MOLECULAR BIOLOGY AND BIOCHEMISTRY OF THE

REGULATION OF HRP/TYPE III IN THE

CORN PANTOEA STEWARTII SUBSP. STEWARTII

DISSERTATION

Presented in Partial Fulfillment of the Requirements for the

Degree Doctor of Philosophy in the

Graduate School of The Ohio State University

By

Massimo Merighi, B. S., M. S.

*****

The Ohio State University

2003

Dissertation Committee:

Professor David L. Coplin, Adviser Approved by

Professor Brian Ahmer ______

Professor Dietz W. Bauer Adviser

Professor Terrence L. Graham Plant Pathology

Copyright by

Massimo Merighi

2003

ii

ABSTRACT

Pantoea stewartii subsp. stewartii is a bacterial pathogen of corn. Its

pathogenicity depends on the expression of a Hrp/type III

secretion/translocation system. The regulatory region of the hrp cluster consists of three adjacent operons: hrpXY encodes a two- component regulatory system, consisting of the response regulator HrpY and sensor PAS-kinase HrpX; hrpS encodes an NtrC-like enhancer-binding protein; and hrpL encodes an ECF sigma factor. In this study, we used genetic and biochemical approaches to delineate the following regulatory cascade: 1) HrpY activates hrpS; 2) HrpS activates hrpL; and 3) HrpL activates secretion and effector genes with ‘Hrp-box’ promoters. This pathway responds to environmental signals and global regulators.

Mutant analysis showed that HrpX is required for full virulence.

Deletion of its individual PAS-sensory domains revealed that they are not redundant and each may have a different role in modulating kinase activity. HrpX probably senses an intracellular signal, possibly related to nitrogen metabolism. pH, osmolarity and nicotinic acid controlled expression of hrpS independently of HrpX/HrpY.

hrpY mutants were completely avirulent. Site-directed mutagenesis of the conserved D57 residue in the HrpY receiver domain showed that it is

ii required for virulence and activation of hrpS and hrpXY in vivo. In vitro studies using purified His6 -HrpY indicated that D57 is the only

phosphorylation site and that His6 -HrpY~P has increased affinity for

PhrpS. Primer extension and deletion analyses located the PhrpS upstream of

an IS-like element and showed that of hrpL from a σ 54 promoter requires an upstream activating sequence. We also discovered that the absence of transcriptional terminators downstream of hrpL and hrpY creates a novel autoregulatory loop, wherein activation of hrpL by

HrpS results in read-through transcription into hrpY and hrpS. Finally, the

EsaR/EsaI quorum sensing system was shown to modulate the expression of hrp genes and virulence. Expression of hrpS was reduced in an esaI background and restored in an esaIR strain.

iii

“...Fatti non foste per viver come bruti,

ma per seguir virtute e canoscenza’.”

(“…(Men) were not created to live like wild ,

but to pursue virtue and knowledge.”)

(Dante’s Inferno, Canto XXVI, 119-120)

Dedicated to my parents, Graziella and Camillo

iv

ACKNOWLEDGMENTS

First of all, I would like to thank my family, dad, mom, and sister, for

their unconditional support and love throughout all these years so far

away from my motherland Italy. I know how much sacrifice that has

meant for you and I hope I made you proud in the end. Secondly, thanks

to my lovely wife, Serena. Simply put, nothing of this could have ever

happened without you. I will never forget your patience and

encouragement during these last six years we have been living together

here in America. Serena also helped me with many of the figures in

Chapter 1.

Obviously, I want to thank Dr. David Coplin, my adviser, for his

mentorship, and the intellectual freedom and support I have enjoyed in the

years spent in his lab. There I fully discovered my vocation as a

molecular bacteriologist and I found the courage to explore many new

areas of research. Secondly, I also have to thank him for the extensive

editing of this dissertation. I also want to express my gratitude to the

members of my SAC, Dr. Brian Ahmer, Dr. Dietz W. Bauer, and Dr.

Terrence L. Graham, for many stimulating discussions and for their

unconditional availability whenever I bugged them for using equipment or

v discussing ideas. In particular, a special thanks goes to Dr. Bauer, for teaching me many things about academic life and beyond, and for trying, unsuccessfully, to make a decent pool player out of me (Sorry Dietz!).

A special acknowledgement goes to current members of the lab, Doris

R. Majerczak and Dr. Jong Hyun Ham. Doris has been a good friend and a wonderful co-worker, always available and extremely helpful. Many of the experiments in this dissertation are a result of the strict cooperation between the two of us. I will always carry with me a good memory of the time spent in the lab together. Jong Hyun has been a good friend with whom I shared many nights in the lab and many discussions about science and our futures.

A particular thanks goes to Dr. Max Teplitski, who has been a dear friend of my wife and me. He has been very close to us both in times of bad and good luck and we will never forget that. He also provided me with samples of purified Salmonella enterica BarA198 before publication for the experiments described in Chapter 4.

Many other friends have helped me to endure the hardships of the

Graduate School, some by discussing science with me, others by giving me different perspectives on the facts of life, some others just by being there when I needed it the most. My dear friends, Maria, Mensheng,

Ricardo, and Dr. Trudy Torto, I will never forget you all and I hope the

Fate will bring great happiness to your lives.

vi I thank also the Ohio Agricultural Research Development Center for

supporting me with a Director’s Fellowship for the first three years, the

Graduate School of the Ohio State University for granting me a

Presidential Fellowship and the Department of Plant Pathology for its support over one year and a half of my Ph.D. program.

Some of the work presented in this dissertation was completed with the support of many people at various levels. In particular, most of the data in

Chapter 2 was published in a peer-reviewed journal before the defense of

this dissertation (Merighi et al., 2003, Mol. Plant Microbe Interact 16:

238-248). For that part, I acknowledge practical contributions from E. H.

Stover (limited to the construction of strains DM701, DM729 and DM733

carrying the pRF205), who also provided the initial observations

on the linear model of regulation and on the hrpS autoregulation several

years before I came in the lab; N. Smith, for the data in Fig. 2.7; D. R.

Majerczak, for her help in performing replicates of enzyme assays, for her

independent construction of strains DM786 and pDM1296,

pDM2560 and for her practical support in various steps leading to the

construction of several plasmids and strains used in that work. I also

acknowledge intellectual contributions by D. L. Coplin in the Discussion

section of Chapter 2. Furthermore I have to acknowledge also the

contribution by D. R. Majerczak of the two Ω interposon mutants used in

Chapter 3, which Doris constructed under my supervision, and her help in

running replicates of the enzyme assays described in Chapters 2 and 6. I

vii finally thank Dr. Jyang Chuyn Jang for his friendship, mentorship and for having me in his lab as radiation user for all the experiments in Chapters

3, 4 and 5 involving radioactive isotopes.

viii

VITA

1971 Born, Italy, European Union

1990-1996: B.S. (“Laurea”), Department of Plant Pathology,

University of Bologna, Italy, Graduated 110/110,

Summa cum Laude. Experimental thesis: “Genetic

fingerprinting by PCR and PFGE of Erwinia amylovora

strains from the Mediterranean basin”. Field of Study:

Plant Pathology, Molecular Phytobacteriology.

Summer 1996: Research Intern, Max Plank Institute für Zellbiologie,

Klaus Geider’s Bacterial Molecular Genetics

Laboratory, Heidelberg, Germany.

1997-1998: Research Assistant (MIRAAF/National Plan of Plant

Biotechnology), Institute of Plant Pathology,

University of Bologna, Italy.

2001: M.S.(non-thesis), The Ohio State University

1998-present: Graduate Student, Graduate School of The Ohio State

University, Department of Plant Pathology and Plant

Molecular Biology/Biotechnology Program:

-O.A.R.D.C. Graduate Fellow (1998-2001);

ix -Graduate Research Associate (2001-2002 and 2003-

present);

-Presidential Fellow of the Ohio State University

(2002-2003)

PUBLICATIONS

Refereed publications

1. Teplitski, M., Chen, H., Rajamani, S., Gao, M., Merighi, M., Sayre,

R. T., Robinson, J. B., Rolfe, B. G., Bauer, W. D. (2004). Chlamydomonas

reinhardtii secretes compounds that mimic bacterial signals and interfere

with quorum sensing regulation in . Plant Physiol. 134:1-10.

2. Merighi, M., Majerczak, D. R., Stover, H. E. and Coplin, D. L.

(2003). The HrpX/HrpY two-component system activates hrpS expression,

the first step in the regulatory cascade controlling the Hrp regulon in

Pantoea stewartii subsp. stewartii. Mol. Plant-Microbe Interact. 16:238-

248.

3. Merighi, M., Sandrini, A., Landini, S., Girotti, S., Ghini, S.,

Malaguti, S., Bazzi, C. (2000). Chemiluminescent and colorimetric

x detection of Erwinia amylovora by immunoenzymatic determination of

PCR amplicons from plasmid pEA29. Plant Dis. 84:49-54.

4. Zhang, Y., Merighi, M., Bazzi, C., K. Geider (1998). Genomic analysis by pulsed-field gel electrophoresis of Erwinia amylovora strains from the Mediterranean region including Italy. J. Plant Path. 80:225-232.

Book Chapters

1. Merighi, M., Majerczak, D. R., and Coplin, D. L. “The hrp genes of

Pantoea stewartii are regulated by a complex system that senses environmental signals”. In: “Plant Pathogenic Bacteria” p. 201-204, S. H.

De Boer Editor, Kluwer Academic Publishers, The Netherlands, 2001.

2. Merighi, M., Sandrini, A., Landini, S., Malaguti, S., Porrini, C.,

Sabatini, A.G., Girotti, S., Ghini, S., and Bazzi, C. "Automated detection of the plant pathogen Erwinia amylovora by chemiluminescent immunoenzymatic determination of PCR amplicons from pEA29: applications to diagnosis and epidemiology". In: "Bioluminescence and

Chemiluminescence: Perspectives for the 21st Century" p. 504-507, A.

Roda, M. Pazzagli, L. J. Kricka and P. E. Stanley Editors, Wiley and

Sons, Chichester, UK, 1998.

xi

FIELD OF STUDY

Major Field: Plant Pathology

Graduate Specialization: Plant Molecular Biology and Biotechnology

Areas of Interest: Bacterial pathogenesis and regulation of virulence;

Bacterial molecular genetics and molecular biology; Biochemistry of two- component systems.

xii

TABLE OF CONTENTS

Abstract…………………………………………………………………………..ii

Dedication……………………………………………………………………….iv

Acknowledgments……………………………………………………………….v

Vita……………………………………………………………………………….ix

List of Tables…………………………………………………………………xxii

List of Figures…………………………………………………………….…xxiii

List of Abbreviations and Symbols………………………………………..xxix

Chapters:

1. LITERATURE REVIEW...... 1

1.1 Stewart’s wilt of corn and its causal agent, Pantoea stewartii subsp. stewartii ...... 1

1.1.1 Historical background ...... 1

1.1.2 Host range and resistance ...... 4

1.1.3 Geographic distribution of the disease ...... 7

1.1.4 Symptoms and signs...... 10

1.1.5 Source of inoculum, disease cycle and histopathology...... 12

1.1.6 Taxonomy ...... 16

1.1.7 Molecular Diagnostics...... 21

1.1.8 Genetics...... 23

xiii 1.1.9 Molecular pathogenesis ...... 28

1.2 Type III secretion systems, hrp genes and their regulation ...... 34

1.2.1 Bacterial secretion systems and the type III secretion apparatus. 34

1.2.2. Phytobacterial Hrp type III secretion systems...... 42

1.2.3 Regulation of type III secretion systems ...... 54

1.2.3.1 Regulation of group I hrp/hrc genes ...... 56

1.2.3.2 Regulation of group II hrp/hrc genes ...... 61

1.3 Mechanisms and strategies of two-component system regulation ..... 63

1.3.1 Regulation of virulence factors ...... 63

1.3.2 Two-component systems...... 65

1.3.2.1 Response regulators ...... 68

1.3.2.2 Sensor histidine kinases...... 70

1.4. Significance, rationale and objectives ...... 75

1.4.1 Significance ...... 75

1.4.2 Rationale ...... 77

1.4.2 Objectives...... 79

2. THE HRPX/HRPY TWO-COMPONENT SYSTEM ACTIVATES

HRPS EXPRESSION, FIRST STEP IN THE REGULATORY

CASCADE CONTROLLING THE HRP REGULON IN PANTOEA

STEWARTII SUBSP. STEWARTII………………………………………….96

2.1 Introduction………………………………………………………………...96

2.2 Materials and methods…………………………………………………..99

xiv 2.2.1 Bacterial strains, plasmids, media and growth conditions……….99

2.2.2 HR and pathogenicity tests…………………………………………100

2.2.3 General recombinant DNA, genetic and bioinformatic techniques……………………………………………………………………101

2.2.5 Construction of Plac-hrpy, Plac-hrpS and Plac-hrpL plasmids……102

2.2.6 Construction of chromosomal P. stewartii hrp insertion mutants……………………………………………………………………….103

2.2.7 Site-directed mutagenesis of hrpY codon D57 by overlap extension

PCR and construction of missense mutant strains………………………104

2.2.9 Construction of P. stewartii strains with in-frame deletions in hrpX……………………………………………………………………….106

2.2.9 Construction of hrp-promoter reporter plasmids…………………105

2.2.10 Measurement of enzymatic reporter gene activity……………...107

2.2.11 Analysis of gfp reporter gene expression by flow cytometry….108

2.3 Results…………………………………………………………………..109

2.3.1 Sequence analysis of the hrp regulatory region of P. stewartii..109

2.3.2 Null mutant phenotypes and epistasis analysis…………………..112

2.3.3 Expression of hrp regulatory genes in different genetic backgrounds…………………………………………………………………114

2.3.4 Role of the conserved aspartate in HrpY………………………….117

2.3.5 Effects of environmental conditions on the expression of hrp genes in vitro……………………………………………………………………….118

xv 2.4 Discussion…………………………………………………………………120

2.5 Summary…………………………………………………………………..128

3. A NOVEL AUTOREGULATORY LOOP ACTIVATED BY HRPS

FROM THE HRPL PROMOTER MODULATES EXPRESSION OF THE

P. STEWARTII TYPE III SECRETION SYSTEM REGULATORS

HRPXY, HRPS AND HRPL...... 151

3.1 Introduction ...... 151

3.2 Materials and Methods ...... 156

3.2.1. Bacterial growth and media ...... 156

3.2.2 Molecular genetics and cloning techniques...... 156

3.2.3 Construction of polar hrpL interposon mutants and of plasmid

pMM191 ...... 157

3.2.4 Analysis of β-glucuronidase or β-galactosidase activity ...... 158

3.2.5 Total RNA isolation...... 158

3.2.6 Primer extension and DNA sequencing ...... 159

3.2.7 RT-PCR analysis ...... 159

3.2.8 Northern hybridizations...... 160

3.3 Results and discussion...... 161

3.3.1 Constitutive expression of HrpS indirectly upregulates hrpS

transcription by acting on the hrpL promoter ...... 161

3.3.2 The upstream region of the hrpL promoter contains a symmetrical

dyad required for HrpS-dependent activation ...... 163

xvi 3.3.3 Primer extension analysis defines the 5’ ends of hrpXY, hrpS and

hrpL transcripts and their potential promoters ...... 164

3.3.4 RT-PCR reveals that the hrpL-hrpXY and hrpXY-hrpS interoperon

regions are transcribed ...... 165

3.3.5 Northern blot analysis reveals the presence of long transcripts

spanning hrpL and hrpXY that are disrupted in a hrpL polar mutant .. 166

3.4 Conclusions...... 168

3.5 Summary...... 172

4. BIOCHEMICAL ANALYSIS OF THE RESPONSE REGULATOR

HRPY AND OF ITS ROLE AS A DUAL ACTIVATOR/REPRESSOR

OF HRP GENES...... 184

4.1 Introduction ...... 184

4.2 Materials and Methods ...... 187

4.2.1 Strains and growth conditions ...... 187

4.2.2 Construction of plasmid-borne gene fusions and of a hrpY[D57E]

allele for plasmid expression...... 188

4.2.3 Construction of plasmids for protein expression and

purification ...... 189

4.2.4 Purification of His6-HrpY ...... 190

4.2.5 Mass spectrometry ...... 192

4.2.6 In vitro phosphorylation of HrpY ...... 193

4.2.7 Electrophoretic mobility shift assay ...... 194

xvii 4.2.8 DNase I footprinting ...... 195

4.2.9 β-glucuronidase enzyme assays ...... 197

4.2.10 Total RNA isolation ...... 197

4.2.11 Primer extension and DNA sequencing...... 198

4.2.12 Sequence analysis and modeling...... 198

4.3 Results...... 199

4.3.1 Deletion analysis of the regulatory elements activated or repressed

by HrpY...... 199

4.3.2 Overexpression and purification of HrpY proteins ...... 202

4.3.3 Phosphorylation of HrpY at D57 by Salmonella enterica

BarA198 ...... 204

4.3.4 Binding of HrpY to 5’ hrpS DNA fragments ...... 206

4.3.5 Binding of HrpY to 5’ hrpL DNA fragments...... 208

4.3.6 Effect of phosphorylation on in vitro HrpY binding activity .... 209

4.3.7 Effect of phosphorylation of HrpY on its in vivo activity ...... 210

4.3.8 Identification of the binding elements of HrpY by DNase I

footprinting ...... 211

4.4 Discussion...... 212

4.5 Summary...... 223

5. GENETIC ANALYSIS OF THE HRPX PAS-KINASE AND PARTIAL

CHARACTERIZATION OF RECOMBINANT HRPX PROTEINS AND

PROTEIN FRAGMENTS ...... 254

xviii 5.1 Introduction ...... 254

5.2 Materials and Methods ...... 255

5.2.1 Media and Strains and general microbial genetic techniques .... 257

5.2.2 Protein modeling ...... 258

5.2.3 Construction of strains with in-frame deletions in various

hrpX domains ...... 259

5.2.4 Construction of plasmids for genetic complementation tests .... 261

5.2.5. Construction of plasmids for expression and purification

of HrpX ...... 262

2+ 5.2.6 Overexpression and purification of His6 -HrpX protein by Ni -

NTA affinity chromatography ...... 264

5.2.7 Overexpression and purification of MBP-HrpX protein by

amylose affinity chromatography...... 266

5.2.8 Overexpression and purification of GST-PAS1-2HrpX protein

by glutathione affinity chromatography...... 268

5.2.9 Mass spectrometry protein fingerprinting...... 269

5.2.10 In vitro autophosphorylation of HrpX and

phosphotransfer reaction...... 270

5.2.11 UV-crosslinking of the HrpX input domain with [γ32P]ATP.... 270

5.2.12 Pathogenicity tests ...... 270

5.3 Results...... 271

5.3.1 The input region of P. stewartii HrpX is formed by two

PAS domains with low similarity between them ...... 271

xix 5.3.2 The two PAS domains of HrpX play different roles in

modulating its activity...... 274

5.3.3 HrpX does not transduce pH, osmolarity or many catabolic

metabolite-associated signals but it appears to sense stimuli related to

nitrogen metabolism and the Krebs cycle ...... 277

5.3.4 Expression and Purification of recombinant HrpX proteins ...... 279

5.3.5 Autophosphorylation kinetics and phosphotransfer ability of HrpX

...... 283

5.3.6 GST-PAS1-2 does not bind ATP...... 285

5.4 Discussion...... 286

5.5 Summary...... 293

6. N-ACYLHOMOSERINE LACTONE-DEPENDENT QUORUM-

SENSING IS REQUIRED FOR FULL EXPRESSION OF THE HRP-

TYPE III SECRETION SYSTEM OF PANTOEA STEWARTII SUBSP.

STEWARTII ...... 327

6.1 Introduction ...... 327

6.2 Materials and Methods ...... 332

6.2.1 Bacterial strains and plasmids...... 332

6.2.2 Genetic and molecular biology techniques...... 333

6.2.3 Plasmid construction...... 333

6.2.4 Enzyme assays...... 334

6.2.5 Plant assays ...... 334

6.3 Results...... 335 xx 6.3.1 An esaI mutant, containing a mutation in the AHL synthase gene,

shows a Hrp- phenotype...... 335

6.3.2 In epistasis experiments, constitutive expression of hrpS, but not

hrpY, suppresses the Hrp phenotype of esaI strains ...... 336

6.3.3 Expression of most hrp/wts genes is dramatically reduced in the N-

acyl-homoserine synthase mutant ...... 336

6.3.4 Expression of hrp secretion genes is population density-dependent

and EsaR mediates the effects of esaI ...... 337

6.4 Discussion...... 339

6.5 Summary...... 346

7. CONCLUSIONS……………………………………………………………………357

LIST OF REFERENCES……………………………………………………364

xxi

LIST OF TABLES

Table Page

1.1 Regulators of type III secretion systems...... 81

2.1 Bacterial strains and plasmids...... 130

2.2 Primers...... 134

2.3 Analysis of DNA and protein sequences ...... 136

2.4 Pathogenicity of wild type and hrp regulatory mutants and complementation analyse ...... 138

2.5 Effects of hrp regulatory genes on plasmid-borne hrp::uidA transcriptional fusions expressed in different genetic backgrounds …139

3.1 Bacterial strains and plasmids...... 173

3.2 Primers ...... 175

4.1 Bacterial strains and plasmids...... 225

4.2 Oligonucleotide primers...... 226

5.1 Bacterial strains and plasmids...... 295

5.2 Oligonucleotide primers...... 298

6.1 Bacterial strains and plasmids...... 348

6.2 Watersoaking symptom severity ratings and hypersensitive response elicitation caused by P. stewartii esaI-esaR mutants and epistatic effects of other Hrp regulatory genes...... 349

xxii

LIST OF FIGURES

Figure Page

1.1 Worldwide distribution of Stewart’s wilt……………………………….82

1.2 Stewart’s wilt disease cycle...... 83

1.3 The quorum sensing regulatory circuit controlling EPS synthesis in P.

stewartii...... 84

1.4 Bacterial secretion systems...... 85

1.5 Schematic representation of and plant pathogen type III

secretion apparatuses and of the flagellar basal body...... 86

1.6 Transcriptional organization of animal and plant type III secretion

systems..…………………………………………………………………………87

1.7 Regulatory circuits controlling type III secretion system transcription

in animal ...... 88

1.8 Phytobacterial Hrp/type III secretion system regulatatory circuits. .. 89

1.9 Modular structure of two-component systems...... 92

1.10 Phosphorylation pathways in two-component systems...... 94

2.1 Physical map of the regulatory region of the P. stewartii DC283 hrp

gene cluster showing the hrpL, hrpX, hrpY and hrpS subclones used in this

study...... 140

xxiii 2.2 Analysis of the secondary structure of the IS element remnant located

in front of hrpS...... 142

2.3 PCR detection of the abortive IS element preceding hrpS in a

collection of P. stewartii strains ...... 143

2.4 β-galactosidase (β-Gal) and β-glucuronidase (GUS) activity of

chromosomal hrp-lacZ or hrpL-uidA fusions in different P. stewartii

genetic backgrounds, with and without various plasmid-borne Plac-hrp

regulatory genes...... 144

2.5 Expression of plasmid-borne hrp-uidA fusions in P. stewartii DC283.

...... 146

2.6 Flow-cytometry analysis of hrp-gfp fusion expression in P. stewartii

DC283 cells from infected corn plants compared to their expression in

LB...... 147

2.7 Effect of environmental conditions on the expression of various hrp-

lacZ fusions...... 149

2.8 Working model for the hrp regulatory cascade in P. stewartii...... 150

3.1. Physical map of the hrp regulatory region, mutants and plasmids. 176

3.2 Effects of HrpS+ constitutive expression on chromosomal hrp-lacZ fusions in various genetic backgrounds...... 177

3.3 The effect of deleting the upstream enhancer sequence (UAS) in the hrpL regulatory region on PhrpL expression...... 178

3.4 Primer extension analysis of hrpL, hrpX and hrpS transcripts...... 179

3.5. RT-PCR analysis of interoperon read-through transcription...... 180

xxiv 3.6 Northern blot analysis of hrp regulatory transcripts...... 181

3.7 Model of the hrp regulatory cascade including the autoregulatory loop triggered by HrpS...... 183

4.1 Expression of plasmid-borne hrpS gene fusions in P. stewartii...... 227

4.2 Expression of plasmid-borne hrpL gene fusions in E. coli SΦ200. . 228

4.3 Nucleotide sequence alignment of the hrpY-hrpS intergenic regions

from several erwinias and salient features of this region

in P. stewartii...... 229

4.4 Computer modeling of the intrinsic curvature of the hrpS promoter

region...... 231

4.5 Primer extension analysis and sequence analysis of the hrpL promoter.

...... 232

4.6 Overexpression and purification of HrpY...... 234

4.7 Mass Spectrometry analysis of recombinant HrpY

and HrpY[D57N]...... 236

4.8 Homology model of the crystal structure of P. stewartii HrpY ...... 237

32 32 4.9 Phospho-transfer of Pi from BarA198~ P to HrpY...... 239

4.10 Map of the hrpS regulatory region and the promoter fragments used as probes in gel shift experiments ...... 240

4.11 Electrophoretic mobility shift analysis of the binding of HrpY to the full length hrpS promoter...... 241

4.12 Electrophoretic mobility shift analysis of the binding of

HrpY to portions of the hrpS promoter...... 242

xxv 4.13 Competition analysis of the binding of HrpY to hrpS fragment C. 243

4.14 Map of the hrpL regulatory region and of the promoter fragments

used as probes in gel shift experiments...... 244

4.15 Electrophoretic mobility shift analysis of the binding of HrpY to

portions of the hrpL promoter...... 245

4.16 Electrophoretic mobility shift analysis of the binding of HrpY to the

full length hrpL promoter fragment...... 246

4.17 Effect of phosphorylation on HrpY binding to the hrpS and hrpL

promoters ...... 247

4.18 β-galactosidase activity of a chromosomal hrpS-lacZ fusion in a P.

stewartii hrpY null mutant carrying various plasmid-borne Plac-hrpY regulatory genes...... 248

4.19 DNase I footprinting assay of HrpY complexes at the

hrpS promoter...... 249

4.20 Model for hrpS activation...... 251

4.21. Model for hrpS activation...... 252

4.22 Model for hrpL regulation...... 253

5.1 Map of primers and PCR fragments used to construct hrpX mutations

(plasmids and chromosomal mutants)...... 299

5.2 Physical maps of the hrpX mutations and plasmids used for complementation studies...... 300

5.3 Plasmids used for expression and purification of HrpX...... 301

5.4 Plasmids used for expression and purification of HrpX domains.... 303

xxvi 5.5 Phylogenetic tree of HrpX homologs...... 305

5.6 PSI-PRED Secondary structure analysis of HrpX...... 306

5.7 Phylogenetic tree of PAS1-like domains in other two component

systems ...... 309

5.8 Threading of the PAS1 amino acid sequence into the FixL PAS

secondary structure...... 310

5.9 Structural model of the P. stewartii PAS1 threaded into the backbone

of B. japonicum FixL PAS...... 311

5.10 Complementation analysis and virulence phenotypes of P. stewartii

hrpX deletion mutants...... 312

5.11 The effect of various hrpX mutations on virulence and the expression

of hrpJ. a typical hrp secretion gene...... 314

5.12 The effect of environmental stimuli on hrpS-lacZ expression and

suppression of these effects by ectopic overexpression of

hrpY[D57N] ...... 315

5.13 Pilot experiment to overexpress polyHis-tagged HrpX proteins.... 317

5.14 Overexpression and purification of His6 -HrpX from E. coli

BL21(DE3) (pMM228)...... 318

5.15 Example of the results from nano LC-MS/MS protein fingerprinting of the tryptic fragment LHSSDTPIK from gel-excised His6 -HrpX...... 321

5.16 Overexpression and purification of MBP-HrpX by amylose affinity

and ion exchange chromatography...... 322

32 5.17 Kinetics of autophosphorylation of His6-HrpX by [γ P]ATP...... 324

xxvii 32 32 5.18 Transphosphorylation of HrpY by BarA~ P and His6-HrpX~ P.. 325

5.19 Purification of the GST-PAS1-2 protein from E. coli DH10B

(pMM374) and attempted direct photoaffinity labeling to [γ32P]ATP. .. 326

6.1 Maps of plasmid-borne reporter gene fusions used in this study. ... 350

6.2 Effect of an esaI mutation on the expression of several hrp regulatory and structural genes...... 351

6.3 Suppression of esaI regulatory effects on hrpJ by addition of

synthetic N-3-oxo-hexanoyl-homoserine lactone (OHHL)...... 352

6.4 Transcriptional expression of hrpJ as a function of density in

esaI, esaR and esaIR mutants...... 353

6.5 Effect of addition of synthetic N-3-oxo-hexanoyl-homoserine lactone

(OHHL) on expression of hrpJ...... 354

6.6 Model describing the RsmA regulatory network in P. carotovorum 355

6.7 Model for quorum sensing regulation of P. stewartii hrpS and

alignment of the putative esaR-like box (HrpS box)...... 356

7.1 Global regulation of hrp genes in P. stewartii…………………………..363

xxviii

LIST OF ABBREVIATIONS AND SYMBOLS

A: adenine

A: alanyl residue (Ala) aa: amino acid

A600: absorbance at 600 nm; see OD600

ATP: adenosine 5’-triphosphate avr: avirulence genes bp: base pair(s)

C: cytosine

CDD: conserved domain database

Ci: Curie (1 Ci=2.22x1012 dpm) cps: capsular polysaccharide synthesis genes cpm: counts per minute

CTD: carboxy-terminal domain

CUR: conserved upstream repeats

D: aspartyl residue (Asp)

DEPC: diethylpyrocarbonate

DPI: day post-inoculation dpm: disintegrations per minute dsp: disease specificity genes

xxix DR: direct repeats

DTT: dithiothreitol

E: E score, expected score

E: glutamyl residue (Glu)

ECF: extra-cytoplasmic function

EDTA: ethylendiaminotetracetic acid

EPS: extracellular polysaccharide esa: Erwinia stewartii autoinducer genes

F1: hybrid generated by crossing two parental lines

G: guanine

G: glycyl residue (Gly)

GC: guanine + cytosine content in molar %

GST: glutathion-S-transferase h: hour(s)

H: histidyl residue (His)

HEPES: 4-(2-hydroxy-ethyl)-1-piperazine-ethanesulfonic acid

HK: histidine kinase hop: Hrp outer protein gene

HR: hypersensitive response hrp: hypersensitive response and pathogenicity genes (pronounced

“harp”)

ICPPB: International Collection of Plant Pathogenic Bacteria

I.D.: columns internal diameter

xxx IHF: integration host factor

IMAC: immobilized metal affinity chromatography

IPTG: isopropyl-D-thiogalactopyranoside

IS: insertion sequence

K: lysyl residue (Lys) kb: 103 base pairs kcal: 103 calories

Kd : dissociation constant (in mol) kV: 103 volts kDa: 103 daltons or atomic mass units

M: in nucleic acid sequences, either A or C

M: in protein sequences, methionine (Met)

MALDI-TOF MS: mass absortion laser desorption ionization time of flight mass spectrometry

MBP: maltose binding protein

MES: 2-(N-morpholino)ethanesulfonic acid min: minutes mol: moles mRNA: messenger RNA

MS/MS: tandem mass spectrometry

MU: 4-methylumbelliferone

MUG: 4-methylumbelliferyl-glucuronide

MUGal: 4-methylumbelliferyl-β-D-galactopyranoside

xxxi m/z: mass over charge ratio

N: any base

N: asparagyl residue

Nano-LC: nanovolume liquid chromatography

NCBI: National Center for Biotechnology Information

NNPP: Neural Network Promoter Prediction nt: nucleotides

OD600: optical density at 600 nm; see A600

PAI: pathogenicity island

PAGE: polyacrylamaide gel electrophoresis

PAS: Per-Arnt-Sim

PCR: polymerase chain reaction

PFGE: pulsed field gel electrophoresis

PMSF: phenylmethylsulphonyl fluoride

QTL: quantitative trait loci/locus

Q-TOF: quadrupole time of glight (see MALDI)

RBS: ribosomal binding site rcs: regulation of capsule synthesis genes

RMSD: root mean square deviation

RR: response regulator rRNA: ribosomal RNA rsm: regulation of secondary metabolites

SD: standard deviation

xxxii SDS: sodium dodecyl sulfate

SK: sensor kinase

SE: standard error

T: thymine

T: in protein sequences, threonine (Thr)

TBE: Tris-borate EDTA buffer

TFA: trifluoroacetic acid

TMS: transmembrane sequence

Tris: tris(hydroxymethyl)aminomethane

tRNA: transfer RNA

TCS: two-component system

TTS: type III secretion

TTSS: type III secretion systems

UAS: upstream activating sequence

UTR: untranslated region

VUR: variable upstream region

W: either adenine or thymine

wts: watersoaking genes

X-Gal: 5-bromo-4-chloro-3-indolyl-β-D-galactoside

Y: in nucleic acid sequences, either cytosine or thymine

Y: in protein sequences, tyrosine (Tyr)

xxxiii

CHAPTER 1

LITERATURE REVIEW

1.1 STEWART’S WILT OF CORN AND ITS CAUSAL AGENT,

PANTOEA STEWARTII SUBSP. STEWARTII

1.1.1 Historical background

The history of bacterial wilt of sweet corn (synonyms: Stewart’s wilt of corn; Stewart’s disease of sweet corn and maize; bacterial leaf blight of maize; maize bacteriosis; Burrill’s disease of maize pro parte) is intimately linked to the dawn of bacteriology and phytopathology. The disease was first described by F. C. Stewart in 1897 in the bulletin of the

Agricultural Experiment Station, New York State, after he initially discovered the microorganism on infected corn plants in a Long Island

(New York) farm in 1895 (Stewart, 1897). A similar bacterial disease of corn was reported by T. J. Burrill eight years earlier (Burrill, 1889). This disease may have been Stewart’s wilt, however, Burrill, who was the first propose that bacteria could cause plant disease, failed to obtain a culture

1 of virulent microbes and his isolates were later identified by Theobald

Smith as nonpathogenic Bacillus cloacae (Smith, 1914). Stewart’s bulletin

is therefore considered to be the first report of the disease. In reality,

Stewart’s claims about the bacterial etiology of corn wilt were more

anecdotal than scientific, since they were merely based on the observation

that bacterial slime was constantly found in infected stems. Stewart was

indeed unable to reproduce symptoms on artificially inoculated yellow

dent corn, popcorn, teosinte and oats (Stewart, 1897) and satisfy what was

then called “Koch’s rules of proof”. His inconclusive experiments were

also plagued by contamination of control plants due to the use of infected

seed.

It was Erwin F. Smith who first proposed Pseudomonas stewartii nova

species as the causal agent of Stewart’s wilt in a paper at a meeting of the

American Association for the Advancement of Science on December 1898

(Smith, 1898). This report was followed by a more accurate description of

the cultural characteristics of this and other “uni-flagellate bacterial

pathogens of plants” (Smith, 1901). The formal proof of P. stewartii as

the causal agent was reported by E. F. Smith only in 1903 with an abstract

in Science (Smith, 1903) summarizing an address delivered in December

1902 before the Society for Plant Morphology and Physiology. For the

first time, Smith was able to artificially inoculate pure cultures of the

microbe on corn plants without wounding, via penetration of leaf stomata

and water-pores. He also conducted a detailed histopathological study,

2 defining the xylem of stems, leaves, husks and cobs as the primary

compartments colonized by the microbe, and parenchyma as secondary

site of . In his classic multi-volume treatise, “Bacteria in

Relation to Plant Diseases” (Smith, 1914), Smith summarizes much of the

work done on this microorganism during the first part of the last century.

Most of the literature following Smith’s pioneering work deals with the discovery of seed and insect transmission of the microbe (Smith, 1909;

Rand and Cash, 1924); attempts to control the disease using “lysins”

(later recognized as bacteriophages) (Thomas, 1938), insecticides (Ayers

et al., 1979; Sands et al., 1979), chemicals and antibiotics (Lockwood and

Williams, 1956; Natti, 1958; Pataky et al., 2000); prediction of disease

outbreaks based on weather variables (Stevens, 1934; Elliott and Poos,

1940; Castor et al., 1975); the use of disease resistant maize and sweet

corn lines (Pataky et al., 1998; Pataky et al., 2000); and the analysis of

the biochemical and quantitative genetic basis of the disease (Ming et al.,

1999). Initial genetic studies during the 1930-40s focused on the selection of virulent vs. avirulent strains with varying color and colony morphology after host passage (Wellhausen, 1937; Lincoln, 1939; Lincoln, 1947) and the analysis of mutation rates upon X-ray mutagenesis (Lincoln and

Gowen, 1942). As a historical curiosity, P. stewartii, along with

Drosophila and TMV, was used in one of the first studies aimed at defining the physical basis of genes and mutations (Gowen, 1941).

Systematic genetic studies of the microbe only began during the mid-

3 1970s with the isolation of temperature-induced, avirulent mutants

(Garibaldi and Gibbins, 1975) and the mobilization of F- and R-factors into P. stewartii (Gibbins et al., 1976; Coplin, 1978). During the 1980s, the introduction of bacteriophage Mu and Tn5 transposon mutagenesis allowed Coplin and associates to isolate capsular polysaccharide and water-soaking defective mutants (Coplin, 1979; McCammon et al., 1985;

Coplin et al., 1986). Such technical advancements were followed by an explosion of studies during the 1990s. The last 15 years have been characterized by the increasing use of genetic engineering and molecular biology to dissect cell biology, virulence and pathogenicity functions of this and several other model pathogens, to revise the taxonomic collocation of the microbe and to diagnose symptomatic and asymptomatic seed with added sensitivity and speed. This has gradually established P. stewartii, as not just an economically important pathogen, but as an extremely useful and powerful model for the analysis of basic gene regulation, cell-to-cell bacterial communication and virulence on corn, which is a monocotyledonous crop of enormous agricultural and industrial importance, with a vast community of researchers and a rich molecular genetics toolbox available.

1.1.2 Host range and resistance

Various species of the genus Zea are the common hosts of the disease, in particular sweet corn (Zea mays var. saccharata; the most susceptible

4 species), popcorn (Zea mays var. everta), and dent corn (Z. mays. var.

indentata; usually more resistant). The host range of P. stewartii includes

other monocotyledous plants. The pathogen has been isolated from

Eastern gama grass (Tripsacum dactyloides) and teosinte (Zea mexicana),

which is the Central-American progenitor of modern corn. Upon artificial

inoculation, P. stewartii is also able to produce symptoms or grow

endophytically in sorghum (Sorghum vulgare), Sudangrass (Sorghum bicolor), millet (Panicum miliatum), wheat (Triticum aestivum), oats

(Avena sativa) and other common grasses, such as yellow foxtail (Setaria glauca), black bent (Agrostis gigantea), Dactylis glomerata, Panicum capillare, Poa pratensis, Euchlena perennis, Setaria pumila, Tripsacum dactyloides, and T. zea. (Pepper, 1967; Ahmad et al., 2001). The only dicotyledonous plants experimentally infected are Golden Cluster beans

(Phaseolus vulgaris) and cucumber (Cucumis sativus) (Pepper, 1967).

Many of these plants may act as natural alternate hosts, contributing to the spread and over-wintering of the disease when sweet corn is not present (Poos, 1939).

Among the more than two thousand commercial dent and sweet corn hybrids, a wide range of disease responses exists, ranging from highly susceptible to highly resistant (Pataky et al., 2000). The use of resistant lines is the major control strategy for Stewart’s wilt. This resistance is

“partial” or “horizontal” (i.e. multigenic). Partial resistance is characterized by restriction of bacterial movement along the vascular

5 system (Braun, 1982). Initial studies identified two major genes and one

minor gene linked to resistance to Stewart’s wilt (Stevenson and Jones,

1953). Recently, one major QTL on chromosome S1 (locus sw1) and a

minor QTL on chromosome 9 controlling partial resistance have been

identified in the resistant line Ki14 (Ming, 1996; Ming et al., 1999). In

other lines, up to three QTLs on three different chromosomes could be

observed (Brown et al., 2001). Monogenic resistance to Stewart’s wilt

(also called “vertical”, or “race-specific” resistance) is not known to exist

in any corn accession (Stevenson and Jones, 1953; Pataky et al., 2000). In field trials, disease reactions of Z. mays inbred lines, F1 hybrids and sweet corn hybrids are classified by agronomists as resistant (R), moderately resistant (MR), moderately susceptible (MS), and susceptible

(S). In general, later maturing lines have higher levels of “horizontal”

(i.e. quantitative) resistance and, in these cases, the disease is confined to

the site of inoculation (in natural conditions, this is the feeding wounds

caused by the insect vector) and systemic wilt does not develop

(Suparyono and Pataky, 1989). Susceptible hybrids may show systemic infection when artificially inoculated as late as the 7- to 9-leaf stages, but

systemic infection of resistant hybrids is not common after the 3-leaf

stage (Pataky et al., 2000). Current national estimates of yearly losses from Stewart’s wilt are not available. In the past century, disease losses

increased progressively until 1930-1932, when the most spectacular

epidemics were recorded. Field losses on sweet corn and popcorn were

6 about 13%. During subsequent years, the increasing use of resistant

hybrids and varieties reduced the losses to a trace, except for occasional

outbreaks (Pepper, 1967). This pattern has continued, except when

individual sweet corn growers plant susceptible, early maturing sweet

corn varieties for marketing reasons. Following mild winters, heavy beetle

populations may overwinter in sheltered fields and these growers still

suffer devastating losses. Control of beetle populations by insecticide sprays or seeds treatments is the second major strategy of disease control currently available (Heichel et al., 1977; Ayers et al., 1979; Munkvold et al., 1996; Pataky et al., 2000; Kuhar et al., 2002). Yield losses can be

significant in susceptible varieties and in early because of the

high incidence of systemic infection in such cases. When plants are

infected systemically, yield is reduced about 1% for each 1% increase in

incidence of systemically infected plants, and this usually translates into a

complete destruction of the individual plant ears. In disease-conducive

years, field incidences of 40-100% can be observed (Pataky, 2003).

