roles of calcium- and calcium/calmodulin-regulated gene expression', Plant Cell, 23(6), pp. 2010-32.

Reina-Pinto, J. J. and Yephremov, A. (2009) 'Surface lipids and plant defenses', Plant Physiol Biochem, 47(6), pp. 540-9.

Reyes, F. C., Buono, R. and Otegui, M. S. (2011) 'Plant endosomal trafficking pathways', Curr Opin Plant Biol, 14(6), pp. 666-73.

Richter, S., Voss, U. and Jürgens, G. (2009) 'Post-Golgi traffic in plants', Traffic, 10(7), pp. 819-28.

Riechmann, J. L., Heard, J., Martin, G., Reuber, L., Jiang, C., Keddie, J., Adam, L., Pineda, O., Ratcliffe, O. J., Samaha, R. R., Creelman, R., Pilgrim, M., Broun, P., Zhang, J. Z., Ghandehari, D., Sherman, B. K. and Yu, G. (2000) 'Arabidopsis transcription factors: genome-wide comparative analysis among eukaryotes', Science, 290(5499), pp. 2105-10.

Rivas, S. (2012) 'Nuclear dynamics during plant innate immunity', Plant Physiol, 158(1), pp. 87-94.

Rivas, S. and Thomas, C. M. (2005) 'Molecular interactions between tomato and the leaf mold pathogen Cladosporium fulvum', Annu Rev Phytopathol, 43, pp. 395-436.

133

Robatzek, S., Chinchilla, D. and Boller, T. (2006) 'Ligand-induced endocytosis of the pattern recognition receptor FLS2 in Arabidopsis', Genes Dev, 20(5), pp. 537-42.

Robert, S., Chary, S. N., Drakakaki, G., Li, S., Yang, Z., Raikhel, N. V. and Hicks, G. R. (2008) 'Endosidin1 defines a compartment involved in endocytosis of the brassinosteroid receptor BRI1 and the auxin transporters PIN2 and AUX1', Proc Natl Acad Sci U S A, 105(24), pp. 8464-9.

Robert-Seilaniantz, A., Grant, M. and Jones, J. D. (2011) 'Hormone crosstalk in plant disease and defense: more than just jasmonate-salicylate antagonism', Annu Rev Phytopathol, 49, pp. 317-43.

Robinson, D. G., Langhans, M., Saint-Jore-Dupas, C. and Hawes, C. (2008) 'BFA effects are tissue and not just plant specific', Trends Plant Sci, 13(8), pp. 405-8.

Rojo, E., Martín, R., Carter, C., Zouhar, J., Pan, S., Plotnikova, J., Jin, H., Paneque, M., Sánchez-Serrano, J. J., Baker, B., Ausubel, F. M. and Raikhel, N. V. (2004) 'VPEgamma exhibits a caspase-like activity that contributes to defense against pathogens', Curr Biol, 14(21), pp. 1897-906.

Romero, I., Fuertes, A., Benito, M. J., Malpica, J. M., Leyva, A. and Paz-Ares, J. (1998) 'More than 80R2R3-MYB regulatory genes in the genome of Arabidopsis thaliana', Plant J, 14(3), pp. 273-84.

Rooney, H. C., Van't Klooster, J. W., van der Hoorn, R. A., Joosten, M. H., Jones, J. D. and de Wit, P. J. (2005) 'Cladosporium Avr2 inhibits tomato Rcr3 protease required for Cf-2-dependent disease resistance', Science, 308(5729), pp. 1783-6.

Rosinski, J. A. and Atchley, W. R. (1998) 'Molecular evolution of the Myb family of transcription factors: evidence for polyphyletic origin', J Mol Evol, 46(1), pp. 74-83.

Roux, M., Schwessinger, B., Albrecht, C., Chinchilla, D., Jones, A., Holton, N., Malinovsky, F. G., Tör, M., de Vries, S. and Zipfel, C. (2011) 'The Arabidopsis leucine-rich repeat receptor-like kinases BAK1/SERK3 and BKK1/SERK4 are required for innate immunity to hemibiotrophic and biotrophic pathogens', Plant Cell, 23(6), pp. 2440-55.

Rozman-Pungercar, J., Kopitar-Jerala, N., Bogyo, M., Turk, D., Vasiljeva, O., Stefe, I., Vandenabeele, P., Brömme, D., Puizdar, V., Fonović , M., Trstenjak-Prebanda, M., Dolenc, I., Turk, V. and Turk, B. (2003) 'Inhibition of papain-like cysteine proteases and legumain by caspase-specific inhibitors: when reaction mechanism is more important than specificity', Cell Death Differ, 10(8), pp. 881-8.

Rushton, P. J., Somssich, I. E., Ringler, P. and Shen, Q. J. (2010) 'WRKY transcription factors', Trends Plant Sci, 15(5), pp. 247-58.

Ryan, C. A. and Pearce, G. (2003) 'Systemins: a functionally defined family of peptide signals that regulate defensive genes in Solanaceae species', Proc Natl Acad Sci U S A, 100 Suppl 2, pp. 14577-80.