1.1.3 Geographic distribution of the disease

Stewart’s wilt probably originated in North America, where it was first

described. Moreover, it was not described (Pepper, 1967) in the centers of

origin of corn (Central America and Mexico) until the mid 1900s.

Considering the ability of P. stewartii to live as an endophyte in grasses

and the wide host range of the corn flea beetle, one may speculate that the

7 introduction of sweet corn into North America by native Americans

provided the occasion for a host species jump, analogous to what is

thought to have happened for Erwinia amylovora (Van der Zwett and Keil,

1979) upon the introduction of cultivated Pomaceae by the first American

settlers during the 1600s. In this respect, and interestingly from an

evolutionary point of view, it is notable that another erwinia, P.

agglomerans pv. gypsophiliae (Cooksey, 1986), also originated in the US

and that all three species share a highly conserved pathogenicity island

located on their chromosome or on a plasmid.

The occurrence of Stewart’s wilt epidemics appears to be strictly linked

to the presence of the most common insect vector, Chaetocnema pulicaria

(Pepper, 1967), or the related species C. denticulata and Diabrotica

undecimpunctata, which are present only in the mid-Atlantic to the

eastern portion of Mid-Western USA. The absence of these vectors is

commonly believed to be the reason that the disease has not spread to

other areas, even though it can be seed-borne. In the USA, the disease is endemic from the Mid-Atlantic states, through the Ohio Valley, to

Missouri and Iowa, with infrequent outbreaks 500 km north and south of the “Corn Belt” (Pataky 2000). This area includes Arkansas, Delaware,

Illinois, Indiana, Kentucky, Maryland, Missouri, New Jersey, New York,

Ohio, Pennsylvania, Tennessee, West Virginia and Virginia. Outbreaks

were also reported in Idaho and Washington state in 1920, but the disease

did not become established, possibly because of the dry climate and/or the

8 absence of the insect vector (Pepper, 1967). From the USA, the pathogen has been spread worldwide during the last 100 years, probably via trading of contaminated corn seed, mainly causing isolated field outbreaks or quarantine interceptions of infected material in Central America, Europe, and Asia (Fig. 1.1). The first case of disease outside of North America occurred in Italy (Pasinetti, 1937), with recurrent outbreaks in 1937,

1948, 1982 and 1983 (Anonymous, 1983), which were all linked to US- imported seed. Based on the European Plant Protection Organization

(EPPO) Data Sheets (EPPO, 1999), the disease has also been reported in

Viet Nam, Thailand, India, China (1942), Malaysia, Austria, Greece,

Poland (1942), Romania, Russia (1932), Switzerland (1950), Canada

(1931), Costa Rica (1949), Puerto Rico (1927), Mexico (1938, 1996,

1999), Guyana, Brazil (1968), Peru and Bolivia, but in most cases

Stewart’s wilt has not become endemic in the new environments. Notable exceptions are the sporadic outbreaks on sweet corn reported in Italy (Po

Valley; C. Bazzi, pers. commun.), Austria (in the regions of

Niederosterreich, Burgenland and Steiermark) and Mexico (Toluca Valley,

Oaxaca, Tabasco, Tlaxcala, Vera Cruz). In these countries, it is not known how the disease spread from the initial foci and became established, but such cases perhaps show a potential ability of P. stewartii to adapt to new secondary host-vector combinations (Pataky, 2000). Based on international agreements, countries with reported outbreaks of bacterial wilt must undergo strict field inspections for seed destined to export

9 markets. The seed from infected fields must be tested prior to export by a

certified lab and certified as pathogen-free. This requirement has

generated much research on diagnostics applied to seed lots. The

regulation of the corn seed trade and the P. stewartii quarantine rules

(European Council Directive 2000/29/EC, May 8, 2000) are major issues

in the turbulent Europe-USA agricultural trade relationships.

1.1.4 Symptoms and signs

Plants may be infected at any developmental stage and the bacterium

can cause a wide range of symptoms, which may be confused with

drought, potassium or nitrogen deficiency, insect damage, or fungal

diseases (e.g. Northern corn leaf blight, Southern corn leaf blight,

Helminthosporium leaf spot). In addition, three other bacteria cause

diseases on corn, including Pseudomonas alboprecipitans (bacterial leaf

blight), P. syringae pv. syringae (holcus spot) and P. andropogonis

(bacterial stripe), but their distribution and importance is extremely

limited. It would be interesting to determine whether or not common virulence factors control the development of water-soaking symptoms in these other gram negative corn pathogens.

Stewart’s wilt typically occurs in two phases. First, infected seedlings

or young plants show leaf streaks and water-soaked lesions with wavy,

irregular margins, from 1 to 10 mm wide and as long as the leaf. These

lesions quickly become necrotic, leading to leaf blight and eventually

10 wilting (Fig. 1.2). Under warm and humid conditions, bacterial ooze may appear on the surface of the lesions. Such early infections can be very severe and often cause complete destruction of the plants. Plants that survive early infections are stunted and later may form premature, abnormal ears, with tassels that are dwarfed, whitened and dead.

In the second phase of the disease, which typically occurs in mature plants that are infected after tasseling, the infected leaves show characteristic streaks parallel to the main veins (Fig. 1.2). Late infections are more common in susceptible dent corn lines. The lesions again have wavy irregular margins and are from yellow to light green in color. With time, the lesions turn brown and coalesce, causing leaf blight. Typically, insect feeding scars are associated with the lesions and they indicate the entry point of the pathogen. The size of such lesions is considerably smaller in resistant varieties, where they often appear as oval spots, 1-2 inches long (Pepper, 1967). Discolored, rotten cavities full of bacterial slime can be observed in the pith near the crown of mature stalks.

Similarly, light microscopy of transverse sections of infected stalks or leaf lesions may reveal yellow bacterial ooze extruding from the vascular bundles. This is a good diagnostic test in field examinations. Severe

Stewart’s wilt infection may predispose plants to stalk rot disease

(Pepper, 1967).

11

1.1.5 Source of inoculum, disease cycle and histopathology

The pathogen can overwinter in the intestine of a few insects of the order Coleopter.. Infected seeds are the other major means of overwintering (Smith, 1909). The importance of secondary hosts in the disease cycle is not known, but given the ability of P. stewartii to grow endophytically in many grasses (Ahmad et al, 2001), we cannot dismiss this possibility. Soil and infested crop residues are not relevant to pathogen survival and overwintering (Pepper, 1967).

The insect vectors belong to the order Coleoptera, family

Chrysomelidae. The most important species are the corn flea beetle (C. pulicaria), the toothed flea beetle (C. denticulata) and the 12-spotted cucumber beetle (D. undecimpunctata). The first experiments pointing to insects as the most important factor in Stewart’s wilt epidemiology date back to the 1920s (Pepper, 1967) and this relationship was proved conclusively by Elliott and Poos (Elliott and Poos, 1934). Extensive experimentation during the 1930s. Thousands of different species were tested as vectors for P. stewartii, other insect species were identified as potential carriers of the disease, including the Northern corn rootworm

(D. logicornis), Western corn rootworm (D. virgifera), the larval stage of

May beetles (Phyllophage sp.), seed corn maggot larvae (Hylemya cilicrura), and the common wheat wireworm (Agriotes mancus) (Elliott and Poos, 1940). However, under field conditions, C. pulicaria appears to

12 be the prevalent vector of the disease (Elliott and Poos, 1940).

Interestingly, C. pulicaria, C. denticulata and D. undecimpunctata have never been reported outside North America, which may explain the limited distribution of the disease outside of the US.

Infected seeds are another known mean of overwintering. A comprehensive review of the literature on seed infection is found in a recent report by Pataky (2000). In corn kernels, bacteria colonize the chalazal region, the aleurone layers and the apoplastic spaces of the endosperm, but are not found in the embryos or seed coat (Rand and Cash,

1924; Ivanoff, 1933). Survival in the seed varies from 5 to 8 months after harvest, depending on the storage temperature (Rand and Cash, 1924).

From infected seed, the bacteria can enter the seedling, with various degree of efficiency depending on the severity of infection in the parent plant and the susceptibility of the host to systemic infection (Pataky,

2000). The early literature mentions plant-to-seed infection rates (kernel infection rate) ranging from 2 to 85% (Pepper, 1967). However, more recent studies have shown that the rate of kernel infection is below

0.025% when the parent is resistant, and ranges from 0.25% to 10% when the parent is rated as susceptible or highly susceptible; the highest rates are limited to cases where the parent plant is systemically infected

(Pataky, 2000).

The disease cycle of Stewart’s wilt is depicted in Fig. 1.2, which summarizes our current understanding of P. stewartii

13 biology/epidemiology. The primary source of inoculum is bacteria over-

wintering in adult corn flea beetles (Fig. 1.2, step 2) or, much less

frequently, infections are started by bacteria surviving in the storage

tissues of maize kernels (Fig. 1.2, step 1). In the first case, the incidence of field infection is correlated with the size of the surviving population of insects. In this respect, studies conducted early last century (Stevens,

1934), based on 35 years of data, showed that Stewart’s wilt infections do not occur when the sum of the monthly mean temperatures for December through February is below 90°F, but disease incidence is high when the sum exceeds 100°F, due to higher populations of surviving vectors.

Penetration of healthy host tissues occurs externally via insect feeding wounds (Fig. 1.2, step 4) or endophytically from the germinating infected seed (Fig. 1.2, step 3). Under experimental conditions, the pathogen can also enter through stomata and hydatodes (Pepper, 1967). It is not clear whether bacteria are transmitted from the insect to the plant via an oral or a fecal route, but bacteria are probably associated with the alimentary tract. In general, very little is known about P. stewartii-insect vector

relationships. With the vast array of molecular tools now available, this

could represent an area of active research in the future. Following

wounding and inoculation, the bacteria localize in the intercellular spaces

of the parenchyma and in the embolic xylem vessels, probably starting

multiplication on the few nutrients available in the cell wall fluids (Fig.

1.2, step 6, 9). In this respect it is worth remembering that like most

14 bacterial plant pathogens, P. stewartii remains in the apoplast and never enters living host cells. This is in contrast to many mammalian pathogens that invade host cells. Based on studies by Braun (Braun, 1982), bacteria inoculated by vacuum infiltration remain singular or in small groups from

1 to 3 h after entering the tissue. Very little EPS seems to be produced during this early phase. A well-formed capsule surrounded by abundant slime could be seen at 12 h (Braun, 1982). Eventually, active multiplication results in the formation of microcolonies in the intercellular spaces of the parenchyma, which can be seen at 30 h post- inoculation. Appearance of watersoaked lesions and blight symptoms

(possibly a form of plant cell death/cytotoxicity) are visible from 48 h to

5 days post-inoculation. Bacteria colonize also the xylem vessels.

Whether bacteria actively force openings into vessels via enzymatic activity or passively break vessel pits by multiplication and mass swelling is not known (Braun, 1982). Production of pectolytic enzymes, which would be required for active movement across the xylem pits, is not observed in vitro (Braun, 1990), but it is not known whether they are produced in vivo. In experiments using 32P-labelled cells, it was shown that the rate of colonization of the xylem was correlated with the rate of transpiration and that the extent of the colonization did not differ between resistant and susceptible plants (Warren, 1951). This result implied that, at least under those experimental conditions, the free mass-flow of bacteria in the xylem was due to physical movement rather than

15 multiplication. Systemic colonization of the vessels causes wilt

symptoms, which is more dramatic in highly susceptible sweet corn

varieties. In susceptible plants, rapid systemic infection, plugging of the

xylem by EPS and fast deterioration and invasion of the adjacent

parenchyma cells occurs. The final result is often the destruction of the

whole vascular bundle and the appearance of wilt symptoms (Wellhausen,

1936; Braun, 1982; Braun, 1990). Bacteria may exit into the intercellular

spaces of the mesophyll through the primary cell wall of broken xylem

vessel pits during the late stages of pathogenesis (Braun, 1982). Work

done in the early 1980s has shown that specific corn agglutinins

(Bradshaw-Rouse et al., 1981) and higher concentrations of methoxy-

benzoxazolinone (Whitney and Mortimore, 1961) may be involved in

restricting massive colonization of parenchyma and xylem in resistant

lines.

1.1.6 Taxonomy

P. stewartii belongs to the domain Bacteria, phylum B.XII

“Proteobacteria phy. nov.”, class III “Gamma-proteobacteria”, order XII

“Enterobacteriales”, family I “Enterobacteriaceae”, genus XXIII

“Pantoea” (based on the 2000 Bergey’s Manual Trust, www.cme.msu.edu/bergeys). The nomenclature of this bacterium has

changed over time as a consequence of evolving methods of bacterial typing and different philosophical approaches to bacterial taxonomy. The

16 current proposed name is Pantoea stewartii (Smith 1898) subsp. stewartii

Mergaert, Verdonck & Kersters 1993, comb. nov., basonym Erwinia stewartii Smith 1898, Dye 1963 (Smith, 1898; Dye, 1963; Mergaert et al.,

1993). The basonym is still considered valid because it was included in the latest “Approved Lists of Bacterial Names” (Skerman et al., 1980).

The type strain defining the species is ATCC 8199 (Holt, 1984), deposited at the American Type Culture Collection as ICPB SS11 (originally classified as Xanthomonas stewartii Smith Dowson and isolated in the

USA from sweet corn by W.H. Burkholder in the 1940s). Other synonyms found in the literature over the last 105 years are "Pseudomonas stewartii" Smith 1898, "Bacterium stewartii" (Smith 1898) Smith 1911,

"Aplanobacter stewartii" (Smith 1898) McCulloch 1918, “Bacillus stewartii” (Smith 1898) Holland 1920, "Phytomonas stewartii" (Smith

1898) Bergey et al. 1923, "Xanthomonas stewartii" (Smith 1898) Dowson

1939, "Pseudobacterium stewartii" (Smith 1898) Krasil’nikov 1949

(Source: List of Bacterial Names with Standing in Nomenclature: Société de Bactériologie Systématique et Vétérinaire, http://www.bacterio.cict.fr/).

In 1898, Erwin F. Smith provided the first description of the cultural characteristics of the bacterium. Originally it was erroneously described as motile with a single polar flagellum and was therefore placed in the genus Pseudomonas (Smith, 1898; Smith, 1901). In 1914, E. F. Smith reclassified the species as Bacterium stewartii (Smith, 1914). After

17 McCulloch reported that it was nonmotile and lacked flagella in 1918, it

was reclassified as a member of the genus Aplanobacter (McCulloch,

1918). An attempt by Holland to compile updated and comprehensive generic indexes of bacteria resulted in Aplanobacter stewartii being moved into the genus Bacillus (Holland, 1920). This lasted until the first edition of Bergey’s Manual was published in 1923 and this bacterium was then included in the genus Phytomonas, which encompassed all gram- negative plant pathogens (Bergey and al., 1923). In 1937, the use of the genus name Phytomonas was questioned by Elliott because it was already in use by zoologists (Elliott, 1937). Then in 1939, Dowson proposed that

P. stewartii be placed in the genus Xanthomonas based on its biochemical properties and the presence of a yellow pigment (Dowson, 1939). This proposal ignored the facts that P. stewartii was nonmotile, did not produce the same yellow carotenoid and was facultatively anaerobic. In

1948 in the 6th edition of Bergey’s Manual, Burkholder classified P.

stewartii as a member of the genus Bacterium, a quite artificial and

heterogeneous taxon (Breed and al., 1948). Only in the 7th edition of

Bergey’s Manual, published in 1957, did Burkholder accept the

nomenclature Xanthomonas stewartii, but under the postil species incerta sedis (Burkholder, 1957). Between 1962 and 1963, the New Zealand scientist Douglas Dye revisited Xanthomonas stewartii taxonomy, accepting early suggestions that the microbe was related to coliform bacteria (Waldee, 1941) and a degenerate member of the

18 Enterobacteriaceae family. Consequently Xanthomonas stewartii was

classified in the genus Erwinia, as member of the so called “herbicola” group III, along with other “yellow” erwinias such as Erwinia herbicola

(syn. Enterobacter agglomerans; comb. nov. Pantoea agglomerans),

Erwinia ananas and Erwinia citri, and apart from the “amylovora” group I

and “carotovora” erwinia group II (Dye, 1963). Dye’s nomenclature is

accepted by the latest “Approved Lists of Bacteria Names” (Skerman et

al., 1980), and is used in the IXth edition of Bergey’s Manual (1984).

More recently, however, this official classification was challenged by

Mergaert et al. (1993). They proposed to move the group III erwinias into

the genus Pantoea based on DNA hybridization data and numeric

taxonomy. Consequently they renamed E. stewartii as Pantoea stewartii

subsp. stewartii. Subsequent analyses of the 16S rRNA gene sequences of

16 type species of the genus Erwinia (Kwon et al., 1997) confirmed the

distribution of most Erwinia species into four clusters, fundamentally

corresponding to Dye’s groups (with the exception of Erwinia salicis,

Erwinia rubrifaciens, and Erwinia nigrifluens, which form a fourth and distantly related clade). E. stewartii grouped with other erwinias of Dye’s

“herbicola” group in cluster I, even though the bootstrap value between it and the “amylovora” branch was only moderate. In the same study, the authors found that other enterobacteria, such as Escherichia coli,

Klebsiella pneumoniae, and Serratia marcescens, were intermixed with species of Erwinia. If nothing else, this finding raises questions on the

19 definition of bacterial species in this family and on the sole use of 16S rDNA analysis to reconstruct phylogenetic relationships. Another 16S rDNA study by Hauben et al. (1998) used 29 enterobacterial plant pathogens and likewise supported the splitting of Dye’s genus Erwinia into four groups, which they proposed to make different genera: Erwinia

(Dye’s group I), Pectobacterium (Dye’s group II), Pantoea (Dye’s group

III), and Brenneria (Kwon’s group 4). They also confirmed the collocation of the Stewart’s wilt agent into the monophyletic genus

Pantoea (Hauben et al., 1998).

P. stewartii is a yellow gram-negative rod, 0.9-2.2 µm in length by 0.4-

0.8 µm in diameter, that is nonmotile and non flagellate (Pepper, 1967;

Kado and Coplin, 2001). In this regard, P. stewartii has been recently reported to be capable of swarming in soft agar motility plates but the cellular basis for this phenotype is not known (S. von Bodman, pers. commun.). P. stewartii is facultatively anaerobic, cytochrome c oxidase negative, nitrate reductase negative, catalase positive and urease negative.

It can assimilate organic and inorganic nitrogen (ammonium ions and nitrates). The bacterium is able to produce acids from galactose, mannose, sucrose, glucose, mannitol, arabinose, xylose, fructose, glycerol, inositol and sometimes from sorbitol, whereas it produces alkali from rhamnose and dulcitol (Pepper, 1967; Kado and Coplin, 2001). Strains of P. stewartii were reported by Dye as delayed lactose fermenters (Dye, 1963), but this phenotype is not usually present (Kado and Coplin, 2001). These

20 differences could reflect strain divergences. P. stewartii shows a weak

endogenous and uncharacterized ß-galactosidase activity, easily visible in

rich media containing X-Gal, but not in McConkey-lactose media. Given that X-Gal is known to be non-specifically hydrolyzed by several

hydrolases (Neidhardt, 1993), such activity does not necessarily imply the

existence of a ß-galactosidase (lacZ) gene. The bacterium is also unable

to hydrolyze starch, casein, aesculin, lipids and pectins, but it can weakly

liquefy gelatin in a strain specific manner (Dye, 1963). It utilizes

gluconate and tartrate, but not propionate, and it can use asparagine as a

sole source of C and N (Dye, 1963). The bacteria are salt tolerant and can

grow in media with up to 7% NaCl (Burkholder and Starr, 1948). The

thermal death point is 53°C, maximum growth temperature is 39°C,

optimum temperature is 28-30°C, and minimum growth temperature is 8°C

(Dye, 1963). P. stewartii is able to survive for 14 months in agar stabs at

15°C but for only 7-10 days on agar plates at 4°C (Pepper, 1967).

1.1.7 Molecular Diagnostics

A comprehensive review of the methods used for the detection and

typing of P. stewartii is available in a recent monograph on phytobacterial

diagnostics (Kado and Coplin, 2001). Common methodologies for

detection and/or identification of P. stewartii involve simple field

inspection of seed plants, grow-out tests of seed, and the use of selective

and semiselective media to isolate pure cultures of bacteria. Cultures are

21 subsequently identified by classic morphological and physiological tests,

analysis of membrane fatty acid methyl-ester derivatives by gas

chromatography (GC-FAME; Wells et al., 1994), and/or enzyme-linked

immunosorbant assays (Lamka et al., 1991). In the last ten years, several

DNA based methods have been developed. One of the first reports described the use of arbitrary primer PCR to produce a species-specific

DNA probe (Blakemore et al., 1992). This approach was subsequently expanded by developing nested primers that amplified the same DNA

fragment (Blakemore et al., 1999). In 1994, a report on the use of the ligase chain reaction and 16S rRNA primers to identify P. stewartii was

published (Wilson et al., 1994), but the use of radioisotopes limited such

methods for routine assays. More recently several other PCR tests have

been reported for the purpose of genetic diversity studies (Fessehaie et al., 2002; Waleron et al., 2002) or detection/identification (Coplin et al.,

2002). In general, P. stewartii strains show a greater homogeneity in their genetic fingerprints than other pantoeas. This is analogous to the situation with E. amylovora, another species that originated from North America.

Genomic analysis of XbaI or SpeI macrorestriction fragments using pulsed field gel electrophoresis showed that P. stewartii strains are highly homogeneous, with Nei and Li similarity coefficients ranging from 0.86 to

0.9 (unpublished; calculated from Coplin et al., 2002). This technique could be used to differentiate P. stewartii from P. agglomerans, P. ananas and Pectobacterium carotovorum.

22

1.1.8 Genetics

The P. stewartii genome has a mol%G+C content of 54.6-55.1%

(estimated for two strains only; Holt, 1984). The size of the P. stewartii

chromosome is estimated to be between 3.5 and 4 Mb by PFGE analysis of

genomic macrorestriction fragments (Coplin et al., 2002). This estimate

does not include most of the 8 to 13 resident plasmids with sizes ranging

from 4 to 318 kb that are present in each of at least 31 virulent strains

from a geographically representative collection (Coplin et al., 1981).

Plasmid DNA may account for up to about 880 kb of the genome

(calculated from data reported in (Coplin et al., 1981). Eight class sizes

(4, 13, 25, 44, 65, 74, 106, and 318 kb) were common among 87% of more of the strains. Campbell-type recombination among replicons could account for “missing” and new plasmids in any given profile (Coplin et al., 1981). The presence of the 318 kb plasmid is somewhat in doubt because the newer plasmid purification techniques presently used in our lab do not isolate it. However, it is of particular interest to us, because several lines of experimentation suggest that the hrp/wts PAI may be located on it. This would be similar to what has been found for the hrp

cluster in P. agglomerans pv. gypsophilae, which is carried on a 150 kb pPATH plasmid (Manulis et al., 1991; Nizan et al., 1997).

In avirulent laboratory strains that have been in culture for over 40 years, the number of plasmids was reduced, but still most of the eight

23 conserved classes were present (Coplin et al., 1981). Several plasmids from strain SW2 have been totally or partially characterized. Plasmid pSW100, a 4.2-kb low copy number (2 copies per cell) mobilizable replicon, has been completely sequenced and found to contain a 0.7 kb origin of replication that was homologous to orgins from p15A, ColE1 and

ColA (Fu et al., 1995). Plasmid pSW500 (a 35-kb low copy plasmid) contains seven 16-bp iteron repeats, a DnaA box and a 1005-bp repA gene

(Fu et al., 1996). The replication origin of pSW800 has been characterized as well (Wu et al., 2001) and the results suggest that a mechanism similar to that regulating plasmid replication in the IncB, IncIa and IncL/M groups also regulates pSW800 replication. Plasmid pSW1200 (106 kb) has an approximate copy number of one and contains a bacteriophage P1-like ori (with a repA gene and eight 19-bp iteron sequences). Plasmid pSW1200 belongs to the IncY group (Fu et al., 1997). The significance of the coexistence of so many replicons in the same cell is both puzzling and intriguing. Perhaps they play a role in adaptation to various niches (such as secondary hosts and insect vectors) (Coplin et al., 1981). No clear

association of any P. stewartii plasmid with virulence or specific

phenotypes is known. The genome of P. stewartii strain DC283 (a

spontaneous NalR derived by D. L. Coplin from the virulent strain ICPPB

SS104 isolated by Art Hooker in 1967) is currently being sequenced at

Baylor College of Medicine and Wisconsin Genetic Center using a random

shot-gun library approach.

24 The genetic techniques available for P. stewartii are quite similar to

that of other enterics. P. stewartii can exchange plasmids with E. coli; it

supports replication of phage Mu (but is not infected by Mu virions) and

of many plasmid replicons, such as ColE1, p15, IncP1, and IncW groups.

However, F-factors are unstable and exhibit fertility inhibition, but not

surface exclusion, suggesting they are probably incompatible with one of

the indigenous plasmids (Coplin, 1978). Lysogenic phages are present in

about half the strains tested (D. L. Coplin, unpublished), but they have

never been genetically characterized. P. stewartii is insensitive to bacteriophage lambda infections, probably due to the absence of the LamB receptor, and infections by phage P1 or P22, two common phages used in generalized transduction of enterics. Although transduction systems are not available for genetic studies of P. stewartii, transposon mutagenesis has been an important tool in dissecting P. stewartii virulence. Delivery of transposons into P. stewartii cells was initially quite difficult because many of the tools available for other enteric bacteria would not work properly in this species. Widely used broad host range, suicide delivery vectors, such as pJB4JI, carrying modified Mu or Tn5, contrary to what observed in P. agglomerans subsp. herbicola, Pectobacterium chrysanthemi and P. carotovorum, replicated well in P. stewartii (Coplin,

1979). Temperature sensitive replicons, such as pUT13::Tn5 (used successfully in P. chrysanthemi) could not be used in P. stewartii because it does not grow at the non-permissive temperatures needed. Direct

25 infection with mini-Mu or MuS::Tn9 was not feasible because P. stewartii

is not sensitive to Mu(G-) phages. One of the first successful strategies for insertion mutagenesis was the use of an RK2::Mu cts62 hybrid plasmid

(pRK210). The replication and transposition of this Mu derivative was

inducible at 42°C after conjugal transfer from E. coli. (Coplin,1979). This

strategy was used to successfully isolate a number of of auxotrophic

mutants (Coplin, 1979). A second approach made use of a hybrid plasmid

comprised of the cryptic P. stewartii conjugative plasmid pDC250

containing a Tn10 insertion (pDC251) and phage Mu kan pf7701. This

genetic strategy allowed the isolation of several avirulent insertion

mutants affected in exopolysaccharide (EPS) synthesis and water-soaking

(Wts) ability for the first time (McCammon, 1985). Later, random

mutagenesis and the isolation of new Wts- mutants was achieved by using pMO75, an IncP-10 suicide replicon from , to

deliver a Tn5 transposon (Coplin et al., 1992). The current standard

procedure for random mutagenesis in P. stewartii makes use of mob+, R6K

ori, pir-dependent suicide replicons in pUT vectors, i.e. the mini-Tn5

series of transposons for polar insertions and gene/protein fusion

generation (de Lorenzo et al., 1990; Majerczak, Merighi and Coplin,

unpublished) or derivatives thereof. The transposons are delivered via

biparental matings from E. coli SM10 or S-17 λpir. Genomic libraries

were constructed in E. coli using broad host range, mobilizable cosmid

vectors, such as pVK100 and pLAFR3 (Coplin et al., 1986; Dolph et al.,

26 1988). These vectors are relatively large (20-22 kb), which limits the

average insert size to 25-30 kb. Following mutagenesis of the cosmids, or subclones thereof, allele exchange is performed by well-established genetic techniques for gram negative bacteria (Maloy et al., 1996). One common approach that we follow when using mutagenized clones with large inserts with at least a few kb of flanking homologous sequences is selectable marker exchange by plasmid incompatibility (Ruvkun and

Ausubel, 1981). The clone, usually a pLAFR derivative (Friedman et al.,

1982), is mobilized by biparental (Simon et al., 1983) or triparental (cit; using pRK2013 or pRK2013::Tn7 as helper plasmids) matings (Ditta et

al., 1980) into P. stewartii. Subsequent destabilization of the plasmid and

forced single cross-over with the homologous region of the chromosome is

done by mobilizing and selecting also for a second incompatible plasmid

(e.g. pJP4JI GmR or pR741 TpR) (Ruvkun and Ausubel, 1981). More recently, we introduced the use of R6K suicide plasmids followed by positive selection in presence of fusaric acid to loose the vector encoded tetracycline resistance (pLD55; Metcalf et al., 1996). This has allowed the construction of unmarked (substitutions and deletions) or marked (Ω or aphA3 insertions) mutations in P. stewartii. As little as 600 bp of flanking sequence homology is needed to exchange the allele into the P. stewartii chromosome (Merighi et al., 2003), obviating the problematic use of sacB-based positive selection vectors, which work poorly in P. stewartii.

27 1.1.9 Molecular pathogenesis

P. stewartii is a vascular pathogen that can also grow in the

intercellular spaces of the parenchymatic tissues of leaves. A major mechanism of virulence (in sensu lato) is represented by the production of exopolysaccharide (EPS) slime, which occludes xylem vessels causing the plant to wilt (Ivanoff et al., 1938; Braun, 1982). EPS is required for biofilm formation and rapid systemic movement in the plant: EPS- deficient strains grow and spread more slowly in the host vascular tissues, are less likely than the wild-type to colonize plants systemically and are unable to cause wilting in infected plants (Braun, 1990). The EPS of P.

stewartii is a very large (~45 MDa), viscous, highly charged

heteropolysaccharide composed of glucose, galactose and glucuronic acid in a 4:2:1 ratio. No other EPSs are made in culture. The EPS forms both a

bound capsule and loose slime. Capsule synthesis is constitutive but

additional slime production is induced by the availability of free sugars in

the growth medium. In addition to its involvement in virulence, EPS

possibly plays a role in protection from phytoagglutinins in the vessels

(Bradshaw-Rouse et al., 1981) and in the colonization of its insect vector.

The genes involved in capsular polysaccharide synthesis were initially cloned in cosmid pES2144 based on complementation of spontaneous non- mucoid mutants (Coplin et al., 1986). Cosmid pES2144 contains five complementation groups (cpsA-E) and the galE gene. Most of these genes are conserved in E. amylovora and E. coli based on sequence homology

28 (Bernhard et al., 1993) and interspecific complementation tests (Bernhard

et al., 1996), except for several glycosyl transferase genes that may account for differences in the structure the two polysaccharides. Cloned

E. amylovora ams genes complemented most P. stewartii cps mutants for slime production and virulence in corn seedling (Bernhard et al., 1996),

but the arrangement of the genes in the two clusters is slightly different.

The P. stewartii cps gene cluster is structurally and functionally related to

the group I capsule biosynthetic gene clusters of other enteric bacteria

(e.g. the colanic acid cps cluster in E. coli) (Dolph and Coplin, 1987;

Dolph et al., 1988). Regulation of the P. stewartii cps genes also appears

to be similar to that for colanic acid in E. coli (Torres-Cabassa et al.,

1987) and amylovoran in E. amylovora. The Rcs (regulator of capsule

synthesis) two-component system is the master regulator of EPS synthesis

in enterics (Leigh and Coplin, 1992). Positive regulators homologous to

E. coli rcsA and the two-component rcsB/C were cloned by

complementation of the corresponding E. coli mutants with P. stewartii

cosmids (Poetter and Coplin, 1991). An RcsA/RcsB heterodimer was later

shown to specifically bind a 23-bp fragment upstream of the cpsA gene of

P. stewartii (Wehland et al., 1999).

EPS production and virulence is also controlled by an acyl-homoserine-

lactone-based quorum sensing system (von Bodman and Farrand, 1995);

reviewed by (Von Bodman et al., 2003) comprising the EsaR/EsaI proteins

(Vibrio fisherii LuxR and LuxI homologs, respectively) (Fig. 1.3). These

29 proteins are encoded by two convergently transcribed genes and

expression of esaR is autoinducible. A putative lux box-like promoter

element is present in the –10 region of the esaR gene (von Bodman and

Farrand, 1995). The major AHL was detected in culture supernatants using

an Agrobacterium tumefaciens NT1(pJM749, pSVB33) reporter strain.

This was purified and identified as N-(-3-oxohexanoyl)-L-homoserine

lactone (OHHL) by electrospray fast atom bombardment mass

spectrometry (von Bodman and Farrand, 1995). In addition, thin layer

chromatography later revealed the production of subnanomolar amounts of

N-(-3-oxooctanoyl)-L-homoserine lactone (OOHL). Expression of esaI is

constitutive (von Bodman and Farrand, 1995) and production of OHHL is

linear with growth, whereas OOHL was detected only at high cell

densities (Von Bodman et al., 1998). The crystal structure of EsaI was

recently solved at a resolution of 1.8 Å (Watson et al., 2002). Catalytic site modeling and site-directed mutagenesis allowed the authors to define critical residues for catalytic activity and acyl-chain specificity involved in the acylation reaction with S-adenosyl-L-methionine (Watson et al.,

2002). EPS production and virulence (using a pin-prick leaf inoculation assay) was abolished in an OHHL synthase (esaI) mutant (von Bodman and Farrand, 1995). This effect was suppressed either by a mutation in the esaR locus or addition of exogenous OHHL. Production of EPS by the esaR and esaRI null mutants was constitutive. Together these results suggested a novel negative regulatory role for EsaR at low concentrations

30 of OHHL (i.e. low cell densities) (Von Bodman et al., 1998). EsaR was later shown to bind to the esaR promoter in the absence of OHHL and

increasingly higher concentrations of OHHL titrated out EsaR from the

protein-target DNA complex (Minogue et al., 2002). No direct binding of

EsaR to any cps promoter or other target genes has yet been reported, so it

is therefore conceivable that esaR may act through an intermediate

regulator(s).

Another global regulatory system controlling virulence traits is the

GacA/GacS two-component system (Heeb and Haas, 2001). GacA belongs to the FixJ class of response regulators and GacS has a domain organization typical of unorthodox sensor kinases (Tsuzuki et al., 1995;

Stock et al., 2000) (See section 1.3.2). First identified in Pseudomonas syringae pv. syringae as a gene required for swarming, lesion formation, and protease and toxin production (LemA=GacA) (Barta et al., 1992;

Hrabak and Willis, 1992; Rich et al., 1992; Rich et al., 1994; Rich and

Willis, 1997; Kinscherf and Willis, 1999), this two-component regulatory system was found to be conserved in many gamma proteobacteria.

GacA/GacS is involved in soft rot development in P. marginalis (Liao et al., 1997), antibiotic and exoenzyme synthesis in Pseudomonas fluorescens (Corbell and Loper, 1995; Whistler et al., 1998; Bull et al.,

2001; Mascher et al., 2002; Zuber et al., 2003), phenazine production in

Pseudomonas aureofaciens (Chancey et al., 1999; Whistler and Pierson,

2003), exoenzyme and siderophore production in Pseudomonas viridiflava

31 (GacA=RepB) and P. marginalis (Liao et al., 1997), biofilm formation

and rpoS expression in P. aeruginosa (Reimmann et al., 1997; Tan et al.,

1999; Parkins et al., 2001), modulation of the ToxR regulon in Vibrio

cholerae (Wong et al., 1998), virulence, exoenzyme and harpin production

in P. carotovorum (Frederick et al., 1997; Eriksson et al., 1998; Cui et

al., 2001; Hyytiainen et al., 2001), alginate and polyhydroxybutyrate

production in Azotobacter vinelandii (Castaneda et al., 2000), perception

of intestinal signals and virulence gene activation in Salmonella enterica

serovar Typhimurium (BarA/SirA) (Johnston et al., 1996; Ahmer et al.,

1999; Prouty and Gunn, 2000; Goodier and Ahmer, 2001; Lawhon et al.,

2002), and switching between glycolytic and gluconeogenic carbon

sources, rpoS expression, and biofilm formation in E. coli (BarA/UvrY)

(Mukhopadhyay et al., 2000; Pernestig et al., 2001; Pernestig et al.,

2003). In P. carotovorum, GacA/GacS(=RpfA) controls transcription of

rsmB (homologous to E. coli csrB), a small regulatory RNA that post-

transcriptionally controls RsmA (homologous to E. coli CsrA) expression

(Cui et al., 2001; Chatterjee et al., 2002). A similar regulatory pathway has been described in E. coli, with UvrY directly controlling csrB transcription in a csrA mutant background (Suzuki et al., 2002). Southern blot hybridizations (Cui et al., 2001) demonstrated the presence of gacA

and gacS homologs in several non-soft-rotting Erwinia/Pantoea species, such as E. amylovora, E. rhapontici, P. agglomerans, P. agglomerans pv.

gypsophilae and P. stewartii, The phenotype of gacA/gacS mutations in P.

32 stewartii is not known, but these genes are likely to play a role in regulating virulence similar to that described in P. carotovorum and P.

syringae.

The most important mechanism of virulence of P. stewartii identified so

far is represented by the expression of a type III secretion (TTS) system

that secretes anti-host effector proteins into the plant apoplast (i.e the

intercellular spaces and the cell walls) or translocates them into the

symplast (i.e. the cytoplasm) (Discussed in section 1.2). Phytobacterial

type III secretion systems are called Hrp (for Hypersensitive response and pathogenicity and pronounced “harp”). These are conserved in many

gram-negative plant pathogens, including P. stewartii, which has a typical hrp cluster that spans over 40-kb. It is required for HR elicitation in tobacco and disease in corn (Coplin et al., 1986; Coplin et al., 1992;

Frederick et al., 1993; Ahmad et al., 2001; Frederick et al., 2001).

Expression of this system is essential for active colonization of the

apoplast and growth to high population densities (>106 CFU/g of wet

tissue). The other major , EPS, is only required for

systemic spread, wilting and maintenance of water-soaking, since EPS

mutants with a functional Hrp TTS system can still colonize the

intercellular spaces of leaves, the xylem and cause initial water-soaking.

33

1.2 TYPE III SECRETION SYSTEMS, HRP GENES AND THEIR

REGULATION

1.2.1 Bacterial secretion systems and the type III secretion apparatus

An impressive number of reviews and book chapters are available on the

topic of bacterial protein secretion in general and type III secretion in particular (Cornelis and Wolf-Watz, 1997; He, 1997; Cornelis, 1998;

Hueck, 1998; Galan and Collmer, 1999; Kaper and Hacker, 1999; Cossart

et al., 2000; Koster et al., 2000; Christie, 2001; Sandkvist, 2001;

Cornelis, 2002; Fischer et al., 2002; Wilson et al., 2002). In this section I

will briefly describe both topics. My objective is to provide a background

for later discussion of phytobacterial Hrp/type III secretion systems and

their regulation, while remanding the reader to the literature for more

details on specific systems (Beer et al., 1991; Hutcheson, 1999; Collmer

et al., 2000; Cornelis and Van Gijsegem, 2000).

Secretion of proteins through membranes is one of the most important

activities of bacterial and eukaryotic cells (Cossart et al., 2000). In

bacteria, it is estimated that about 20% of the proteins synthesized are

targeted outside the cytoplasm (Wilson et al., 2002). Bacteria need

protein secretion to synthesize extracellular structures -such as flagella,

pili and fimbriae-, to degrade and transport into the cell various classes of

macromolecules, to build functional membranes, and to target toxins to

34 the extracellular or intracellular compartments of eukaryotic cells, while they are acting as partners in symbiotic or parasitic interactions.

Gram-negative bacteria have several compartments that include the cytosol, the cytoplasmic membrane, the periplasm and the outer membrane. Secreted proteins synthesized in the cytoplasm may be targeted to another compartment or to the extracellular environment via various pathways. Known secretory pathways include: i) the type I secretion systems, which are part of the complex family of ABC transporters, ii) the twin-arginine translocation (Tat) pathway, iii) the type III secretion systems, and iv) the general secretory pathway (GSP), with its four terminal branches, namely the chaperone-usher system, type

II secretion pathway, type IV secretion and type V autotransporters (Fig.

1.4) (Wilson et al., 2002).

Type I secretion system substrates do not have N-terminal cleavable signal sequences, but instead possess targeting sequences in their C- terminal domains. Type I apparatuses are composed of three proteins and are energized by ATP hydrolysis, as are all the other secretion systems.

Substrates of such widespread systems include the RTX haemolysin toxin of E. coli, antibacterial peptides (such as bacteriocins and antibiotics), and the Bordetella pertussis cyclolysin toxin. Proteins traveling the general secretory pathway in gram-negative bacteria have an 18- to 26- amino acid N-terminal signal sequence that is cleaved by a signal peptidase during secretion across the inner membrane. This pathway is

35 exemplified by the Sec system of E. coli, which is composed of a Sec

translocase complex located on the cytoplasmic membrane and one or

more aspecific cytoplasmic chaperones (such as SecB, DnaK, DnaJ or

GrpE), which keep the substrate unfolded. The GSP only translocates its

substrates to the periplasmic compartment. In many cases this is not the

final destination, so that other secretion pathways, collectively called

“terminal branches” of the GSP, are involved in translocation across the outer membrane. The main terminal branch (MTB) or type II secretion is the most commonly used pathway in gram-negative bacteria. This system is composed of 12 to 14 proteins, including an outer membrane multimeric

(10-12 subunits) secretin (e.g. Klebsiella oxytoca PulD), which has a

central pore large enough (90 Å in diameter) to accomodate folded

proteins (Wilson et al., 2002). Type II secretion is involved in the export

of many enzymes involved in degradation of macromolecules and/or in

virulence. Examples include the P. aeruginosa phospholipase C, alkaline

phosphatase, and endotoxin A (via the xcp-endoded system), Vibrio

cholerae Cholera Toxin, Type IV pili involved in twitching motility and adhesion of many gram-negative pathogens, pectic enzymes and cellulases though the P. carotovorum out system, and polygalactorunases via the

Xanthomonas campestris xps-encoded system (reviewed in (Hobbs and

Mattick, 1993). The chaperone-usher pathway is involved in the secretion of specific external appendices, such as pili and fimbriae. The type V systems involve protein substrates, whose C-terminal portions include a

36 pore-forming domain, which is able to fold into a 14-strand beta-barrel structure in the outer membrane that directs the translocation of the N- terminal “passenger” domain (Wilson et al., 2002). Type IV pathways secrete either proteins or nucleoprotein complexes. In at least two cases, type IV systems vectorially translocate DNA and/or proteins into eukaryotic cells. The first case is exemplified by the vir system of

Agrobacterium tumefaciens, which uses a dedicated inner membrane

secretion machine as an alternative to the GSP. The second example is the

cag system of Helicobacter pilori (Backert et al., 2002). Other examples

of substrates traveling this pathway are Bordetella pertussis Pertussis

Toxin and the Legionella pneumophila LcmX toxin (Fischer et al., 2002).

Type IV systems are thought to have evolved from plasmid conjugation systems (Christie, 2001; Fischer et al., 2002). The twin-arginine secretion system has been recently described as a distinct general secretory pathway that recognizes specific, cleaved targeting signals and is able to secrete fully folded substrates (Wilson et al., 2002).

TTS systems are unique in that they always mediate secretion/translocation of proteins not just across bacterial membranes, but also across eukaryotic membranes (cytoplasmic and/or phagosomal) in a vectorial fashion. In this respect, they truly function as molecular syringes injecting proteins into a eukaryotic host cell’s cytosol. These systems are always found in bacteria that spend some part of their life cycle in close interaction with either plant or animal cells. The bacterial

37 proteins injected into the host cells act as “powerful effectors” able to

subvert the host’s physiology to the bacterium’s advantage (Cornelis,

2002), thereby initiating “biochemical cross-talk” between the two bionts

(as defined by J. E. Galan; Galan and Bliska, 1996). Substrates of TTS systems do not have cleavable N-terminal sequences and the secretion step is believed to occur without periplasmic intermediates, with the exception of P. syringae (Charkowski et al., 1997). The nature of the secretion signal is still controversial and N-terminal polypeptide amphipathicity/primary sequence composition (Lloyd et al., 2001; Lloyd et al., 2002) a 5’-mRNA signal (Anderson and Schneewind, 1997;

Anderson et al., 1999; Anderson and Schneewind, 1999; Hienonen et al.,

2002) and/or chaperone-mediated targeting (Cheng and Schneewind,

1999), have been shown in various systems.

TTS systems were described for the first time in Yersinia enterocolitica,

Yersinia pseudotubercolosis and between 1991 and 1994

(Michiels and Cornelis, 1991; Michiels et al., 1991; Allaoui et al., 1994;

Sory and Cornelis, 1994; Woestyn et al., 1994). Since then, more than 20

TTS systems have been reported in animal, plant and insect pathogens and symbionts (Table 1.1). Besides Yersinia spp. (reviewed by (Cornelis,

2002), TTS apparatuses have been described in many other gram-negative animal pathogens, such as Salmonella spp. (reviewed in (Galan and

Collmer, 1999), Shigella spp. (Van Nhieu, 1997), enteropathogenic E.coli

(EPEC) and enterohemorragic E. coli (Frankel et al., 1998; Torres and

38 Kaper, 2002), P. aeruginosa, Chlamydia trachomatis, Chlamydia psittaci,

Chlamydia pneumoniae, Bordetella bronchiseptica, Bordetella pertussis,

and Burkholderia pseudomallei. The contribution of the TTS system to the

disease cycles of these pathogens at the cellular level varies and often

spans opposite behaviors in different species, ranging from inhibiting

phagocytic cells (e.g. in Yersinia spp.) to stimulating internalization by

nonphagocytic cells (in Salmonella enterica and Shigella spp.). In general, the net effect is to favor entry or survival of the invading pathogen by disrupting the host metabolism or disrupting intracellular signal transduction pathways, thereby causing a variety of symptoms, such as mild/severe diarrhea, septicemia, septic shock, and chronic infections

of the lung (Muller et al., 2001). TTS systems have also been described in

all gram-negative, non-tumorigenic phytobacteria surveyed so far, namely

P. syringae pathovars (1985-86) (Lindgren et al., 1986), P. stewartii

(Coplin et al., 1986), P. agglomerans pv. gypsophilae (Nizan et al.,

1997), E. amylovora (Steinberger and Beer, 1988; Barny et al., 1990; Beer

et al., 1991), P. carotovorum (Bell et al., 2002), P. chrysanthemi (Bauer

et al., 1994), X. campestris pv. campestris (Arlat et al., 1991) and pv.

vesicatoria (Arlat et al., 1991) and Ralstonia solanacearum (Boucher et

al., 1987). TTS apparatuses in phytobacteria are encoded by a subset of

the so-called hrp genes, which are discussed in greater detail in the next

section. TTS systems have also been described in the plant symbiont

Rhizobium leguminosarum (Viprey et al., 1998; Perret et al., 1999; Krause

39 et al., 2002) the plant growth promoting P. fluorescens strain SBW25

(Preston et al., 2001) and in the insect symbiont Sodalis glossinidius and

the entomopathogenic bacterium Photorhabdus luminescens (Ffrench-

Constant et al., 2000; Dale et al., 2001). In phytobacterial pathogens, TTS systems promote bacterial growth in plant tissues and subsequent disease symptom development, especially water-soaking or edema-like lesions and necrosis. They are also required for elicitation of rapid programmed defense responses that are induced in resistant plants in response to the specific elicitor proteins injected by the invading bacteria. These then trigger plant cell death, thereby restricting the infection (Staskawicz et al., 2001).

By far, the best-characterized TTS system is the Ysc (Yop secretion) apparatus of Yersinia spp. (Cornelis, 2002), also called the Ysc injectisome. The proteins traveling this system are called Yops (for

Yersinia outer proteins). The Ysc apparatus is an organelle made of a basal body spanning the inner membrane, the peptidoglycan layer and the outer membrane. The basal body is extended toward the external environment with a rigid needle-like structure. The basal body architecture resembles the flagellum basal body (Fig. 1.5), with the proximal part of the structure being formed by a large cylinder (YscQ) and an inner membrane-spanning ring (YscJ), corresponding to the flagellar C and MS rings, respectively. This portion also contains the protein-pumping apparatus (analogous to the flagellum-specific secretion

40 apparatus) comprised of YscV, YscU, YscR, YscT, YscS and the ATPase motor YscN. The distal part of the body, which does not have homologous proteins in the remaining structure of the flagellar apparatus, forms a 200-

Å ring-shaped structure with a hollow center of about 50 Å in diameter that is made of YscC monomers. YscC belongs to the PulD/pIV superfamily of outer membrane secretins. This secretin-like multimer is probably stabilized by the lipoprotein YscW. The outer ring is most likely connected to the basal body proximal cylinder by the lipoprotein YscJ.

However, YscJ has not yet been localized in Yersinia spp. (Cornelis,

2002). The Ysc needle is a rigid polymeric rod about 1200 Å long and 70

Å wide, mainly formed by the 6-kDa YscF monomer and three other proteins and containing a hollow conduit 20 Å in diameter. This needle- like structure can be isolated by physical methods and is usually associated with the basal body secretin-like protein and YscJ. The injectisome can release Yops into the external environment, but that alone is not sufficient for the translocation of proteins into eukaryotic cells.

Three additional Yop translocators (YopB, YopD and LcrV) are required for this step. Therefore, the description of the injectosome as a syringe directly poking holes into eukaryotic membranes is somewhat misleading

(Cornelis, 2002). An important property of TTS machines is that, at least in some animal pathogens (Yersinia spp., P. aeruginosa and Shigella spp.), release of TTS substrates is dependent on contact with host cell surfaces, which in some systems may be mimicked in vitro by modulating

41 divalent cation concentration and temperature (e.g. Yersinia spp.) or by

adding amphipathic dyes, such as Congo Red, or serum proteins (e.g.

Shigella spp.) (Francis et al., 2002). In addition, cell contact also

increases TTS system gene transcription upon secretion of negative

regulators, such as LcrQ (Petterson et al., 1996). For one plant pathogen, cell-contact regulation of the hrp-TTS genes has been suggested (Aldon et

al., 2000).

1.2.2. Phytobacterial Hrp type III secretion systems

Many reviews on the discovery of hrp genes and their structural and

functional characterization have been published in the last decade (Willis

et al., 1991; Bonas, 1994; van Gijsegem et al., 1995; Alfano and Collmer,

1997; Lindgren, 1997; He, 1998; Hutcheson, 1999; Cornelis and Van

Gijsegem, 2000). More recently the emphasis in this area has shifted to the functional analysis of substrates of the Hrp Type III secretion system

(such as Avr proteins and other effectors) in various host-pathogen model systems (Bonas and Van den Ackerveken, 1999; Buttner and Bonas,

2002). In this section, I will present a brief overview of the properties of this class of secretion machines and their substrates.

hrp genes were first identified in P. syringae pv. syringae in 1985-1986

(Lindgren et al., 1986), well before the identification of mammalian pathogen TTS systems. Only in 1992, did studies on the DNA sequences of hrp genes from X. campestris pv. vesicatoria (Fenselau et al., 1992)

42 and R. solanacearum (Gough et al., 1992) first reveal homologies to some of the proteins in TTS systems of Yersinia spp. and Shigella flexneri.

Initially, hrp genes were functionally defined based on their mutant phenotype; hrp genes had to be dispensable for growth in minimal media, but essential for pathogenicity in susceptible hosts and for eliciting the hypersensitive response (HR) in incompatible hosts and non-hosts

(Lindgren et al., 1986). The HR is a rapid and localized defense response of the plant that involves localized programmed cell death, active oxygen species production, and synthesis of phenolics and antimicrobial compounds at and around the site of infection (Klement, 1982; Klement et al., 2003). During natural infections, the reaction is usually not visible because too few host and bacterial cells are involved, but, under experimental conditions, high doses of inoculum (>107 /ml) will cause

massive cell collapse and confluent necrosis that is visible to the naked

eye (Klement, 1982). This reaction has the net effect of restricting

pathogen multiplication and further progress of the disease. HR is not

elicited by true non-pathogens, including many saprophytic bacteria and

the molecular cloning host E. coli, even at high inoculum doses.

hrp genes are clustered, spanning up to 41 kb of DNA (van Gijsegem et

al., 1995) and are located either on the chromosome or large plasmids

(Boucher et al., 1987). In some cases, cosmid clones containing nearly

complete hrp clusters enable non-pathogenic bacteria, such as E. coli K12

and P. fluorescens, to cause an HR, but not disease (Lindgren, 1997).

43 Subsequent DNA sequencing of overlapping cosmid clones spanning hrp

clusters and mutational analysis showed that hrp clusters encoded not

only the proteins involved in secretion, but also proteins involved in the

specific regulation of hrp and effector genes, elicitors of the HR in non-

host plants, and disease-specific (Dsp) proteins involved in pathogenicity

but not HR elicitation in Erwinia/Pantoea spp. (Bogdanove et al., 1998;

Bogdanove et al., 1998; Mor et al., 2001). Sequencing of the flanking

regions of hrp clusters (especially Pseudomonas syringae; (Alfano et al.,

2000; Collmer et al., 2000; Noel et al., 2002; Deng et al., 2003) and the completion of several genome projects (Bell et al., 2002; da Silva et al.,

2002; Salanoubat et al., 2002; Buell et al., 2003) have revealed that sets of putative substrates of the Hrp secretion apparatus are commonly located nearby. Another realization from these studies is the concept of hrp clusters as horizontally mobile pathogenicity islands (PAIs)

(Hutcheson, 1999). The presence of remnants of transposons and insertion sequences and, in some cases, tRNA genes (Alfano et al., 2000) at the extreme ends of the hrp/effector cluster is indicative of previous mobilization via plasmids, conjugative transposons or phages followed by integration into the chromosome or plasmids of the recipient strain (Kaper and Hacker, 1999). Homing of the PAI during the evolution of the species supposedly degraded its ability to mobilize further (Kaper and Hacker,

44 1999). In some species, a few loci showing Hrp phenotypes were also

found to be unlinked to the major hrp cluster, e.g. Xanthomonas hrpX and

P. syringae hrpM (Reviewed in van Gijsegem et al., 1995).

Phytobacterial TTS systems are usually encoded by three to four operons for a total of 19 to 21 genes (Hutcheson, 1999). Up to 12 of the encoded proteins show similarities with Ysc proteins that represent the core components of the secretion apparatus. Of these, eight share similarities with flagellum apparatus proteins and nine are conserved among all phytobacterial TTS systems (He, 1997; He, 1998). The nine conserved genes have been named hrc, which stands for hrp conserved

(Bogdanove et al., 1996), and each gene uses the same fourth letter as its

corresponding ysc homolog.

hrp gene clusters have been classified into two groups based on their genetic organization and manner of regulation (Alfano and Collmer, 1997)

(Hutcheson, 1999). Group I hrp clusters include those initially described in P. syringae pathovars, reviewed in (Hutcheson, 1999) and (Collmer et

al., 2000), and Erwinia amylovora (Steinberger and Beer, 1988; Beer et

al., 1991; Laby and Beer, 1992; Wei et al., 1992; Wei and Beer, 1993;

Wei and Beer, 1995; Kim et al., 1997; Kim and Beer, 1998; Wei et al.,

2000b). Later studies found nearly identical clusters throughout the

erwinias (Bauer et al., 1994; Frederick et al., 2001; Mor et al., 2001) and

NCBI unfinished microbial genomes). In erwinias, the hrpJ (containing

hrcV-hrcN), hrpU (containing hrcQRSTU), hrpA (containing hrcJ) and

45 hrpC (containing hrcC) operons are collinear with the corresponding transcriptional units in P. syringae, with the minor difference that the hrpU-J operon is split in two different transcriptional units (Fig. 1.6)

(Huang et al., 1993; Lidell and Hutcheson, 1994; Preston et al., 1995;

Kim et al., 1997; Deng et al., 1998; Kim et al., 1998). In erwinia hrp clusters, the hrpA-hrpC operons and hrpJ-hrpU operons are divergently transcribed with the master regulatory genes (hrpL, hrpXY and hrpS) located in between (Bauer et al., 1994; Wei et al., 2000b; Frederick et al.,

2001; Nizan-Koren et al., 2003). In P. syringae, the above operons are convergently transcribed and the regulatory genes are located on each end

(hrpL is on one side and hrpRS on the other). Unlinked candidate orthologs for hrpX and hrpY have been found in P. syringae and in P. aeruginosa, but their role is not known. Many TTS system effector genes are located in the regions flanking the core cluster (Fig. 1.6). These strong similarities suggest horizontal transfer and subsequent homing- related recombination/transposition events (Hutcheson, 1999).

The second group of hrp clusters is found in Xanthomonas spp. and R. solanacearum. These group II clusters are basically collinear to each other, with the only difference being the position of the regulatory genes hrpXXc vs. hrpBRsol (Arlat et al., 1991; Van Gijsegem et al., 1995). The

secretion genes are organized into four operons; hrcC, hrcJNT, hrcUV and hrcQRS, the last two being transcribed in the opposite direction. The sequence similarity of group II hrc genes to their mammalian TTS system

46 homologs is less than that found for group I hrc genes (Hutcheson, 1999).

The regulatory systems for group I and II hrp gene are completely different; group I hrp cluster are regulated by the alternative sigma factor

HrpL, whereas group II are regulated by AraC-like transcriptional activators (HrpB for R. solanacearum and HrpX for Xanthomonas spp.).

Further details are discussed below.

The proteins secreted by the hrp/hrc TTS systems have been named

Hops (for Hrp outer proteins), to parallel the designation of Yersinia outer proteins as Yops (Alfano and Collmer, 1997). At least four classes of

Hops travel the secretion apparatus: Harpins, Hrp pilins, translocators and effector proteins. Harpins, Hrp pilins and translocators are collectively defined as helper proteins for their presumed or demonstrated roles in aiding the secretion/translocation of other TTS system substrates (Collmer

et al., 2002).

Harpins are hydrophilic, glycine-rich, heat-stable, cysteine-free proteins

that are able to elicit an HR when infiltrated at high concentration into the

plant apoplast (Alfano and Collmer, 1997). Harpins travel through the

TTS apparatus but are released into the apoplastic spaces, rather than

being injected into host cells (Wei et al., 1992; Bogdanove et al., 1996;

Pike et al., 1998; Perino et al., 1999). Harpins have been described in E.

amylovora (harpinEa or HrpN and HrpW; Wei et al., 1992; Gaudriault et

al., 1998; Kim and Beer, 1998) in other erwinias (Bauer et al., 1995;

Mukherjee et al., 1997; Ahmad et al., 2001), in all P. syringae pathovars

47 surveyed (harpinPss or HrpZ; He et al., 1993; Charkowski et al., 1998), and in R. solanacearum (PopA1; Arlat et al., 1994), but no sequence similarity exists among the HrpN, HrpW, HrpZ and PopA1 proteins. The precise role of harpins in pathogenesis is currently unknown. In general, they are abundantly secreted into the growth medium and therefore were the first Hops to be characterized. Initially harpins were thought to be the universal HR elicitor molecule (Wei et al., 1992; Beer et al., 1993; He et al., 1993; Bauer et al., 1995), but subsequent studies in Pseudomonas syringae questioned this conclusion (Alfano et al., 1996). In P. stewartii and E. amylovora, hrpN mutations greatly affect HR elicitation in tobacco but not virulence (Barny, 1995; Ahmad et al., 2001). The reason for these differences in phenotypes could be due to functional redundancy in P. syringae or to differences in the secretion mechanism. HrpZ is able to bind plant cell walls (Hoyos and al., 1996) and create transient pores in lipid bilayers (Lee et al., 2001). P. syringae and Erwinia amylovora HrpW proteins have a C-terminal pectate lyase domain that binds pectin, a major component of cell walls (Charkowski et al., 1998; Kim and Beer, 1998).

These properties may play a role in facilitating translocation of effectors

(Collmer et al., 2002) or in the release of nutrients from host cells (Lee et al., 2001). Harpins are also of commercial interest because they induce systemic resistance responses when sprayed on plants (Dong et al., 1999;

Kariola et al., 2003).

48 Hrp pilins, such as P. syringae HrpA (Roine et al., 1997), E. amylovora

HrpA (Wei et al., 2000a), X. campestris HrpE1 (He and Jin, 2003), R.

solanacearum HrpY, (Van Gijsegem et al., 2000) are small proteins abundantly secreted into the growth medium (for a recent review see He and Jin, 2003). In contrast to animal pathogens (Knutton, 1998), Hrp pili are essential for secretion of other Hops (Van Gijsegem et al., 2000; Wei et al., 2000a). P. syringae HrpA has been shown to spontaneously assemble into a pilus-like structure in vitro (Roine et al., 1997). Unlike the short rigid needles of mammalian pathogens, Hrp pili have been observed in vivo as flexuous filaments of ca. 80 Å x 2 µm. They appear to be extended by condensation of monomers at the distal end (Jin and He,

2001). Hrp pilus assembly in vivo is hrp/hrc dependent. Hrp pili have been shown to cross the thick (100-200 nm) plant cell wall (Hu et al.,

2001).

Xanthomonas HrpF is the only Hop shown by genetic and biophysical approaches to behave as a putative translocator (Buttner et al., 2002).

HrpF is a predominantly hydrophilic protein with two hydrophobic domains at the C-terminus. It is highly similar to Rhizobium fredii NolX, which is required for legume host specificity (Huguet and Bonas, 1997).

Non-polar mutations in hrpF produce a Hrp phenotype but do not alter secretion of other Hops in vitro (Rossier et al., 2000). Experiments with lipid bilayers showed binding of HrpF and HrpF-dependent pore formation

(Buttner et al., 2002).

49 The fourth group of proteins traveling the Hrp secretion apparatus

encompasses the Hop-effectors. This is by far the largest class of Hops,

comprising more than 30 proteins per species in some cases (Collmer et

al., 2002). Secretion levels of Hops in culture and in planta are much

lower than for harpins or Hrp pilins. This delayed their discovery as TTS

system substrates. All Hop effectors share a few common properties. They

have no effect on the plant cell when applied to the apoplast and they are

vectorially delivered into the host cell. This has been shown by direct

approaches using immunocytochemical and biochemical assays, e.g.

AvrBs2 (Casper-Lindley et al., 2002) and AvrBs3 (Szurek et al., 2002), or by indirect approaches using transgenic expression and/or two-hybrid analysis (Gopalan et al., 1996; Van den Ackerveken et al., 1996; Mudgett

and Staskawicz, 1999; Kim et al., 2002). In most cases, mutagenesis of

hop genes has either no effect or only quantitatively affects virulence

(Ritter and Dangl, 1995; Alfano et al., 2000; Lavie et al., 2002; Deng et

al., 2003), suggesting functional redundancy or additivity. Some effectors

have functional nuclear localization signals and/or features of eukaryotic

transcription factors (Gueneron et al., 2000; Szurek et al., 2001; Marois

et al., 2002; Deslandes et al., 2003). Therefore, they are presumed to act

by directly altering gene expression patterns of the host. Enzymatic

activities have been shown or proposed for a few of them, as discussed

below.

50 Three subclasses of Hop effectors have been recognized from an

operational viewpoint: Avr (avirulence) proteins, Dsp/Pth (disease

specificity/pathogenesis) proteins and candidate Hop effectors. This

division is somewhat artificial, because at a molecular level their

mechanisms of action probably overlap in many cases, but it reflects the

historical and practical approach still found in the literature.

Avr (avirulence) proteins are effectors that are recognized by a resistant

plant’s surveillance system, which is represented by R gene products and

other accessory proteins (reviewed by McDowell and Dangl, 2000; Dangl

and Jones, 2001; Staskawicz et al., 2001; Holt et al., 2003). Their existence was predicted by the gene-for-gene hypothesis elaborated during the late 1930s-early 1940’s (reviewed by Flor, 1971). avr genes in bacteria were initially cloned by screening collections of virulent strains

(able to cause disease on compatible hosts) carrying cosmid libraries established from avirulent strains (able to cause HR on incompatible hosts). These transcojugants were assayed for gain of incompatibility phenotype on host cultivars carrying specific R genes (Staskawicz et al.,

1984; Gabriel et al., 1986; Staskawicz et al., 1987). Mutations in this subclass of effectors result in loss of HR elicitation in resistant/incompatible cultivars (i.e. expressing cognate R proteins) of a given host species, but usually they have no major effects on disease development in susceptible/compatible cultivars (i.e. lacking the cognate

R gene). Avr proteins are a result of the coevolution of pathogens with

51 their host resistance gene system. Genes encoding this class of proteins are frequently found in Xanthomonas and Pseudomonas, two genera of phytobacteria able to infect a great variety of hosts. Typically these species have several dozens of hop genes per genome. For some Avr proteins a great deal of information on their interaction with the R-protein system is available, but their primary function in virulence is often unknown. In a few cases, clear effects on virulence are observed.

Transposon mutations in P. syringae avrPto, avrE and avrRpm1 and

Xanthomonas avrBs2 caused a dramatic reduction of growth in planta and symptom development in specific hosts (Ritter and Dangl, 1995; Swords et al., 1996; Chang et al., 2000). Suppression of programmed cell death is another function that has been attributed to Avr/Hop effectors. Evidence for this often comes from studies where transgenic expression of Avr proteins is susceptible genetic backgrounds. For instance, AvrPtoB suppresses Pto-AvrPto-based HR defense responses leading to increased susceptibility (Abramovitch et al., 2003); similarly, cell wall based defenses (e.g. callose depositions) were suppressed in a salicylic acid- dependent manner by ectopic expression of AvrPto in Arabidopsis thaliana (Hauck et al., 2003).

Dsp/Pth proteins are operationally defined as effector proteins that contribute to virulence and symptom development for which no cognate

R-gene has been identified. One subgroup of Dsp proteins, which have high sequence similarity, has been described in erwinias. It includes E.

52 amylovora DspA/E, P. stewartii WtsE, P. agglomerans pv. gypsophilae

DspE (Gaudriault et al., 1997; Frederick et al., 2001; Mor et al., 2001) and P. carotovorum DspE (Yang et al., 2002). These all share amino acid similarity to P. syringae AvrE (Lorang and Keen, 1995). A second example of avr-like genes that are essential for virulence is X. citri PthA, a member of the large avrBs3 family (Dangl, 1994). pthA expression in X. campestris changes the pathogen’s host-specificity (Swarup et al., 1991).

Mutations in pthA eliminate virulence on its normal host citrus (Swarup et al., 1992), and transient expression of pthA in susceptible host cells led to elicitation of cell death and hypertrophic symptoms (Duan et al., 1999).

Candidate effectors are putative Hops for which a virulence or avirulence function has not been demonstrated. They have been identified by sequence homology to known Hops or by various screening procedures.

These include in silico screens of genomic sequences for Hrp promoters and secretion motifs or functional screens for Hrp regulated promoters,

Hrp-dependent secretion or in planta expression (Boch et al., 2002; Fouts et al., 2002; Petnicki-Ocwieja et al., 2002), reviewed in (Collmer et al.,

2002). They are the fastest growing class of effectors. Some Hop effectors are homologous to known avirulence proteins. For example, HopPtoJ is the P. syringae homolog of X. vesicatoria AvrXv3 and HopPtoC is the homolog of AvrPpiC2 (Petnicki-Ocwieja et al., 2002). For some, Hrp- dependent secretion and or translocation has been demonstrated (e.g.

HopPtoD1, HopPtoE, HopPtoG, HopPtoH, HopPtoJ, HopPtoK and many

53 others, Petnicki-Ocwieja et al., 2002; HopPtoB1, Alfano et al., 2000;

Deng et al., 2003; HopPtoA1 and HopPtoA2, Badel et al., 2002). Others have enzymatic properties that enable them to modify plant defense responses. For example, HopPtoD2 is a tyrosine phosphatase (Espinosa et al., 2003) and AvrPphB is a Cys-protease (Shao et al., 2002). Sequence homology suggests that enzymatic activities may be associated with other

Hops. For example, HopPmaG may be a transglycosylase; HopPmaH may be a pectin lyase; HopPtoC and HopPtoN may be cysteine-proteases; and

HopPtoS3 may be an ADP-ribosyltransferase (reviewed in Collmer et al.,

2002). Recently, another candidate hop gene, HopPtoA1, has been shown to contribute to P. syringae colony formation in the apoplast of a compatible host (Badel et al., 2002).

1.2.3 Regulation of type III secretion systems

During the infection process, TTS systems are transcriptionally and post-translationally activated in a highly regulated fashion. The pathogen responds to initial cues provided by the host, adapts to subsequent host responses, and thereby creates a new niche while modulating its “invasion program” over time and space. In mammalian pathogens, two distinct steps can be recognized (Francis et al., 2002): an initial phase where components of the TTS system are made and assembled and then a second phase where the effector proteins are produced and targeted for secretion.

In reality, some temporal overlapping occurs, which may contribute to

54 differences in the timing and ordered secretion of various classes of effectors. In plant pathogens, the secretion apparatus and the effector genes always appear to be co-regulated at the transcriptional level. In general, one important difference between animal and plant pathogens is that gram-negative plant pathogens always remain extracellular. By adhering to the cell wall, some 100-200 Å away from the host plasma membrane, they experience a relatively homogeneous environment inside the host tissues. On the other hand, a mammalian pathogen must be able to deal with a diverse range of host compartments until it reaches an appropriate site of infection. Moreover, mammalian pathogens have direct access to the host membranes and some are even able to internalize and multiply in phagosomal vescicles (e.g. Salmonella and Shigella spp.)

(Cossart et al., 2000).

All TTS systems are controlled by complex regulatory networks that are able to integrate various environmental cues, including chemical, nutritional and host signals, i.e. pH; oxygen tension; redox potential;

3- 2+ 2+ ammonium, PO4 , Mg , Ca and nucleotide concentrations; growth phase; and availability of specific nutrients (Hueck, 1998), as well as microbial signals, such as quorum sensing molecules (Sperandio et al.,

1999). The necessity to minimize energy consumption or to express the secretion system only at the right time and place are often cited as the main reasons for such underlying complexity (Hueck, 1998). Among the regulators involved in controlling the transcription of TTS system

55 apparatuses and their effectors are: two-component systems (Grimm and

Panopoulos, 1989; Bernardini et al., 1990; Behlau and Miller, 1993;

Frederick et al., 1993; Xiao et al., 1994; Bajaj et al., 1995; Grimm et al.,

1995; Nakayama and Watanabe, 1995; Johnston et al., 1996; Ochman et al., 1996; Shea et al., 1996; Wengelnik et al., 1996; Merighi et al., 2003;

Nizan-Koren et al., 2003), AraC-like activators (Sakai et al., 1986;

Yother et al., 1986; Cornelis et al., 1987; Forsberg and Wolf-Watz, 1988;

Cornelis et al., 1989; Frank and Iglewski, 1991; Genin et al., 1992;

Kaniga et al., 1994; Gomez-Duarte and Kaper, 1995; Tobe et al., 1996;

Wengelnik and Bonas, 1996), alternative sigma factors of various classes

(Wei et al., 1995; Xiao et al., 1994; Cui et al., 2002), global regulators affecting DNA topology (Cornelis et al., 1991; Hromockyj et al., 1992;

Cui et al, 2002), effector protein chaperones (Darwin et al., 2001), TTS pilus components (Wei et al., 2000a), and housekeeping proteases (Bretz et al., 2002). Table 1.1 summarizes some of the known regulators of TTS systems. Examples of complex regulatory pathways described in mammalian pathogens are shown in Fig. 1.7.

1.2.3.1 Regulation of group I hrp/hrc genes

In Group I hrp clusters, the alternative sigma factor HrpL enables the recognition and transcriptional activation of promoters containing “Hrp boxes” in their –10/-35 regions (Fellay et al., 1991) (Fig. 1.8). Hrp box promoters have been found in front of all known hrp/hrc secretion

56 operons of group I Hrp clusters (Huynh et al., 1989; Innes et al., 1993;

Salmeron and Staskawicz, 1993; Shen and Keen, 1993; Xiao and

Hutcheson, 1994; Wei and Beer, 1995; Chatterjee et al., 2002; Fouts et

al., 2002) and most, but not all (Shen and Keen, 1993), avr genes. HrpL

belongs to the ECF (extracytoplasmic function) family of sigma factors

(Lonetto, 1991) but, in contrast to other members of this family (Hughes and Mathee, 1998; Raivio and Silhavy, 2001; Helmann, 2002), no anti- sigma factor has been described.

Transcription of hrpL is under the control of a σ 54 promoter based on

three observations: i) the hrpL promoter regions always contain a putative

–12/-24 σ 54 promoter consensus sequence; ii) in P. syringae, expression of hrpL is strongly reduced in an rpoN mutant (Fellay et al., 1991;

Hendrickson et al., 2000; Hendrickson et al., 2000); and iii) expression of

P. stewartii hrp genes is reduced in an E. coli rpoN strain (Frederick et

al., 1993). In P. stewartii, a region with high A+T content and an IHF-

like consensus sequence is present just upstream of the σ 54 promoter

(Merighi et al., 2001). Similarly, A. Chatterjee and coworkers (Chatterjee

et al., 2002) found that expression of Pectobacterium carotovorum hrpL

in E. coli is dependent on himA (=IHF).

Transcription initiation and elongation at σ 54 promoters typically

requires NtrC-class of enhancer-binding proteins (Hirschman et al., 1985;

Weiss et al., 1991). In this regard, all group I hrp clusters use NtrC-like

transcriptional enhancers (Grimm and Panopoulos, 1989), such as hrpR

57 and hrpS, to activate hrpL (Rahme et al., 1991; Frederick et al., 1993;

Xiao et al., 1994). This csubclass of enhancer-binding proteins have a C-

terminal DNA binding domain carrying a helix-turn-helix domain of the

Fis-like family and a σ 54-activating ATPase domain, but they lack an N- terminal phosphorylatable receiver domain (Grimm and Panopoulos,

1989). Response regulators with this structure are expected to be constitutively proficient for binding to DNA at tandem symmetrical dyad

sequences (Klose et al., 1994), even if oligomerization, which usually

follows phosphorylation of the receiver domain, may not occur. In P.

syringae, the HrpS and HrpR orthologs have been shown to interact in vitro and both proteins are required for full activation of hrpL (Hutcheson

et al., 2001). A symmetrical dyad located 90-110 bp upstream of the σ 54-

promoter is present in P. syringae (Hutcheson, 1999) and P. stewartii

(Merighi et al., 2003). In Chapter 3, we will show that this dyad is

required for activation of hrpL by HrpS in P. stewartii .

In P. syringae, expression of hrpR, hrpS and downstream genes is

environmentally regulated. Complex amino acid sources, such as peptone,

suppress expression of hrpR, hrpS and other hrp genes in P. syringae pv.

phaseolicola (Rahme et al., 1992) and other pathovars (Xiao et al., 1992;

Xiao et al., 1994). M9 minimal medium induces hrpR and hrpS to low

levels, while growth in tobacco increases their expression about 1000-fold

(Rahme et al., 1992) in as little as 1 h post-infiltration (Xiao et al., 1992).

A modified minimal medium with optimized salts, nitrogen, carbon

58 sources and pH further increases expression over M9. Interestingly, Krebs cycle intermediates, such as succinate and citrate repress hrp genes in P. syringae, probably via catabolite repression (Huynh et al., 1989). How environmental signals are coupled to hrpRS and hrpL activation is not known.

In addition to HrpS and HrpL, the erwinias also have a two-component signal transduction system encoded by the hrpXY operon, which is located downstream of hrpL. In Erwinia amylovora, Wei et al. (2001) suggested that HrpY directly controls expression of hrpL in concert with HrpS.

Growth in an acidic, low ionic strength, low phosphate, minimal medium containing mannitol and 5 mM ammonium sulfate can induce hrp genes to levels comparable to those measured in tobacco tissues and higher than those observed in pear tissues (Wei et al., 1992). Addition of nicotinic acid, certain amino acids (His, Asp, Trp and Gly), 0.1% casamino acids, or some carbon sources (glucose and maltose) repress hrp genes (Wei et al., 1992). The nature of the primary signals transduced by the two- component system and the regulatory circuits controlling hrpS are not known. In Pectobacterium spp., where hrp genes play a marginal role in virulence (Rantakari et al., 2001), hrpS is hypoyhesized to be constitutively expressed (Chatterjee et al., 2002) and the upstream pathways of hrp regulation are not characterized. P. carovovorum hrpL appears to be regulated by RpoN, HrpS and IHF at the transcriptional

59 level and by the RsmA/rsmB system at the post-transcriptional level

(Chatterjee et al., 2002). More details on this global regulatory system are

given in the Introduction to Chapter 6.

An interesting hrp-regulatory effect has been attributed to HrpA in P.

syringae. HrpA is the major component of the Hrp pilus, In addition to

being absolutely required for secretion of two effectors, the hrpA gene

appeared to be essential for full expression of genes encoding other Hrp

regulatory, secretion, and effector proteins. In particular, hrpA-mediated

gene regulation was shown to affect levels of hrpR and hrpS mRNA (Wei

et al., 2000a). The effect of HrpA appears to be unrelated to its role in

protein secretion because it is possible to isolate HrpA missense mutants

that show reduced hrpRS expression, but exhibit normal secretion of Hops

(Wei et al., 2000a).

Most of the regulatory circuits described above involve transcriptional

activation. Only one example of negative regulation of hrp genes is

known so far. A negative feed back loop triggered by HrpL activation of

the hrpC secretion operon has been described in P. syringae (Preston et

al., 1998). This may be due to over-production of HrpV, because non-

polar mutations in hrpV resulted in increased secretion of harpin and upregulation of three secretion operon reporter gene fusions. Epistasis analysis further suggested that hrpV acts as a hrp repressor upstream of hrpRS (Preston et al., 1998).

60 The Hrp regulatory cascade is also subject to post-translational regulation. Work in S. Hutcheson’s lab has recently shown that the Lon protease has a negative regulatory effect on the expression of hrpL via post-translational degradation of HrpR (Bretz et al., 2002). P. syringae lon::Tz mutants are upregulated for hrpL transcription, exhibit enhanced

AvrPto secretion, and elicit faster plant responses. Immunoblot analysis showed that the half-life of HrpR is increased in a lon mutant under hrp- inducing conditions in vitro. The authors further suggest that transcription of hrpRS is constitutive, but this contradicts other workers’ results (Wei et al., 2000a) and may be due to differences in the techniques used to measure gene expression (Northern analysis vs. RT-PCR).

1.2.3.2 Regulation of group II hrp/hrc genes

Group II hrp clusters are regulated in response to environmental stimuli similar to those affecting expression of group I hrp genes, such as growth in minimal media. For X. campestris, minimal media with glucose and/or fructose are stimulatory, but for X. vesicatoria either a tomato conditioned medium (Schulte and Bonas, 1992) or a specialized minimal medium (XVM1 and XVM2) containing methionine, sucrose and low phosphate concentrations is necessary to activate hrp genes (Wengelnik and Bonas, 1996). For R. solanacearum, pyruvate and glutamate are the most effective carbon sources for inducing hrp genes in vitro (Arlat et al.,

1992).

61 Expression of group II hrp/hrc genes is usually controlled directly by

AraC-like regulators (Fig. 1.8). In R. solanacearum the AraC-like

activator is encoded by hrpB, which was the first member of this class of hrp regulators to be discovered (Genin et al., 1992), and in X. vesicatoria, it is encoded by hrpX. Unlike hrpBRsol, hrpXXv is not linked to the hrp

cluster (Wengelnik and Bonas, 1996). These AraC-like transcription

factors directly activate expression of most of the hrp secretion operons

by recognizing specific promoter elements (Bonas, 1994; Wengelnik and

Bonas, 1996). With the exception of avrBs3 (Knoop et al., 1991), most

avr genes in Xanthomonas are also part of the Hrp regulon (Noel et al.,

2001).

Additional regulators control expression of the AraC-like genes. In R.

solanacearum, the OmpR-like response regulator HrpGRsol was shown to

control hrpBRsol expression (Brito et al., 1999). More recent results indicate that HrpGRsol may also independently control other unrelated

functions important for vascular colonization (Vasse et al., 2000).

hrpGRsol itself is controlled by a complex regulatory cascade. The TonB- like receptor PrhA (plant regulator of hrp genes) (Marenda et al., 1998) is

located in the outer membrane and perceives cell-contact signals (Aldon

et al., 2000) but not chemical/nutritional stimuli (Marenda et al., 1998).

Upon perceiving the plant signal, PrhA activates PhrI/PhrR (Brito et al.,

2002), which is probably an ECF sigma factor/antisigma factor complex.

Mutations in these regulators delay symptom development (Brito et al.,

62 2002). The PhrI sigma factor activates transcription of the FixJ-class

response regulator PrhJ (Brito et al., 1999), which in turn activates

hrpGRsol in presence of plant cells. HrpGRsol is required both for induction

in minimal medium and in planta, whereas PrhJ was only needed for

transduction of plant signals (Brito et al., 1999). This suggests that

hrpGRsol is regulated by two independent pathways, one dependent on

plant signals (perhaps cell contact) and the other on nutritional and

physiological variables.

In X. campestris, the OmpR-class, orphan response regulator HrpGXc activates transcription of hrpXXc (Schulte and Bonas, 1992). Interestingly, activation of the hrcC secretion gene is dependent on hrpGXc, but not on

hrpXXc (Wengelnik and Bonas, 1996). Certain amino acid substitutions in

HrpGXc that constitutively activate hrp gene expression also enhance the

timing and intensity of disease symptoms (Wengelnik et al., 1999). How

HrpGXc is regulated at a post-translational level and how plant and

metabolic signals are transduced is not known.

1.3 MECHANISMS AND STRATEGIES OF TWO-COMPONENT

SYSTEM REGULATION

1.3.1 Regulation of virulence factors

Bacterial pathogens activate their virulence mechanisms in response to

external cues signaling the presence of specific host compartments. This

implies tight regulation of the corresponding virulence genes. Gene

63 expression is modulated by multiple signals, including salt concentration,

pH, redox potentials, iron availability, carbon sources, quorum sensing

molecules, temperature, small organic compounds, and growth phase.

(Finlay and Falkow, 1997; Wilson et al., 2002). In only a few cases is the

precise mechanism of signal perception and transduction known. These

signals may turn on or off the various virulence functions at any level of

regulation, i.e. transcriptional, post-transcriptional, translational and

post-translational. Transcriptional regulation appears to be the most

frequently used because it is the most energetically cost-effective for the

microbe. It is also the more commonly studied because is easier to identify by genetics means. A great variety of regulatory circuits have

been described, from systems where one master regulator is controlling

many virulence genes to pathways where many regulators are involved in

controlling one or a few virulence functions. The biochemical cross-talk

that develops during pathogenic interactions implies that many bacterial

genes are required at various stages of the infection and multi-component

regulatory cascades are needed to integrate the many signals to which

they respond.

Various classes of regulators have been described in bacterial pathogens

(Table 1.1). Among these we find two-component signal transduction

systems, alternative sigma factors (σ 54, σ 32, σ F, σ 28 and σ E-like ECF

sigmas), quorum sensing transcriptional regulators, AraC-like activators,

LysR-type regulators (e.g. Salmonella SpvR, Agrobacterium OccR, and

64 Rhizobium NodD), other activator/repressors directly controlled by small

molecules or ions (e.g. CAP and Fur), and global regulators affecting the

local DNA topology (IHF and H-NS). The remainder of this section will

focus on the generalities of two-component systems.