Sakuma, Y., Liu, Q., Dubouzet, J. G., Abe, H., Shinozaki, K. and Yamaguchi-Shinozaki, K. (2002) 'DNA-binding specificity of the ERF/AP2 domain of Arabidopsis DREBs, transcription factors involved in dehydration- and cold-inducible gene expression', Biochem Biophys Res Commun, 290(3), pp. 998-1009.

Salomon, S. and Robatzek, S. (2006) 'Induced Endocytosis of the Receptor Kinase FLS2', Plant Signal Behav, 1(6), pp. 293-5.

Sanderfoot, A. A., Kovaleva, V., Bassham, D. C. and Raikhel, N. V. (2001) 'Interactions between syntaxins identify at least five SNARE complexes within the Golgi/prevacuolar system of the Arabidopsis cell', Mol Biol Cell, 12(12), pp. 3733-43.

Sarkar, A. and Wintrode, P. L. (2011) 'Effects of glycosylation on the stability and flexibility of a metastable protein: the human serpin α (1)-antitrypsin', Int J Mass Spectrom, 302(1-3), pp. 69-75.

Sarris, P. F., Duxbury, Z., Huh, S. U., Ma, Y., Segonzac, C., Sklenar, J., Derbyshire, P., Cevik, V., Rallapalli, G., Saucet, S. B., Wirthmueller, L., Menke, F. L., Sohn, K. H. and Jones, J. D. (2015) 'A Plant Immune Receptor Detects Pathogen Effectors that Target WRKY Transcription Factors', Cell, 161(5), pp. 1089-100.

Sawinski, K., Mersmann, S., Robatzek, S. and Böhmer, M. (2013) 'Guarding the green: pathways to stomatal immunity', Mol Plant Microbe Interact, 26(6), pp. 626-32.

134

Schaffer, R., Landgraf, J., Accerbi, M., Simon, V., Larson, M. and Wisman, E. (2001) 'Microarray analysis of diurnal and circadian-regulated genes in Arabidopsis', Plant Cell, 13(1), pp. 113-23.

Schaller, A., Stintzi, A. and Graff, L. (2012) 'Subtilases - versatile tools for protein turnover, plant development, and interactions with the environment', Physiologia Plantarum, 145(1), pp. 52-66.

Scheuring, D., Viotti, C., Krüger, F., Künzl, F., Sturm, S., Bubeck, J., Hillmer, S., Frigerio, L., Robinson, D. G., Pimpl, P. and Schumacher, K. (2011) 'Multivesicular bodies mature from the trans-Golgi network/early endosome in Arabidopsis', Plant Cell, 23(9), pp. 3463-81.

Schiermeyer, A., Hartenstein, H., Mandal, M. K., Otte, B., Wahner, V. and Schillberg, S. (2009) 'A membrane-bound matrix-metalloproteinase from Nicotiana tabacum cv. BY-2 is induced by bacterial pathogens', BMC Plant Biol, 9, pp. 83.

Schlüter, U., Muschak, M., Berger, D. and Altmann, T. (2003) 'Photosynthetic performance of an Arabidopsis mutant with elevated stomatal density (sdd1-1) under different light regimes', J Exp Bot, 54(383), pp. 867-74.

Schwab, F., Zhai, G., Kern, M., Turner, A., Schnoor, J. L. and Wiesner, M. R. (2015) 'Barriers, pathways and processes for uptake, translocation and accumulation of nanomaterials in plants - Critical review', Nanotoxicology, pp. 1-22.

Schwessinger, B., Roux, M., Kadota, Y., Ntoukakis, V., Sklenar, J., Jones, A. and Zipfel, C. (2011) 'Phosphorylation-dependent differential regulation of plant growth, cell death, and innate immunity by the regulatory receptor-like kinase BAK1', PLoS Genet, 7(4), pp. e1002046.

Schähs, M., Strasser, R., Stadlmann, J., Kunert, R., Rademacher, T. and Steinkellner, H. (2007) 'Production of a monoclonal antibody in plants with a humanized N-glycosylation pattern', Plant Biotechnol J, 5(5), pp. 657-63.

Senthil-Kumar, M. and Mysore, K. S. (2013) 'Nonhost resistance against bacterial pathogens: retrospectives and prospects', Annu Rev Phytopathol, 51, pp. 407-27.

Seo, E., Choi, D. and Choi (2015) 'Functional studies of transcription factors involved in plant defenses in the genomics era', Brief Funct Genomics, 14(4), pp. 260-7.

Seo, P. J., Hong, S. Y., Kim, S. G. and Park, C. M. (2011a) 'Competitive inhibition of transcription factors by small interfering peptides', Trends Plant Sci, 16(10), pp. 541-9.

Seo, P. J., Kim, M. J., Park, J. Y., Kim, S. Y., Jeon, J., Lee, Y. H., Kim, J. and Park, C. M. (2010a) 'Cold activation of a plasma membrane-tethered NAC transcription factor induces a pathogen resistance response in Arabidopsis', Plant J, 61(4), pp. 661-71.

Seo, P. J., Kim, M. J., Song, J. S., Kim, Y. S., Kim, H. J. and Park, C. M. (2010b) 'Proteolytic processing of an Arabidopsis membrane-bound NAC transcription factor is triggered by cold-induced changes in membrane fluidity', Biochem J, 427(3), pp. 359-67.