1.3.2 Two-component systems

Two-component systems (TCS) were first described in the mid-1980s

(Nixon et al., 1986; Ninfa, 1986; Hess, 1988). Since then, hundreds of

TCSs have been described in eubacteria and archea, and to a lesser extent in yeasts, filamentous fungi, and plants (Stock et al., 2000). In bacterial species for which genomic information is available, there are from none

(Mycoplasma genitalium) to 80 (Synechocystis spp.) TCSs. In E. coli, there are at least 30 TCSs (Mizuno, 1997). Many reviews (Stock, 1989;

Bourret et al., 1989; Parkinson, 1992; Stock et al., 1994; Hoch, 2000;

Stock et al., 2000) and books (Hoch et al., 1995) are available on the topic. Here I will briefly introduce some of the structural and functional properties of bacterial TCSs.

TCSs couple signal perception to transcriptional activation of genes (or post-translational activation of enzymes and molecular machines) as a way for bacteria to adapt to diverse niches. TCSs are pairs of proteins with a modular structure (Fig. 1.9). In the simplest case, a sensory histidine kinase (HK) protein, with an N-terminal, sensing (input) domain and a C-terminal conserved kinase core (transmitter) domain, interacts

65 with a response regulator protein (RR), composed of a conserved N-

terminal, regulatory (receiver) and a variable C-terminal effector (output) domain. The RR protein is usually a transcription factor, but there are

exceptions to this general rule. The common theme is that activation of

the RR triggers specific downstream responses. Transduction of stimuli

occurs via phosphorylation reactions involving conserved amino acid

residues in the transmitter and receiver domains. The chemistry of TCSs

includes three phosphotransfer reactions involving His and Asp residues

(Hoch and Silhavy, 1995; Stock et al., 2000):

Autophosphorylation: HK-His + ATP ⇔ HK-His~P + ADP

Phosphotransfer: HK-His~P + RR-Asp ⇔ HK-His + RR-Asp~P

Dephosphorylation: RR-Asp~P + H2 O ⇔ RR-Asp+ Pi

All of these reactions require divalent cations (Mg2+ in particular) for

proper geometry at the catalytic site. HKs catalyze the formation of

phosphoramidate (N~P) bonds, which have a highly negative free energy

of hydrolysis. N~P bonds are acid labile and alkali stable. HKs are

preferentially unphosphorylated due to the thermodynamic properties of

the N~P bond and the normally low ATP/ADP ratio in the cell (Stock et

al., 2000). Perception of the stimuli by the input domain of the HK

changes this equilibrium and the direction of the phosphoryl group flow.

RRs receive the phosphoryl group from their cognate HK forming a high-

energy acyl phosphate bond. Asp~P bonds are acid and alkali labile. The

energy of the Asp~P residues in the context of the

66 increases the free energy of hydrolysis and this is thought to drive the

long-range conformational changes used by RRs for activation of their

function (Stock et al., 2000). The half-life of the Asp~P moiety of

bacterial RRs in vitro ranges from a few seconds to several hours. It is

often further reduced by the intrinsic autophosphatase activity of the RR

protein, by the presence of phosphatase activity in the cognate HK, or by

the action of cellular regulatory phosphatases (Hoch et al., 1995) (Fig.

1.10).

In addition to the prototypical His-to-Asp phosphotransfer, more complex pathways may be employed to transfer Pi from ATP to the

response regulator. In particular, so called hybrid sensor kinases

(described below) and multicomponent phosphorelay systems use a three-

step His-to-Asp-to-His-to-Asp cascade, which implies the existence of

five distinct reactions (autophosphorylation, three phosphotransfer

reactions and a dephosphorylation step) (Stock et al., 2000). In

phosphorelay systems, the His- and Asp-containing signaling modules are

on separate proteins, instead of being covalently linked. This kind of

signaling molecule is exemplified by the KinA/Spo0F/Spo0B/Spo0A

sporulation cascade of Bacillus subtilis and they are very common in

eukaryotic TCS transduction systems (Stock et al., 2000).

67

1.3.2.1 Response regulators

Response regulators are the “terminal effectors” of the adaptive

response (Stock et al., 2000). They can catalyze the phosphotransfer independently of the HK, as shown by the ability of acetyl phosphate, carbamoyl phosphate, imidazole phosphate and phosphoramidate to function as phosphodonors to RRs (Lukat et al., 1992). Most RRs are transcription factors with a DNA-binding domain in the output region.

Some have enzymatic activities associated with the output domain (e.g. E.

coli CheB methylesterase) and others have no output domain, but are able

to interact with other proteins modulating their activity (e.g. E. coli CheY

interacting with FliM in the flagellar motor) (Fig. 1.9).

The receiver domain is conserved in all RRs. The crystal structures of

several Rrs, such as CheY, NarL and OmpR, show a doubly-wound (α/β) 5 fold. Conserved amino acid residues are located around the catalytic site: a phosphorylated Asp is present at the end of β-strand 3, two Asp residues are in loop 1, a Lys is in loop 5, and a Thr is in loop 4. (In CheY, these would correspond to D12, D13, D57, T87 and K109.) The three conserved

Asp side chains form a catalytic triad coordinating Mg2+ and they are

involved in the phosphorylation and dephosphorylation reactions. Usually,

the most C-terminal Asp residue forms the acyl~P bond. Invariant Thr and

Lys residues also appear to be involved triggering in phosphorylation-

induced conformational changes (Stock et al., 2000).

68 The output domains of RRs with DNA-binding activities are classified

into three different families. OmpR-like regulators form the so-called

“winged-helix” subfamily, the largest group of RRs, based on the analysis

of all the sequenced bacterial genomes (Hoch and Varughese, 2001). This

subfamily is characterized by a DNA-binding fold containing a

recognition helix that binds the major groove of DNA and two flanking

loops (the “wings”) that recognize the minor groove. The mode of action

of different RRs varies, with some recognizing and activating different

portions of the RNA polymerase holoenzyme/promoter complex (e.g.

αCTD or σ 70). In the case of PhoB, phosphorylation promotes

dimerization of the protein (McCleary, 1996; Fiedler, 1996). Direct DNA

repeats are often the binding site for this class of RR and these proteins

bind DNA as asymmetrical dimers (Rampersaud et al., 1989; Harlocker et

al., 1995; Harrison-McMonagle et al., 1999). FixJ-like regulators have

helix-turn-helix DNA-binding motifs in their output domains similar to

the LuxR/GerR-subfamily of transcription factors (Henikoff et al., 1990).

For some members of this class (e.g. UhpA and NarL/NarP),

phosphorylation has been shown to promote symmetrical dimerization and

binding to inverted repeat DNA sequences (Stewart, 1993; Olekhnovich and Kadner, 2002), but direct repeats flanking symmetrical dyads are also targets for cooperative binding to large promoter regions (Li et al., 1994).

Finally, the third subfamily includes NtrC-like RRs, which have output regions composed of an ATPase domain and a helix-turn-helix DNA

69 binding domain (Hoch and Silhavy, 1995). The mechanism of activation

for NtrC has been extensively studied (Keener and Kustu, 1988; Su et al.,

1990; Weiss et al., 1991; Klose et al., 1993; Porter et al., 1993; Klose et

al., 1994; Wedel and Kustu, 1995; Wyman et al., 1997; Pelton et al.,

1999; Rombel et al., 1999). Unphosphorylated NtrC dimers can bind DNA

and the major dimerization domain is located in the C-terminal region.

Phosphorylation, by the NrtB kinase or small phosphodonors, promotes

dimer-oligomerization to larger complexes. Mutations at the conserved

D54, which is the site of phosphorylation, to glutamate mimic the D~P

residue. Finally, NtrC oligomers tethered to the enhancer sites stimulate

ATP hydrolysis, which supplies free energy for the RNA polymerase

holoenzyme/σ 54-promoter isomerization from a closed to an open

complex, activating transcription.

1.3.2.2 Sensor histidine kinases

Sensor HKs are modular proteins located either in the cytoplasm or,

more commonly, transmembrane (Fig. 1.9). They perceive

environmental/cellular signal(s) and transduce them to a cognate response

regulator located in the cytoplasm. HKs range in size from 40 to 200 kDa

and contain from two to six distinguishable domains. In the simplest case, the input domain of the HK controls the phosphorylation events occurring at the kinase core domain by altering its rate of autophosphorylation. In some cases, separate His phosphotransfer (HPt) domains are present.

70 These are able to undergo phosphorylation, but unable to show kinase or phosphatase activities, when separated from the remaining domains of the

HK. In transmembrane sensor kinases, there are often linker domains of

40 to 120 amino acids connecting the sensing/transmembrane domains to the cytoplasmic transmitter and these are important for signal propagation

(HAMP; Appleman et al., 2003; Appleman and Stewart, 2003).

Autophosphorylation is a bimolecular reaction involving cross- phosphorylation of HK dimers at conserved His residues located in their kinase core domains (Stock et al., 2000). In some cases, the HK core also catalyzes the dephosphorylation reaction (e.g. NtrC and FixL), which accelerates the return to the RR “ground” state.

The kinase core of prototypical HKs is about 350 amino acids long and is composed of a dimerization domain and an ATP/ADP-binding phosphotransfer domain or catalytic domain (Stock et al., 2000). The kinase core contains five recognizable amino acid motifs (Hoch and

Silhavy, 1995): the phosphorylation site (H box) and the N, D, F and G boxes in the nucleotide binding site. The dimerization domains form antiparallel four-helix bundles and each usually includes the H box in helix 1. In a few kinases, such as CheA, the H box is separated from the dimerization domain and located in a more N-terminal His- phosphotransfer domain (HPt) (Fig. 1.9). The catalytic domain contains the N-D-F-G boxes and binds ATP/ADP at a highly flexible cleft. Most structural information is based on the E. coli CheA and EnvZ protein

71 fragments derived from X-ray diffraction crystal structures and/or NMR

solution structures, but a complete structural analysis for a full length HK

is not available yet.

The kinase core interacts with the receiver domain in the RR during the

phosphotransfer step. How these two domains interact has been partially

revealed by cocrystalization of the Spo0F response regulator with the

cognate Spo0B phosphotransferase (Zapf et al., 2000). The interaction surface involves residues on the loops connecting the α-helices and β- strands of the RR and the α1 and α2’helices in the dimerization/H box of

the HK (Hoch and Varughese, 2001). Invariant residues are involved in

the catalytic process. Such residues are brought into the proper geometry

upon the interaction of conserved anchor residues and variable

recognition residues at the interaction surface. The variable recognition

residues are responsible for preventing heterologous interactions among

non-cognate HK-RR pairs, thereby ensuring specificity in signal

propagation (Zapf et al., 2000).

Based on the domain organization, at least two major classes of HK

have been recognized. Orthodox histidine kinases are either soluble or,

more commonly, transmembrane proteins composed simply of an input

sensory domain and a cytoplasmic transmitter domain (Fig. 1.10). Most

HKs fall into this class. Hybrid sensor kinases are composed of a set of

additional RR-like receiver and HPt domains at the C-terminus of an

orthodox HK structure. This domain architecture allows for an

72 intramolecular His-to-Asp-to-His phosphotransfer cascade. This general

organization is found in several hybrid kinases such as E. coli ArcB,

Salmonella spp. BarA, and Bordetella BvgS, but in a few cases either the

HPt domain is missing (E. coli RcsC, Mixococcus spp. FrzE) or

duplicated/triplicated receiver domains are present, with (Vibrio cholerae

NP_230993) or without (X. campestris NP_642807; P. putida NP_744503)

an HPt domain (CDART analysis at the NCBI database).

Input sensory domains are essential components of HKs but in general

they are poorly characterized. As mentioned above, in the majority of the

cases, the nature of the specific signals perceived at the molecular level

are not known. Moreover, the mechanisms by which the sensory domain

controls the autophosphorylation rate at the transmitter domain are also

not well understood. In some cases, deletion of the input domain releases

the autokinase activity of the sensor protein (Monson et al., 1992;

Scholten and Tommassen, 1993; Kramer and Weiss, 1999), whereas in

other systems reduces the autophosphorylation activity (KinA; Wang et

al., 2001) or specifically increases the phosphatase activity (E. coli

KdpD; Jung and Altendorf, 1998). Input domains generally show little

sequence conservation, even if a few families of signaling motifs have

been recognized. The PAS (Per-Arnt-Sim) domain superfamily (Pfam

00989) is perhaps the best characterized (Zhulin, 1997; Taylor, 1999;).

PAS domains have been found in over 1100 proteins encoded by both prokaryotic and eukaryotic genomes. PAS domains are cytoplasmic, small

73 (100-120 amino acids) protein motifs that fold into a conserved PAS core,

containing a β-scaffold and a helical connector (See Chapter 5 for more

details on their structure). The structural elements of the mixed α/β fold

are designated A , B , C , D , E , F , G , H , and I and a standardized

numbering system has been proposed to facilitate comparison of the

various PAS motifs (Amezcua et al., 2002). PAS domains are commonly

involved in protein/protein interactions and intermolecular association of

macromolecules (Amezcua et al., 2002). They also function as sensory

modules. Various cofactors, which determine signal specificity, are

accommodated within the hydrophobic PAS core. PAS domains have been

shown to bind many different types of cofactors, such as heme, to sense

O2 (FixL), 4-hydroxycinnamic acid to sense blue light (PYP), FAD, to

sense redox levels (NifL), FMN, to sense ble light (plant phototropins)

and ATP, to sense the energy charge (KinA) (Stephenson and Hoch,

2001). In bacteria, PAS motifs are exclusively found in TCSs, as single,

tandem or multiple domains (up to six). The presence of multiple PAS

motifs in a sensor raises the interesting possibility of multiple ligands

being perceived by a single protein. Due to their importance both as

universal sensory modules and as key components of important bacterial

virulence regulatory systems, PAS domains are being characterized in

great detail using biophysical methods. Intramolecular conformational

changes upon ligand binding have been recently observed in solution

using NMR (Amezcua et al., 2002). In particular the FG loop has been

74 shown to be critically important for the inhibition of the kinase core activity in the absence of a ligand in the PAS core pocket. Binding of hydrophobic ligands would trigger conformational changes that propagate to the Fα helix, the FG loop and the Hβ strand, thereby relieving the

inhibition of the kinase core (Amezcua et al., 2002).

Other families of motifs are indentifiable in input sensory domains of

TCS kinases (reviewed in more detail by Galperin et al., 2001). For some, the ligand is known or hypothesized. These motifs include the FliY-like periplasmic binding protein (PBPb) domain (~220 aa; Pfam 00497), often involved in amino acid binding, the Cache domain (~120 aa; Pfam02743), the MHYT module (~200 aa; COG3300), the GAF sensor domain (~150 aa; Pfam 01590), similar to a PAS fold but binding cGMP, the HAMP linker domain (~50 aa; Pfam00672) and the CBS domains (Pfam 00517).

1.4. SIGNIFICANCE, RATIONALE AND OBJECTIVES

1.4.1 Significance

Stewart’s wilt is an economically important disease of sweet corn in the

eastern part of the U.S. “Corn Belt”, especially after mild winters. It can

cause up to 100% yield losses in the field. Virtually no direct control

measures exist and disease management is mainly based on pesticide

treatments to control the insect vector and the use of tolerant sweet corn

cultivars. P. stewartii is a quarantine pathogen for more than 100

countries importing U.S. hybrid corn seed and frequent commercial

75 controversies arise between the U.S.A. and the European Community or

Far East countries on issues related to the presence of P. stewartii in corn

seed stocks. Understanding the virulence mechanism of P. stewartii at a

molecular level will eventually open new avenues for the rational

management of Stewart’s wilt and perhaps for the elimination of

transmission of the pathogen via seed.

P. stewartii-corn pathosystem is a good model for studying hrp type III

secretion apparatuses and their regulation because: (a) P. stewartii is a

good endophyte that can colonize a wide range of grasses without causing

symptoms in many of them, it causes disease in a few hosts, including

cultivated corn, and it multiplies in the intestinal tract of insect vectors.

In contrast to P. agglomerans, P. stewartii is not found in environmental

samples or on the surfaces of plants because it survives poorly under

water stress and in sunlight. (b) P. stewartii uses two of the major

virulence mechanisms of phytobacteria, i.e the production of EPS to plug

water-conducting xylem vessels and the production of effector proteins to

cause watersoaking of parenchymatic tissues. This offers the possibility to

study the coregulation of two important pathogenicity mechanisms (cps

and hrp genes). (c) The master regulators of hrp genes are conserved in

all erwinias. Therefore, we hope that any information gained from P.

stewartii may be applicable to all other members of this group. (d) Most

genetic tools developed for E. coli genetics are applicable to P. stewartii.

(e) The P. stewartii genome is being sequenced at the Baylor College of

76 Medicine and the University of Wisconsin at Madison and a draft sequence should be available at the begining of 2004. This will greatly facilitate functional analysis of pathogencity genes in this species and the study of their regulatory networks. (f) Corn is an ideal experimental host because it is easy to grow and and most of the Stewart’s wilt symptoms can be observed on young seedlings. (g) Corn is also one of the best genetically characterized crops; detailed genetic maps, vast EST libraries and numerous molecular reagents, including BAC and yeast two-hybrid libraries and transposon mutant collections, are publicly available.

1.4.2 Rationale

hrp genes are essential for the pathogenicity of P. stewartii as they are for many other necrogenic phytobacteria. Several observations make the study of hrp regulation particularly attractive in P. stewartii.

In many non-host plants, P. stewartii often remains latent without eliciting apparent host defense responses even at high inoculum doses.

For example, P. stewartii cells infiltrated at high cell density into tobacco will not cause hypersensitive plant cell death unless the bacteria are pre- grown in minimal media that mimics the metabolic signals that activate hrp/hrc type III secretion genes in planta. Many plants, such as tobacco and various grasses, are apparently unable to provide adequate signals to alert P. stewartii that it is in a potential host. Understanding the molecular mechanisms by which hosts and the environment control the

77 expression of virulence determinants may open the way to new strategies

for disease control, for instance by interfering with the sensory system of

the pathogen before an R-Avr interaction is even established.

Metabolic control of virulence gene expression plays an important role

in group I hrp genes but little is known about the earliest steps of signal

sensing. The hypothesis put forward in this dissertation is that the HrpX

PAS-kinase and the HrpY response regulator form a signal transduction

system that senses nutritional/energy stresses. The discovery of potential

HrpX/HrpY orthologs in the unfinished genome sequences of X. campestris and P. syringae DC3000 further suggests that control by starvation signals may be common to all group I and II hrp systems and perhaps to other pathogens. Once the genome of P. stewartii is sequenced, the regulatory networks that cross-talk with the HrpX/HrpY system and other downstream master regulators may be dissected. The molecular characterization of the regulatory cascade involving HrpX/HrpY, HrpS and HrpL will provide a solid framework for future studies at the genome- wide level.

Quorum sensing regulation of P. stewartii capsule polysaccharide production has been well characterized (Von Bodman, et al., 1998).

Mutants unable to produce OHHL produce very little EPS and are severely affected in their ability to cause wilting and watersoaking in corn seedlings. While the original report on esaI mutants focused on EPS production, the reduction in watersoaking due to this mutation suggested

78 that quorum sensing may also affect hrp gene expression, because cps mutants are still able to form lesions. This may imply a novel modality of host-pathogen interaction in the early phases of pathogenesis. At the same time it opens new avenues for controlling Stewart’s wilt by targeting the quorum sensing signaling mechanism.

1.4.2 Objectives

This dissertation deals with several aspects of the signal transduction cascade that activates and/or represses hrp regulatory genes in P. stewartii. In particular, I focused primarily on the regulation of hrpS by the HrpX/HrpY two component and the EsaR/EsaI quorum sensing systems and secondarily on the regulation of hrpL by HrpY. HrpS and

HrpL are two key transcription factors that control the activation of all known hrp/hrc/wts genes in P. stewartii. The HrpX/HrpY system is predicted to mediate a typical adaptive response to environmental signals and the EsaR/EsaI quorum sensing system mediates cell-to-cell communications.

The major objectives of my research were:

1. To genetically dissect the regulatory cascade controlling

hrp/hrc/wts gene expression in P. stewartii, which involves the

HrpX sensor kinase, the HrpY response regulator, the HrpS

enhancer-binding protein and the HrpL alternative sigma factor

(Chapter 2; published in Merighi et al., 2003).

79 2. To analyze at the genetic and biochemical levels a novel

autoregulatory loop triggered by HrpS (Chapter 3).

3. To analyze the biochemical properties of the response regulator

HrpY and its novel dual role as activator of hrpS and a repressor of

hrpL (Chapter 4).

4. To genetically dissect the sensory domains of HrpX and establish

reagents for the future study of its biochemical functions (Chapter

5).

5. To define the classes of signals perceived by the various

components of the regulatory cascade controlling hrp genes

(Chapter 5).

6. To determine the role of AHL-based quorum sensing in modulating

hrp gene expression in P. stewartii (Chapter 6).

While the data presented in Chapters 2, 3 and 6 can be considered completed studies, Chapter 4 and especially Chapter 5 contain much preliminary data that will be valuable for future studies.

80

Organism AraC-like Two- Alternative Histone- Other proteins component sigma like factors systems factors proteins Yersinia spp. VirF=LcrF YmoA S. typhimurium InvF HilA SPI-1 BarA/SirA PhoQ/PhoP S. typhimurium SpiR/SsrA, SPI-2 EnvZ/OmpR

E. amylovora HrpS HrpL HrpX/HrpY σ54? P. syringae pv. HrpR, HrpS HrpL syringae σ54 R. solanacearum HrpB HrpG PhrI PhrR PhrJ PhrA X. campestris HrpXv HrpG

E. carotovora HrpS HrpL RsmA/B, GacS/GacA σ54 IHF

Table 1.1 Regulators of type III secretion systems.

81 . y endemic; Black dots indicate interceptions and/or interceptions indicate dots Black endemic; ort of the disease in the countr the in disease of the ort p

oradic outbreaks. The dates refer to the first re p

. Figure 1.1. Worldwide distribution of Stewart’s wilt. is Red (grey) dots indicates areas were the disease s

82 6

4

5 7 9

3

1 8 2

2 mm

1

Figure 1.2. Stewart’s wilt disease cycle.

Bacteria overwinter in the flea’s beetles (2) alimentary tract or in corn seeds (1).

Survival in asymptomatic hosts may also occur. Corn seedlings can be infected via seed transmission, leading to early stunting, wilting and leaf streaking (5). Beetle infestations transmit the disease through feeding wounds on the leaves (4). The vascular bundles (6) and the intercellular spaces (9) are colonized, watersoaking occurs and dense population can be seen in xylem vessels (7). Bacteria can be seen streaming out of xylem vessels in transversal sections of infected leaves (8). Leaves show elongated, lens-shaped lesions that become necrotic. Secondary infections can be started by contaminated beetles.

83

Apo-EsaR OHHL EsaR active

inactive

Acyl-carrier protein SAM EsaI

EPS

esa I esaR

esaR-Box

Figure 1.3. The quorum sensing regulatory circuit controlling EPS synthesis in P.

stewartii.

The LuxR-homolog EsaR receptor-regulator binds target DNA elements as an

apoprotein. EsaI is a synthase catalyzing the formation of OHHL from SAM and acyl-

carrier protein. Increased OHHL concentration at high bacterial cell densities titrates

out EsaR. The OHHL-conjugated EsaR has much lower affinity for the target promoter

elements. The model is analogous to the lacI/lacZ operon model. OHHL: oxo- hexanoyl-homoserine lactone; SAM: S-adenosyl-methionine.

84 N C N C N C N C

OM PulD

N C Periplasm N C N

IM Sec

N C N C N C N C Chaperones Chaperones Type I II III IV V

Fig. 1.4. Bacterial secretion systems.

Five major classes of bacterial secretion systems are recognized. Type I systems are members of the ABC transporter superfamily. They recognize substrates via C- terminal signals ans translocated them in a single step. Type II secretion systems utilize the inner membrane Sec apparatus (general secretory pathway) to recognize N- terminal cleavable signal pepetides in their substrates. Secretion involve a periplasmic intermediate. Translocation through the outer membrane requires a multimeric secretin-like complex. Type III secretion systems secrete and translocate proteins recognized via various targeting mechanisms (chaperones, 5’-UTR-mRNA and non-cleavable signal peptides). Type III secretion systems are required for host- bacteria symbiotic/pathogenic interactions of many gram-negative species. Type IV systems vectorially secrete proteins or protein/DNA complexes. Type V systems are the so-called autotransporters. Substrates traveling this system cross the inner membrane via the sec pathway. IM inner membrane; OM, outer membrane (Adapted from Cossart et al., 1999),

.

85 YopB

YscD

20 Å 1200 Å 70 Å DISTAL ROD L RING

o.m. o.m.

YscC P. 50 Å PROXIMAL P RING 200 Å BASAL ROD BODY MS RING YscJ*

i.m. i.m.

C RING YscQ

YscN Flagellum YscU Specific secretion YscT YscR apparatus YscS LcvD

Hrp Pilus

o.m.

HrcC P.

HrcJ*

i.m.

HrcQ

HrcN o.m.= outer membrane i.m. = inner membrane p. = peptidoglycan layer HrcU HrcT HrcS HrcR HrcV

Figure 1.5. Schematic representation of animal and plant pathogen type III secretion apparatuses and of the flagellar basal body. Adapted from Cossart et al., (1999) and He, (1998).

86

s are regulatory genes, blue boxes are conserved genes, blue s are regulatory al and plant pathogen type III secretion genes. secretion genes, white boxes are other secretion genes secretion boxes are other genes, white secretion Figure 1.6. Transcriptional rganization of anim red boxe boxes Black proteins, are genes for secreted

87 Yersinia spp. 37ºC yop/ysc/ lcrGV/sycD

virF (escape macrophage endocytosis) LcrQ 26ºC

YmoA=H-NS

S. enterica Typhimurium (SPI-1) ? Supercoiling ~P +

barA sirA hilA invF sip/ssp inv-spa-prg (epithelium invasion)

Low pH High Fe3+ 2+ Low Mg RcsA/B/C

~P

phoQ phoP pmrD pmrB pmrA ugd (LPS modification)

Low pH ? Low osmolarity

~P (SPI-2) ~P

ompR envZ ssrBA spi- ssa- sse

(macrophage survival & proliferation)

Figure 1.7. Regulatory circuits controlling type III secretion system transcription in animal pathogens.

Red (dark grey): sensor kinases; blue (light grey): response regulators; white: AraC- like activators; black arrows: secretion genes.

88

Figure 1.8. Phytobacterial Hrp/type III secretion system regulatatory circuits.

Red (dark grey): sensor kinases; blue (light grey): response regulators; white: AraC- like activators; yellow (stripes), alternative sigma factors; black arrows: secretion genes; i.m. inner membrane; o.m., outer membrane.

89

P. syringae Lon RpoN IHF ENVIRONMENTAL SIGNALS Hrp Box +/ hrpR hrpS hrpL hrp/hrc ? ? HrpA HrpV

E. amylovora ?

Hrp Box

hrp/hrc hrpX hrpY hrpS hrpL

P. carotovorum Constitutive RpoN IHF ? Hrp Box

hrp/hrc hrpX hrpY hrpS hrpL GacA/S RsmA OHHL

X. campestris ?

(PIP) hrcC UNKNOWN hrpG hrpX hrp/hrc

(Figure 1.8 continued)

90 Figure 1.8 (continued)

Metabolic signals R. solanacearum PLANT CELL CONTACT PhrI/PhrR

PrhA (TonB-like) hrp/hrc i.m. phrJ hrpG hrpB on the o.m.

91

Figure 1.9. Modular structure of two-component systems.

Two-component sensor kinases and response regulators are formed by modular domains performing the discrete tasks of signal perception, transmission, reception and output response. Sensor kinases can be composed of more that two modules allowing for intramolecular phosphotransfer cascades. In class I kinases the H-box containing the phosphrylatable Histidine residue is located nearby the glycine-rich

ATP binding domain; in class two kinases the Hrp domain is physically separated from the transmitter domain. Pfam and COG databases are available at www.sanger.ac.uk/software/ Pfam and www.ncbi.nlm.nih.gov/cog. (Figure adapted from Swanson et al., 1994).

92

D CheY, Spo0F

OmpR, and FixJ classes D (e.g. HrpY, NarL, UhpA, SirA) Receiver D NtrC class D (enhancer-binding ) ~120 aa D Methyl-transferases; GGDEF, EAL, HD-GYP regulators Response regulators

~ H D H ArcB, BarA

Hybrid kinases

~ H D RcsC

Class I H NarX

H EnvZ H Kinase core H Rm FixL ~250-350 aa

NtrB, KinA, DegS, HrpX H Bj FixL

H CheA Class II

Orthodox kinases Figure 1.9 (continued)

93

Figure 1.9 (continued)

Legend:

D Receiver (Pfam00072)

σ54-dependent ATPase (Pfam00158)

GGBEF (Pfam01590), EAL (Pfam00990), HD-GYP (COG2206)

DNA binding : GerR-like HTH (Pfam00196), Fis-like HTH (Pfam02954.9); winged

helix (COG0745)

H HisKA or HATPase (Pfam00512 and Pfam02518) H Hpt (Pfam01627) Periplasmic domain (e.g. Pfam00497, Pfam02743, COG3300; etc.)

Cytoplasmic input domain (often with PAS/PAC domains; COG2202)

H-kinase dimerization domain (Pfam 2895.9)

HAMP linker (Pfam00672)

Rm, Rhizobium meliloti; Bj, Bradyrhizobium japonicum; H, Histidine; D, aspartate; Hpt, histidine phosphotransfer; HTH, helix-turn-helix

94

P P

D PhoP H NDFG ATP PhoQ ADP Pi H2O (Orthodox SK) P Class I D CheY P

H NGNDFGFG P 1 2 CheA D CheB (Orthodox SK) Class II

P P P P

H NDFG D H D BvgA NG1FG2 BvgS (hybrid SK) H PPP P

H NDFG D H D RcsB NG1FG2 RcsC YojN (hybrid SK) P

H NDFG P P P KinA D H D Spo0A P Spo0F Spo0B H NDFG (Phosphorelay) KinB

Figure 1.10. Phosphorylation pathways in two-component systems (Adapted from

Stock et al, 2001).

95 CHAPTER 2

THE HRPX/HRPY TWO-COMPONENT SYSTEM ACTIVATES HRPS

EXPRESSION, FIRST STEP IN THE REGULATORY CASCADE

CONTROLLING THE HRP REGULON IN PANTOEA STEWARTII

SUBSP. STEWARTII

2.1 INTRODUCTION

Pantoea stewartii subsp. stewartii (Basonym Erwinia stewartii) causes

Stewart’s bacterial wilt and leaf blight of sweet corn and maize (Zea mays

L.). Stewart’s wilt is the most serious bacterial disease of sweet corn in the north-central and eastern United States. The pathogen is transmitted by the corn beetle (Chaetocnema pulicaria) through feeding wounds

(Claflin, 1999). Bacteria grow in the xylem vessels of the host plant causing wilt and in the intercellular spaces of the leaves causing “water- soaked” lesions. The primary virulence factor responsible for wilting is the extracellular polysaccharide (EPS; Dolph et al., 1988) and the genes required for its production and regulation have been characterized

(Torres-Cabassa et al., 1987; Dolph et al., 1988; Coplin and Majerczak

96 1990; Poetter and Coplin, 1991). Water-soaking (Wts) is the result of cell

membrane damage and accumulation of fluids in the apoplastic spaces of

the leaf parenchyma.

A cluster of hrp genes (Alfano and Collmer, 1997) is required by many

gram-negative plant-pathogenic bacteria for elicitation of the

hypersensitive response (HR; Klement 1982) in nonhost plants and for

pathogenicity in a compatible host. hrp genes encode a type III secretion

system that is conserved among plant and animal bacterial pathogens

(Galan and Collmer 1999). They also encode several regulatory proteins

and the secreted effector proteins that determine interactions with

susceptible and resistant hosts. P. stewartii has a typical hrp cluster that

spans over 28-kb and is required for HR elicitation in tobacco and disease

in corn (Coplin et al., 1986; Coplin et al., 1992; Ahmad et al., 2001;

Frederick et al., 2001). In contrast to more necrogenic pathogens, such as

Pseudomonas syringae and Erwinia amylovora, P. stewartii is only able to elicit an HR in tobacco leaves when the bacteria are pre-grown in a Hrp- inducing, minimal medium or hrpS is constitutively expressed (Ahmad et al., 2001). Thus unlike sweet corn and maize, tobacco does not provide all of the signals necessary to induce hrp gene expression in this species.

In gram-negative plant pathogenic bacteria, there are two lineages of hrp genes that differ significantly in their mode of regulation (Alfano and

Collmer 1997). In general, expression of hrp genes is environmentally

regulated. Both plant factors (Schulte and Bonas 1992; Aldon et al., 2000)

97 and chemical/metabolic stimuli mimicking the plant apoplast (Rahme et

al., 1992; Wei et al., 1992; Xiao et al., 1992) have been shown to activate

their expression, but the mechanism by which these signals are perceived

is still unknown. Group I hrp genes in P. syringae, Erwinia spp.,

Pectobacterium spp. and Pantoea spp. are regulated by HrpS and the

alternative sigma factor HrpL (Frederick et al., 1993; Xiao et al., 1994;

Xiao and Hutcheson 1994; Wei and Beer 1995). Expression of hrpL

requires the alternate sigma factor σ 54 in addition to the NtrC-like

transcriptional activator HrpS (Frederick et al.,, 1993; Hendrickson et al.,

2000a). In P. syringae, HrpR, which is homologous to HrpS, is also

required for full activation of the hrpL promoter and may form

heteromeric complexes with HrpS (Hutcheson et al., 2001). In E.

amylovora, Wei et al., (2000b) described a two-component signal transduction system, HrpX/HrpY, which is orthologous to the one described in this work. They proposed that both HrpS and HrpY regulate

hrpL and their effects are additive. In contrast, Group II hrp genes in

Xanthomonas and Ralstonia spp. are regulated by AraC-like

transcriptional activators. In Ralstonia solanacearum, a complex

regulatory cascade transduces a signal arising from contact with the plant

cell wall and activates hrp gene expression (Aldon et al., 2000; Brito et

al., 2002).

In this report, we describe the sequences and function of genes included

in P. stewartii hrp complementation groups V-VII, which were previously

98 identified as regulatory loci (Frederick et al., 2001). hrpS (formerly designated wtsA) was characterized by Frederick et al., (1993) and its sequence is reanalyzed here in the context of new data. Using mutational and epistasis analysis, we determined the Hrp phenotype of individual regulatory genes and the order of the steps in the regulatory cascade. The results of this study support a model in which phosphorylated HrpY activates hrpS in response to environmental stimuli. This is different from the pathway reported for E. amylovora, but similar to that concurrently discovered in Pantoea agglomerans pv. gypsophilae (Nizan-Koren et al.,

2003).

2.2 MATERIALS AND METHODS

2.2.1 Bacterial strains, plasmids, media and growth conditions

E. coli and P. stewartii strains and plasmids used in this study are listed in Table 2.1. Luria-Bertani (LB) broth and agar (Ausubel et al., 1987) were used for strain maintenance and for growth of P. stewartii under

Hrp-repressing conditions. Fusaric acid-containing tetracycline sensitive selection (TSS) medium was prepared as described (Metcalf et al., 1996).

To induce hrp genes in P. stewartii strains, overnight LB broth cultures were washed twice in inducing medium (IM; Frederick et al., 2001), inoculated into 2 ml of the same medium at OD600nm = 0.05, grown for 12-

16 h and adjusted to final OD600nm ca. 0.5. In the pH experiment, HEPES and Tris-HCl buffers were used for the neutral and basic pH ranges,

99 respectively, in place of MES. Liquid cultures were grown in flasks or tubes shaken at 200 rpm at 37°C for E. coli or 29°C for P. stewartii.

When appropriate, antibiotics were supplied at the following

concentrations: ampicillin, 200 µg ml-1; kanamycin, 50 µg ml-1; and tetracycline, 20 µg ml-1 (or 15 µg ml-1 for selection of single copy Tetr cointegrates).

2.2.2 HR and pathogenicity tests

To test for HR elicitation on tobacco (Nicotiana tabacum L., cv. “White

Burley”) plants, P. stewartii strains were grown overnight in IM liquid

9 medium, washed and resuspended in water at OD600nm = 0.52 (ca. 1x10

CFU ml-1). Cells were infiltrated into tobacco leaves using four replicates

per strain as previously described (Frederick et al., 2001) and necrosis

was rated 24-48 h after infiltration. Pathogenicity tests were performed by

inoculating the whorls of 5-day-old sweet corn seedlings (Zea mays var.

saccharata, cv. “Seneca Horizon”) as previously described (Coplin et al.,

1986). The plants were held in growth chambers at 29°C (photoperiod 16

h, 15000 lux, relative humidity 99%). After 3 days, disease severity was

rated using a 0 to 3 scale (0= no symptoms, 1=scattered small lesions, 2 = numerous lesions, and 3 = extensive lesions that remained water-soaked with ooze forming on leaf surfaces).

100

2.2.3 General recombinant DNA, genetic and bioinformatic techniques

DNA isolation, agarose gel electrophoresis, restriction enzyme

digestion and ligation were done following standard procedures (Ausubel

et al., 1987). PCRs used either native Taq (for colony-PCR screens and

construct mapping) or cloned Pfu (for molecular cloning) DNA

polymerases (Invitrogen, Carlsbad, CA, U.S.A and Clontech, Palo Alto,

CA, U.S.A., respectively) and standard amplification protocols (Ausubel

et al., 1987). For colony-PCR, we used cells lysed in 50 mM NaOH for 10

min at 100°C as templates. For pBluescript and pRK415 vectors, the

reactions were performed using one primer anchored on either the lacI

vector gene (primer LACF) or the αlacZ gene fragment (primer LACR)

and a second primer conveniently located in the recombinant insert (Table

2.2). Standard deprotected oligonucleotide primers were ordered from

IDT-DNA. Probes for Southern blot hybridizations were labeled with 11-

Dig-dUTP and colorimetric immunodetection was done as recommended by the manufacturer (Roche Biochemicals). Plasmids were introduced into

E. coli and P. stewartii strains by electroporation, chemical transformation (Inoue et al., 1990) or mobilized by triparental mating using pRK2013::Tn7 as a helper plasmid (Ditta et al., 1980).

DNA sequencing was performed using cycle sequencing with fluorescent dye terminator chemistry and the ABI 3700 capillary electrophoresis sequencer (PE-Applied Biosystems) at The Ohio State

101 University Plant-Microbe Genomic Facility. Database analyses of

predicted proteins were performed using BLAST/PSI-BLAST (Altschul et al., 1997), whereas conserved domain analysis used the CD-search service

(ver. 1.54; Marchler-Bauer et al., 2002), both available through the

National Center for Biotechnology Information

(http://www.ncbi.nlm.nih.gov). Statistical significance of the matches are reported as E-values. Promoter searches were performed using a neural

network prediction algorithm (NNPP; Reese et al., 1996) and the

IHF/Sigma54 promoter-finder program (Goodrich et al., 1990). RNA

secondary structure predictions were performed at the

www.genebee.msu.su web server. Other sequence manipulation and

analyses were executed using software procedures in the GCG package

Ver. 9 (Oxford Molecular).

2.2.5 Construction of Plac-hrpY, Plac-hrpS and Plac-hrpL plasmids

For complementation and over-expression experiments, the hrpX, hrpY and hrpL ORFs, including their ribosomal binding sites, were amplified by PCR and cloned into high and low copy number cloning vectors such that they were expressed from Plac (Fig. 2.1). E. coli DH10B was used as

a cloning host. In all cases, cosmid pES1044 or plasmid pMM58 were

used as templates for the amplification reactions. PCR products were gel-

purified, digested with appropriate restriction enzymes, and then ligated

into pBluescript KS(+) or pBluescript SK(+) (Stratagene). Clones

102 containing recombinant plasmids were screened by colony-PCR using the

LacF primer and the reverse primer used for the PCR cloning. Individual

genes were subsequently subcloned into the low copy, mobilizable vector,

pRK415 (Keen et al., 1988), such that they were transcribed from Plac.

Vectors and recombinant plasmids are listed in Table 2.1 and primers in

Table 2.2.

hrpY was amplified using primers YFH and YRH. The 675-bp amplicon

was cloned into pBluescript KS(+) to produce pMM46, and then subcloned

into pRK415, to produce pMM52. hrpL, without its σ 54-dependent

promoter, was amplified using primers LFB and LRE. The 613 bp PCR

fragment was cloned into pBluescript SK(+) to produce pMM22 and then

subcloned into pRK415 to produce pMM6. The construction of Plac-hrpS

expression plasmids pRF205 (low copy) and pRF8 (high copy) and the

hrpXY plasmids pRF201 was previously described (Frederick et al., 1993).

Plasmid pMM58, containing the entire regulatory region, was constructed

by ligating the 1.7 kb HindIII insert of pRF8 into pRF201SK digested

with HindIII.

2.2.6 Construction of chromosomal P. stewartii hrp insertion mutants

Construction of Tn5 and Tn3HoHoI mutants with insertions in hrpX,

hrpY, hrpS and hrpL has been previously described (Frederick et al.,

2001). Double mutants containing selected null mutations and lacZ gene fusions were constructed by allele exchange of Tn5-mutagenized cosmids

103 as previously reported (Frederick et al., 2001) by using an incompatible

plasmid, pR751 (Ruvkun and and Ausubel 1981), to force the first cross-

over event. Nonpolar hrpL mutations were constructed by ligating the

aphA-3 cassette from pUC18K (Menard et al., 1993) into hrpL in pRF201.