Seo, P. J., Kim, S. G. and Park, C. M. (2008) 'Membrane-bound transcription factors in plants', Trends Plant Sci, 13(10), pp. 550-6.

Seo, P. J., Lee, S. B., Suh, M. C., Park, M. J., Go, Y. S. and Park, C. M. (2011b) 'The MYB96 transcription factor regulates cuticular wax biosynthesis under drought conditions in Arabidopsis', Plant Cell, 23(3), pp. 1138-52.

Seo, P. J. and Park, C. M. (2010) 'MYB96-mediated abscisic acid signals induce pathogen resistance response by promoting salicylic acid biosynthesis in Arabidopsis', New Phytol, 186(2), pp. 471-83.

Seo, P. J., Xiang, F., Qiao, M., Park, J. Y., Lee, Y. N., Kim, S. G., Lee, Y. H., Park, W. J. and Park, C. M. (2009) 'The MYB96 transcription factor mediates abscisic acid signaling during drought stress response in Arabidopsis', Plant Physiol, 151(1), pp. 275-89.

Serrano, M., Coluccia, F., Torres, M., L'Haridon, F. and Métraux, J. P. (2014) 'The cuticle and plant defense to pathogens', Front Plant Sci, 5, pp. 274.

Sharfman, M., Bar, M., Ehrlich, M., Schuster, S., Melech-Bonfil, S., Ezer, R., Sessa, G. and Avni, A. (2011) 'Endosomal signaling of the tomato leucine-rich repeat receptor-like protein LeEix2', Plant J, 68(3), pp. 413-23.

135

Shim, J. S. and Choi, Y. D. (2013) 'Direct regulation of WRKY70 by AtMYB44 in plant defense responses', Plant Signal Behav, 8(6), pp. e20783.

Shin, R., Burch, A. Y., Huppert, K. A., Tiwari, S. B., Murphy, A. S., Guilfoyle, T. J. and Schachtman, D. P. (2007) 'The Arabidopsis transcription factor MYB77 modulates auxin signal transduction', Plant Cell, 19(8), pp. 2440-53.

Shindo, T., Misas-Villamil, J. C., Hörger, A. C., Song, J. and van der Hoorn, R. A. (2012) 'A role in immunity for Arabidopsis cysteine protease RD21, the ortholog of the tomato immune protease C14', PLoS One, 7(1), pp. e29317.

Shirley, A. M. and Chapple, C. (2003) 'Biochemical characterization of sinapoylglucose:choline sinapoyltransferase, a serine carboxypeptidase-like protein that functions as an acyltransferase in plant secondary metabolism', J Biol Chem, 278(22), pp. 19870-7.

Shirley, A. M., McMichael, C. M. and Chapple, C. (2001) 'The sng2 mutant of Arabidopsis is defective in the gene encoding the serine carboxypeptidase-like protein sinapoylglucose:choline sinapoyltransferase', Plant J, 28(1), pp. 83-94.

Siezen, R. J. and Leunissen, J. A. (1997) 'Subtilases: the superfamily of subtilisin-like serine proteases', Protein Sci, 6(3), pp. 501-23.

Simon, M., Lee, M. M., Lin, Y., Gish, L. and Schiefelbein, J. (2007) 'Distinct and overlapping roles of single-repeat MYB genes in root epidermal patterning', Dev Biol, 311(2), pp. 566-78.

Solomon, M., Belenghi, B., Delledonne, M., Menachem, E. and Levine, A. (1999) 'The involvement of cysteine proteases and protease inhibitor genes in the regulation of programmed cell death in plants', Plant Cell, 11(3), pp. 431-44.

Song, J., Win, J., Tian, M., Schornack, S., Kaschani, F., Ilyas, M., van der Hoorn, R. A. and Kamoun, S. (2009) 'Apoplastic effectors secreted by two unrelated eukaryotic plant pathogens target the tomato defense protease Rcr3', Proc Natl Acad Sci U S A, 106(5), pp. 1654-9.

Spoel, S. H., Johnson, J. S. and Dong, X. (2007) 'Regulation of tradeoffs between plant defenses against pathogens with different lifestyles', Proc Natl Acad Sci U S A, 104(47), pp. 18842-7.

Staiger, D. and Brown, J. W. (2013) 'Alternative splicing at the intersection of biological timing, development, and stress responses', Plant Cell, 25(10), pp. 3640-56.

Stracke, R., Werber, M. and Weisshaar, B. (2001) 'The R2R3-MYB gene family in Arabidopsis thaliana', Curr Opin Plant Biol, 4(5), pp. 447-56.

Sugimoto, K., Takeda, S. and Hirochika, H. (2000) 'MYB-related transcription factor NtMYB2 induced by wounding and elicitors is a regulator of the tobacco retrotransposon Tto1 and defense-related genes', Plant Cell, 12(12), pp. 2511-2528.

Syed, N. H., Kalyna, M., Marquez, Y., Barta, A. and Brown, J. W. (2012) 'Alternative splicing in plants--coming of age', Trends Plant Sci, 17(10), pp. 616-23.

Takagi, H., Koga, M., Katsurada, S., Yabuta, Y., Shinde, U., Inouye, M. and Nakamori, S. (2001) 'Functional analysis of the propeptides of subtilisin E and aqualysin I as intramolecular chaperones', FEBS Lett, 508(2), pp. 210-4.

Takemoto, D., Rafiqi, M., Hurley, U., Lawrence, G. J., Bernoux, M., Hardham, A. R., Ellis, J. G., Dodds, P. N. and Jones, D. A. (2012) 'N-terminal motifs in some plant disease resistance proteins function in membrane attachment and contribute to disease resistance', Mol Plant Microbe Interact, 25(3), pp. 379-92.

Taki, N., Sasaki-Sekimoto, Y., Obayashi, T., Kikuta, A., Kobayashi, K., Ainai, T., Yagi, K., Sakurai, N., Suzuki, H., Masuda, T., Takamiya, K., Shibata, D., Kobayashi, Y. and Ohta, H. (2005) '12-oxo-phytodienoic acid triggers expression of a distinct set of genes and plays a role in wound-induced gene expression in Arabidopsis', Plant Physiol, 139(3), pp. 1268-83.

Takken, F. L. and Goverse, A. (2012) 'How to build a pathogen detector: structural basis of NB-LRR function', Curr Opin Plant Biol, 15(4), pp. 375-84.

Tanaka, H., Onouchi, H., Kondo, M., Hara-Nishimura, I., Nishimura, M., Machida, C. and Machida, Y. (2001) 'A subtilisin-like serine protease is required for epidermal surface

136

formation in Arabidopsis embryos and juvenile plants', Development, 128(23), pp. 4681-9.

Tang, X., Xiao, Y. and Zhou, J. M. (2006) 'Regulation of the type III secretion system in phytopathogenic bacteria', Mol Plant Microbe Interact, 19(11), pp. 1159-66.

Tavares, C. P., Vernal, J., Delena, R. A., Lamattina, L., Cassia, R. and Terenzi, H. (2014) 'S-nitrosylation influences the structure and DNA binding activity of AtMYB30 transcription factor from Arabidopsis thaliana', Biochim Biophys Acta, 1844(4), pp. 810-7.

Taylor, M. A., Baker, K. C., Briggs, G. S., Connerton, I. F., Cummings, N. J., Pratt, K. A., Revell, D. F., Freedman, R. B. and Goodenough, P. W. (1995) 'Recombinant pro-regions from papain and papaya proteinase IV-are selective high affinity inhibitors of the mature papaya enzymes', Protein Eng, 8(1), pp. 59-62.

Taylor, N. A., Shennan, K. I., Cutler, D. F. and Docherty, K. (1997) 'Mutations within the propeptide, the primary cleavage site or the catalytic site, or deletion of C-terminal sequences, prevents secretion of proPC2 from transfected COS-7 cells', Biochem J, 321 ( Pt 2), pp. 367-73.

Thomma, B. P., Eggermont, K., Penninckx, I. A., Mauch-Mani, B., Vogelsang, R., Cammue, B. P. and Broekaert, W. F. (1998) 'Separate jasmonate-dependent and salicylate-dependent defense-response pathways in Arabidopsis are essential for resistance to distinct microbial pathogens', Proc Natl Acad Sci U S A, 95(25), pp. 15107-11.

Thomma, B. P., Eggermont, K., Tierens, K. F. and Broekaert, W. F. (1999) 'Requirement of functional ethylene-insensitive 2 gene for efficient resistance of Arabidopsis to infection by Botrytis cinerea', Plant Physiol, 121(4), pp. 1093-102.

Tian, M., Benedetti, B. and Kamoun, S. (2005) 'A Second Kazal-like protease inhibitor from Phytophthora infestans inhibits and interacts with the apoplastic pathogenesis-related protease P69B of tomato', Plant Physiol, 138(3), pp. 1785-93.

Tian, M., Huitema, E., Da Cunha, L., Torto-Alalibo, T. and Kamoun, S. (2004) 'A Kazal-like extracellular serine protease inhibitor from Phytophthora infestans targets the tomato pathogenesis-related protease P69B', J Biol Chem, 279(25), pp. 26370-7.

Tian, M. and Kamoun, S. (2005) 'A two disulfide bridge Kazal domain from Phytophthora exhibits stable inhibitory activity against serine proteases of the subtilisin family', BMC Biochem, 6, pp. 15.

Tian, M., Win, J., Song, J., van der Hoorn, R., van der Knaap, E. and Kamoun, S. (2007) 'A Phytophthora infestans cystatin-like protein targets a novel tomato papain-like apoplastic protease', Plant Physiol, 143(1), pp. 364-77.

Tornero, P., Conejero, V. and Vera, P. (1996) 'Primary structure and expression of a pathogen-induced protease (PR-P69) in tomato plants: Similarity of functional domains to subtilisin-like endoproteases', Proc Natl Acad Sci U S A, 93(13), pp. 6332-7.

Tornero, P., Conejero, V. and Vera, P. (1997) 'Identification of a new pathogen-induced member of the subtilisin-like processing protease family from plants', J Biol Chem, 272(22), pp. 14412-9.

Torres, M. A., Dangl, J. L. and Jones, J. D. (2002) 'Arabidopsis gp91phox homologues AtrbohD and AtrbohF are required for accumulation of reactive oxygen intermediates in the plant defense response', Proc Natl Acad Sci U S A, 99(1), pp. 517-22.