The resulting plasmid, pDM2560, was used to exchange the hrpL::aphA-3

allele into the P. stewartii chromosome as described above. Several

strategies to construct strains with usable chromosomal hrpL-lacZ

transcriptional fusions were unsuccessful. Therefore, we created a strain

with a single-copy chromosomal hrpL-uidA transcriptional fusion by a

Campbell-type single crossover event. Plasmid pDM2785, which contains

hrpLp cloned into pPL6GUSC, was forced to recombine with the hrpL locus on the chromosome following introduction of an incompatible plasmid, pR751, to produce hrpL-uidA/hrpL+ Tetr Tpr strains. This

plasmid was recombined into wild-type (DC283), hrpY (DM064) and hrpS

(DM758) backgrounds to produce strains DM2831, DM2837 and DM2844,

respectively.

2.2.7 Site-directed mutagenesis of hrpY codon D57 by overlap

extension PCR and construction of missense mutant strains

To evaluate the role of the conserved aspartate residue at position 57 in

HrpY, we performed site-directed mutagenesis by a modification of the

megaprimer overlap extension PCR technique (Aiyar and Leis 1993). D57

was changed by nonconservative substitution to alanine [D57A] or by

104 conservative substitution to asparagine [D57N]. Five primers were

designed: YFH and YRH delimited the hrpY ORF; YR2 was the antisense

primer; and YD57N and YD57A were the sense-orientation mutagenic

primers for each substitution. The mutagenic primers contained missense

base changes at D57 (bold font in Table 2.2) and a near silent restriction

site (NheI and DraI for YD57A and YD57N, respectively). PCRs were

performed using pMM46 as template DNA. The first round of reactions

used primer pairs YFH+YR2 and YRH+YD57A (or YD57N) to produce

350 bp and 500 bp (or 508 bp) overlapping amplicons. Gel purified PCR

products from the two reactions were mixed, annealed and used as templates for a second round of PCR using primers YFH and YRH. PCR products were gel-purified and cloned into pBluescript KS(+).

Recombinant colonies were screened by PCR and by restriction enzyme analysis for the presence of the marker restriction site and the inserts were sequenced. Plasmids pMM57 and pMM92 contain the hrpY[D57A] and hrpY[D57N] alleles, respectively. These inserts were then subcloned into pRK415 to produce plasmids pMM74 and pMM118, respectively. To construct missense hrpY mutants, the hrpY[D57A] and hrpY[D57N] ORFs

were subcloned from pMM57 and pMM92 into suicide vector pLD55

(Metcalf et al., 1996) as XhoI/SstI inserts. The unmarked mutations were

introduced into the P. stewartii genome using the procedure described by

Metcalf et al., (1996) and confirmed by Southern blot hybridization.

105 2.2.9 Construction of P. stewartii strains with in-frame deletions in

hrpX

Two unmarked in-frame deletions in hrpX were constructed in vitro by

PCR. The deletions span amino acids 17 to 414 and 17 to 470 as indicated in the subscripts. To construct ∆hrpX(17-414), DNA fragments

corresponding to the ends of hrpX and its flanking sequences were

amplified by PCR from cosmid pES1044 using primer pairs EdX1+BdX1

and BdX4+XdX3 (Table 2.2). ∆hrpX(17-470) was similarly constructed

using the same primers, except BdX3 was substituted for BdX4. PCR

fragments were ligated into pBluescript KS(+) as EcoRI-BamHI and

BamHI-XbaI contiguous fragments. The deleted alleles were then ligated into pLD55 as XhoI/SstI inserts, transformed into E. coli BW20339

(Metcalf et al., 1996), mobilized to P. stewartii and introduced into the chromosome as described above. Cosmid pDM1296 hrpX+ hrpY::Tn5, which expresses hrpX from its native promoter, restored HR elicitation to these mutants, thereby genetically confirming that these strains are in fact bona fide nonpolar hrpX mutants.

2.2.9 Construction of hrp-promoter reporter plasmids

hrp promoter fragments containing the 5’ regulatory regions of hrpX, hrpS, hrpL, hrpJ and wtsE were either produced by PCR using the primers described in Table 2.1 or isolated as restriction fragments from cosmids pES411 or pES1044. To amplify the hrpL promoter region from -207 to

106 +182 (coordinates refer to the first base of the ORF), primers JFBG and

LRBG were used (Table 2.2). An hrpX promoter fragment containing from

–198 to + 46 bp was produced with primers XFHG and XRBG. The hrpS

promoter region from nucleotides -928 to +384 bp was amplified with primers SFG and SR3BG. To isolate two promoter regions containing hrp

boxes, a 0.9 kb BamHI insert containing PhrpJ and a 1.1 kb HindIII-BamHI

fragment containing PwtsE were purified from plasmid pRF201::Tn5seq118

and cosmid pES411, respectively. The former plasmid has a Tn5seq

insertion at +252 bp in hrpL (R. Frederick and D. Coplin, unpublished).

The DNA fragments, digested with BamHI and/or HindIII as required,

were ligated into the low-copy, promoterless, uidA fusion vector

pL6GUSC (Knoop et al., 1991). hrp-uidA reporter plasmids are listed in

Table 2.1. The same inserts used for hrp-uidA fusions were cloned into

the gfp reporter vectors pPROBE-KT or pPROBEgfp[AAV] (Miller et al.,

2000) to produce plasmids pMM165 hrpS-gfp[AAV], pDM2760 hrpJ-

gfp[AAV], pDM2756 hrpL-gfp[AAV] and pDM2724 hrpX-gfp.

2.2.10 Measurement of enzymatic reporter gene activity

β-glucuronidase (GUS) activity was assayed fluorometrically using 4-

methyl-umbelliferyl-β-D-glucuronide as described by Jefferson (1987),

but the assays were scaled down to fit microtiter plates and analyzed

using a Victor 1420-2 multilabel reader (PE-Applied Biosystems). Net

GUS activity of each strain was corrected for the basal fluorescence of P.

107 stewartii DC283 or E. coli SΦ200 carrying pL6GUSC without an insert.

-1 -1 -1 Specific activity was expressed in GUS units (pmol MU min OD600 ml of culture at 37°C). β-galactosidase activity was measured fluorometrically as described by Miller (1992) with the following modifications to adapt the protocol to the multiplate reader. Aliquots of cultures (100 µl) were mixed with 100 µl of a 2X β-galactosidase assay buffer (2.5 mM 4-methyl-umbelliferyl-β-D-galactopyranoside, 30%

DMSO, 200 mM NaCl in 200 mM potassium phosphate buffer, pH 7) and data collected as described above. β-galactosidase activity units were

-1 -1 -1 pmol MU min OD600nm ml of culture at 25°C.

2.2.11 Analysis of gfp reporter gene expression by flow cytometry

To quantify gfp expression, bacterial cells, collected by centrifugation

from growth media or from ooze extracted from plant tissues, were

resuspended and fixed in sterile filtered 0.5% paraformaldehyde at a final

OD600nm of 0.1-0.2. Suspensions were scanned in a FACS-Calibur

benchtop flow cytometer (Bencton-Dickinson) at the Ohio State

University Flow Core facility. The argon ion laser was tuned at 488 nm

and equipped with a 515/40 nm band-pass emission filter. GFP expression

was detected by forward and 90° side fluorescent light scattering. About

20,000 bacteria were counted per sample, Side scattering was

preferentially used to reduce the background noise. Data are reported as

distribution histograms of fluorescence vs. the number of cells (“events”). 108

2.3 RESULTS

2.3.1 Sequence analysis of the hrp regulatory region of P. stewartii

The hrp regulatory region was identified by transposon mutagenesis and

shown to contain three complementation groups (Frederick et al., 2001).

The region was subcloned from the cosmid pES1044 into two contiguous

plasmids, pRF8 and pRF201SK (Fig. 2.1). Prior sequencing of pRF8

(Frederick et al., 1993) revealed a single ORF, hrpS, which encodes an

NtrC-like enhancer protein homologous to HrpS from P. syringae pathovars (Rahme et al., 1991; Xiao et al., 1994), E. amylovora (Wei et al., 2000b) and P. agglomerans pv. gypsophilae (Mor et al., 2001). The main features of hrpS have been previously described (Frederick et al.,

1993). The sequence of the pRF201SK insert revealed three ORFs between hrpJ and hrpS, which have been designated hrpL, hrpX and hrpY. All of the ORFs in the regulatory region are transcribed from right to left as shown in Fig. 2.1. A detailed summary of the sequence analysis of the hrpL-hrpXY-hrpS region is shown in Table 2.3. Sequence analysis of the

54 region upstream of hrpL revealed a σ consensus box (CTGGCA-N6 -

TTGCT; probability score = 0.92) at -51 bp from the start codon of hrpL.

This was preceded by an A+T rich region containing putative integration

host factor (IHF) binding elements. A single dyad-symmetrical site

(TGCAA-N4 -TTGCA), weakly matching the NtrC-dependent upstream

activating site (TGCACC-N5 -GGTGCA; Porter et al., 1995), was located

109 at –163 bp. The predicted 21.3-kDa protein encoded by hrpL is

homologous to alternate sigma factors of the ECF σ 70 family (Lonetto et

al., 1994), including HrpL of E. amylovora (E value = 3e-74, 75%

identity; Wei and Beer 1995), P. agglomerans pv. gypsophilae (E = 2e-79,

80% identity; Mor et al., 2001) and P. syringae pathovars (E = 6e-30, 47

to 48% identity; Rahme et al., 1991; Xiao and Hutcheson 1994).

The hrpX ORF is located 144 bp downstream from the stop codon of

70 hrpL. A weak σ promoter consensus sequence (TGTGGA-N17-TAACAA,

probability score = 0.87) is located at -22 bp. Only 30 bp separates hrpX

from the downstream hrpY gene. The predicted proteins encoded by hrpX

(485 aa) and hrpY (213 aa) have homology to bacterial two-component

signal transduction systems, with HrpX similar to “orthodox” sensor

kinases and HrpY similar to bacterial response regulators (Hoch and

Silhavy, 1995). HrpX has two tandem PAS domains (Taylor and Zhulin,

1999) in the input region. HrpY belongs to the FixJ class of response

regulators with a conserved aspartate (D57) that is the putative site of

phosphorylation as inferred from multiple sequence alignments (not

shown). The closest PSI-BLAST matches in the GenBank database for these two genes were the ortholog HrpX/HrpY proteins of P. agglomerans pv. gypsophilae (E = 0, 79% identity for HrpX; E = e-103, 88% identity for HrpY; Mor et al.,, 2001) and E. amylovora (E = 0, 79% identity for

HrpX; E = 2e-96, 84% identity for HrpY; Wei et al., 2000b).

Interestingly, a putative HrpX and HrpY orthologous pair (E = 3e-40 for

110 HrpX; E = 9e-39 for HrpY) is also present in P. syringae DC3000

(Unfinished Microbial Genomes BLAST page at NCBI http://www.ncbi.nlm.nih.gov).

Our sequence analysis did not reveal any Rho-independent terminators

immediately following the end of hrpY, and multiple translational stops

are present in the 842-bp hrpY-hrpS intergenic region. BLAST searches of

the GenBank database indicated that a 483-bp IS-like element is present at

-23 bp from the start of hrpS (Fig. 2.2). This putative IS element is

flanked by two 8-bp direct repeats and contains two 75-bp imperfect

inverted repeats. The latter match sequences flanking unidentified

prophage genes and integrases present in the chromosomes and virulence

plasmids of Salmonella typhimurium and S. typhi (Genbank accessions

AE006471, AE008744, AL627268) and also flanking virulence ORFs 2

and 3 in P. syringae pv. phaseolicola plasmid pAV511 (accession

AF141883).Analysis of the secondary structure of the IS element using

the mFold algorithm revealed a strong hairpin with a free energy of

dissociation of –149.2 kcal mol-1 (Fig. 2.2). However, the IS-like element

in P. stewartii does not contain an ORF. Several putative σ 70 promoter

sequences are present upstream of and inside the IS-like element. PCR

analysis of 12 P. stewartii strains from our lab collection revealed that

this IS remnant was present in the same physical position in all strains

tested (Fig. 2.3) but not in a highly related strain of P. ananas.

111 2.3.2 Null mutant phenotypes and epistasis analysis

In a previous study, we isolated Tn5 and Tn3HoHoI insertion mutations in hrpL, hrpX, hrpY and hrpS and recombined them into the genome of P.

stewartii wild-type strain DC283, producing strains hrpL::Tn3HoHoI

(MEX105), hrpX::Tn3HoHoI (MEX116), hrpY::Tn5 (DM064) and

hrpS::Tn5 (DM758) (Fig. 2.1; Frederick et al., 2001). All of these mutants

have a Hrp- phenotype, cannot synthesize harpin and exhibit down-

regulation of several cosmid-borne hrp-lacZ fusions in secretion genes.

Since transposon mutations in hrpX are polar on hrpY, we constructed two

new nonpolar, in-frame, deletion mutants of hrpX, ∆hrpX(17-414) (DM2790)

and ∆hrpX(17-470) (DM2788) (Fig. 2.1). In tobacco, the newly constructed

∆hrpX mutants were unable to produce HR symptoms, even when

infiltrated at high cell densities (>109 CFU ml-1). On sweet corn

seedlings, mean disease severity ratings at 3 days after inoculation were

2.9 ± 0.2 and 1.9 ± 1.0 for DC283 and DM2790, respectively. Although

these means are not significantly different at P=0.05, we observed large

variations in both infection rates and symptom severity with the mutant

but not with the wild-type strain. In general, the ∆hrpX mutants were

capable of attaining wild-type levels of virulence, but with variable

response times and incidence. This may mean that the hrpX mutants are

much more sensitive to physiological differences between individual

seedlings.

112 Complementation analysis was done to determine the epistatic

interactions between different genes and rule out the possibility that some

phenotypes were due to second-site mutations. Low copy number plasmids

(pRK415 or pLAFR3) carrying hrpXY+, hrpY+, hrpS+ and hrpL+ (Fig. 2.1)

were constructed and introduced into different mutants. The ability of

these plasmids to restore HR elicitation in tobacco and virulence on sweet

corn to the mutants was tested (Table 2.4). In all cases where virulence

was restored, the complemented mutants were fully virulent (disease

severity rating ≥ 2.5 in at least two experiments) and the bacteria did not

need to be preinduced in IM to cause the HR. The Plac-hrpL plasmid,

pMM6, restored HR elicitation in tobacco and virulence on sweet corn to

all the insertion and deletion mutants in the regulatory region. Next, the

+ Plac-hrpS plasmid, pRF205, was able to restore the Hrp phenotype to

strains with mutations in hrpX and/or hrpY (strains MEX116, DM2788,

DM2790, DM064) and hrpS (DM758) and to a double hrpY hrpS mutant

(DM733), but not to an hrpL mutant (MEX105). Finally, plasmid pMM52, which contains Plac-hrpY, was able to restore HR elicitation and full virulence to ∆hrpX (DM2788, DM2790) and hrpY (DM064) mutants, but

not to hrpS (DM758) and hrpL (MEX105) mutants. Based on the

observations that i) mutations in hrpS were not suppressed by hrpY+, ii)

constitutive expression of hrpS+ or hrpL+ restored the Hrp+ phenotype to

hrpS and hrpY hrpS mutants, and iii) constitutive expression of hrpY+ and

113 hrpS+ alleles did not overcome the Hrp- phenotype of hrpL mutants, we

propose the following basic model for the Hrp regulatory cascade: hrpY Æ

hrpS Æ hrpL Æ hrp box promoters.

2.3.3 Expression of hrp regulatory genes in different genetic

backgrounds

To confirm the order of the regulatory cascade, hrpY::Tn5, hrpS::Tn5

and hrpL::aphA3 insertions were introduced into the chromosome of

selected hrp-lacZ or hrp-uidA strains to create double regulatory mutants.

In addition, we tested the effects of constitutively expressing various

plasmid-borne regulatory genes on chromosomally located lacZ and uidA

transcriptional fusions to hrpX, hrpS, hrpL and hrpJ promoters. Low copy number plasmids carrying Plac-hrpY, Plac-hrpS and Plac-hrpL were used to

over-express these genes and β-galactosidase or GUS activity of the

reporter fusions was measured following induction in IM liquid medium.

Results from these experiments are summarized in Fig. 2.4.

A second-site mutation in hrpY abolished expression of hrpS-lacZ and

hrpL-uidA (Fig. 2.4C) and a mutation hrpS decreased expression of hrpL- uidA (Fig. 2.4B), as expected from the epistatic analysis. Over-expressing hrpY turned up the hrpX-lacZ fusion ca. two-fold (Fig. 2.4A), showing

that hrpXY is a potentially autoregulated operon. Plac-hrpY stimulated the

hrpS-lacZ fusion ca. 8-fold in a hrpY+ background and ca. 20-fold in a

hrpY hrpS-lacZ double mutant (Fig. 2.4B). Plac-hrpY also up-regulated

114 hrpL ca. 7-fold in the hrpY+ chromosomal background and ca. 5-fold in

the hrpY hrpL-uidA double mutant, but it repressed hrpL expression in the

hrpS hrpL-uidA double mutant (Fig. 2.4C). The inability of HrpY to

increase hrpJ expression in strains containing a second-site hrpS or hrpL mutation (strains DM701 and DM786) further indicates that it does not directly activate hrpLp or hrp-box promoters, respectively (Fig. 2.4D). In

+ + a hrpS hrpXY background, the Plac-hrpS plasmid up-regulated both hrpL-

uidA and hrpJ-lacZ fusions 12- and three-fold, respectively (Figs. 4C and

4D). This effect was even more apparent in the hrpS hrpJ-lacZ and hrpY

hrpJ-lacZ double mutants, where expression increased 44- and 82-fold, respectively (Fig. 2.4D). However, Plac-hrpS did not affect hrpJ

expression in a hrpL hrpJ-lacZ double mutant (Fig. 2.4D), demonstrating

that it does not directly affect genes with hrp box promoters. Finally, the

Plac-hrpL plasmid up-regulated hrpJ-lacZ fusions in all genetic

backgrounds (hrpXY, hrpY, hrpS and hrpL; Fig. 2.4D), thereby confirming

that hrpL is the last gene in the hrp regulatory cascade. Significantly, the

expression of the hrpJ-lac fusions in this experiment mirrored the

virulence of the same strains in the above mutational and epistasis

analysis.

Since the lack of apparent terminators following the hrpL and hrpXY

operons means that transposon-induced mutations could have partial polar

effects on the level of expression of downstream genes, we repeated the

above experiments with plasmid-borne, reporter gene fusions in trans to

115 the tested regulatory mutations. We therefore constructed a series of

reporter fusions in pPL6GUS. These plasmids contained the regulatory

regions upstream of the hrpX, hrpS and hrpL ORFs fused to a uidA

reporter gene with its own ribosomal binding site (Fig. 2.1). Following induction in IM liquid medium, expression of each promoter was assayed in wild-type strain DC283 and in mutants lacking a hrp regulatory gene

(Table 2.5). Transcription from the hrpXp region, between –198 and +46

bp from the hrpX start, was high and constitutive in DC283. Expression of the hrpS-uidA fusion, containing the region between –928 and +384 bp,

was reduced up to 10-fold in hrpY and ∆hrpX mutants as compared to

DC283. The hrpL-uidA fusion was reduced ca. 250-fold in hrpS, ∆hrpX

and hrpY mutants.

Since Wei et al., (2000b) proposed that HrpS and HrpY both directly

activate hrpL in E. amylovora, we recreated each step of the P. stewartii

regulatory cascade in E. coli SΦ200, a genetic background that would lack other known hrp-specific regulatory molecules. We constructed E. coli

strains that harbored one of the above low copy hrp-uidA reporter

plasmids along with a high copy number plasmid that over-expressed a

regulatory gene (plasmids pMM46, pRF8, pMM22 carrying hrpY, hrpS and

hrpL, respectively). These strains were grown in LB broth, the regulatory

gene was induced with IPTG, and the cells were assayed for GUS activity.

(The following data are expressed as means ± S.D. They from

representative experiments with at least three replicates each and

116 statistical significance P was analyzed by ANOVA). The hrpXp-uidA

fusion in pMM25 was constitutively expressed in E. coli, but was further

stimulated as much as 6-fold by over-expression of hrpY (from 26 ± 6 to

127 ± 10 units; P <0.05). The hrpS-uidA fusion carried by pMM50 was up-

regulated 8-fold by HrpY (from 5 ± 2 to 42 ± 5 units; P<0.05) and the

hrpL-uidA fusion in pDM2785 was activated by HrpS (30-fold from 79 ±

20 to 2370 ± 817 units; P<0.001). In contrast, over-expression of hrpY did

not enhance hrpLp expression, but instead it consistently repressed its

basal level by ca. 6-fold (from 79 ± 20 to 12 ± 3 units; P<0.001). Finally,

the hrpJp-uidA fusion, which contains a promoter with an hrp-box, was

strongly activated by hrpL expressed from plasmid pMM22 (2000-fold

induction from 2 to 3970 units; P<0.001).

Overall, our data support a model where HrpY autoregulates hrpXY

expression, stimulates hrpS transcription and represses hrpL. These

results further confirm that HrpS activates hrpL and then HrpL alone

controls expression of hrp box promoters.

2.3.4 Role of the conserved aspartate in HrpY

Based on sequence homology and protein modeling (analysis not

shown), the conserved aspartate residue at position 57 is the most likely

phosphorylation site in HrpY. To test if D57 is required for activation of

HrpY, we mutated this codon to alanine and asparagine. Both

substitutions were expected to prevent phosphorylation, while minimizing

117 structural effects on the protein, as shown for similar response regulators

(Jin et al., 1990; Moore et al., 1993; Hoch and Silhavy 1995). The

missense mutant alleles were constructed in vitro by spliced overlap

extension PCR (SOE-PCR) and exchanged for the wild-type locus in the

chromosome of DC283 using the positive selection, suicide vector pLD55.

Both the hrpY[D57N] and the hrpY[D57A] chromosomal mutants were

completely unable to infect corn whorls or to elicit the HR when

infiltrated into tobacco leaves; they were complemented for both

phenotypes by the Plac-hrpY plasmid, pMM52. Unexpectedly, low copy

number plasmids carrying the same mutant alleles, pMM74 Plac-

hrpY[D57A] and pMM118 Plac-hrpY[D57N], were also able to complement

the null phenotype of the D57A, D57N and Tn5 hrpY chromosomal

mutants. As discussed below, this type of overexpression artifact is

common with two-component regulators and necessitates the use of single

copy mutant alleles for assessing phenotypes.

2.3.5 Effects of environmental conditions on the expression of hrp genes in vitro

The hrp/wts genes of P. stewartii are repressed in rich media (i.e. media with complex nitrogen sources) and induced in IM minimal inducing medium (Coplin et al., 1992, Frederick et al., 2001). To determine if the promoters of all the hrp regulatory genes would behave similarly to media repression, we compared the expression of the plasmid-borne hrpJp-uidA,

118 hrpLp-uidA, hrpSp-uidA and hrpXp-uidA fusions in wild-type strain

DC283 strain grown in LB broth and IM pH5.5. The results showed that all of the uidA gene fusions were dramatically repressed (up to ca. 3000- fold) in LB broth except for hrpXp-uidA, which was constitutively expressed (Fig. 2.5). The same promoter regions were cloned into the reporter plasmid pPROBEgfp[AAV] to form gfp fusions. This new set of

gene fusions allows the study of gene expression in vivo. Here we report

preliminary results that represent a proof of concept for the use of

unstable gfp fusions in analyzing virulence gene expression in plants. The

basal activity of the gfp fusions was measured in LB using flow-

cytometry. Gene expression data are presented as histograms showing the

distribution of the intensity of side (90º) light scattering (“90ºLS Log”)

upon excitation of the labels at 488 nm. Growth in corn increased the

proportion of cells fluorescing above the threshold from 17 to 36 fold

(Fig. 2.6). Upshifts of the median values was not as dramatic. It could be

that at +3 DPI only a small fraction of bacteria are still interacting with

intact cells, whereas the majority has turned down the expression of the

type III secretion system.

The regulatory effects of pH, phosphate, ammonia, and sodium chloride

were tested in vitro using chromosomal hrpX-lacZ, hrpS-lacZ and hrpJ-

lacZ fusions (Fig. 2.7), to avoid copy number artifacts while analyzing

the effect of individual environmental signals. IM was used as a basal

medium to test each variable. Expression of hrpS-lacZ and hrpJ-lacZ was

119 decreased by neutral to basic pH values (with maximal expression at pH

5.5) and by high potassium phosphate, ammonium chloride and sodium chloride levels. When we plotted the dose-response curves for each salt, expressed as its osmotic pressure (calculated by approximating activities to concentration values), the curves overlapped (data not shown), implying that part of the ability of IM to induce hrp genes could be due to its low osmolarity per se, rather than to nitrogen or phosphate starvation.

In contrast, expression of hrpX-lacZ was not affected by pH and was only slightly reduced by addition of the three salts (Fig. 2.7). In rich culture media, plasmid pRF205 Plac-hrpS relieved repression of promoters

downstream of hrpS in the cascade (data not shown). This was consistent

with the ability of pRF205 to enable P. stewartii to elicit an HR in

tobacco irrespective of the media used to grow the inoculum. We also

tested the effect of 2 mM nicotinic acid on wild-type DC283 cells

carrying pDM2531 hrpJ-uidA during growth in IM pH5.5. Expression of

this fusion in IM was reduced 5-fold by nicotinic acid (from 7357 ± 138

to 1439 ± 60 units ± S.E. in two experiments).

2.4 DISCUSSION

Bacterial virulence factors are usually tightly regulated and controlled

by environmental stimuli (Finlay and Falkow 1997). Mediating these

signals are complex, often hierarchical, regulatory cascades that integrate

multiple inputs and finely tune the infection process. The cascades

120 controlling type III and type IV secretion systems in several mammalian

and plant pathogens have been shown to include complete or partial two-

component signal transduction pairs. These constitute a common strategy

for linking the expression of energy-consuming secretion apparatuses to signals indicating the presence of a host. In this study, we present

evidence for a regulatory cascade that activates hrp/hrc type III

secretion/effector genes in P. stewartii in response to in vitro stimuli that

mimic the host environment. The cascade involves, from the top down

(Fig. 2.8), a two-component signal transduction pair (HrpX and HrpY), a

transcriptional activator (HrpS), and finally an alternative sigma factor

(HrpL). In this study, we first confirmed that HrpS and HrpL serve the

same functions in P. stewartii as they do in E. amylovora and P. syringae

(Xiao et al., 1992; Rahme et al., 1991; Wei et al., 1995, 2000;

Hendrickson et al., 2000ab; Hutcheson et al., 2001) and then we focused on the role of the HrpX/HrpY two-component system.

The occurrence of orthologs of HrpX/HrpY in E. amylovora (Wei et al.,

2000b), P. agglomerans pv. gypsophilae (Mor et al., 2001), P.

chrysanthemum (Ham, J.H., Rojas, C.M., Collmer, A., unpublished;

GenBank accession no. AF448202) and possibly in P. syringae pv. tomato

DC3000 (Unfinished Microbial Genomes database, available on line at the

NCBI or TIGR web sites) suggests that this two-component system may

represent a common mechanism for the environmental regulation of group

I hrp genes. However, only the functional and genetic analysis of E.

121 amylovora hrpXY has been published (Wei et al., 2000b) to date. The role

proposed by Wei et al., (2000b) for E. amylovora HrpY is different from

the one we present here for P. stewartii in that HrpY appears to serve as a modulator of HrpS activation of hrpL, instead of activating hrpS. This conclusion was based on their finding that a nonpolar hrpY::TnphoA mutation did not affect a chromosomal hrpS::Tn5-gusA1 reporter and on an earlier observation (Wei et al., 1995) that a hrpS mutation only partially regulates hrpL. A similar, independent study on hrp gene regulation in P. agglomerans pv. gypsophilae has been published

concurrently with ours (Nizan-Koren et. al. 2002). Their results with

HrpY are in complete agreement with our findings for P. stewartii. Other key data in Wei et al., (2000b), such as the inability of hrpY to complement a hrpS mutation, the downregulation of hrpL in a hrpY background and the autoregulation of the hrpXY operon by HrpY, are consistent with both our results and those in P. agglomerans pv.

gypsophilae. This suggests that common regulatory elements may interact

differently in the pantoeas and some other erwinias than they do in E.

amylovora. An extreme example of this is found in P. carotovorum,

which has both hrpS and hrpL. However, hrpL expression in this species

is constitutive and the hrp genes are post-transcriptionally regulated by

RsmA, which affects the stability of hrpL mRNA (Chatterjee et al., 2002).

In view of the apparent differences in the role of HrpY between P. stewartii and E. amylovora, we verified our hypothetical model using

122 multiple approaches. In particular, the regulatory effect of HrpY on hrpS

transcription was demonstrated repeatedly by measuring plasmid-borne

reporter gene fusions in hrp regulatory mutants (Table 2.5) by using

complementation tests with plasmid-borne hrpY+ in hrpS-lacZ or

hrpY::Tn5 hrpS-lacZ strains (Fig. 2.4B), and by reconstructing individual

steps of the cascade in E. coli. Moreover, over-expression of hrpY+

repressed the hrpL-uidA reporter in both a hrpS background (Fig. 2.4C)

and in E. coli genetic reconstruction experiments, further indicating that

HrpY cannot activate hrpL. It is possible that repression of PhrpL by unphosphorylated HrpY could provide a way of keeping the basal level of hrp gene expression low under noninducing conditions and/or provide a mechanism to turn it back down late in infection. In chapter 4 we present biochemical evidence for the function of HrpY as activator of hrpS and repressor of hrpL.

In the course of our complementation experiments, we found that a Plac-

hrpY plasmid activated hrpSp in the absence of hrpX and complemented

polar hrpXY mutants. These observations raised the question as to whether

or not HrpY always needs to be phosphorylated in order to activate hrp

genes. To address this question, we carried out site-directed mutagenesis

of the conserved aspartate (D57) in HrpY, which is the putative

phosphorylation site, and found that P. stewartii mutants carrying either

hrpY[D57N] or hrpY[D57A] chromosomal alleles in single copy were

completely Hrp-. This result indicated that D57 is required for HrpY to

123 function as a positive regulator and is genetic evidence that phosphorylation is required for HrpY activation. We believe that the avirulence of these mutants reflects the role of D57 in phosphorylation, rather than any nonspecific structural changes brought about by the amino acid substitutions. This role is supported by the solvent-exposed position of this residue (based on 3D protein modeling; not shown), its conserved position and the type of substitutions performed as shown for other well studied receiver domains (Jin et al., 1990; Moore et al., 1993).

Consequently, we attribute our finding that plasmid-borne wild-type HrpY and D57 mutant alleles were active in the absence of hrpX to an over- expression artefact that is commonly encountered in genetic studies of two-component systems (Hoch and Silhavy 1995). For example, in E. coli, overexpression of UhpA (a related FixJ-class response regulator) activated its target promoter uhpTp. This over-expression effect was observed with both wild-type UhpA and a D54A mutant protein (Shattuck-

Eidens and Kadner 1983; Kadner 1995). Two common explanations for this phenomenon are cross-talk with paralog kinases and the ability of the unphosphorylated response regulator to bind promoters at abnormally high concentrations (Hoch and Silhavy 1995).

In general, hrp genes are induced in minimal media that mimic conditions found in planta (Lindgren 1997), but these can vary considerably between species. For example, hrp genes in E. amylovora are stimulated by low pH and repressed by glucose, ammonium salts,

124 asparagine, histidine and nicotinic acid, but are unaffected by osmolarity

(Wei et al., 1992). P. syringae hrp genes are likewise induced at low pH in minimal media, but they are nonspecifically repressed by high salt concentrations (Rahme et al., 1992). The results of our study showed that expression of many hrp genes in P. stewartii was similarly repressed by rich media, activated in planta and regulated by pH and nitrogen sources, but, in contrast to E. amylovora, they were also affected by salt concentration (Fig. 2.7). hrpS-lacZ and hrpJ-lacZ chromosomal fusions were inhibited at neutral to basic pH and at high concentrations of potassium phosphate, sodium chloride and ammonium sulfate. In a previous study, we found that nitrates are also strongly repressing, but casamino acids increase expression rather than repress it (Coplin et al.,

1992).

For our studies on the role of the HrpX/HrpY two-component system in

sensing environmental signals, we constructed two P. stewartii mutants

with different in-frame deletions within the kinase domain of HrpX. In these strains, the expression of hrpS, hrpL and hrpJ was greatly reduced

(to levels comparable to a null hrpY mutant) when assayed in IM (Table

2.4) and the ∆hrpX mutants could not cause an HR on tobacco, which in

P. stewartii requires pre-induction in IM. This is the first clear indication

in any system that HrpX is needed to sense the chemical/physiological

signals provided in vitro by an apoplast-mimicking Hrp-inducing medium.

The hrpXY operon itself maintained a high basal level of expression in

125 both rich media (Fig. 2.5) and in IM medium at neutral pH with added

salts (Fig. 2.7), whereas the hrpS promoter was repressed under these

conditions. Thus, hrpS appears to be the first gene in the regulatory

cascade that is under strong environmental control and its transcriptional

regulation by phosphorylated HrpY and global regulators (M. Merighi and

D. Coplin, unpublished) may therefore be the key point in the cascade

where environmental signals are integrated.

The exact nature of the signals detected by HrpX is currently under

investigation in our laboratory. At this point, however, the sequence

analysis of HrpX offers a possible explanation as to how the hrp regulon

is able to respond to such a wide range of environmental conditions. The

absence of transmembrane regions indicates that HrpX is probably a

cytoplasmic protein and the presence of duplicated PAS domains in the N-

terminal portion suggests that its signal may be internal. PAS domains are

versatile cytosolic protein modules involved in monitoring light, redox potential, oxygen, small ligands and the energy level inside cells (Taylor and Zhulin 1999; Stephenson and Hoch 2001). Cluster analysis of a

diverse set of PAS domains from prokaryotic sensor proteins revealed that the PAS domains of HrpX are most closely related to those of Azotobacter vinelandii NifL (analysis not shown), which is a flavoprotein with FAD as its prosthetic group (reviewed in Taylor and Zhulin, 1999). Therefore by analogy to NifL, HrpX might sense redox potentials and energy changes within the cell. This hypothesis would explain how the Hrp regulon is

126 able to respond to most of the conditions known to induce hrp genes, i.e

acidity, phosphate and nitrogen starvation, all of which could affect

cellular redox potentials.

The importance of HrpX during infection appears similar in

Pantoea/Erwinia species in that it is not absolutely required for virulence.

In E. amylovora, a nonpolar hrpX::TnphoA mutant was greatly reduced in

virulence, response time and growth in planta and the expression of hrpL was reduced five-fold in inducing medium (Wei et al., 2000b). In contrast, a nonpolar hrpX::mini-Tn10-Km mutant of P. agglomerans pv. gypsophilae remained fully virulent and did not show any decrease in downstream hrp gene expression in planta, but the effect of this mutation was not determined in an inducing medium (Nizan-Koren et al., 2002).

Since the basis for the nonpolarity of TnphoA and mini-Tn10 transposons

in Erwinia spp. is not understood, it is possible that hrpY was not

expressed at normal levels in these experiments. The two P. stewartii

∆hrpX mutants used in this study were only slightly reduced in virulence,

infectivity and response time, but they exhibited much greater variability

in symptom severity and incidence on corn seedlings. This suggests that

the ability of the pathogen to sense the nutritional status of its host is

very important to insure consistent infection. On the other hand, given

that ∆hrpX mutants still retained significant virulence on sweet corn,

while nonphosphorylatable mutants of HrpY were totally nonpathogenic,

we hypothesize that there is an alternative HrpX-independent pathway

127 leading to HrpY activation in planta. This could reflect either the increased accumulation of small phosphodonor molecules (McCleary et al., 1993) under starvation conditions or the action of an alternative bacterial kinase specifically activated in corn (but not in tobacco or IM).

It will be interesting to learn if this alternate phosphorylation pathway responds to host cell contact or specific plant inducer molecules.

2.5 SUMMARY

A regulatory cascade activating hrp/hrc type III secretion and effector genes was delineated in Pantoea stewartii subsp. stewartii, a bacterial pathogen of corn. Four hrp regulatory genes were characterized: hrpX and hrpY encode the sensor-kinase and response regulator, respectively, of a two-component signal transduction system; hrpS encodes an NtrC-like transcriptional enhancer; and hrpL encodes an alternative sigma factor.

Epistasis analysis, expression studies using gene fusions, and genetic reconstruction of each step in Escherichia coli were used to delineate the following pathway: HrpY activates hrpS and also positively autoregulates the hrpXY operon. In turn, HrpS is required for full activation of the σ 54- dependent hrpL promoter. Finally, HrpL controls expression of all known hrp and wts genes. In vitro, hrpS and all downstream hrp genes were regulated by pH and salt concentration. Mutants with in-frame deletions in hrpX were still partially virulent on corn, but unable to sense the chemical and/or metabolic signals that induce hrp genes in vitro. Site-

128 directed mutagenesis of HrpY indicated that aspartate 57 is the probable phosphorylation site and that it is needed for activity. These findings suggest that both HrpX and an alternate mechanism are involved in the activation of HrpY in planta.

129 Strain, or plasmid Relevant phenotype(s) and/or genotype(s) Source or reference

Escherichia coli

r q + + BW20339 F' 128::Tn10-12(Kan ) lacI ∆lacZM15 pro(BA) /DE3(lac)X74 uidA(∆MluI)::pir recA1

∆phoA532 ∆(phnC∆DEFGHIJKLMNOP)33-30 Metcalf et al., ( 1996)

DH10B F- mcrA ∆(mrr-hsdRMS-mcrBC) φ80dlacZ∆M15 ∆lacX74 endA1 recA1 deoR

∆(ara,leu)7697 araD139 galU galK nupG rpsL λ- Gibco BRL

- - - r HB101 F thi-1 hsd20 (r Bm B) sup E44 recA13 ara-14 leuB6 proA2 lacY1 rpsL20 (Sm )

xyl-5 mtl-1 Boyer and

Roulland-Dussoix (1969)

HB101Rif spontaneous Rifr strain Lab collection

Sφ200Rif metB strA purB ∆(agg-uidA-man) Rifr Wei et al., (2000)b

Pantoea stewarti subsp. stewartiii

SS104 Wild type, Wts+, HR+ ICPPBa

DC283 SS104 Nalr Coplin et al., (1986)

DC102 P. stewartii corn isolate A. Vidaver Lab

DC110 P. stewartii SF8-2. isolated in Urbana, OH Hooker, 1967

DC119 P. stewartii corn isolate S. Walker, 1976

DC145 P. stewartii ATCC29227 Lab collection

DC146 P. stewartii ATCC 29228 Lab collection

DC147 P. stewartii ATCC 29229 Lab collection

DC162 P. stewartii corn isolate, Hudson Valley, N.Y. Lab collection, 1975

DM064 DC283 hrpY1296::Tn5 (Kanr) Frederick et al., (2001)

DM701 DC356 ∆hrpS hrpJ268::Tn3HoHoI (Ampr, Rifr) This study

DM711 DC283 hrpJ79::Tn3HoHoI (Ampr) Frederick et al., (2001)

DM729 DC283 hrpY1296::Tn5 hrpJ79::Tn3HoHo (Kanr, Ampr) This study

DM733 DC283 hrpY64::Tn5 hrpS1::Tn3HoHoI (Kanr, Ampr) This study

DM758 DC283 hrpS224::Tn5 (Kanr) Frederick et al., (2001)

DM786 DC283 hrpJ79::Tn3HoHoI hrpL2560::aphA-3 (Kanr, Ampr) This study

DM2788 DC283 ∆hrpX(17-470) This study

DM2790 DC283 ∆hrpX(17-414) This study

DM2831 DC283 hrpL-uidA (Tetr) This study

Table 2.1. (continued)

130 Table 2.1 (continued)

DM2837 DC283 hrpY hrpL-uidA (Kanr Tetr) This study

DM2844 DC283 hrpS hrpL-uidA (Kanr, Tetr) This study

MEX1 DC283 hrpS1::Tn3HoHoI (Ampr) Frederick et al., (2001)

MEX105 DC283 hrpL105::Tn3HoHo (Ampr) Frederick et al., (2001)

MEX116 DC283 hrpX116::Tn3HoHoI (Ampr) Frederick et al., (2001)

MM211 DC283 hrpY[D57A] This study

MM254 DC283 hrpY[D57N] This study

SW2 P. stewartii corn isolate, Wooster, OH Lab collection, 1974

SW20 P. stewartii corn isolate, Fayette Co., KY Lab collection, 1975

SW21 P. stewartii corn isolate, Grayson Co., KY Lab collection, 1975

SW4 P. ananas corn isolate, Portsmonth, OH Lab collection, 1975

Plasmids pBluescript

KS and SK (+) ColE1 αlacZ (Apr) Stratagene pLAFR3 IncP cos αlacZ (Tcr) Staskawicz et al., (1987)

pRK415 IncP αlacZ (Tcr) Keen et al., (1988)

pVK100 IncP cos (Tcr) Knauf and Nester (1982)

pLD55 R6K ori, αlacZ, tetAR (Apr) Metcalf et al., (1996) pUC18K promoter/terminator-less aphA-3 cassette Menard et al., (1993) pR751 IncP, (Tpr, Sp/Smr) Ruvkun and and Ausubel (1981)

pSU19 pACYC187 replicon (Cmr) Bartolome et al., 1991

pPL6GUSC pLAFR6 derivative carrying a promoterless uidA gene (Tcr) Knoop et al., (1991) pPROBE-KT pVS1 derivative (KmR) carrying a promoterless gfpmut3 gene Miller et al.,, 2000 pPROBEgfp[AAV] pBROBE derivative carrying a promoterless gfp[AAV] Miller et al.,, 2000

+ - r r r pRK2013::Tn7 ColE1 mob traRK2 ∆ repRK2 repE kan::Tn7 (Tp , Sm , Sp ) Ditta et al., (1980)

pDM1296 pES1044 hrpY::Tn5 Coplin et al., (1992)

pDM2560 pRF201 hrpL::aphA-3 This study

pDM2756 pDM2785 insert in pPROBEgfp[AAV] to create a hrpL-gfp[AAV] fusion This study

pDM2760 pDM2531 insert in pPROBEgfp[AAV] to create a hrpJ-gfp[AAV] fusion This study

pDM2724 pMM25 insert in pPROBEKT to create a hrpX-gfp fusion This study

pDM2842 pBKS (+) with PCR-made hrpX allele deleted from aa 17 to 414 This study

Table 2.1 (continued) 131 Table 2.1 (continued) pDM2846 pBKS (+) with PCR-made hrpX allele deleted from aa 17 to 470 This study pDM2848 pLD55 with the 0.98 kb SstI-XhoI fragment from pDM2842 This study pDM2851 pLD55 with the 1.1 kb SstI-XhoI fragment from pDM2846 This study pES1044 cosmid clone in pVK100 containing left half of hrp cluster Coplin et al., (1986) pES411 cosmid clone in pVK100 containing right half of hrp cluster Coplin et al., (1992)

+ pMM6 pRK415 with a 0.6 kb BamHI-EcoRI hrpL PCR fragment, transcribed from Plac This study

+ pMM22 pBSK(+) with a 0.6 kb BamHI-EcoRI hrpL PCR fragment, transcribed from Plac This study

pMM46 pBKS(+) with a 0.7 kb HindIII hrpY fragment made by PCR, transcribed from Plac This study pMM52 pRK415 with the 0.7 kb BamHI-EcoRI hrpY+ fragment from pMM46, transcribed

from Plac This study

pMM57 pBKS(+) with a 0.7 kb HindIII hrpY[D57A] fragment made by PCR-SOE This study

pMM58 pRF201SK with the 1.7 kb HindIII fragment from pRF8 and containing hrpL+,

hrpXY+, hrpS+ This study

pMM74 pRK415 with the 0.7 kb HindIII hrpY[D57A] fragment from pMM74,transcribed from PlacThis study

pMM92 pBKS(+) with a 0.7 kb HindIII hrpY[D57N] fragment made by PCR-SOE This study

pMM118 pRK415 with a 0.7 kb HindIII hrpY[D57N] fragment from pMM92, transcribed from Plac This study pMM165 pMM50 insert in pPROBEgfp[AAV] This study pMM245 pLD55 with the 0.7 kb SstI-XhoI fragment from pMM57 This study pMM246 pLD55 with the 0.7 kb SstI-XhoI fragment from pMM92 This study pMM330 pSU19 with the 0.7 kb SalI-SstI hrpY+ fragment from pMM46 transcribed

from Plac This study pRF201 pLAFR3 with the 3.6 kb BamHI-HindIII fragment containing hrpL+, hrpX+ and hrpY+

from pES1044 Coplin et al., (1992) pRF201SK pBSK(+) with the 3.6 kb BamHI-HindIII fragment from pRF201 This study pRF205 pVK100 with the 1.7 kb HindIII fragment containing hrpS from pES1044 Frederick et al., (1993) pRF8 pBSK(+) with the 1.7 kb HindIII fragment containing hrpS from pRF205 Frederick et al., (1993) pDM2531 pPL6GUSC with a 0.9 kb BamHI fragment to create a hrpJ-uidA fusion This study pDM2785 pPL6GUSC with a 389 bp BamHI PCR fragment to create a hrpL-uidA fusion This study pDM2791 pPL6GUSC with a 1.1 kb HindIII-BamHI fragment from pES411 to create

a wtsE-uidA fusion This study

Table 2.1 (continued)

132

Table 2.1 (continued) pMM25 pPL6GUSC with a 244 bp HindIII-BamHI PCR fragment to create a

hrpX-uidA fusion This study pMM50 pPL6GUSC with a 1.3 kb BamHI PCR fragment to create a hrpS-uidA fusion This study

Table 2.1. Bacterial strains and plasmids.

133

Primer Sewuence (5’Æ3’) Coordinates Source pBluescript and pPL6GUSC vectorsb

LacF ACCATGATTACGCCAAGCTC 813-794 This study

LacR GGCGATTAAGTTGGGTAACG 577-558 This study

5GUS CATTTCACGGGTTGGGGTTTCT 15-36 Jefferson (1987)

P. stewartii subsp. stewartiib

XFH ACGGAAGCTTACTTCAAGGAGGCGAGATGC 1364-1383 This study

4XRH ACGGAAGCTTGTCCTTCGCGCATGATGACG 2924-2905 This study

YFH CCGAAGCTTTTCGATAACAATATGGATAACAC 2855-2877 This study

YRH CCGAAGCTTATTAAGAGCAATCCTAAGAGA 3521-3501 This study

LFB CGCGGATCCACATTAAGCCAAACGGCAAAA 662-642 This study

LRE CCGGAATTCTGTTTTACCCCGTTCAGTT 1255-1237 This study

201.3 CCTGGCGAACCTCATACAGG 411-430 This study

201.1 CGCTATCATCTCGCCCTGTG 1691-1672 This study

YD57A TCTTTTACTGCTAGCTATGTGTATGCCTGG 3022-3051 This study

YD57N CAGGTCGATCTTTTACTTTTAAATATGTGTATGCCTGG 3014-3051 This study

YR2 GTTTCAGGATCGTTATCTTTCGTGATG 3204-3178 This study

EdX1 CGCGAATTCTTCGGTATTGCCCTGAACCT 919-938 This study

BdX1 CGCGGATCCTTTGATTGGCGTATCACTGCT 1427-1407 This study

BdX2 CGCGGATCCAACGAGGTTCTCATAGAGGT 2622-2641 This study

BdX3 CGCGGATCCACCAATGGAACCCGTGTGACT 2794-2813 This study

XdX3 CGCTCTAGATTTGCTCGGCGAGCATTTGG 3272-3253 This study

XFHG CCCAAGCTTACGTTCGCGTTTGTCACGGA 1182-1201 This study

XRBG CGCGGATCCTGATTGGCGTATCACTGCTA 1425-1406 This study

SFBG CGCGGATCCATAAAGCCCGCCTGATGGAA 3423-3442 This study

SR3BG CGCGGATCCTCTAGATGGCATGCTATCGATCTCGT 4735-4715 This study

JFBG CGCGGATCCGAGCTCTTTCCCAGGCAACAGAGTTAA 481-501 This study

LRBG CGCGGATCCTCTAGAATGACTTCCAGCCAGGTCAT 869-850 This study

ISR CGCGGATCCTCGGAATGAGTGCTCATTAT 4370-4389 This study

Table 2.2. (Continued)

134

Table 2.2 (continued)

a Positions of LacF/R and 5GUS primers are according to GenBank accession X52331

(pBluescriptKS+) and U12639 (uidA locus) respectively; positions for P. stewartii

primers correspond to GenBank accession AF282857 and do not include the 5’-end

restriction sites and anchoring tail. b Underlined bases represent engineered restriction sites; mutant codons are shown in

bold

Table 2.2. Primers

135

Regulatory elements/Protein features Protein homologies (E value; % similarity; % identity)b Accession # hrpL gene RBS (ACGGAG) at -6 54 σ promoter (CTGGCA-N6-TTGCT) at –51 IHF binding site (A+T rich region) at –55 UAS (TGCAA-N4-TTGCA) at –163 No terminators at 3’-end 21 kDa ECF sigma factor P.agglomerans pv. gypsophilae HrpL (2e-79; 80%; 89%) AAF76211 E. amylovora HrpL (1e-74; 75%; 86%) AAA91801 P. chrysanthemi HrpL (8e-49; 54%; 72%) AAM46690 P. syringae pv. tomato HrpL (6e-33; 48%; 66%) BAA82036 E. coli σE (3e-11; 28%; 50%) P34086 hrpX hrpY operon RBS (AGGAG) at -5 70 σ promoter (TGTGGA-N17-TAACAA) at –22 No terminators at 3’-end

HrpX, 56 kDa sensor kinase E. amylovora HrpX (0; 79%; 89%) AAD24682 No transmembrane domains P.agglomerans pv. gypsophilae HrpX (0; 72%; 85%) AAF76212 PAS-1 input domain (aa 16 to 65 P. chrysanthemi HrpX (e-146; 60%; 74%) AAM46691 PAS-2 input domain (aa 143 to 198) P. syringae pv. tomato DC3000 TCSc (3e-34; 31%; 50%) TIGR_317_5668 His-kinase domain (aa 296 to 485) X. campestris pv. campestris (1e-30; 30%; 49%) AAM41248 H296 putative phosphorylation site

HrpY: 24 kDa response regulator P.agglomerans subsp. gypsophilae HrpY (e-103; 88%; 93%) AAF76213 Receiver domain (aa 1 to 118) E. amylovora HrpY (6e-96; 83%; 89%) AAD24683 D57 putative phosphorylation site Pectobacterium chrysanthemi HrpY (8e-77; 70%; 80%) AAM46692 Output domain (aa 149 to 213) P. syringae pv. tomato DC3000 TCRc (3e-51; 49% ; 69%) TIGR_317_5668 HTH domain (LuxR homolog; aa Pectobacterium carotovorum ExpA (3e-39; 32%; 56%) CAA64809 69 to 194) Salmonella typhimurium SirA (2e-98; 33%; 58%) AAC44801 E. coli UhpA (2e-26; 33%; 55%) BVECAU

hrpS gene RBD (AGGGGC) 483-bp IS like at –23 Rho-independent terminator at 3’ end

36 kDa NtrC-like response regulator No N-terminal receiver domain E. amylovora HrpS (e-146; 79%; 90%) AAD24684 σ54-dependent ATPase domain P.agglomerans pv. gypsophilae HrpS (e-145; 79%; 91% AAF76214 C-terminal HTH domain Pectobacterium chrysanthemi HrpS (e-101; 59%; 73%) AAM46693 P. syringae pv. phaseolicola HrpS (3e-83; 57%; 70%) CAA54728 P. syringae pv. phaseolicola HrpR (2e-81; 56%; 70%) CAA54727

Table 2.3. (continued)

136 Table 2.3. (continued)

a Putative DNA regulatory elements, homologous proteins and conserved protein

domains were identified by computerized sequence analysis as described in Material

and Methods. RBS = ribosomal binding site; IHF = integration host factor; UAS=

upstream activating sequence (weakly matches NtrC UAS described by Porter et al.,

1995); HTH= helix-turn-helix motif; IS= insertion sequence; TCS = two-component

sensor; TCR = two-component regulator b Within parenthesis are the following Genbank statistics (in order): E value based on the July 8 2002 version of the Genbank nr database; % identity and % similarity (= positives) for BLASTP local alignments. c P. syringae pv. tomato. The hrpX hrpY homologs found by TBLASTP search form a

two-gene operon in the non-annotated contig TIGR_317_5668, recently renamed

PSPTO0559 and PSPTO0558 (Buell et al., 2003). The HrpX homolog in P. syringae

has the same domain organization of P. stewartii HrpX except for a periplasmic

binding protein domain in the N-terminal region (sequences courtesy of The Institute

for Genomic Research).

Table 2.3. Analysis of DNA and protein sequences

137

,

-hrpY lac -hrpL lac , pMM52 P - - - - hrpY + Path HR Path P hrpX hrpS

lac- HR n in IM pH 5.5 overnight.

Path P

hrpY no symptoms; +++ = many lesions; V variable a

lac- HR regulatory gene regulatory , - = no HR. Cells were grow

trans P + Path

HR hrpX uppression tests were low-copy number replicons: pDM1296 number replicons: were low-copy tests uppression

regulatory mutants and complementation analyse ys after inoculation of the whorls: - = hrp -hrpL + +++ + +++ + +++ + +++ + +++ + +++ + +++ + lac c V Path b ------+ +++ + +++ + +++ - + ------+ +++ + +++ + +++ + +++ + +++ + - - - -

------, and pMM6 P -hrpS lac HR + +++ + +++ + +++ + +++ + +++ + +++ + +++ + +++ + +++ + None 5 ------+ +++ - + - - - - - 5 3 - +++ 3 Tn ::Tn ::Tn ::Tn hrpX ∆ wild-type hrpS hrpY:: hrpJ hrpL c. Disease ratings on corn leaves at 3 da pRF205 P pRF205 a. complementation/s used for the plasmids All b. Hypersensitive response in tobacco. HR ratings: + = full ______DM758 type and of wild 2.4. Pathogenicity Table DM064 DC283 DC283 DM2790 _ MEX1 DM711

Strain Genotype

138

GUS activity (Units ± S.E.)c in different genetic backgroundsd Plasmid / Promoter fusionb Wild type ∆hrpX hrpY hrpS hrpL

pMM25 -198/+46 hrpXp 307±53 370±10 206±2 ND ND pMM50 -928/+384 hrpSp 135±9 20±9 21±4 119±12 ND pDM2785 -207/+182 hrpLp 6053±959 164±5 67±12 36±1 ND

pDM2531 -516/+425 hrpJp 4851±704 94±13 17±5 3±1 0±0 pDM2791 –914/+22 wtsEp 399±125 7±5 1±1 0±0 4±2

a Strains were grown in IM pH 5.5 broth to late exponential-early stationary phase and assayed for GUS activity as described in Material and Methods. b Transcriptional fusion names are preceded by the promoter fragment coordinates in bp (+1 is first base of ORF). c -1 -1 -1 One unit = 1 pmol MU min OD600 ml of culture at 37°C. S.E.= standard error of

two to six independent experiments, each with at least two replicates. d Strains: wild type = DC283; ∆hrpX = DM2790 (mutant DM2788 gave similar

results); hrpY = DM064; hrpS = DM758 or MEX1; hrpL = MEX105.

Table 2.5. Effects of hrp regulatory genes on plasmid-borne hrp::uidA

transcriptional fusions expressed in different genetic backgrounds.

139

Figure 2.1. Physical map of the regulatory region of the P. stewartii DC283 hrp gene cluster showing the hrpL, hrpX, hrpY and hrpS subclones used in this study.

Open boxes represent regulatory ORFs, the cross-hatched box is hrpJ, a secretion

operon, and the solid box represents an IS remnant in the hrpS 5’ regulatory region.

The direction of transcription is shown by arrows. Solid circles indicate Tn5

insertions, the open circles are point mutations, flags indicate Tn3HoHoI insertions

and their orientation, squares denote aphA-3 insertions, and ∆ indicates in-frame

deletions. Open triangles in front of genes represent putative σ 70 promoters, the

closed semicircle is a σ 54 promoter, and the open semicircle is a hrp box promoter.

P lac indicates the position of the constitutive lac promoter in the plasmid vector. Solid boxes labeled “uidA” indicate transcriptional fusions to the GUS gene in pL6GUSC.

Restriction enzyme abbreviations: H, HindIII, B, BamHI.

140

141

Figure 2.2. Analysis of the secondary structure of the IS element remnant located in front of hrpS.

The secondary structure corresponding to the transcript spanning the 483-bp IS-

element located in front of hrpS was performed using the Genebee server

(www.genebee.msu.su). An energy minimize 2D-folding model is shown. The

calculated free energy of folding was –149.2 Kcal mol-1.

142

1 2 3 4 5 6 7 8 9 10 11 12 13 14

1 kb

Figure 2.3. PCR detection in a collection of P. stewartii strains of the abortive IS element preceding hrpS.

Primers ISR and SFBG flanking the 483-bp IS-like element in front of hrpS were used to amplify by “colony PCR” a ca. 1-kb product from a collection of Pantoea stewartii strains. Only strain SW4 tested negative and was later identified as Pantoea ananas by biochemical and nutritional assays. Lane 1, 1-kb (+) molecular weight ladder; lane 2, P. stewartii strain SW2; lane 3, P. stewartii strain SW20; lane 4, P. stewartii strain DC102; lane 5, P. stewartii strain DC110; lane 6, P. stewartii strain

DC119; lane 7, P. stewartii strain DC145; lane 8, P. stewartii strain DC146; lane 9,

P. stewartii strain DC147; lane 10, P. stewartii strain DC162; lane 11, P. ananas strain SW4; lane 12, P. stewartii strain SW21; and lane 13, P. stewartii strain DC283.

143

Figure 2.4. β-galactosidase (β-Gal) and β-glucuronidase (GUS) activity of chromosomal hrp-lacZ or hrpL-uidA fusions in different P. stewartii genetic backgrounds, with and without various plasmid-borne Plac-hrp regulatory genes.

Cells were grown in IM pH 5.5 broth to late log phase and β-Gal specific activity

-1 -1 -1 was measured (1 unit = 1 pmol MU min OD600 ml of culture at 25°C). The first

column indicates which regulatory gene was constitutively expressed in trans (if any);

the second column indicates the second hrp mutation (if any) in the strain. Regulatory

gene clones: pMM52 Plac-hrpY and pMM46 Plac-hrpY (panel C; low copy pMM330 was

used as well with similar results); pRF205 Plac-hrpS (panel D) and pRF8 Plac-hrpS

(panel C); and pMM6 Plac-hrpL (panel D). Controls consisting of the vector plasmids

without inserts had no effect on expression of any of the gene fusions.

Panel A: β-Gal activity in hrpX-lacZ strain MEX116; Panel B: β-Gal activity in the

hrpS-lacZ strain MEX1 and hrpY hrpS-lacZ strain DM733. Panel C: GUS activity in

hrpL-uidA strain DM2831, hrpY hrpL-uidA strain DM2837 and hrpS hrpL-uidA strain

DM2844. Panel D: β-Gal activity in the hrpJ-lacZ strain DM711, hrpY hrpJ-lacZ

strain DM729, hrpL hrpJ-lacZ strain DM786, and ∆hrpS hrpJ-lacZ strain DM701.

144

145

10000

1000 rich medium 100 IMpH5.5 10 specific GUS activity (pmol MU min-1 OD-1) 1 DC283/pMM25 DC283/pMM50

DC283/pDM2531 DC283/pDM2785

Figure 2.5. Expression of plasmid-borne hrp-uidA fusions in P. stewartii DC283.

P. stewartii strain DC283 carrying hrp-uidA plasmid-borne fusions was grown in LB broth and IM pH5.5 liquid medium and GUS activity was measured using a fluorometric protocol as described in Materials and Methods. One unit = 1 pmol MU

-1 -1 -1 min OD600 ml of culture at 37°C. Error bars represent standard deviations from single experiments with at least two replicates. Strains: wild type = DC283; plasmid pMM25 contains -198/+46 hrpXp, pMM50 contains -928/+384 hrpSp, pDM2785

contains 207/+182 hrpLp, and pDM2531 contains -516/+425 hrpJp.

146

Figure 2.6. Flow-cytometry analysis of hrp-gfp fusion expression in P. stewartii

DC283 cells from infected corn plants compared to their expression in LB.

P. stewartii DC283 cells carrying hrp-gfp plasmid-borne fusions were either grown in

LB broth or recovered as ooze from the surface of inoculated in corn seedlings. The whorls of 8-day-old potted sweet corn seedlings were inoculated with bacteria supsended in 0.2% Tween-40. After two days the tops of the plants were cut off, placed in moist chambers and incubated for an additional 24 h; bacterial ooze was then collected from the leaves. Bacteria were washed in sterile saline solution and resuspended in 0.5% paraformaldehyde to fix the cells. Samples were analyzed in a benchtop flow cytometer fitted with an argon laser tuned at 488 nm. Side Light scattering fluorescence (“90ºLS Log”) of ca. 20,000 cells per sample was read at

515/40 nm band-pass filter. Gene expression values are reported as % of gated cells above the fluorescence threshold (M) in LB vs. corn. Red solid histogram = cells grown in LB. Black histogram: cells collected corn leaves at 3 DPI; pDM2724 is hrpX-gfp; pMM165 is hrpS-gfp[AAV]; pDM2756 is hrpS-gfp[AAV]; pDM2760 is hrpJ-gfp[AAV]

147

M=0.73% vs. 20% M=0.7% vs. 17.9%

M M

pDM2724 pMM165

M=0.7% vs. 12% M=0.7% vs. 24.9%

MM

pDM2756 pDM2760

M=% of cells above the fluorescence threshold upon growth in LB vs. in planta

148

Figure 2.7. Effect of environmental conditions on the expression of various hrp-

lacZ fusions.

The effects of neutral pH, 50 mM sodium chloride, 25 mM ammonium sulfate and

25 mM potassium phosphate are shown as a percentage of the activity in standard IM

pH 5.5 liquid medium (control). Cells were grown in modified IM, with one

component changed as indicated in the Z-axis of the graph, for 16 h and β-Gal activity

was measured. hrpX-lacZ = strain MEX116, hrpS-lacZ = strain MEX1 and hrpJ-lacZ =

DM711.

149

Figure 2.8. Working model for the hrp regulatory cascade in P. stewartii.

The HrpX sensor perceives undefined metabolic and/or stress signals and activates

HrpY via phosphorylation (indicated by ~P). A parallel phosphorylation mechanism

may also activate HrpY in planta. HrpY then activates hrpS and its own operon; HrpS

and σ 54 activate hrpL; and finally HrpL enables RNA polymerase core enzyme (E) to activate promoters containing hrp boxes (e.g. the harpin, wts and hrp secretion genes).

150

CHAPTER 3

A NOVEL AUTOREGULATORY LOOP ACTIVATED BY HRPS

FROM THE HRPL PROMOTER MODULATES EXPRESSION OF THE

P. STEWARTII TYPE III SECRETION SYSTEM REGULATORS

HRPXY, HRPS AND HRPL

3.1 INTRODUCTION

Pantoea stewartii subsp. stewartii is a bacterial pathogen of sweet corn and maize that causes wilting and leaf necrosis, preceded by a water- soaking phase (Pepper, 1967). Although extracellular polysaccharide

(EPS) is a major virulence factor contributing to wilting symptoms (Dolph et al., 1988; Coplin and Majerczak, 1990), the Hrp type III secretion system is absolutely required for growth in the host to high cell density

(>108 CFU/g tissue) and for development of both lesions and wilt symptoms (Coplin et al., 1986; Coplin et al., 1987). hrp genes are organized in horizontally mobile, pathogenicity islands (PAI) (Hutcheson,

1999; Kaper and Hacker, 1999) located, in different pathogens, either on the chromosome (Erwinia amylovora, Pectobacterium carotovorum,

Pectobacterium chrysanthemi, Pseudomonas syringae, Xanthomonas 151 campestris, Xanthomonas vesicatoria) or plasmids (Pantoea agglomerans

pv. gypsophiliea, Ralstonia solanacearum) (Rosenberg et al., 1982;

Boucher et al., 1987; Valinsky et al., 1998; Ezra et al., 2000). Hrp PAIs usually carry their specific regulatory genes linked to the secretion/effector loci, but additional unlinked genes are often required for hrp/hrc expression. All pathogens carrying group I hrp clusters (P.

syringae and erwinias) require the chromosomally encoded, nitrogen

starvation sigma factor σ 54 (RpoN) for expression of hrp secretion and

effector genes (Frederick et al., 1993; Hendrickson et al., 2000).

Furthermore, expression of P. carotovorum hrpL is reduced in a himA

(IHF) E. coli mutant (Chatterjee et al., 2002), implying a positive role for

IHF in hrp expression. On the other hand, in group II pathogens, X. vesicatoria hrpX, which encodes an AraC-like activator, is unlinked to the main PAI (Wengelnik and Bonas, 1996) and the OmpR-class response regulator HrpG lacks a physically linked cognate sensor kinase

(Wengelnik et al., 1996).

During the homing process for horizontally acquired DNA (Kaper and

Hacker, 1999), it is indeed reasonable to envisage the fortuitous interaction between regulatory elements in the PAI and various housekeeping regulatory genes already present in the recipient strain.

Some of these new interactions may mimic those present in the donor

152 strain, whereas others may be novel and selected for or against. This may adapt the expression pattern of the virulence genes to fit the new host environment.

Regulation of group I hrp clusters always involves three common regulatory proteins, HrpS, HrpL and RpoN, and probably IHF as well. As mentioned above, IHF regulates hrp genes in P. carotovorum, but its role is only assumed for the other group I pathogens, based on the presence of similar A+T rich regions 60-80 bp upstream of hrpL (Hutcheson, 1999;

Merighi et al., 2003). The Lon protease may act as an additional negative regulator of hrp gene expression as shown by a recent study on

Pseudomonas syringae (Bretz et al., 2002) and work from our lab (Ham and Coplin, unpublished) on P. stewartii. Phytobacterial pathogens of the erwinia group also use the HrpX/HrpY two-component system, which is not found in P. syringae.

In spite of the high similarity (around 80% identity) of the regulatory proteins, the 5’- and 3’-end region of their genes are quite different. This divergence is reflected in different basal levels of expression and different epistatic interactions of the regulatory genes in the various species. For instance, in E. amylovora, Wei and colleagues concluded that

HrpS is environmentally regulated in a HrpX/HrpY-independent manner and that both HrpY and HrpS act additively to control hrpL (Wei et al.,

2000b). On the other hand, the role of HrpX/HrpY in Pectobacterium spp.

153 is unclear, because hrpS expression appears to be constitutive (Chatterjee et al., 2002) and the phenotype and regulatory effects of polar hrpXY

mutations are not known.

In P. stewartii and P. agglomerans pv. gypsophiliae, HrpY has been

shown in genetic reconstruction experiments to activate hrpS, but not

hrpL. Moreover, in a E. coli rpoN+ himA+ background, HrpS alone appears

to be sufficient for full activation of hrpL (Chapter 2; Merighi et al.,

2003; Nizan-Koren et al., 2003). A significant difference between P.

stewartii and the other erwinias is the abortive IS element located just

upstream of the RBS of hrpS in P. stewartii (Merighi et al., 2003). This element might form a strong hairpin-like secondary structure that could

potentially alter expression of hrpS itself. In addition, it could also carry

additional regulatory elements that may explain some of the unique features of hrpS regulation in P. stewartii, e.g. its regulation by quorum

sensing as discussed in Chapter 6.

In all erwinias, the hrpL, hrpXY and hrpS regulatory operons are

colinear and transcribed in the same direction, from hrpL to hrpS (Fig. 1),

whereas in P. syringae, hrpL and hrpRS are located on opposite ends of

the hrp/hrc cluster, probably due to a translocation event during homing of the PAI (Hutcheson, 1999). In P. stewartii, differently from other erwinias, preliminary results indicated that hrpS was autoregulated in some manner. Ectopic expression of hrpS upregulated hrpS::lacZ and

hrpX::lacZ fusions, but the effect on hrpS was abolished by a polar 154 insertion in hrpY (Coplin, unpublished observation). The particular

transcriptional organization of the three regulatory operons and the

apparent absence of potential terminators downstream of hrpL and hrpY

(Merighi et al., 2003) suggested that transcription from either or both

PhrpL and PhrpX could continue into hrpS. Alternatively, increased

expression of hrpXY and trans-activation of hrpS by HrpY could also

cause this autoregulatory effect. Both scenarios would create a novel

autoregulatory loop in the regulatory pathway.

The purpose of this study was to analyze the molecular basis of the

positive autoregulatory loop triggered by HrpS and to identify the

regulatory elements, promoters and transcripts involved. In particular we

wanted to determine: (i) if the HrpS-dependent autoregulatory effect is

triggered by activation of the hrpL and/or hrpXY promoters; (ii) whether

autoregulation acts by the effect in trans of HrpY on hrpS or it is instead

due to the cis-effect of transcription continuing from hrpXY and/or hrpL

into hrpS.

We initially studied the effect of ectopic expression of hrpS in double

mutants with a chromosomal polar insertion in one regulatory gene and a

reporter gene fusion in another. After we defined the candidate promoters

involved in hrpS autoregulation, we characterized them by analyzing the

5’-ends of the various mRNAs and then we utilized this information to

demonstrate inter-operon transcription using RT-PCR. The existence of

155 long transcripts starting from hrpL and spanning hrpL and hrpXY and part

of 5’ hrpS was further shown by Northern blot analysis. The possible

roles of hrpS autoregulation in P. stewartii biology are discussed.

3.2 MATERIALS AND METHODS

3.2.1. Bacterial growth and media

The bacterial strains and plasmids used in this study are given in Table

3.1. Strains were grown in Luria-Bertani (LB) broth and agar (Ausubel et

al., 1987) for inoculum preparation and cloning procedures or in IM

liquid medium at pH5.5 (IM5.5) (Frederick et al., 2001) for induction of

hrp gene expression as described (Merighi et al., 2003). Selection for tetracycline sensitive strains was done in the fusaric acid medium

(Buchner selection) described by Metcalf et al. (1996). When required, antibiotics were added at the following concentrations: tetracycline at 20

µg/ml, nalidixic acid at 10 µg/ml, kanamycin at 50 µg/ml,

chloramphenicol at 50 µg/ml, and ampicillin at 200 µg/ml.

3.2.2 Molecular genetics and cloning techniques

Plasmids were transferred among strains by biparental mating or

triparental mating using pRK2013::Tn7 as helper plasmid (Ditta et al.,

1980) or introduced by electroporation. Double mutants were constructed

by allele exchange of mutations cloned in the suicide vector pLD55 with

positive selection of double recombinants in fusaric acid media as 156 described previously (Metcalf et al., 1996; Merighi et al., 2003).

Preparation of cell lysates, DNA purification, molecular cloning

experiments, Southern blot analysis and PCR reactions were performed

following standard procedures (Ausubel et al., 1987).

3.2.3 Construction of polar hrpL interposon mutants and of plasmid

pMM191

To construct a polar mutation in hrpL to use for allele exchange in P.

stewartii, the interposon Ω/Spr was purified as a 2 kb SmaI fragment from

plasmid pHP45::Ω and ligated in the NvuI site of plasmid pDM2533,

which contains the hrpL gene in a 1.3 kb PCR fragment. The resulting

plasmid, pDM2862, was digested with KpnI/SpeI and the 3.3 kb hrpL::Ω

fragment was subcloned into the suicide plasmid pLD55 to create

pDM2868. E. coli strain BW20339 (pDM2868) was crossed with strains

MEX1 hrpS-lacZ and MEX116 hrpX-lacZ by triparental mating (Frederick et al., 2001) and transconjugants were selected for Tetr. Single cross-over

mutants were streaked on fusaric acid medium to select for the second

recombination step as previously described (Metcalf et al., 1996).

Putative Tets double-recombinants strains were screened by PCR and

named DM2824 hrpL::Ω hrpS-lacZ and DM2830 hrpL::Ω hrpX-lacZ.

Interposon-chromosome junctions were confirmed by Southern blot

analysis. Plasmid pMM191, containing a 5’deletion of the hrpL regulatory

157 region, was constructed by amplifying a 315 bp fragment of hrpL with

primers LRBG and LFU and pMM58 as template DNA (Table 3.2). The

insert was cloned as a BamHI insert in pPL6GUSC and sequenced.

3.2.4 Analysis of β-glucuronidase or β-galactosidase activity

β-glucuronidase and β-Galactosidase activities in bacterial cultures

were measured using the fluorogenic substrates 4-methyl-umbelliferyl (4-

MU)-glucuronide (MUG) and 4-MU-galactoside (MUGal), respectively.

Assay were done in 96-well microtiter plates following standard protocols

(Miller, 1992; Jefferson, 1985) modified as previously described (Merighi

et al., 2003). One unit of enzyme activity is equal to 1 pmol 4-methyl-

umbelliferone (MU) per min per OD600.

3.2.5 Total RNA isolation

Total bacterial RNA was isolated from 5 ml of overnight cultures of

cells grown in IM pH5.5 liquid medium or 5 ml of exponentially growing

cultures in LB broth. Cells were pelleted and resuspended in 0.1 ml of

TEN100 buffer (10 mM Tris-HCl, 100 mM NaCl, 1 mM EDTA, pH 7.0).

Five volumes of Trizol (Invitrogen) at 65°C were added to the

resuspended cells, which were then vortexed, shaken at 65°C for 10 min

and centrifuged at 10,000 x g for 30 min at 4°C. The upper aqueous phase was extracted with 0.1 ml of water-saturated chloroform and the RNA was precipitated by adding one volume of isopropanol and centrifuging as 158 above. RNA was resuspended in 50 µl of DEPC-treated water and treated with DNase I (Invitrogen) according to the manufacturer’s instructions.

To assess the quality of the RNA preparations, 10 µl aliquots were analyzed by standard agarose gel electrophoresis with ethidium bromide staining.

3.2.6 Primer extension and DNA sequencing

Analysis of 5’-ends of transcripts was performed by primer extension

(Sambrook et al., 1989). Primers of the “PE” series described in Table 3.2 were labeled with ([γ32P] ATP and cDNA synthesis was carried out using

M-MLV reverse transcriptase (Promega) with 50 µg of RNA per reaction following the manufacturer’s instructions. DNA sequencing was performed with the same primers used for cDNA synthesis following the fmolTM cycle sequencing protocol from Promega. The labeled extension products were visualized by autoradiography with Kodak XOMAT films after electrophoresis in 6% denaturing acrylamide gels at 55 Watts in a

Bio-Rad sequencing apparatus. For weakly labelled cDNAs, a second exposure was performed, and the DNA ladder and cDNAs were merged electronically using Photoshop Elements (Adobe).

3.2.7 RT-PCR analysis

Analysis of inter-operon transcripts was performed using the Titan-One-

Tube RT-PCR protocol from Roche. Sequence and coordinates of the

159 primers are given in Table 3.2. For the synthesis/amplification of cDNAs

from internal portions of hrpS (Fragment A; Fig. 3.5) and hrpL (Fragment

D; Fig. 3.5), the primer pairs RTSR/RTSF and RTL/RTLF, respectively, were used. For the reverse transcription/amplification of hrpS-hrpY

(Fragment B; Fig. 3.5) and hrpX-hrpL (Fragment C; Fig. 3.5) inter-operon

cDNAs, primers RTYSR/RTYSF and RTLXR/RTLXF were used,

respectively. Per each 50-µl reaction, 2 µg of DNA-free total RNA was

mixed with 1 µM of each primer, 1X Titan RT buffer (with 1.5 mM

MgCl2 ), 0.2 mM dNTP mix, 5 mM DTT, 10 U RNase inhibitor

(Invitrogen), and 1 µl Titan enzyme mix (AMV and Expand High Fidelity;

Roche). Samples were incubated at 55°C for 30 min for reverse transcription, followed by PCR amplification. Initial denaturation was performed at 94°C for 2 min, followed by 30 cycles of denaturation at

94°C for 1 min, annealing at 55°C for 1 min, extension at 68°C for 2 min; final extension was at 68°C for 10 min. The “No RT” control was run following a standard PCR protocol using Taq DNA polymerase (Ausubel et al., 1987).

3.2.8 Northern hybridizations

For each strain, ca. 10 µg of total bacterial RNA was electrophoresed in a 0.9% agarose gel in 1X MOPS/formamide buffer as previously described (Ausubel et al., 1987). RNA gels were blotted overnight onto a

Nytran (S&S) nylon filter in 5X SSC buffer and fixed to the filter by UV

160 cross-linking following standard protocols (Sambrook et al., 1989). The

probes used to analyze the various transcripts were synthesized by PCR

using plasmid pMM58 as a template. Primers LFB and LRE (Merighi et

al., 2003) were used to make the probe for the hrpL transcript, primers X-

pMal-F and X-pMal-R for the hrpXY transcript and primers ET15SF and

ET15SR for the hrpS transcript. PCR products were labeled by random

priming using ([α32P] dATP and the Klenow fragment of DNA polymerase

I (Ausubel et al., 1987). About 2 x 107 cpm (Cerenkov) of probe were

used for each hybridization experiment. Prehybridization-hybridization

and post-hybridization wash conditions were performed at high stringency

conditions. The hybridization signals from the blots were detected by exposure to X-ray films (Kodak Biomax MS) at –80°C between two

intensifying screens for 24 to 48 h.

3.3 RESULTS AND DISCUSSION

3.3.1 Constitutive expression of HrpS indirectly upregulates hrpS

transcription by acting on the hrpL promoter

To confirm our preliminary results on hrpS-lacZ autoregulation, we analyzed gene expression in the strain MEX1 hrpS-lacZ carrying pRF205

+ Plac-hrpS after growth in IM5.5 for 16 h. (A map with relevant mutants

and plasmids is shown in Fig. 3.1.) β-galactosidase levels showed a modest but consistent upregulation of hrpS ranging from 1.5- to 2.5-fold

(Fig. 3.2) upon overexpression of hrpS from pRF205. A similar 161 experiment using strain DM733 hrpY::Tn5 hrpS-lacZ was performed. The results showed that the presence of the hrpY::Tn5 mutation strongly reduced expression of hrpS-lacZ in strain DM733 as compared to MEX1 and that the hrpS-lacZ fusion was not stimulated by overexpression of hrpS (Fig. 3.2). This finding is consistent with the model of regulation discussed in Chapter 2 and this latter result implies that HrpS is not acting directly on the hrpS promoter, but may require either HrpY or transcription from the hrpXY and/or hrpL promoters.

To verify whether the autoregulatory effect was due to transcriptional effects on the hrpXY promoter, we first assayed its expression in the presence and absence of pRF205 using strain MEX116 hrpX-lacZ as the genetic background. The results showed that hrpXY expression was upregulated ca. 4-fold by ectopic expression of HrpS (Fig. 3.2).

Upregulation of hrpX-lacZ by pRF205 was blocked in a strain carrying a second polar mutation in hrpL (strain DM2824 hrpL::Ω hrpX-lacZ) (Fig.

3.2), which was constructed by allele exchange of a hrpL::Ω gene into

MEX116 hrpX-lacZ. It is unlikely that HrpL is required for hrpXY upregulation by HrpS, because there is not a recognizable Hrp box in front of hrpXY, and the previous result suggests that the hrpXY promoter is not directly regulated by HrpS. Therefore, we hypothesized that the hrpL promoter may be required for the HrpS-dependent effect. This hypothesis was confirmed by assaying hrpS expression in strain DM2830

162 hrpL::Ω hrpS-lacZ in the presence and absence of pRF205 (Fig. 3.2).

Again, the autoregulation of hrpS observed in strain MEX1 hrpS-lacZ was

blocked by a polar mutation in hrpL.

Collectively these results suggest that HrpS increases hrpS-lacZ

expression via its action on the hrpL σ 54 promoter, which is part of the

linear cascade model described in Chapter 2. To explain how

autoregulation occurs, two possible scenarios can be hypothesized: (i) the

HrpS effect depends on increased expression of hrpXY via the hrpL σ 54 promoter, with subsequent activation of hrpS by HrpY as described in

Chapter 2 (“trans-effect”); or (ii) transcription from the hrpL σ 54 promoter could continue through hrpXY into hrpS (“cis-effect”). The first possibility is partially ruled out by the observation reported in Chapter 2

(Merighi et al., 2003) that a plasmid-borne hrpS-uidA gene fusion is not affected by hrpS mutations. In this experiment, we would have expected some downregulation of hrpS-uidA if the “trans-effect” hypothesis was correct, because reduced expression of hrpL, and consequently hrpXY, in the hrpS mutant would have ultimately reduced HrpY levels.

3.3.2 The upstream region of the hrpL promoter contains a symmetrical dyad required for HrpS-dependent activation

We wanted to test if the putative UAS site located at –163 from hrpL is required for the HrpS-dependent upregulation of this gene. The dyad-

163 symmetrical DNA element (TGCAA-N4 -TTGCA) weakly matches the E.

coli glnA enhancer site (Chapter 2). PCR fragments containing 134 and

208 bp upstream of hrpL were cloned into the GUS reporter plasmid pPL6GUSC to create plasmids pDM2785 and pMM191, respectively. The resulting plasmids were conjugated into strains DC283 and DM758 hrpS::Tn5 and then PhrpL-uidA expression was measured after growth in

IM5.5. The control reporter fusion in pDM2785, which contains all the

hrpL elements required for HrpS-dependent regulation, was highly

expressed in DC283 and reduced to basal levels in the hrpS background

(Fig. 3.3). However, the expression of the fusion in pMM191, which is

missing the putative UAS, was reduced to basal levels in both genetic

backgrounds (Fig. 3.3), thereby confirming that the deleted 74 bp region

contains an essential cis-acting sequence needed for normal regulation of

hrpL.

3.3.3 Primer extension analysis defines the 5’ ends of hrpXY, hrpS and

hrpL transcripts and their potential promoters

To further test the “cis-effect” model, we located the 5’ ends of the

hrpXY, hrpS and hrpL transcripts by primer extension analysis and

identified their potential promoters. As shown in Chapter 2, sequence

analysis revealed potential σ 70 promoters in front of hrpXY and hrpS and a

σ 54 promoter in front of hrpL. Primer extension analysis of mRNA

expressed under Hrp-inducing conditions showed that the hrpS mRNA 164 started 611 nt upstream of the ORF and included the entire abortive IS element (Fig. 3.4). Such a long untranslated leader sequence may form a strong (-149.2 kcal mol-1) hairpin structure, as discussed in Chapter 2.

This highly stable structure could complicate the analysis of transcripts by RT-PCR, reducing the efficiency of DNA polymerization or cause template switching. A putative σ 70 promoter is centered at –25 bp from the hrpS transcription start and has sequence (5’-GATGGTT-N17-TATTT-

3’). Primer extension of the hrpL transcript consistently produced a strong signal mapping at -39 nt from the hrpL start codon (Fig. 3.4). Visual

54 inspection identified a σ promoter (5’-CTGGCA-N6 -TTGCT-3’) centered at –19.5 bp from the transcript start. The hrpXY transcript start was mapped at –12 from the start codon, with a σ 70 consensus promoter

(5’-TGTGGA-N16-TTAA-3’) centered - 24.5 bp from the transcriptional start nucleotide(Fig. 3.4).