Toth, I. K., Bell, K. S., Holeva, M. C. and Birch, P. R. (2003) 'Soft rot erwiniae: from genes to genomes', Mol Plant Pathol, 4(1), pp. 17-30.

Toth, I. K. and Birch, P. R. (2005) 'Rotting softly and stealthily', Curr Opin Plant Biol, 8(4), pp. 424-9.

Traw, M. B. and Bergelson, J. (2003) 'Interactive effects of jasmonic acid, salicylic acid, and gibberellin on induction of trichomes in Arabidopsis', Plant Physiol, 133(3), pp. 1367-75.

Tripathi, L. P. and Sowdhamini, R. (2006) 'Cross genome comparisons of serine proteases in Arabidopsis and rice', BMC Genomics, 7, pp. 200.

137

Tse, Y. C., Mo, B., Hillmer, S., Zhao, M., Lo, S. W., Robinson, D. G. and Jiang, L. (2004) 'Identification of multivesicular bodies as prevacuolar compartments in Nicotiana tabacum BY-2 cells', Plant Cell, 16(3), pp. 672-93.

Tsuda, K. and Katagiri, F. (2010) 'Comparing signaling mechanisms engaged in pattern-triggered and effector-triggered immunity', Curr Opin Plant Biol, 13(4), pp. 459-65.

Tsuda, K. and Somssich, I. E. (2015) 'Transcriptional networks in plant immunity', New Phytol, 206(3), pp. 932-47.

Ueda, T., Uemura, T., Sato, M. H. and Nakano, A. (2004) 'Functional differentiation of endosomes in Arabidopsis cells', Plant J, 40(5), pp. 783-9.

Ueda, T., Yamaguchi, M., Uchimiya, H. and Nakano, A. (2001) 'Ara6, a plant-unique novel type Rab GTPase, functions in the endocytic pathway of Arabidopsis thaliana', EMBO J, 20(17), pp. 4730-41.

Ulker, B. and Somssich, I. E. (2004) 'WRKY transcription factors: from DNA binding towards biological function', Curr Opin Plant Biol, 7(5), pp. 491-8.

Vailleau, F., Daniel, X., Tronchet, M., Montillet, J. L., Triantaphylidès, C. and Roby, D. (2002) 'A R2R3-MYB gene, AtMYB30, acts as a positive regulator of the hypersensitive cell death program in plants in response to pathogen attack', Proc Natl Acad Sci U S A, 99(15), pp. 10179-84.

VAN Baarlen, P., Woltering, E. J., Staats, M. and VAN Kan, J. A. (2007) 'Histochemical and genetic analysis of host and non-host interactions of Arabidopsis with three Botrytis species: an important role for cell death control', Mol Plant Pathol, 8(1), pp. 41-54.

Van der Ent, S., Verhagen, B. W., Van Doorn, R., Bakker, D., Verlaan, M. G., Pel, M. J., Joosten, R. G., Proveniers, M. C., Van Loon, L. C., Ton, J. and Pieterse, C. M. (2008) 'MYB72 is required in early signaling steps of rhizobacteria-induced systemic resistance in Arabidopsis', Plant Physiol, 146(3), pp. 1293-304.

van der Hoorn, R. A. (2008) 'Plant proteases: from phenotypes to molecular mechanisms', Annu Rev Plant Biol, 59, pp. 191-223.

van der Hoorn, R. A. and Jones, J. D. (2004) 'The plant proteolytic machinery and its role in defence', Curr Opin Plant Biol, 7(4), pp. 400-7.

van der Hoorn, R. A. and Kamoun, S. (2008) 'From Guard to Decoy: a new model for perception of plant pathogen effectors', Plant Cell, 20(8), pp. 2009-17.

van der Hoorn, R. A. L. and Kaiser, M. (2012) 'Probes for activity-based profiling of plant proteases', Physiologia Plantarum, 145(1), pp. 18-27.

Van Hautegem, T., Waters, A. J., Goodrich, J. and Nowack, M. K. (2015) 'Only in dying, life: programmed cell death during plant development', Trends Plant Sci, 20(2), pp. 102-13.

Van't Klooster, J. W., Van der Kamp, M. W., Vervoort, J., Beekwilder, J., Boeren, S., Joosten, M. H., Thomma, B. P. and De Wit, P. J. (2011) 'Affinity of Avr2 for tomato cysteine protease Rcr3 correlates with the Avr2-triggered Cf-2-mediated hypersensitive response', Mol Plant Pathol, 12(1), pp. 21-30.

Verhagen, B. W., Glazebrook, J., Zhu, T., Chang, H. S., van Loon, L. C. and Pieterse, C. M. (2004) 'The transcriptome of rhizobacteria-induced systemic resistance in arabidopsis', Mol Plant Microbe Interact, 17(8), pp. 895-908.

Von Groll, U., Berger, D. and Altmann, T. (2002) 'The subtilisin-like serine protease SDD1 mediates cell-to-cell signaling during Arabidopsis stomatal development', Plant Cell, 14(7), pp. 1527-39.

Wang, C., Eufemi, M., Turano, C. and Giartosio, A. (1996) 'Influence of the carbohydrate moiety on the stability of glycoproteins', Biochemistry, 35(23), pp. 7299-307.