3.3.4 RT-PCR reveals that the hrpL-hrpXY and hrpXY-hrpS interoperon regions are transcribed

Oligonucleotide primers were designed to amplify cDNA transcripts spanning the hrpL-hrpXY and hrpY-hrpS intergenic regions and internal portions of hrpL and hrpS using a “one-tube” RT-PCR reaction (Fig.

3.5A; fragments C, B, D and A respectively). Total RNA was purified from P. stewartii DC283 grown in IM5.5 to early stationary phase. The control products from within the hrpL and hrpS ORFs (254 and 232 bp for 165 fragments A and D, respectively) were obtained, indicating that the hrp

regulatory genes were expressed under the experimental conditions used.

PCR products were also obtained using primers designed to amplify either the entire hrpXY upstream regulatory region (256 bp, fragment C; lane 2,

Fig. 3.5B) or the part of the hrpS regulatory region located upstream of the IS-like element (271 bp, fragment B; lane 4, Fig. 3.5B). Attempts to obtain PCR products with primers amplifying a product spanning the whole 5’hrpS region (fragment E; not shown in the gel figure) repeatedly failed, perhaps due to the strong secondary structure of the IS-like mRNA.

No product was amplified from the control without the reverse transcriptase enzyme, which was run using primers for the internal portion of hrpS and DNA-free total RNA as template (lane 6, Fig. 3.5B). These results indirectly show that no termination occurs downstream of hrpL and hrpXY.

3.3.5 Northern blot analysis reveals the presence of long transcripts spanning hrpL and hrpXY that are disrupted in a hrpL polar mutant

We attempted to confirm the results of the RT-PCR analysis by demonstrating the existence of long transcripts starting from PhrpL and spanning hrpXY and hrpS. We analyzed the transcripts of the hrp regulatory regions by Northern blotting using probes homologous to the hrpL, hrpX and hrpS genes (Fig. 3.6). DNA-free total RNA was prepared from strains DC283 (wild-type) and MEX105 hrpL::Tn3HoHoI grown in 166 IM5.5 to early stationary phase. Northern blot signals were generally

weak, probably due to the high-turnover rate of hrp messages, and

extended exposure for several days were required. In DC283, the hrpS

probe identified at least four transcripts of sizes 2.4 kb, 1.6 kb, 1.2 kb and

1.0 kb. The 2.4 kb and 1.6 kb signals were absent in the MEX105 hrpL::Tn3HoHoI strain, which still had the 1.2 and 1.0 kb transcripts

(Fig. 3.6). The negative control strain DM758 hrpS::Tn5 did not produce any signals with the hrpS or hrpL probes (not shown). The hrpL probe detected many transcripts in DC283 ranging in size from 2.9 kb to 0.7 kb

(Fig. 3.6), but only a 0.5 kb transcript was detected in strain MEX105 hrpL::Tn3HoHoI. Even if the long transcripts have a size shorter than expected, which could be due to rapid degradation of the mRNAs, these results are consistent with a regulatory model whereby transcription from the hrpL promoter continues into hrpXY and part of the hrpS regulatory region. This is consistent with the absence of clearly identifiable rho- independent terminators downstream of hrpL or hrpY. In contrast, potential terminators are present in the nucleotide sequence of this region from the related species E. amylovora, but a detailed transcriptional analysis has not been done in this bacterium. Therefore, this form of autoregulation may be a specific adaptation of P. stewartii.

167

3.4 CONCLUSIONS

The majority of the genetic and biochemical data presented in this

chapter support a model where hrpS is autoregulated at the transcriptional level via a DNA site located 134 bp upstream of hrpL, almost 3 kb from the hrpS 5’ regulatory region. This autoregulation does not appear to

depend upon trans direct activation of hrpS by either HrpS or HrpY,

because it was observed only when reporter fusions to hrpS were located in cis to the other regulatory genes. The effect of HrpS on its own

expression seems to be due instead to the production of a novel transcript

spanning the entire hrpL-hrpXY-hrpS region.

The results obtained from the RT-PCR analysis were consistent with the absence of transcriptional terminators downstream of hrpL and hrpY, thereby confirming the sequence analysis presented in Chapter 2. This result implies that interoperon transcription is possible in this system. We were not able to demonstrate the existence of the predicted 4.2 kb transcript originating from the hrpL promoter, but a 2.9 kb mRNA was identified in DC283 using the hrpL probe, which was absent in polar hrpL

mutants. Similarly, a 2.4 kb mRNA was detected in DC283 with the hrpS

probe and long transcripts were absent in the hrpL mutant. The estimated

sizes of the long transcripts in these experiments may not be accurate because our size standards did not contain fragments >3.5 kb. Also,

168 bacterial mRNAs undergo rapid turnover, so that the isolation of large,

undegraded transcripts is technically difficult. The model in Fig. 3.7

summarizes these findings.

Autoregulatory feedback loops are common in biological systems and

extensive literature on their mathematical modeling is available (Glass

and Kauffman, 1972; Thomas, 1973; Thomas, 1978). Most simple positive

feedback mechanisms lead to bistable systems, where two steady states

are observed. Modest perturbations of the system have little effect on the

level of expression, whereas major stimuli shift the system back and forth

between the two extreme levels of expression (e.g. on and off). The

regulatory loop triggered by HrpS appears to be a pure feedback

mechanism, at least at first approximation. If we consider the other

additional regulators entering the loop, the system may theoretically allow

for stable cycles or multiple steady states depending on whether negative

or positive regulators are involved in the feedback (Thomas, 1978). For

instance, additional known positive regulators are HrpY acting on hrpS

and HrpY acting on hrpXY. Negative regulators of hrpS are hypothesized

or shown to exist in Chapter 4 and Chapter 6. These negative regulators

may act by merely (re?)-inforcing the “off” state of the bi-stable switch more than being feedback controllers. Therefore the net effect is probably a system with at least two steady states.

Whether HrpY plays a direct role in reinforcing the HrpS-autoregulatory loop is not clear yet. The absence of any effect of hrpS mutations on a 169 plasmid-borne hrpS::uidA reporter (Table 2.5) suggests that changes in

the level of hrpXY expression arising from the feedback loop do not affect

hrpS significantly, possibly because the two-component system is already

at maximal expression under the growth conditions used. Therefore, one

may conclude that HrpY trans-activation of hrpS may not reinforce the loop significantly in the experimental conditions used in this study.

Alternatively, the fact that the hrpS-gusA fusion is plasmid-borne may alter the expected down shift in the hrpS mutant if one of the roles of

HrpY is to function as an antirepressor for hrpS transcriptional activation, as discussed in Chapter 4. In other words, a putative repressor of hrpS

may be titrated out by the plasmid-borne promoter fusion, therefore

masking the effect of HrpY.

An additional level of regulation of hrpS may involve a repressor of the interoperon transcription from hrpXY into hrpS. Indeed, hrpXY

transcription has a high, constitutive basal level in rich media.

Considering the absence of rho-independent terminators downstream of

hrpY, constitutive expression of hrpS would also be expected. However,

hrpS is repressed in rich media and only activated in planta or in IM5.5.

This issue is partially addressed in Chapter 4.

Determining whether the regulatory system works as an on/off switch or

as a rheostat with multiple levels of expression may be relevant to the

biology of the pathogen when it switches from living in the insect vector

to colonizing a corn leaf. From an experimental viewpoint, determining 170 this would require accurate quantification of the level of RNA and protein

expression over time and space using highly unstable reporters able to

quickly respond to the changing environment. The GUS reporters used in

this study are either not suitable for in planta expression analysis or are

too stable to indicate rapid decreases in expression. To this respect, the high turnover gfp[AAV] fusions described in Chapter 2 may offer new

tools to tackle these questions.

One possible biological role of the autoregulatory loop may be to

compensate for high turnover rates of the hrpS mRNA. Anecdotically, we had difficulties in amplifying by RT-PCR hrpS fragments from some RNA preparations, whereas the amplification of hrpL and hrpXY cDNAs was

always successful. This could reflect a greater instability for hrpS

transcripts. Measurements of turnover rates were not performed, so this

remains only a working hypothesis, but RNA stability has been shown to

play a role in regulation of hrp genes in different erwinias (Chatterjee et

al., 2002). As a negative note, an additional result previously published

by our lab (Frederick et al., 2001) seems to contradict the importance of

this autoregulatory loop during the infection of the plant. Indeed, a

Tn3HoHoI polar insertion between hrpL and hrpXY (strain MEX7) did not

affect virulence as measured by water-soaking production. This insertion

is expected to disrupt the autoregulatory loop at the transcriptional level.

One possible explanation is that under experimental infections, which

present less demanding and variable conditions than in natural field 171 infections, the autoregulatory loop may be optional. Future experiments should test whether virulence of MEX7 under field conditions is affected or not.

3.5 SUMMARY

Due to the particular arrangement of the hrpL, hrpXY and hrpS operons and the absence of clearly identifiable transcription terminators downstream of hrpL and hrpXY, activation of hrpL by HrpS can increase expression of HrpS leading to its apparent autoregulation. In this study, we used genetic and molecular biology approaches to demonstrate the existence of this autoregulatory loop and its dependence on a UAS located upstream of hrpL. The role for autoregulation in the disease cycle/biology of the disease and future studies are discussed.

172 Strain, or plasmid Relevant phenotype(s) and/or genotype(s) Source or reference

Escherichia coli

r q + + BW20339 F' 128::Tn10-12(Kan ) lacI ∆lacZM15 pro(BA) /DE3(lac)X74 uidA(∆MluI)::pir recA1

∆phoA532 ∆(phnC∆DEFGHIJKLMNOP)33-30 Metcalf et al. ( 1996)

DH10B F- mcrA ∆(mrr-hsdRMS-mcrBC) φ80dlacZ∆M15 ∆lacX74 endA1 recA1 deoR

∆(ara,leu)7697 araD139 galU galK nupG rpsL λ- Gibco BRL

- - - r HB101 F thi-1 hsd20 (r Bm B) sup E44 recA13 ara-14 leuB6 proA2 lacY1 rpsL20 (Sm )

xyl-5 mtl-1 Boyer and Iates Roulland

-Dussoix (1969)

HB101Rif spontaneous Rifr strain Lab collection

Sφ200Rif metB strA purB ∆(agg-uidA-man) Rifr Wei et al. (2000)b

P. stewarti subsp. stewartiii

SS104 Wild type, Wts+, HR+ ICPPBa

DC283 SS104 Nalr Coplin et al. (1986)

DM064 DC283 hrpY1296::Tn5 (Kanr) Frederick et al. (2001)

DM733 DC283 hrpY64::Tn5 hrpS1::Tn3HoHoI (Kanr, Ampr) Merighi et al. (2003)

DM758 DC283 hrpS224::Tn5 (Kanr) Frederick et al. (2001)

MEX1 DC283 hrpS1::Tn3HoHoI (Ampr) Frederick et al. (2001)

MEX116 DC283 hrpX116::Tn3HoHoI (Ampr) Frederick et al. (2001)

Plasmids pBluescript

KS and SK (+) ColE1 αlacZ (Apr) Stratagene pLAFR3 IncP cos αlacZ (Tcr) Staskawicz et al. (1987)

pRK415 IncP αlacZ (Tcr) Keen et al. (1988)

pHP45::Ω ColE1, (Apr) Frey and Krisch, 1985

pLD55 R6K ori, αlacZ, tetAR (Apr) Metcalf et al. (1996)

Table 3.1 (continued)

173 Table 3.1 (continued)

pPL6GUSC pLAFR6 derivative carrying a promoterless uidA gene (Tcr) Knoop et al. (1991)

+ - r r r pRK2013::Tn7 ColE1 mob traRK2 ∆ repRK2 repE kan::Tn7 (Tp , Sm , Sp ) Ditta et al. (1980)

pDM2533 pBSK carrying a 1.2 kb PCR fragment containing hrpL+ This study

pDM2862 pDM2533 with an Ω/SpR cassette at the NruI site of hrpL This study

pDM2868 pLD55 with the pDM2862 insert as XhoI/SstI fragment This study

pDM2785 pPL6GUSC with a 389 bp BamHI PCR fragment to create a hrpL-uidA fusion Merighi et al., (2003)

pDM1296 pES1044 hrpY1296::Tn5 (Kanr) Coplin et al., (1992)

pRF205 pVK100 with the 1.7 kb HindIII fragment containing hrpS from pES1044 Frederick et al. (1993)

pMM25 pPL6GUSC with a 244 bp HindIII-BamHI PCR fragment to create a

hrpX-uidA fusion Merighi et al. (2003) pMM50 pPL6GUSC with a 1.3 kb BamHI PCR fragment to create a hrpS-uidA fusion Merighi et al. (2003) pMM191 pPL6GUSC with a +220 to –95 hrpL PCR fragment fused to uidA This study

a International Collection of Plant Pathogenic Bacteria

Table 3.1. Bacterial strains and plasmids.

174 Primer Sequence (5'Æ3') Sequence coordinates (5') Source

RTSR4650 GTAAGCTAAGTTGTTGAC 4650-4633 This study

RTSF4397 ACTCATTTCGACCACATC 4397-4414 This study

RTLR931 GGGCAATACCAGATACCC 931-914 This study

RTLF700 AACCTGCACCCTACTGAG 700-683 This study

RTYSR3692 GCTGATATCGGGTAAGGA 3692-3675 This study

RTYSF3422 CATAAAGCCCGCCTGATG 3422-3439 This study

RTLXR1424 GATTGGCGTATCACTGCT 1424-1407 This study

RTLXF1169 CTGTTGGCACAGTACGTT 1169-1152 his study

LFB CGCGGATCCACATTAAGCCAAACGGCAAAA 662-642 Merighi et al., 2003

LRE CCGGAATTCTGTTTTACCCCGTTCAGTT 1255-1237 Merighi et al., 2003

XPMALR CGGAATTCTCACTCAGGGTATTAAAGGAATGGTC 2816-2837 This study

XPMALF GGGAATTCATGCAGTATTTGGACAGAC 1380-1398 This study

ET15SF CTACATATGAATATTGAAAATAATGAGCACTCATTCCGA 4354-4383 This study

ET15SR CGCGGATCCCAATCAGTCAGACAATGATGCGTTGCGTT 5301-5327 This study

LRBG CGCGGATCCTCTAGAATGACTTCCAGCCAGGTCAT 869-850 Merighi et al., 2003

LFU- CGCGGATCCGAGCTCAAAGTTGCGAAAAAATGAAGG 550-570 This study

PEX GCACGCTTCGTCATTGACAT 1498-1517 This study

PEL TATGTGGGGATGGAGGGTTAACGTG 724-748 This study

PES-1 CGATAGGTTGCTCTTTCGTGAATG 4412-4435 This study

PES-2 TCACCGACTCGCAATAATCC 4286-4305 This study

PES-3 TGGCTGGCTCACAATAACTG 4168-4187 This study

PES-4 TAGACAGGCAGCATGTGTTC 4024-4043 This study

PES-5 TGCGAGCCGCTACACTTATC 3839-3858 This study

PES-6 CGGGTAAGGAATTGATGGAG 3665-3684 This study

Table 3.2. Primers

175

hrpY1296::Tn5 hrpL2868::Ω

hrpS hrpY hrpX hrpL

MEX1 MEX116

HindIII HindIII XhoI SstI Plac pRF205 pDM2868

BamHI

pDM1296

Figure 3.1. Physical map of the hrp regulatory region, mutants and plasmids.

Open boxes represent open reading frames and the solid box represents the IS element remnant. Solid circles indicate Tn5 insertions, open circles indicate Ω interposon insertions, and arrows indicate Tn3HoHoI-lacZYA insertions. The inserts of the plasmids used in this study are shown beneath the map. The arrows also indicate the direction of transcription. MEX1 is DC283 hrpS-lacZ and MEX116 is

DC283 hrpX-lacZ.

176 3500 3000 2500

2000 none 1500 +pRF205 1000 500

B-Galacoside activity (pmol MU/min/OD600) 0

830 EX1 2 M DM733 M2824 M D MEX116 D

Figure 3.2. Effects of HrpS+ constitutive expression on chromosomal hrp-lacZ

fusions in various genetic backgrounds.

P. stewartii strains were grown in IM5.5 to OD600> 0.7 and ß-galactosidase assays

were performed as described in Materials and Methods. Enzyme units are pmol MU

per min per OD600. Strain MEX1 is DC283 hrpS-lacZ, strain DM733 is DC283 hrpY::Tn5 hrpS-lacZ, strain DM2824 is DC283 hrpL::Ω hrpS-lacZ, strain MEX116 is

DC283 hrpX-lacZ, and strain DM2830 is DC283 hrpL::Ω hrpX-lacZ.

177

54 DC283 DM758 σ IHF UAS hrpL hrpJ

5793±10 62±2 GUS pDM2785

71±2 54±1 GUS pMM191

100 bp

Figure 3.3. The effect of deleting the upstream enhancer sequence (UAS) in the hrpL regulatory region on PhrpL expression.

Expression of reporter plasmids pDM2785 and pMM191, carrying 208 bp and 134 bp, respectively, of the 5’ hrpL region fused to uidA, was analyzed in strains DC283 and DM758 after growth in IM5.5. pMM191 lacks the putative UAS. A map of the hrpL-hrpJ intergenic region and the inserts cloned into the reporter plasmids is shown. Data are means from two experiments with two replicates each. GUS activities

-1 –1 are reported as pmol MU min OD600 ± standard deviation.

178

TGCA T G C A 5’ T G C A 5’ C T C G G 5’ C +1’ T T A A T G T -10 G -12 A T C C T T C C A G T T G T C T A G T T C -10 A G C A C A C C T A C T A C A T T A G T G +1 G +1 T T C T A C A A A G A C +1 C A T A C T C G C T G A T C A C 3’ G G hrpL hrpX hrpS G 3’ 3’

Figure 3.4. Primer extension analysis of hrpL, hrpX and hrpS transcripts.

Total RNA from DC283 cells grown in IM5.5 was annealed with the 32P-end labeled primers PEL, PEX and PES-5 shown in Table 3.2 and extended as described in

Material and Methods. Lanes A, T, G, and C show the dideoxy sequencing ladders for

the hrpL, hrpX and hrpS genes generated with the same primers. The transcriptional

start sites are indicated by solid arrows. Mapping of the hrpXY 5’ mRNA end always

produced very weak signal (+1), with often a secondary signal locate 21 nt further

upstream (+1’). The lane corresponding to the hrpX cDNA was enhanced with

Photoshop (Adobe). The start codons (ATG) are highlighted. RBS = ribosomal binding

site.

179 A 254 bp 271 bp 256 bp 232 bp A B C D RTSR → RTYSR → ← RTYSF RTLXR → ← RTLF

← RTSF RTLXR → 1229 bp ← RTLXF E

hrpS hrpY hrpX hrpL

B 1 2 3 4 5 6 7

1 kb

0.5 kb

Figure 3.5. RT-PCR analysis of interoperon read-through transcription.

(A) The products amplified by RT-PCR and their sizes are shown as black bars above the genetic map (amplicons A, B, C, D, E). The black box upstream of hrpS represents the abortive IS element. The long transcript hypothesized in this study and mRNAs from each operon are shown below the map as arrows. (B) To demonstrate inter-operon transcription, total RNA was isolated from strain DC283 grown in IM5.5. mRNAs were reverse transcribed and cDNA was amplified using specific primer pairs to obtain the PCR products shown in the gel. Lane 1: 1 kb-plus molecular weight markers (Invitrogen); Lane 2: control without reverse transcriptase; lane 3: fragment

A; Lane 4: fragment B: lane 5: fragment C; lane 6: fragment D; lane 7: 1 kb-plus molecular weight markers.

180

Figure 3.6. Northern blot analysis of hrp regulatory transcripts.

(A) Northern blot autoradiographs. Total RNA was isolated from DC283 and

MEX105 hrpL::Tn3HoHoI grown in IM5.5. About 10 µg of RNA was separated in a formamide/MOPS gel and blotted on a nylon filter. The filter was then hybridized at high stringency with either a hrpS or a hrpL probe, labeled by random priming with

[α 32P]dATP. The washed filters were exposed to X-ray film for 48 hrs. Sizes of the transcripts in kb are indicated next to arrows and were determined using E. coli and

Arabidopsis thaliana rRNA as molecular mass standards. (B) Gene map with the probable position of the observed transcripts and the map position of the probes used

(other transcripts could by hypothesized if one considers hybridizations not spanning the whole probe). The symbol indicates a Tn3HoHoI transposon insertion.

181

182

HrpS

~P

hrpS hrpY hrpX hrpL

Activation Hrp regulon

Figure 3.7. Model of the hrp regulatory cascade including the autoregulatory loop

triggered by HrpS

This regulatory model is an extension of the linear model discussed in Chapter 2.

The activation of hrpL by HrpS is not only part of the basic cascade that activates the

Hrp regulon, but also triggers a positive autoregulatory loop that increases the levels

of hrpXY and hrpS transcripts. This loop may reinforce the autoregulatoy loop

triggered by HrpX/HrpY, thereby stabilizing the “on” state of the regulatory circuit.

183

CHAPTER 4

BIOCHEMICAL ANALYSIS OF THE RESPONSE REGULATOR

HRPY AND OF ITS ROLE AS A DUAL ACTIVATOR/REPRESSOR

OF HRP GENES

4.1 INTRODUCTION

hrp/hrc genes are required for infection and colonization of plants by necrogenic bacteria of the genera Pseudomonas, Xanthomonas, Erwinia

and Pantoea (Alfano and Collmer, 1997). They also play a role in rapid

establishment of soft rot erwinias, where enzymatic tissue disintegration

is only delayed in a hrp mutant (Rantakari et al., 2001; Perombelon,

2002). P. stewartii has a group I hrp gene cluster (Frederick et al., 2001), similar to those in Pseudomonas, Erwinia, Pectobacterium and other

Pantoea species, which is regulated by the HrpS enhancer and the ECF alternative sigma factor HrpL (Hutcheson, 1999). As shown for Pantoea agglomerans pv. gypsophilae and Erwinia amylovora, P. stewartii hrp genes are also regulated by the HrpX/Y two-component system (Wei et

al., 2000b; Merighi et al., 2003; Nizan-Koren et al., 2003). In all cases where mutational analysis has been performed, the HrpY response

184 regulator is essential for pathogenicity, whereas nonpolar mutations in the

hrpX sensor kinase gene have modest or no effects on pathogenicity (Wei

et al., 2000b; Merighi et al., 2003; Nizan-Koren et al., 2003). HrpY has

sequence similarity to response regulators of the FixJ class, which

includes E. coli NarL and UhpA and Bacillus subtilis DegU (Merighi et

al., 2003). Members of this class share a typical modular organization

(Stock et al., 2000), with a conserved N-terminal receiver domain (shared

by all classes of response regulators), a flexible linker, and a C-terminal

domain that contains a LuxR/GerR superfamily DNA-binding motif

(Heinikoff, 1995). The N-terminal domain of P. stewartii HrpY contains

three conserved aspartyl residues (D11, D12 and D57). By homology to

other response regulators, D57 is expected to be the phosphorylation site.

In agreement with this prediction, conservative and structurally neutral

amino acid substitutions at this position, such as D57N and D57A, abolish

the activity of hrpY when expressed in single copy. However, these

mutant proteins can fully complement hrpY null mutants for virulence or

HR elicitation when overexpressed from plasmids (Merighi et al., 2003).

Similar findings have been reported for E. coli UhpA (Kadner, 1983;

Kadner, 1995; Webber and Kadner, 1995). In all these cases, the

implication is that both the phosphorylation status of the response

regulator and its concentration in the cell are important parameters for its

activity and that the requirement for phosphorylation can be bypassed by

overexpressing it.

185 Using a genetic approach, we previously showed that P. stewartii HrpY

activates expression of hrpS and autoregulates the hrpXY operon (Merighi et al., 2003; and Chapter 2). This was confirmed by genetic reconstruction of each step in E. coli, which suggests that the effects of HrpY could be direct, unless HrpY is trans-regulating conserved factors in E. coli

(Merighi et al., 2003). However, the same set of genetic reconstruction experiments unexpectedly revealed that HrpY also represses the basal level of hrpL transcription when HrpS is absent. Inspection of the sequence of the regulatory regions of hrpS and hrpL did not reveal any apparent common sites that might explain the opposite effect of HrpY on these two promoters, but the regulatory sites bound by FixJ-class regulators are often quite degenerate (Tyson et al., 1994; Dahl et al.,

1997).

Although extensive genetic data is available for the regulatory role of

HrpY in P. stewartii, P. agglomerans pv. gypsophiliae and E. amylovora, a biochemical analysis of the target regulatory regions has not been done and it is not known how HrpY activates transcription. In this study, we define the portions of the hrpS and hrpL 5’ regions required for HrpY- dependent regulation and we provide biochemical evidence for the interaction of recombinant HrpY with defined regions of the two hrp promoters. We also show that HrpY has a single phosphorylation site located at D57 and that its phosphorylation increases the affinity of HrpY for the DNA sites in front of the hrpS promoter. Specific binding of HrpY

186 to the hrpL promoter at low affinity was also demonstrated in vitro.

Preliminary DNase I footprints for the hrpS regulatory region and a

working model explaining most of our genetic and biochemical data are presented.

4.2 MATERIALS AND METHODS

4.2.1 Strains and growth conditions

E. coli and P. stewartii strains and plasmids used in this study are listed in Table 4.1. Luria-Bertani (LB) broth and agar were used for strain

maintenance and clonings. To induce hrp genes in P. stewartii strains, we

followed the procedure described in Chapter 2 using the IM pH5.5

(IM5.5) liquid medium. Liquid cultures were grown in flasks or tubes with

shaking at 200 rpm at 37°C for E. coli or 29°C for P. stewartii. Special

growth conditions were used for growing E. coli BL21(DE3) strains

during expression purification of various recombinant proteins, as

described below. When appropriate, antibiotics were supplied at the

following concentrations: ampicillin, 200 µg ml-1; kanamycin, 50 µg ml-1; and tetracycline, 20 µg ml-1; chloramphenicol, 34 µg ml-1.

187 4.2.2 Construction of plasmid-borne gene fusions and of a hrpY[D57E] allele for plasmid expression

Unless otherwise stated, isolation and manipulation of recombinant

DNA molecules used standard procedures (Ausubel et al., 1987) or the various product manufacturers’ instructions. Fragments containing deletions of the 5’ regulatory region of hrpS were produced by PCR using the primers given in Table 4.2. Oligonucleotide primers were synthesized at IDT-DNA (Iowa). All PCR products were generated with 15 to 20 cycles of amplification using Pfu DNA polymerase (Clontech) and 500 ng of plasmid pMM58 as a template. Gel-purified products were digested with BamHI and ligated into pPL6GUSC to create transcriptional fusions to a uidA (GUS) reporter gene (Fig. 4.1). In particular, the hrpS fragment from–231 to +628 was amplified with primers SF3509B and SR4368B and the corresponding reporter plasmid was named pMM400. Region –165 to

+628 was amplified with primers SF3575B and SR4368B and the reporter plasmid constructed was named pMM396. Region –90 to +628 was amplified with primers SF3650B and SR4368B and the corresponding plasmid was named pMM393. Region –51 to +628 was amplified with primers SF3689B and SR4368B and the corresponding plasmid was named pMM403. Finally, region –5 to +628 was amplified with primers SF3735B and SR4368B and the corresponding plasmid was named pMM392.

Construction of the two reporter plasmids carrying different fragments of the hrpL regulatory region is described in Chapters 2 and 3. The insert

188 in plasmid pDM2785 spans from –148 to +220 of hrpL (with +1 being the

transcription start at the σ 54 promoter identified in Chapter 3) and plasmid

pMM191 carries the –94 to +220 fragment. A third hrpL reporter gene

54 fusion lacking sequences downstream from the σ -promoter region

coordinates –148 to –12) was constructed by PCR, using primers JFGB

(a.k.a. HRPLP1) and L637R. The resulting plasmid was named pMM428

(Fig. 4.2).

4.2.3 Construction of plasmids for protein expression and purification

PCR was performed using Pfu DNA polymerase (Clontech), primers ET-

YF-NdeI and ET-YR-NdeI (Table 4.2), and plasmids pMM46 hrpY+ or

pMM94 hrpY[D57N] as DNA templates (Merighi et al., 2003). Primer ET-

YF-NdeI introduced an NdeI site at the start codon of hrpY, which maintains the reading frame with the vector’s N-terminal His6 -tag and

primer ET-YR-NdeI introduced three stop codons downstream of the hrpY

ORF. The ca. 670-bp PCR fragment, spanning hrpY+ or hrpY[D57N], was

gel-purified, digested with NdeI, and ligated into an NdeI-linearized pET-

15b plasmid (Novagen). The resulting plasmids, pMM221 and pMM222,

carried hrpY and hrpY[D57N], respectively, fused to an N-terminal His6 -

tag and expressed from a T7lac inducible promoter. The inserts in

plasmids pMM221 and pMM222 were sequenced using a cycle sequencing

protocol with fluorescent terminators and an ABI3700 capillary sequencer

at the OSU Plant-Microbe Genomics Facility. 189 4.2.4 Purification of His6 -HrpY proteins

Plasmids pMM221 and pMM222 were transformed by electroporation

into E. coli BL21(DE3) carrying pLysS. Several independent Apr Cmr isolates were stored at –20 and -80°C in 40% glycerol, 10 mM MgSO4 , 50 mM Tris-HCl pH 7. For protein expression, bacteria were grown overnight in 5 ml LB broth with antibiotic selection. A 3-l flask containing 1000 ml

LB broth was inoculated with 2.5 ml of overnight culture. The flask was shaken at 37°C until the culture reached A600=0.8 (measured in a B&L

Spectronic 20 colorimeter with 13 mm x 100 mm tubes). IPTG was added

to a final concentration of 1 mM and the culture was grown at 15°C for 18

to 19 h until A600=1.4. Cells were harvested by centrifugation in a Sorval

GSA rotor at 8000 rpm for 20 min at 4°C and then resuspended in 35 ml

of ice-cold lysis buffer (10 mM imidazole, 500 mM NaCl, 1 mM MgCl2 , 1 mM PMSF, 20 mM Tris-HCl pH 7.9 at 4°C), containing 50 µl of protease inhibitor cocktail (Sigma Cat. P8849). The cell suspension was incubated in 0.1 mg ml-1 lysozyme for 20 min on ice and then passed two times

through a 40-K French pressure cell (Thermo IEC) at 10,000 psi. Ten 5-s

bursts were applied with a Biologics Model 150V/T sonicator at 20% duty

cycle to further reduce the viscosity. The lysate was centrifuged twice at

20,000 x g for 15 min at 4°C. . Column chromatography was performed at

4°C. The supernatant was then applied to a 1.5 ml Ni2+-IDA column (I.D.=

1 cm; Sigma Cat. 56000) packed using His-Bind™ resin (Novagen),

190 previously loaded with 50 mM NiS04 and equilibrated with lysis buffer.

The column was washed with ten column-volumes of lysis buffer and ten

column-volumes of washing buffer (40 mM imidazole, 500 mM NaCl, 1

mM MgCl2 , 0.5 mM PMSF, 20 mM Tris-HCl pH 7.9 at 4°C). The target

His6 -HrpY proteins were eluted with 6 ml each of 100 mM and 250 mM

imidazole buffers (100 mM or 250 mM imidazole, 500 mM NaCl, 20 mM

Tris-HCl pH 7.9 at 4°C, 1 mM MgCl2 , 0.5 mM PMSF) in 1 ml fractions.

Fractions containing higher concentrations of the 26.1 kDa recombinant

protein, as assayed by SDS-PAGE, were pooled and dialyzed overnight

against 2 l of storage buffer (67 mM potassium glutamate, 125 mM

HEPES pH 7.9 at 4°C, 0.5 mM DTT, 0.5 mM PMSF, 20% glycerol). The

semipurified protein preparation was stored at –80°C until the next step.

The protein was further purified by anion exchange chromatography.

Preparations were diluted 1:10 in buffer A (20 mM Tris-HCl pH8.0 at

4°C, 25 mM NaCl, 0.25 mM EDTA, 0.2 mM PMSF), dialyzed twice for 2

h against 1 l of buffer A and loaded onto a 5-ml Q-Sepharose (Amersham-

Pharmacia) column prepared according to the manufacturer’s instructions.

The column was washed with 15 ml buffer A, proteins were eluted in a

linear buffer gradient prepared by mixing 30 ml buffer A with 30 ml

buffer B (20 mM Tris-HCl pH8.0 at 4°C, 500 mM NaCl, 0.25 mM EDTA,

0.2 mM PMSF) using a 100-ml gradient former (GIBCO) and 1 ml

fractions were collected.. Fractions containing highly purified HrpY were pooled and the buffer was exchanged for 1X storage buffer (10 mM Tris- 191 HCl pH 7.9, 125 mM NaCl, 1 mM DTT, 3 mM MgCl2 , 38% glycerol) by

repeated centrifugation in a Vivaspin 15R ultrafiltration tube (Viva

Science) with a molecular weight cut-off of 10 kDa following the

manufacturer’s instructions. Protein concentration was determined by the

dye-binding method of Bradford (Bradford, 1976). The degree of

purification was estimated by Coomassie Brilliant Blue staining

(Laemmli, 1970).

4.2.5 Mass spectrometry

Protein molecular weight measurements were performed by electrospray

ionization mass spectrometry using a Micromass Q-TOFTM II (Micromass,

Wythenshawe, UK) mass spectrometer at the OSU CCIC-MS Facility.

About 300 pmol of protein in storage buffer was concentrated and desalted in C18-solid phase manual syringe traps (Michrom BioResources) following the manufacturer’s instructions and then eluted in 50% acetonitrile, 0.1% TFA. Protein samples were introduced by infusion at 10

µl/min. Electrospray ionization was performed with a capillary voltage of

3.5 kV, source temperature of 110°C, cone voltage of 60 V, and a coaxial nitrogen flow. The quadrupole 1 range was set from m/z 100-1800 and all transmitted ions were scanned over m/z 100-3000. Data were acquired in continuum mode for 2 to 5 min. m/z spectral data were deconvoluted in the 22 to 30 kDa region using MaxEnt software (Micromass).

192 4.2.6 In vitro phosphorylation of HrpY

HrpY was phosphorylated using two alternative methods. To perform

chemical phosphorylation, a mixture containing 110 µM HrpY and 50 mM

carbamoyl phosphate (Sigma Cat. C5625) in phosphorylation buffer (125

mM potassium glutamate, 125 mM HEPES pH 8.0, 5 mM MgCl2 , 0.1 mM

DTT, 5% glycerol) was incubated at 25°C. For gel shift assays,

phosphorylated HrpY preparations were serially diluted in

phosphorylation buffer containing 50 mM NaCl. For enzymatic

phosphorylation, 15 µl of a mixture containing 25 µM HrpY, 75 µM

Salmonella enterica BarA198 (Teplitski et al., in press), 40 µM [γ32P]ATP

(10 Ci/mmol; NEN) and 0.1 mM cold ATP in phosphorylation buffer was incubated at 25°C for 90 min. Reactions were terminated with 5 µl of 4X stop buffer (125 mM Tris-HCl pH 6.8, 8 mM EDTA, 4% SDS, 8% β- mercapto-ethanol, 20% glycerol, 0.02% bromophenol blue), separated by

SDS-PAGE (12% monomer gel), and visualized by by Coomassie Brilliant

Blue staining (Laemmli, 1970). Gels were dried under vacuum at 60°C for

30 min. The radiolabeled proteins were visualized with a phosphor storage screen, analyzed on a Molecular Dynamics PhosphorImager and quantified with ImageQuant software.

193 4.2.7 Electrophoretic mobility shift assay

DNA fragments used as probes for electrophoretic mobility shift assays

were made by PCR using plasmid pMM58 (Merighi et al., 2003) as

template. Six probes (A, B, C, C1, C2 and C3) were designed to amplify

various fragments of the hrpS regulatory region located upstream of the

IS-like element. Primers SF3509 and SR3845 were used for amplification

of fragment A (327 bp); SF3509 and SR3575 for fragment B (76 bp);

SF3575 and SR3645 for fragment C (261 bp); SF3575 and SR3650 for

fragment C1 (76 bp); SF3650 and SR3689 for fragment C2 (86 bp); and

SF3689 and SR3845 for fragment C3 (101 bp). Six probes were also

synthesized to span the hrpL regulatory region (F, G, H, I, L and M). PCR

reactions used the following primer pairs: LRBG and L638F for probe F

(232 bp); LRBG and L620F for probe G (250 bp); LRBG and L578F for

probe H (292 bp); LRBG and L555 for probe I (315 bp); LRBG and JFBG

for probe L (389 bp); and L638R and JFBG for probe M (156 bp). PCR

products were purified from agarose or acrylamide gels and quantified by

the ethidium bromide staining spot method (Ausubel, 1997). DNA

fragments were labeled at the 5’ ends by incubation with T4

polynucleotide kinase (USB) and [γ32P]ATP (3000 Ci/mmol; NEN).

Unincorporated nucleotides were removed by Sephadex G-25 spin

chromatography (Sambrook et al., 1989) and labeled DNA was diluted to

10,000 cpm µl -1 (Cerenkov counts). DNA binding reactions (20 µl) contained 15 to 50 fmol 32P-labeled DNA probe, different amounts of 194 HrpY, 200 ng acetylated BSA (NEB), 1 µg poly(dI-dC) (Roche) in binding

buffer (125 mM potassium glutamate, 125 mM HEPES pH 7.9, 75 mM

NaCl, 5 mM MgCl2 , 0.1 mM EDTA, 0.2 mM DTT). Unlabeled specific

competitor DNA was added before the addition of HrpY. The reactions

were incubated at 25°C for 30-40 min. Reaction mixtures were loaded,

without loading buffer, onto a 1.5 mm-thick 6% acrylamide nondenaturing

gels (19:1 mono:bis ratio) in a V16 vertical gel apparatus (Waterman), run

into the gel at 25 V, and then separated at 200 V for 2-2.5 h. Gel and

running buffers were 1X TBE (97 mM Tris-HCl, 90 mM borate, 3 mM

EDTA, pH 8.0). The gels were dried under vacuum at 80°C and the

radioactive fragments were visualized either by autoradiography at –80°C

with Kodak MS films or on a phosphor storage screen and scanned on a

Molecular Dynamics PhosphorImager for quantification.

4.2.8 DNase I footprinting

Primer SF3509 was end-labeled with [γ32P]ATP (3000 Ci/mmol; NEN) using T4 polynucleotide kinase (USB) and unincorporated nucleotides were removed by Sephadex G25 spin chromatography. HrpS promoter fragments basically corresponding to probe A were amplified with SF3423 and SR3835 using pMM58 as a template. PCR products were purified from a 1% low-melting point agarose gel and recovered DNA was resuspended in TE buffer. DNA-protein binding reactions were performed as described for gel-shift assays using increasing concentrations of HrpY and 10 ng of 195 hrpS promoter DNA. All reactions contained 5 mM CaCl2 . The equilibrated complexes were subjected to partial digestion with protease- free DNase I (Worthington Biochemicals Cat. LS006342) diluted to 10

-1 units ml in binding buffer with 0.01% NP40 and 40 mM CaCl2 .

Digestion was performed at 25°C for 2 min and terminated with 3.5-vol stop buffer (400 mM sodium acetate pH 5.5, 0.2% SDS, 10 mM EDTA, 50

µl ml-1 yeast tRNA) and DNA was precipitated by adding 2.5 vol 100%

ethanol. Partially digested complexes were amplified by linear PCR in a

15 µL reaction mixture containing 1X PCR reaction buffer (Invitrogen), 2

32 mM MgCl2 , 2 pmol P-labeled SF3423 primer, and 0.2 U Taq DNA

polymerase (Invitrogen). The amplification cycle consisted of

denaturation at 95°C for 2 min, followed by 30 cycles of denaturation at

95°C for 30 s, annealing at 50°C for 30 s, extension at 72°C for 30 s.

Labeled DNA was separated on 6 % polyacrylamide gels in 1X TBE buffer with 8 M urea. A and G sequencing ladders were produced from fragment

A by PCR cycle sequencing using dideoxyterminators from the fmolTM

Cycle sequencing kit (Promega) and run alongside the DNase I digestion patterns. Gels were dried under vacuum at 80°C and exposed to X-ray films at –80°C for several days.

196 4.2.9 β-glucuronidase enzyme assays

β-glucuronidase (GUS) activity was assayed fluorometrically using 4-

methyl-umbelliferyl-β-D-glucuronide as described by Jefferson (1987),

but the assays were scaled down to fit microtiter plates and analyzed

using a Victor 1420-2 multilabel reader (PE-Applied Biosystems). Net

GUS activity of each strain was corrected for the basal fluorescence of P.

stewartii DC283 or E. coli SΦ200 carrying pPL6GUSC without an insert.

-1 -1 - Specific activity was expressed in GUS units (pmol MU min OD600 ml

1 of culture at 37°C).

4.2.10 Total RNA isolation

Total bacterial RNA was isolated from 5 ml of overnight cultures of

cells grown in IM pH5.5 liquid medium or 5 ml of exponentially growing

cultures in LB broth. Cells were pelleted and resuspended in 0.1 ml of

TEN100 buffer (10 mM Tris-HCl, 100 mM NaCl, 1 mM EDTA, pH 7.0).

Five volumes of Trizol (Invitrogen) at 65°C was added to the resuspended

cells, which were then vortexed, shaken at 65°C for 10 min and

centrifuged at 10,000 x g for 30 min at 4°C. The upper aqueous phase was

extracted with 0.1 ml of water-saturated chloroform and the RNA was

precipitated by adding one volume of isopropanol and centrifuging as

above. RNA was resuspended in 50 µl of DEPC-treated water and treated

with DNase I (Invitrogen) according to the manufacturer’s instructions.