Wang, F., Shang, Y., Fan, B., Yu, J. Q. and Chen, Z. (2014) 'Arabidopsis LIP5, a positive regulator of multivesicular body biogenesis, is a critical target of pathogen-responsive MAPK cascade in plant basal defense', PLoS Pathog, 10(7), pp. e1004243.

Wang, J., Cai, Y., Miao, Y., Lam, S. K. and Jiang, L. (2009) 'Wortmannin induces homotypic fusion of plant prevacuolar compartments', J Exp Bot, 60(11), pp. 3075-83.

Wang, X., Feng, H., Tang, C., Bai, P., Wei, G., Huang, L. and Kang, Z. (2012) 'TaMCA4, a novel wheat metacaspase gene functions in programmed cell death induced by the

138

fungal pathogen Puccinia striiformis f. sp. tritici', Mol Plant Microbe Interact, 25(6), pp. 755-64.

Watanabe, N. and Lam, E. (2011) 'Calcium-dependent activation and autolysis of Arabidopsis metacaspase 2d', J Biol Chem, 286(12), pp. 10027-40.

Wildermuth, M. C., Dewdney, J., Wu, G. and Ausubel, F. M. (2001) 'Isochorismate synthase is required to synthesize salicylic acid for plant defence', Nature, 414(6863), pp. 562-5.

Wilkins, O., Nahal, H., Foong, J., Provart, N. J. and Campbell, M. M. (2009) 'Expansion and diversification of the Populus R2R3-MYB family of transcription factors', Plant Physiol, 149(2), pp. 981-93.

Win, J., Chaparro-Garcia, A., Belhaj, K., Saunders, D. G., Yoshida, K., Dong, S., Schornack, S., Zipfel, C., Robatzek, S., Hogenhout, S. A. and Kamoun, S. (2012) 'Effector biology of plant-associated organisms: concepts and perspectives', Cold Spring Harb Symp Quant Biol, 77, pp. 235-47.

Wu, L., Chen, H., Curtis, C. and Fu, Z. Q. (2014) 'Go in for the kill', Virulence, 5(7), pp. 710-21.

Wu, S., Lu, D., Kabbage, M., Wei, H. L., Swingle, B., Records, A. R., Dickman, M., He, P. and Shan, L. (2011) 'Bacterial effector HopF2 suppresses arabidopsis innate immunity at the plasma membrane', Mol Plant Microbe Interact, 24(5), pp. 585-93.

Xia, Y. (2004) 'Proteases in pathogenesis and plant defence', Cell Microbiol, 6(10), pp. 905-13.

Xia, Y., Suzuki, H., Borevitz, J., Blount, J., Guo, Z., Patel, K., Dixon, R. A. and Lamb, C. (2004) 'An extracellular aspartic protease functions in Arabidopsis disease resistance signaling', EMBO J, 23(4), pp. 980-8.

Xu, C. and Ng, D. T. (2015) 'Glycosylation-directed quality control of protein folding', Nat Rev Mol Cell Biol, 16(12), pp. 742-52.

Xu, F., Xu, S., Wiermer, M., Zhang, Y. and Li, X. (2012) 'The cyclin L homolog MOS12 and the MOS4-associated complex are required for the proper splicing of plant resistance genes', Plant J, 70(6), pp. 916-28.

Xu, X., Chen, C., Fan, B. and Chen, Z. (2006) 'Physical and functional interactions between pathogen-induced Arabidopsis WRKY18, WRKY40, and WRKY60 transcription factors', Plant Cell, 18(5), pp. 1310-26.

Yamada, K., Matsushima, R., Nishimura, M. and Hara-Nishimura, I. (2001) 'A slow maturation of a cysteine protease with a granulin domain in the vacuoles of senescing Arabidopsis leaves', Plant Physiol, 127(4), pp. 1626-34.

Yamagata, H., Masuzawa, T., Nagaoka, Y., Ohnishi, T. and Iwasaki, T. (1994) 'Cucumisin, a serine protease from melon fruits, shares structural homology with subtilisin and is generated from a large precursor', J Biol Chem, 269(52), pp. 32725-31.

Yamagata, H., Uesugi, M., Saka, K., Iwasaki, T. and Aizono, Y. (2000) 'Molecular cloning and characterization of a cDNA and a gene for subtilisin-like serine proteases from rice (Oryza sativa L.) and Arabidopsis thaliana', Biosci Biotechnol Biochem, 64(9), pp. 1947-57.

Yamaguchi, Y., Barona, G., Ryan, C. A. and Pearce, G. (2011) 'GmPep914, an eight-amino acid peptide isolated from soybean leaves, activates defense-related genes', Plant Physiol, 156(2), pp. 932-42.

Yamaguchi, Y., Huffaker, A., Bryan, A. C., Tax, F. E. and Ryan, C. A. (2010) 'PEPR2 is a second receptor for the Pep1 and Pep2 peptides and contributes to defense responses in Arabidopsis', Plant Cell, 22(2), pp. 508-22.

Yamaguchi, Y., Pearce, G. and Ryan, C. A. (2006) 'The cell surface leucine-rich repeat receptor for AtPep1, an endogenous peptide elicitor in Arabidopsis, is functional in transgenic tobacco cells', Proc Natl Acad Sci U S A, 103(26), pp. 10104-9.

Yang, C., Lu, X., Ma, B., Chen, S. Y. and Zhang, J. S. (2015) 'Ethylene signaling in rice and Arabidopsis: conserved and diverged aspects', Mol Plant, 8(4), pp. 495-505.

Yang, S., Tang, F. and Zhu, H. (2014) 'Alternative splicing in plant immunity', Int J Mol Sci, 15(6), pp. 10424-45.

139

Zander, M., Thurow, C. and Gatz, C. (2014) 'TGA Transcription Factors Activate the Salicylic Acid-Suppressible Branch of the Ethylene-Induced Defense Program by Regulating ORA59 Expression', Plant Physiol, 165(4), pp. 1671-1683.

Zeng, W. and He, S. Y. (2010) 'A prominent role of the flagellin receptor FLAGELLIN-SENSING2 in mediating stomatal response to Pseudomonas syringae pv tomato DC3000 in Arabidopsis', Plant Physiol, 153(3), pp. 1188-98.

Zhang, J., Li, W., Xiang, T., Liu, Z., Laluk, K., Ding, X., Zou, Y., Gao, M., Zhang, X., Chen, S., Mengiste, T., Zhang, Y. and Zhou, J. M. (2010) 'Receptor-like cytoplasmic kinases integrate signaling from multiple plant immune receptors and are targeted by a Pseudomonas syringae effector', Cell Host Microbe, 7(4), pp. 290-301.

Zhang, J., Shao, F., Li, Y., Cui, H., Chen, L., Li, H., Zou, Y., Long, C., Lan, L., Chai, J., Chen, S., Tang, X. and Zhou, J. M. (2007) 'A Pseudomonas syringae effector inactivates MAPKs to suppress PAMP-induced immunity in plants', Cell Host Microbe, 1(3), pp. 175-85.

Zhang, J. H., Barr, V. A., Mo, Y., Rojkova, A. M., Liu, S. and Simonds, W. F. (2001) 'Nuclear localization of G protein beta 5 and regulator of G protein signaling 7 in neurons and brain', J Biol Chem, 276(13), pp. 10284-9.

Zhang, X. C. and Gassmann, W. (2003) 'RPS4-mediated disease resistance requires the combined presence of RPS4 transcripts with full-length and truncated open reading frames', Plant Cell, 15(10), pp. 2333-42.

Zhang, X. C. and Gassmann, W. (2007) 'Alternative splicing and mRNA levels of the disease resistance gene RPS4 are induced during defense responses', Plant Physiol, 145(4), pp. 1577-87.

Zhang, Z., Liu, Y., Ding, P., Li, Y., Kong, Q. and Zhang, Y. (2014) 'Splicing of receptor-like kinase-encoding SNC4 and CERK1 is regulated by two conserved splicing factors that are required for plant immunity', Mol Plant, 7(12), pp. 1766-75.

Zhao, C., Johnson, B. J., Kositsup, B. and Beers, E. P. (2000) 'Exploiting secondary growth in Arabidopsis. Construction of xylem and bark cDNA libraries and cloning of three xylem endopeptidases', Plant Physiol, 123(3), pp. 1185-96.

Zheng, Y., Schumaker, K. S. and Guo, Y. (2012) 'Sumoylation of transcription factor MYB30 by the small ubiquitin-like modifier E3 ligase SIZ1 mediates abscisic acid response in Arabidopsis thaliana', Proc Natl Acad Sci U S A, 109(31), pp. 12822-7.

Zhou, A. and Li, J. (2005) 'Arabidopsis BRS1 is a secreted and active serine carboxypeptidase', J Biol Chem, 280(42), pp. 35554-61.

Zhou, J., Lee, C., Zhong, R. and Ye, Z. H. (2009) 'MYB58 and MYB63 are transcriptional activators of the lignin biosynthetic pathway during secondary cell wall formation in Arabidopsis', Plant Cell, 21(1), pp. 248-66.

Ziemienowicz, A., Merkle, T., Schoumacher, F., Hohn, B. and Rossi, L. (2001) 'Import of Agrobacterium T-DNA into plant nuclei: two distinct functions of VirD2 and VirE2 proteins', Plant Cell, 13(2), pp. 369-83.

Zipfel, C. (2013) 'Combined roles of ethylene and endogenous peptides in regulating plant immunity and growth', Proc Natl Acad Sci U S A, 110(15), pp. 5748-9.

Zipfel, C. (2014) 'Plant pattern-recognition receptors', Trends Immunol, 35(7), pp. 345-51. Zipfel, C., Kunze, G., Chinchilla, D., Caniard, A., Jones, J. D., Boller, T. and Felix, G. (2006)

'Perception of the bacterial PAMP EF-Tu by the receptor EFR restricts Agrobacterium-mediated transformation', Cell, 125(4), pp. 749-60.

Zipfel, C., Robatzek, S., Navarro, L., Oakeley, E. J., Jones, J. D., Felix, G. and Boller, T. (2004) 'Bacterial disease resistance in Arabidopsis through flagellin perception', Nature, 428(6984), pp. 764-7.

Author: Pierre BUSCAILL

Title: A protease of the subtilase family negatively regulates plant defence through its interaction with the Arabidopsis transcription factor AtMYB30.

PhD Supervisor: Susana RIVAS

__________________________________________________________________________

Abstract:

Plants defence responses are often associated with the development of the so-called hypersensitive response (HR), a form of PCD that confines the pathogen to the infection site. The sharp boundary of the HR suggests the existence of efficient mechanisms that control cell death and survival. The Arabidopsis transcription factor AtMYB30 positively regulates plant defence and HR responses by enhancing the synthesis of sphingolipid-containing Very Long Chain Fatty Acids (VLCFA) after bacterial infection. The activity of AtMYB30 is tightly controlled inside plant cells through protein-protein interactions and post-translational modifications. During my PhD, we identified a protease of the subtilase family (AtSBT5.2) as a AtMYB30-interacting partner. Interestingly, we have shown that the AtSBT5.2 transcript is alternatively spliced, leading to the production of two distinct gene products that encode either a secreted [AtSBT5.2(a)] or an intracellular [AtSBT5.2(b)] protein. The specific interaction between AtMYB30 and AtSBT5.2(b), but not AtSBT5.2(a), leads to AtMYB30 specific retention outside of the nucleus in small intracellular vesicles. atsbt5.2 Arabidopsis mutant plants, in which both AtSBT5.2(a) and AtSBT5.2(b) expression was abolished, displayed enhanced HR and defence responses. The fact that this phenotype is abolished in an atmyb30 mutant background suggests that AtSBT5.2 is a negative regulator of AtMYB30-mediated disease resistance. Importantly, overexpression of the AtSBT5.2(b), but not the AtSBT5.2(a), isoform in the atsbt5.2 mutant background reverts the phenotypes displayed by atsbt5.2 mutant plants, suggesting that AtSBT5.2(b) specifically represses AtMYB30-mediated defence.

__________________________________________________________________________

Keywords: Arabidopsis thaliana, hypersenstive response, Pseudomonas syringae, subtilase, endosomes, transcription factor.

__________________________________________________________________________

Discipline: Plant-pathogen interactions

Laboratory of Plant-Microbe Interactions (LIPM)

UMR CNRS/INRA 2594/441, 24 Chemin de Borde Rouge – Auzeville, CS 52627, 31326 Castanet-Tolosan cedex, France.

Auteur : Pierre BUSCAILL

Titre : Une protéase de la famille des subtilases régule négativement les réactions de défense à travers son interaction avec le facteur de transcription d’Arabidopsis AtMYB30.

Directrice de thèse : Susana RIVAS

__________________________________________________________________________

Résumé :

Les réactions de défense végétales sont souvent associées au développement de la réponse hypersensible (HR), une forme de mort cellulaire programmée qui confine l'agent pathogène au niveau du site d'infection. La frontière nette de la HR suggère l'existence de mécanismes efficaces qui contrôlent la frontière entre mort cellulaire et survie. Le facteur de transcription d'Arabidopsis AtMYB30 régule positivement la HR et les réponses de défense de la plante en augmentant la synthèse des acides gras à très longue chaîne (VLCFA) après infection bactérienne. L'activité d’AtMYB30 est étroitement contrôlée à l'intérieur des cellules végétales par des interactions protéine-protéine et des modifications post-traductionnelles. Au cours de mes travaux de thèse, nous avons identifié une protéase de la famille des subtilases (AtSBT5.2) en tant que partenaire protéique d’AtMYB30. Chose intéressante, nous avons montré que le transcrit d’AtSBT5.2 est épissée de façon alternative, conduisant à la production de deux produits de gènes distincts codant soit pour une isoforme sécrétée [AtSBT5.2 (a)] soit une isoforme intracellulaire [AtSBT5.2 (b)]. L'interaction spécifique d’AtMYB30 avec AtSBT5.2(b), mais pas avec AtSBT5.2(a), conduit à une rétention d’AtMYB30 à l'extérieur du noyau au sein de petites vésicules intracellulaires. Des plantes d’Arabidopsis mutantes atsbt5.2, ne montrant ni expression d’AtSBT5.2(a) ni d’AtSBT5.2(b), présentent des réactions de défense et de HR accrues. Ce phénotype étant abolie dans un fond génétique mutant atmyb30, AtSBT5.2 est donc un régulateur négatif de la résistance aux maladies induites par AtMYB30. Fait important, la surexpression de l’isoforme AtSBT5.2(b), mais pas celle de l’isoforme AtSBT5.2(a), dans le fond mutant atsbt5.2 rétablit les phénotypes présentés par les plantes mutantes atsbt5.2, ce qui suggère qu’AtSBT5.2(b) réprime spécifiquement la réponse de défense induite par AtMYB30.

__________________________________________________________________________

Mot clés : Arabidopsis thaliana, endosomes, facteur de transcription, Pseudomonas syringae, réponse hypersensible, subtilase.

__________________________________________________________________________

Discipline : Interactions plantes-microorganismes pathogènes

Laboratoire des Interaction Plantes-Microorganismes (LIPM)

UMR CNRS/INRA 2594/441, 24 Chemin de Borde Rouge – Auzeville, CS 52627, 31326 Castanet-Tolosan cedex, France.


Recommended