197 4.2.11 Primer extension and DNA sequencing

Analysis of the 5’-ends of transcripts was performed by primer

extension (Sambrook et al., 1989). Primer PEL (Table 4.2) was labeled

with [γ32P]ATP and cDNA synthesis was carried out using M-MLV reverse transcriptase (Promega) with 50 µg of RNA per reaction following

the manufacturer’s instructions. DNA sequencing was performed with the

same primers used for cDNA synthesis following the fmolTM cycle sequencing protocol from Promega. The labeled extension products were visualized by autoradiography with Kodak XOMAT films after electrophoresis in 6% denaturing acrylamide gels at 55 Watts in a Bio-Rad sequencing apparatus (50 x 22 cm gel plate). For weakly labelled cDNAs, a second autoradiography exposure was performed, and the DNA ladder and cDNAs were merged electronically using Photoshop Elements

(Adobe).

4.2.12 Sequence analysis and modeling

DNA and protein sequences were aligned using ClustalX. Modeling of the HrpY tertiary structure was performed using a homology approach with the Swiss-Model algorithm at www.expasy.ch. The model was

checked for structural errors with the protein verification tool

WhatCheck. DNA curvature propensity and bendability was calculated

using Bend-It at www.icgeb.trieste.it.

198 4.3 RESULTS

4.3.1 Deletion analysis of the regulatory elements activated or

repressed by HrpY

To identify the sequences required for activation of hrpS and

repression of hrpL by HrpY, we constructed nested deletions of the

corresponding 5’-regulatory regions of each gene by PCR and fused the

resulting promoter fragments to the reporter gene uidA. The region

between hrpY and hrpS spans 844 nucleotides and includes a 483-bp-long

IS-like element located at –23 bp from the hrpS start codon. Primer

extension experiments described in Chapter 3 revealed a transcript

starting 106 nt from the 5’ end of the IS element (Fig. 3.4) with a σ 70 promoter nearby. We therefore focused our attention on only the region upstream of the IS-like element.

The DNA sequence of the region located between the end of hrpY and the start of the IS-like element (region +95 to –231 from the transcription start) in P. stewartii was aligned with the corresponding regions from P. agglomerans pv. gypsophilae, E. amylovora and Pectobacterium chrysanthemi to identify conserved and divergent sequence motifs. Cluster analysis grouped the sequences according to phylogenetic relationships

(not shown), with highest similarity between P. stewartii and P. agglomerans. Several blocks of highly conserved sequences were found in the most 5’ part of the region (here referred as CUR, conserved upstream region), followed by a more divergent short region (VUR = variable

199 upstream region) and then a putative IHF site (Fig. 4.3). Downstream of

the IHF site, two direct repeats were conserved among pantoeas (PDR =

Pantoea direct repeats). These were followed by a putative, highly

conserved σ 70-promoter, centered at –24 bp from the mapped transcription

start (Fig. 4.3). The curvature propensity and bendability of the

P.stewartii hrpS promoter region was also analyzed using the Bend-It algorithm (Fig. 4.4). Two peaks of potential intrinsic curvature were found around positions –129 and -147.

We tested in first approximation the function of these hypothetical elements by deletion analysis and construction of the reporter plasmids

described in Fig. 4.1. The plasmids were tested in wild-type and

hrpY[D57N] genetic backgrounds after growth in IM5.5. Plasmids

containing the entire hrpS 5’ regulatory region, such as pMM50 and

pMM400 (Fig. 4.1), required HrpY for activation. Removal of the region

from –165 to –231 (comprising the CUR) totally abolished hrpS

expression in both genetic backgrounds. Further deletion of the 5’ regions comprising the VUR (plasmid pMM393; Fig. 4.1) restored some

measurable basal activity, but the gene fusion showed much lower activity

compared to pMM400 or pMM50, and the expression was HrpY-

independent. A similar result was observed for pMM403, which carries a

further deletion spanning the two PDRs (Fig. 4.1) but contained the σ 70-

promoter. Plasmid pMM392 did not contain the σ 70-promoter and showed

even lower basal expression levels. 200 A detailed sequence analysis of the hrpL promoter region of P. stewartii is described in Chapter 2. Two hrpL reporter plasmids, pMM191 and pDM2785 were constructed for the studies described in Chapter 2 and

3. pMM191 lacks the UAS element required for activation by HrpS (see

Chapter 3). The plasmids were moved into E. coli SΦ200, the strains were grown in a rich medium and then the fusions were tested for a response to pMM46, which expresses hrpY+. The results shown in Fig. 4.2 confirmed the ability of hrpY to repress the full-length hrpL promoter (plasmid pDM2785; see also Chapter 2). A similar result was obtained with pMM191, which contains only the region up to –95 bp, without the UAS.

This implies that HrpY does not repress hrpL basal expression by interfering with the binding of HrpS (in P. stewartii) or its paralog NtrC

(in the E. coli SΦ200 reconstruction system) to the conserved UAS. While performing the primer extension experiments described in Chapter 3, we observed the presence of a second transcript starting at –58±1 from the

σ 54-promoter transcription start (Fig. 4.5). This transcript was not consistently found in all P. stewartii RNA preparations, but it was made in rich media. This implies the existence of a tandem promoter upstream of the σ 54-box. Sequence analysis using the NNPP algorithm indeed identified a putative σ 70 promoter centered at –68 (Fig. 4.5). We hypothesized that this promoter was responsible for the basal levels of hrpL expression and that it was repressed by HrpY. A third reporter plasmid, pMM428, containing the UAS element, but lacking the region 201 downstream of the σ 54-promoter, was constructed (Fig. 4.2). This hrpL- uidA fusion had a lower basal expression compared to pDM2785, and was no longer repressed by HrpY. Instead, a reproducible increase in hrpL expression was observed (Fig. 4.2). These results imply that sequences important for the repression of PhrpL by HrpY reside downstream of base –

12.

4.3.2 Overexpression and purification of HrpY proteins

HrpY and HrpY[D57N] were overexpressed using the pET-15b

expression plasmid. Inserts, produced by PCR, were engineered to maintain the correct translation frame with a vector-encoded N-terminal polyhistidine tag (ca. 2 kDa in size). To show that the his6 tag did not

affect biological activity, the his6 -hrpY gene was subcloned into pRK415.

Although this plasmid had an adverse effect on the growth of P. stewartii transconjugants, it weakly complemented a hrpY null mutant (data not shown). We decided to use the recombinant form of the protein for further studies due to the convenience of purification. Induction of the T7lac promoter of plasmids pMM221 (hrpY) and pMM222 (hrpY[D57N]) in E. coli BL21(DE3)(pLysS) by IPTG resulted in the production of 26-kDa polypeptides, which were not present in uninduced cells. (For HrpY see

Fig. 4.6A, lane 2 vs. 1; HrpY[D57N] is not shown). Fig. 4.6 (A to C) shows recovery of His6 -HrpY after various steps in its purification by

immobilized metal affinity chromatography using step-elution gradients

202 (Fig. 4.6 B) and subsequent anion-exchange chromatography of the

semipurified preparation in a linear salt gradient. Highly purified His6 -

HrpY was recovered from the 136 to 250 mM NaCl fractions (Fig. 4.6 C).

In some cases, ion-exchange chromatography was not necessary to obtain

highly purified HrpY, especially if washing buffers with higher concentrations of imidazole (60 to 80 mM) were used before the final elution step (100 to 150 mM). In general, the purity of the samples was assessed by densitometric analysis of Coomassie-stained SDS-PAGE gels.

Samples were also tested for the absence of background kinase activity using [γ32P]ATP. In all cases purity was > 95% and kinase activity was

not visible after autoradiography for 24 h. The average yield was of 8 mg

of purified protein per liter of culture. The protein stocks had

concentrations ranging from 150 to 350 µM in storage buffer and were

kept at –20°C until use.

The molecular weight of the polypeptides was confirmed by Q-TOF

electro-spray ionization mass spectrometry (Fig. 4.7). For the HrpY

sample, the major protein species had a molecular weight of 26051 ± 58

Da (Fig. 4.7A-B), whereas HrpY[D57N] was 26122 ± 76 Da (Fig. 4.7C-

D), both in close agreement with the expected mass of 26.1 kDa. Finally,

western blot analysis using anti-His6 monoclonal antibody confirmed the

identity of the recombinant polypeptides (not shown).

203 4.3.3 Phosphorylation of HrpY at D57 by Salmonella enterica BarA198

Response regulators are usually phosphorylated at a single Asp residue in the receiver domain, even if secondary phosphorylation sites are sometimes present (Delgado et al., 1993). The tertiary structure of HrpY was modeled using a homology approach based on the experimental crystallographic data from NarL and Spo0F (Fig. 4.8A) crystals or solution structures as templates (Protein Data Base files 1404.pdb,

1RNL.pdb, 1FSP.pdb, 1NAT.pdb). An E-score of 0.0001 was needed for the ProMod II algorithm to construct an approximated model of HrpY.

The model was energy-minimized with GroMol96 (Structure free energy =

-5742.091 KJ/mol) and the structure was checked with the What-If algorithm for detecting major structural problems. Modeling was successful from residue 3 to 206 out of the 216 making up the HrpY polypeptide. Several side-chain clashes and improper dihedral angles were found, but they did not involve the catalytic site (data not shown). The model predicts that three conserved aspartyl residues (D11, D12 and D57) at the top of a β–strand form the conserved catalytic triad (Fig. 4.8B).

Mutagenesis at D57 was simulated with SpdbViewer v. 7.0. Substitution of D57 with N, A or E and selection of the most energetically favorable rotamer did not cause clashes with nearby residues, implying that major structural changes are not expected upon mutagenesis (not shown).

We decided to directly test if D57 was the sole phosphorylation site in the response regulator. The putative cognate sensor kinase for HrpY is

204 encoded by HrpX because they are in the same operon and null mutations

in HrpX alter the Hrp phenotype (as shown in Chapter 2). Attempts to

purify sufficient amounts of active full-length HrpX for reproducible in

vitro trans-phosphorylation assays repeatedly failed (See Chapter 5).

Instead we used recombinant Salmonella enterica BarA198 (a truncated

version of BarA engineered with a N-terminal polyHis tag) for our

experiments. We selected this kinase because it was available from B.

Ahmer lab, a BarA ortholog (GacS) is present in all erwinias, and SirA

(the cognate response regulator for BarA) belongs to the same class of

regulators (FixJ) as HrpY. Fig. 4.9 shows that HrpY alone did not

catalyze an autophosphorylation reaction in the presence of [γ32P]ATP. On

the other hand, BarA198 autophosphorylated in presence of [γ32P]ATP

(Fig. 4.9, Lane 2) and phospho~BarA198 efficiently transferred 32P to

HrpY (Lane 3). The HrpY/BarA198 complex appeared more active than

BarA198 alone in catalyzing the autophosphorylation reaction, at least under the conditions used here (molar ratio kinase:response regulator monomers = ca. 3:1). Finally, the HrpY[D57N] mutant protein was not phosphorylated in presence of BarA198 and [γ32P]ATP (Lane 4),

demonstrating that D57 is the only phosphorylation site in HrpY.

Complete kinetic studies of the phosphotransfer and dephosphorylation

reaction were not conducted because we do not know whether or not the

205 use of BarA is a good model for the interaction of HrpX with HrpY. More information on Salmonella BarA biochemistry was recently published

(Teplitski et al., 2003)

4.3.4 Binding of HrpY to 5’ hrpS DNA fragments

The binding of HrpY to various DNA fragments spanning the hrpS regulatory region located between the IS-like element and hrpY was studied by electrophoretic mobility shift assays. Again, we focused on this region because primer extension and genetic experiments indicate that it contains the hrpS promoter. The fragments were synthesized by PCR and end-labeled with 32P. Their map positions are shown in Fig. 4.10.

Increasing concentrations of unphosphorylated HrpY (≤625 nM) were incubated with 25 fmol 32P-labeled fragment A (327 bp) for 40 min and then the protein-DNA complexes were separated from the free probe by electrophoresis in native 6% acrylamide gels (Fig. 4.11). A peculiar shift pattern was observed, with complexes of intermediate mobility migrating between the free probe and the samples with maximal concentration of

HrpY. The intermediate complexes formed wavy bands and may have been produced upon dissociation of the binding complex at lower HrpY concentrations. This has been reported in other systems (Dahl et al.,

1997). A similar result was obtained with the HrpY[D57N] variant (data not shown). All the reactions were performed in a large molar excess of the aspecific competitor poly(dI:dC). In a specificity control, binding of

206 HrpY to the hrpS fragment (A probe) was dramatically reduced by competition with a 100-fold molar excess of cold fragment A (Fig. 4.11).

These results demonstrate specific binding of HrpY to the full-length hrpS

promoter region.

To identify the HrpY-binding sites in fragment A (coordinates +95 to –

231), PCR probes carrying different portions of the hrpS promoter region

were synthesized and end-labeled with 32P. Binding assays were

performed with: probe B (-165 to –231) containing the CUR; probe C1 (-

90 to 165) carrying the CUR; probe C2 (-5 to –90) containing the DRs and

the σ 70-promoter; probe C3 (+95 to –5) containing the 5’UTR upstream of the IS-like element; and probe C (+95 to –165) that spanned C1, C2 and

C3 (Fig. 4.10). In each assay, 10-15 fmol of probe was incubated with 5

µM of HrpY with or without cold specific competitor DNA. HrpY formed high affinity complexes with fragment C, similar to what observed for probe A (Fig. 4.12). Significant amounts of fragment C2 were also retarded by HrpY and three complexes were observed. In a few experiments, fragment B was weakly bound by HrpY, but the sites must have had low affinity, because this binding was eliminated by competition with cold fragment C. Fragment C1 did not bind HrpY, whereas occasional binding was shown for probe C3.

To demonstrate the specificity of the binding of HrpY to fragment C, excess unlabeled fragments B, C1, C2 and C3 were incubated with HrpY and 32P-labeled probe C (Fig. 4.13). Up to 250-fold molar excess of 207 unlabeled fragments B, C1 and C3 did not block the binding of HrpY to

probe C, whereas binding was reduced in the control assay using cold

fragment C (Fig. 4.13). Interestingly, the use of cold C2 fragments caused

a supershift of the HrpY-probe C complex, suggesting the formation of

HrpY oligomeric complexes exposing multiple DNA-binding surfaces.

4.3.5 Binding of HrpY to 5’ hrpL DNA fragments

The binding of HrpY to the hrpL regulatory region was also studied by

electrophoretic mobility shift assays. Six DNA probes, spanning the

region from +221 to –148 (coordinates are relative to the transcriptional

start site of the σ 54-promoter), were synthesized by PCR (Fig. 4.14).

Probes were end-labeled with [γ32P]ATP, incubated with 20 µM HrpY in the presence of 1 µg of poly(dI:dC), and complexes at equilibrium were analyzed by native acrylamide gel electrophoresis (Fig. 4.15). HrpY retarded the mobility of probes L (+221 to –148) and M (-12 to –148), and to a lesser extent it bound to probes I (+221 to –94), H (+221 to –71), G

(+221 to –29). However, only a very small amount of probe F (+221 to –

11) was retarded. Titration experiments to determine the first approximation Kd were performed with the full-length hrpL promoter

(probe L). These results indicate that there is at least one strong operator site for HrpY, located between –12 and –29, and possibly a second weaker site(s) downstream of base -12. The specificity of the binding of HrpY to probe L was evaluated by competition experiments using specific and non- 208 specific competitor DNA (Fig. 4.16). As expected for specific binding, a

150-fold excess of poly(dI:dC) did not compete for the binding of HrpY to

probe L in the absence of poly(dI:dC), but a similar molar excess of cold probe L DNA did. The specificity of complex formation with the subfragments of probe L was not evaluated.

4.3.6 Effect of phosphorylation on in vitro HrpY binding activity

The effect of phosphorylation on the first approximation affinity of

HrpY for the target promoter sequences was analyzed by gel shift assays.

Probe C (Fig.C) was used to study the affinity for the hrpS promoter.

HrpY was phosphorylated either by BarA198 or by 50 mM carbamoyl phosphate, which is a small phosphodonor molecule often used as an alternative to acetyl phosphate. Increasing concentrations of phosphorylated (P) or unphosphorylated HrpY (from 5 nM to 80 µM) were incubated with a constant amount of probe C at nanomolar concentrations.

Protein-DNA complexes were separated on a 6% acrylamide gel and the bound fraction (ν) was measured by quantifying the amount of free probe

with a phosphor-imager scanner and comparing it to the no-protein

controls. The difference was used to calculate the bound fraction. Raw

data were transformed to linearize the binding isotherm and plotted in a

Scatchard graph (ν vs. ν/P) (Fig. 4.17A) and the apparent Kd was

calculated as the Y-axis intercept of the linear regression curve. When

HrpY was phosphorylated by BarA198, the Kd for binding to probe C was 209 decreased ca. 12-fold (from 20 nM to 1.7 nM). Similar changes in Kd

were obtained using carbamoyl phosphate. (Data not shown). These

results indicate that, at least in vitro, phosphorylation of HrpY can increase its affinity for target sites in the hrpS promoter region.

Similar experiments were done to measure the affinity of HrpY for the the hrpL promoter (probe L). However, it was necessary to use greater

concentrations of HrpY to obtain binding. The results in Fig. 4.17B show

that the Kd for binding to PhrpL is about three orders of magnitude higher

than that for binding to PhrpS. Interestingly, phosphorylation did not alter

the affinity of HrpY for its target sites in probe L.

4.3.7 Effect of phosphorylation of HrpY on its in vivo activity

The effects of conservative and structurally neutral substitutions at D57

in HrpY on hrpS expression were studied in P. stewartii. pRK415 plasmid

derivatives carrying Plac-hrpY[D57A] and Plac-hrpY[D57N] alleles were constructed as described in Chapter 2 (plasmids pMM74 and pMM118, respectively). A new hrpY mutant with a D57E substitution was engineered in vitro, following the SOE-PCR procedure described in

Chapter 2, and cloned into pRK415 to produce pMM99. The plasmids were mobilized into P. stewartii DM733, a strain carrying a chromosomal hrpS::lacZ fusion in a null hrpY background. The resulting strains were grown in IM5.5 and ß-galactosidase was measured (Fig. 4.18). Strain

+ DM733(pMM52), expressing Plac-hrpY , was used as positive control. The

210 HrpY[D57A] and HrpY[D57N] proteins were still able to activate hrpS, but its expression was only 33% and 25% that of the wild-type.

Interestingly, the HrpY[D57E] protein was completely unable to upregulate hrpS-lacZ.

4.3.8 Identification of the binding elements of HrpY by DNase I footprinting

The locations of the HrpY-binding sequences were identified by testing the ability of unphosphorylated HrpY to protect against DNase I cleavage of the of the hrpS promoter. Linear PCR with a 32P-labeled primer was used to amplify the partially digested complexes. Bottom strand-labeled fragments have not been tested yet. The experiments have not been repeated yet and the following results must be considered with caution.

Probe A (Fig. 4.10) was used for these experiments. Double autoradiography exposures were needed because the DNase I digestion ladders with the protein were underloaded (Fig. 4.19). A readable digestion ladder was obtained for sequences with coordinates +10 to –182.

Therefore most of the regions corresponding to probes B and C3 (Fig.

4.10) were not analyzed. Decreased DNase I sensitivity was shown for regions –50 to -71, -78 to –102, -125 to –165 and –174 to –182 (Fig. 4.19 black boxes). Hypersensitive sites were located around –50, -71, -104, -

106, -111, -123, and –138 (Fig. 4.19, solid arrows). These sites may indicate changes in the local bending angle of the hrpS promoter upon

211 binding of HrpY. The sequences protected by HrpY varied greatly.

Consequently chemical reagents with greater accessibility to the stacked

bases and/or the phosphate backbone of the DNA will be required to

identify a clear binding site. A tentative consensus site, MAATTYY,

shared by all the footprints is proposed (vertical arrowheads in Fig. 4.6).

This consensus exhibits dyad symmetry in the footprint located between –

78 to –102 compared to footprint at –165 to –148. When confirmed, these

results should support the specific binding observed with probe A and the

formation of three complexes between probe C2 and HrpY, but they may

contradict the absence of a mobility shift with probe C1. Potential

changes in the protection patterns upon binding by HrpY~P, which may be

critical in establishing a model for HrpY-dependent regulation of hrpS,

were not determined.

4.4 DISCUSSION

Data presented in this chapter provide the first detailed biochemical

analysis of the interaction of a hrp regulatory protein with its target

promoters. In particular, we provide evidence supporting the role of HrpY

as both an activator of hrpS and a repressor of hrpL. Possible models of

regulation are shown in Figures 4.20, 4.21 and 4.22.

As shown in Chapter 2 and Fig. 4.1, activation of hrpS is hrpY- dependent. Deletion of region –165 to –231 (corresponding to fragment B in Fig. 4.10) renders the hrpS-uidA fusion in pMM396 totally repressed

212 compared to the full-length fusion (e.g. pMM50 and pMM400). This

implies that DNA sites in fragment B, alone or in concert with other DNA

elements, are critically important for the activation of hrpS by HrpY. This is also consistent with the previous observation that hrpS loci lacking

fragment B are unable to complement null hrpS mutants, unless the gene

is driven by a strong vector promoter, such as Plac in pRF205 (Frederick

et al., 1993).

Electrophoretic mobility shift (EMS) assays were performed using

purified recombinant HrpY because crude cell extracts from P. stewartii

or E. coli showed nonspecific activity in negative control reactions (i.e.

null mutant strains or strains carrying the vector alone). This suggests

that bacterial proteins other than HrpY are able to bind to the probes (data

not shown). EMS assays showed weak, inconsistent binding of HrpY to

sites in fragment B (-165 to –231), no binding to fragment C1 (-90 to –

165), weak, erratic binding to C3 (+95 to –5) and stronger binding,

possibly at multiple sites, to fragment C2 (-5 to –90) (Fig. 4.10 and Fig.

4.12). Interestingly, probes A and C have much greater affinity for HrpY than the individual subfragments (approximate Kd <1 nM and 20 nM, vs.

micromolar ranges, respectively), as estimated visually from the bound

fractions in Fig. 4.11 and Fig. 4.12 or by Scatchard plot analysis (Fig.

4.17). This may be due to multiple discrete sites of interaction acting cooperatively. The fact that competition experiments using fragment C as

a probe and fragment C2 as the cold competitor created a supershift may

213 be interpreted as oligomerization of HrpY and interaction of some of the

subunits with non-contiguous DNA fragments. This also suggests that a

looping mechanism may be involved in hrpS activation in vivo (Fig.

4.21). Our preliminary DNase I footprinting experiments seem to indicate

that HrpY binds at multiple sites in the hrpS promoter region. These

experiments should be interpreted with caution at this point because we

have only examined the top strand protection and have not repeated the

experiment. Analyzing the top strand of probe A identified five footprints.

These sites span sequences corresponding to probes C2 and C1 and part of probe B (whose footprints were not well resolved by the experiments performed so far) (Fig. 4.19). Interestingly, one of the putative protected sites was split between probe C1 and probe C2, which may explain why the gel shift with probe C1 was negative (Fig. 4.12). Also, probe C2 contained two of the DNase I footprints and this finding is consistent with the multiple complexes seen in EMS gels. It is unclear at this point whether the weak binding to fragments B and C3 is specific or non- specific. The MAATYYY sites found in the regions protected by HrpY are quite degenerate, but for other FixJ-class response regulators, similar A+T rich sites have been reported, e.g. RAAAYY for UhpA (Dahl et al., 1997) and TACYNMT for NarL (Tyson et al., 1994). A more comprehensive footprinting analysis involving the use of chemical probes for major/minor groove and backbone interactions will be necessary to precisely define a clearer consensus binding site, especially in A+T rich

214 regions. Complex binding patterns at multiple sites were likewise

observed for E. coli NarL, a FixJ-class regulators, and OmpR, a “winged

helix” class regulator. In particular, NarL dimers (Maris et al., 2002) bind

four heptamers located in front of the fndG operon and to eight heptamers

in front of the narG operon in a cooperative and hierarchical manner (Li

et al., 1994). There may be different orientation and spacing among

heptamers in the various promoters regulated by NarL, but cooperative binding usually requires a central site with a 7-2-7 dyad symmetry

(Darwin et al., 1996). Another FixJ-class regulator, E. coli UhpA, binds as dimers to inverted repeats of the uhpT promoter, which were mapped by iron chelated hydroxy-radical footprinting (Olekhnovich and Kadner,

2002). On the other hand, asymmetrical OmpR dimers bind, in a cooperative fashion, multiple tandemly arranged 10-bp sites (Harrison-

McMonagle et al., 1999).

One possible working model consistent with the genetic and biochemical data invokes a repressor of hrpS acting on sequences located in fragment C1 (Figs. 4.20 and 4.21). This repressor would be required for maximal downregulation of the hrpS promoter. Its existence is hypothesized mainly to explain the observation that plasmid pMM396, containing the hypothetical operator site, has absolutely no promoter activity when compared to plasmids pMM393, which exhibits a low basal level, and pMM400, which is normally regulated by HrpY (Fig. 4.1). The absence of transcription termination demonstrated in Chapter 3 also calls

215 for a mechanism to downregulate read-through transcription of hrpY into hrpS under Hrp-repressing conditions. In this model, HrpY would then act as an antirepressor by binding to weak affinity sites in fragment B and to higher affinity sites in fragments C2 and C1, as shown by EMS assays

(Fig. 4.12) and the preliminary DNA footprintings (Fig. 4.19), respectively. The weak sites in fragment B have not yet been formally shown by footprinting but are suggested by the weak binding of HrpY to this probe. HrpY may therefore activate hrpS by forming a DNA loop involving both enhancer sites (at least three or four sites) and the site near the σ 70 promoter (Fig. 4.20). The presence of an IHF site

(AtgtAAccgttTTt) centered at –118 in fragment C1 further supports this looping model. IHF belongs to a class of DNA bending proteins that are able to impose strong curvatures (up to 180°) in A+T rich DNA sequences

(consensus WATCAAN4 TTR). Moreover, computer modeling suggests that

the hrpS promoter has as intrinsic curvature in the absence of any bound

factor due to the low G+C content and its estimated bending angle is 100°

(Fig. 4.4). In E. coli, IHF has also been shown to act as a repressor of

NarL-regulated genes (Browning et al., 2000). By analogy to the E. coli

Fis anti-activator (Browning et al., 2000), the hypothetical repressor

(perhaps IHF itself) may act upstream of the hrpS promoter and HrpY

would be the equivalent of the NarL antirepressor. (Wu et al., 1998).

Looping has been hypothesized for NarL dependent regulation of narK

(Kolesnikow et al., 1992), but probably it occurs via the formation of 216 higher order complexes of multiple transcription factors more than by

simple DNA bending (Dong et al., 1992). Alternatively, one might hypothesize that a second site, downstream of the σ 70 promoter, is

required for repression, which is typical of repressors in general (Wagner,

2000). This repression might likewise involve the formation of a DNA

loop (Fig. 4.21). Deletion of fragment B (-231 to –165) and C1 (–90 to –

165) would affect both the repressing mechanisms and the HrpY-

dependent activation of hrpS, as observed in Fig. 4.1 with plasmid

pMM393. The presence of repressor-binding sites in multiple upstream

positions has been described for the ompF gene in E. coli. In this system

both the –42 to –52 site and the –360 to –380 sites are important for

negative regulation of ompF, while sites in the –70 to –100 region are

important for activation of ompF by OmpR (Tsung, et al., 1989; Mizuno

et al., 1988; Ostrow et al., 1986). Both hypotheses are plausible, but other

explanations could be possible. Further research is needed to clarify the

mechanism used by HrpY. For example, in vitro transcription studies

and/or electron microscopy analysis of the hrpS promoter region using

purified HrpY and fractionated P. stewartii cell extracts may directly

demonstrate the existence of the repressor and the loops. Linker-scanning

or SOE-PCR mutagenesis could also be attempted in order to substitute or

delete the hypothesized operator site, ultimately confirming the DNA

binding data.

217 HrpY was also shown to repress transcription of hrpL (Fig. 4.22). The

basal expression of this gene was repressed about 7-fold by HrpY in an E.

coli reconstituted system (Fig. 4.2; pDM2785) and this repression did not require the UAS element recognized by HrpS (Fig. 4.2; pMM191).

Deletion analysis of the regulatory region of hrpL is not complete, but our

data to date suggest that an operator site downstream of base –12 is

required for the repression by HrpY (Fig. 4.2; pMM428). The

upregulation observed with plasmid pMM428 in the presence of hrpY is

not understood at this point. These data partially agree with the gel shift results, in that at least two sets of binding sites could be recognized: a high affinity site located between –29 and –12 and a potential weaker site located downstream of base –12. The specificity of the binding of HrpY to the latter site was not tested and the weakly shifted complex shown in lane two of Fig. 4.15 needs to be confirmed by further analysis. These two sites may act in concert and their affinity for HrpY may be decreased when they are located in separate probes. DNA footprinting of this region should provide more definite information on the location and specificity of the binding sites protected by HrpY and determine whether or not this is an overexpression artifact. What is notable is that the affinity of HrpY for the hrpL promoter is about 1000-fold lower than for the hrpS promoter

(Fig. 4.17). Assuming that binding constants measured in vitro are relevant to what happens in vivo, repression of hrpL would only occur

when HrpY reaches ca. 20,000 to 50,000 copies per cell, as compared to

218 the 20 to 50 copies needed to activate hrpS. From a biological point of view, hrpY may have acquired the additional function of fine-tuning the level of HrpL, a critical sigma factor coupling the transcription of all the type III secretion genes and their effectors. By indirectly activating hrpL via HrpS or directly repressing it, HrpY could quickly adapt the Hrp response to the environmental conditions sensed by HrpX. In P. stewartii, the HrpY regulator may also need to repress hrpL to act as a brake or the governor for the unique autoregulatory loop under non-inducing conditions. However, for this model to work, HrpY would also have to switch to repressing hrpS at the elevated concentrations needed for it to repress hrp. This mechanism of activation/repression would be similar to the regulation of ompF by OmpR (Rampersaud et al., 1994).

This study also explored the role of phosphorylation in modulating the activity of HrpY as a transcription factor. Protein phosphorylation gels

(Fig. 4.9) showed that the D57 residue is probably the sole phosphorylation site. More sensitive techniques, such as phospho-aspartyl derivatization followed by MS/MS spectrometry (Sanders et al., 1992), were not attempted. For other response regulators, phosphorylation affects their activity as transcription factors in various and often overlapping ways; it may change their quaternary structure, by promoting dimerization or oligomerization (Han and Winans, 1994; Li et al., 1994; Fiedler and

Weiss, 1995; Mettke et al., 1995), their DNA binding affinity to single or multiple cooperatively-acting sites (Galinier et al., 1994; Rampersaud et

219 al., 1994; Huang and Igo, 1996; Dahl et al., 1997; Huang et al., 1997;

Bergstrom et al., 1998). Our results showed that phosphorylation of HrpY increases its affinity for DNA sites in the hrpS promoter (Fig. 4.17 A), but not in the hrpL promoter. More precise assessments of the binding affinities using filter-binding assays or electron-plasmon resonance were not pursued. Consequently, we consider the Kds calculated from Scatchard plots as “first approximation” constants. The importance of D57 in transcription activation was also tested in vivo using unphosphorylatable hrpY[D57] alleles expressed from multicopy plasmids. The reduction in hrpS activation due to the hrpY[D57N] and hrpY[D57A] mutations, as compared to wild type hrpY, are probably due to the effect of these mutations on the Kd , or, in other words, to the inability of HrpY[D57N] to oligomerize at the hrpS promoter to the same extent as HrpY~P. However, it could alternatively reflect the inability of the unphosphorylated form to fully activate the RNA polymerase holoenzyme. Genetic screening for positive control mutations in hrpY, i.e. those affecting activation but not

DNA binding, together with screenings for suppressing mutations in components of the RNA polymerase holoenzyme may be helpful in separating these two effects more clearly (Aiba et al., 1994). In this regard, the hrpY[D57E] mutant may be an interesting mutant to study in more depth with biochemical approaches because molecular modeling predicts it should produce a correctly folded protein as competent for

DNA binding as HrpY[D57N], but yet, it is unable to activate hrpS, even

220 when overexpressed. In other systems, similar substitutions at the

phosphrylation site render the response regulator constitutively active

(Klose et al., 1993).

Salmonella enterica sensor kinase truncated mutant BarA198 (Teplitski

et al., 2003) was used to phosphorylate HrpY in vitro, both in protein labeling experiments and in gel shift assays. BarA orthologs are found in many γ-proteobacteria, including the genera Pseudomonas, Pantoea,

Erwinia and Pectobacterium. Chatterjee and coworkers (Cui et al., 2001) used low stringency Southern blot hybridization to show that P. stewartii also has a barA/gacS-like gene. Usually there is a certain degree of specificity between response regulators and their cognate kinases in order to achieve proper regulation (Hoch and Varughese, 2001). In particular, E. coli BarA is quite specific for its cognate UvrY regulator (Pernestig et al., 2001). Whether the phosphorylation of HrpY by BarA198 described in this study is an artifact of the elevated protein concentration used in vitro or it has physiological relevance in P. stewartii cells is not known. It is worth noting that BarA/SirA and GacS/GacA have important roles in virulence regulation of salmonellae, erwinias and pseudomonads.

Small phosphodonors can phosphorylate response regulators in vitro and in vivo (Lukat et al., 1992). In this study, carbamoyl phosphate was shown to modify the affinity of HrpY for hrpS promoters and this was interpreted as an indication of its ability to phosphorylate HrpY. Acetyl phosphate did not show the same effect (not shown). Formal

221 demonstration of this reaction using radiolabeled carbamoyl phosphate was not attempted because we did not have the means to synthesize it.

Future studies on the alteration of HrpY footprints at the hrpS and hrpL promoters might use this convenient and cheap small phosphodonor instead of recombinant kinases.

In conclusion, the dual role of HrpY as an activator of hrpS and a repressor of hrpL was supported in this study by independent approaches.

However, the biological significance of and the inherent logic behind its role as a repressor remains somewhat puzzling and open to different interpretations. Our preliminary data on the interaction of HrpY with specific sites in the hrpS promoter represents an important step toward the future in vitro analysis of other global and specific transcription factors in controlling this key step of the hrp regulatory cascade. These factors include: a) the DNA-bending protein IHF; b) the quorum sensing response regulator EsaR; c) the DNA-binding protein H-NS, which we have shown binds to probe C (Merighi and Coplin, unpublished data); and d) other novel regulators, such as a NikR-homolog, a member of the ribbon-helix- helix family of transcription factors recently implicated in the negative regulation of hrp genes (Ham and Coplin, unpublished data).

222

4.5 SUMMARY

Pantoea stewartii subsp. stewartii is a bacterial pathogen of corn,

whose pathogenicity depends on translocation of effector proteins into host cells by the Hrp/type III secretion system. Using genetic approaches, we delineated a regulatory cascade that controls the hrp/hrc secretion and wts effector genes in response to environmental signals and various global regulators. The cytoplasmic HrpX/HrpY PAS kinase/response regulator pair senses an unknown internal signal resulting in the phosphorylation of

HrpY. The most likely phosphorylation site in HrpY is a conserved D57 residue that is required for virulence. HrpY activates hrpS transcription

and HrpS, in turn, enhances transcription of hrpL. At the end of the

cascade, HrpL is an ECF alternative sigma factor that activates the genes

in the Hrp regulon.

In this study, we characterized the regulatory elements in the hrpS and

hrpL 5’ regions using primer extension analysis, electrophoretic mobility

shift assays and a series of overlapping subclones creating uidA reporter

gene fusions. Progressive deletions from the 5’ end of the hrpS promoter

region first abolished transcription and then restored it, but it was no

longer HrpY-dependent. This finding is consistent with the deletion of an

operator for an unknown transcriptional repressor and a model in which

HrpY acts as an antirepressor. A similar analysis of the hrpL promoter has

yet to be completed, but our experiments to date suggest the presence of

223 an operator required for HrpY-dependent repression downstream of the

54 σ promoter. Purified, unphosphorylated His6 -HrpY bound specifically to portions of the 5' regulatory regions of hrpS and hrpL in gel retardation assays. In vitro phosphorylation of HrpY increased its binding affinity for the hrpS promoter, but not for the hrpL promoter. Preliminary electrophoretic mobility shift assays and DNase I footprinting experiments suggest that HrpY may bind to multiple sites within each promoter region.

224 Strain or plasmid Relevant phenotypes and genotypes Source or reference

Escherichia coli

DH10B F- mcrA ∆(mrr-hsdRMS-mcrBC) φ80dlacZ∆M15 ∆lacX74 endA1 recA1 deoR

∆(ara,leu)7697 araD139 galU galK nupG rpsL λ- Invitrogen

- - - r HB101 F thi-1 hsd20 (r Bm B) sup E44 recA13 ara-14 leuB6 proA2 lacY1 rpsL20 (Sm )

xyl-5 mtl-1 Boyer and Roulland-Dussoix (1969)

Sφ200Rif metB strA purB ∆(agg-uidA-man) Rifr Wei et al. (2000)b

BL21(DE3)/pLysS F- ompT hsdSB (rB- mB-) gal dcm (DE3) caarying pLysS Novagen

P. stewarti subsp. stewartiii

DC283 Wild type, Wts+, HR+ Nalr Coplin et al. (1986)

DM064 DC283 hrpY1296::Tn5 (Kanr) Frederick et al. (2001)

MM254 DC283 hrpY(D57N) Merighi et al. (2003)

Plasmids pBluescript

KS and SK (+) ColE1 αlacZ (Apr) Stratagene pRK415 IncP αlacZ (Tcr) Keen et al. (1988)

pPL6GUSC pLAFR6 derivative carrying a promoterless uidA gene (Tcr) Knoop et al. (1991)

pMM221 pET15b with a 0.7 kb NdeI hrpY PCR fragment This study pMM222 pET15b with a 0.7 kb NdeI hrpY D57N PCR fragment This study

pMM46 pBKS(+) with a 0.7 kb HindIII hrpY fragment made by PCR, transcribed from Plac Chapter 2 pMM92 pBKS(+) with a 0.7 kb HindIII hrpY(D57N) fragment made by PCR-SOE This study pMM50 pPL6GUSC with a +994 to –317 hrpS PCR fragment fused to uidA Chapter 2 pMM400 pPL6GUSC with a +629 to –231 hrpS PCR fragment fused to uidA This study pMM396 pPL6GUSC with a +629 to –165 hrpS PCR fragment fused to uidA This study pMM393 pPL6GUSC with a +629 to –90 hrpS PCR fragment fused to uidA This study pMM403 pPL6GUSC with a +629 to –51 hrpS PCR fragment fused to uidA This study pMM392 pPL6GUSC with a +629 to –5 hrpS PCR fragment fused to uidA This study pDM2785 pPL6GUSC with a +220 to –148 hrpL PCR fragment fused to uidA Chapter 2 pMM191 pPL6GUSC with a +220 to –95 hrpL PCR fragment fused to uidA Chapter 3 pMM428 pPL6GUSC with a -12 to –148 hrpL PCR fragment fused to uidA This study

Table 4.1. Bacterial strains and plasmids.

225

Sequence coordinates Primer Sequence (5’Æ 3’) (5’Æ3’)a Source

ET-YF-NdeI GGGAATTCCATATGGATAACACGATTCGAATA 2866-2887 This study

ET-YR-NdeI GGGAATTCCATATGCAATCAGTCAATTAAGAGCAATCCTAAGAGA 3501-3525 This study

JFBG aka HrpLP1 CGCGGATCCGAGCTCTTTCCCAGGCAACAGAGTTAA 481-501 Merighi et al., 2003

L578F CGCGGATCCAATGTAAATCGTTAAATTTC 578-597 This study

L597R CGGGGATCCTGAAATTTAACGATTTAC 581-597 This study

L620F CGCGGATCCGCTGGCACAGACTTTGCTAT 620-639 This study

L637R CGGGGATCCAGCAAAGTCTGTGCCAGCGA 618-637 This study

L638F CGGGGATCCATCGCCATTAAGCCAAACGG 638-657 This study

LRBG CGCGGATCCTCTAGAATGACTTCCAGCCAGGTCAT 869-850 Merighi et al., 2003

PEL TATGTGGGGATGGAGGGTTAACGTG 724-748 This study

SF3509B CGCGGATCCGATTGCTCTTAATTTACAAAT 3509-3529 This study

SF3575B CGCGGATCCAAAAAAGCTTTTTAAACAGC 3575-3594 This study

SF3650B CGCGGATCCGAAATCCTTACAATCCTC 3650-3667 This study

SF3689B CGCGGATCCCAGCATATGACGATGGTTT 3689-3707 This study

SF3735B CGCGGATCCCCCCTGTAAGCACTTCGTAA 3735-3754 This study

SFBG CGCGGATCCATAAAGCCCGCCTGATGGAA 3423-3442 Merighi et al., 2003

SR3575 GCATAAGAAATACCATGTC 3717-3735 This study

SR3650 CCTGACTAACCCTGTGTGAAATG 3628-3650 This study

SR3735 GCATAAGAAATACCATGTCA 3817-3835 This study

SR4368B CGGGATCCGCTCACTCACTCACATTATTTTCAATATTCAT 4355-4369 This study

YD57E TCTTTTACTTCTAGAAATGTGTATGCCTGG 3022-3051 This study

YFH CCGAAGCTTTTCGATAACAATATGGATAACAC 2855-2877 Merighi et al., 2003

YR2 GTTTCAGGATCGTTATCTTTCGTGATG 3204-3178 Merighi et al., 2003

YRH CCGAAGCTTATTAAGAGCAATCCTAAGAGA 3521-3501Merighi et al., 2003

Table 4.2. Oligonucleotide primers.

226

RBS +1 σ70 +611 GUS activity HindIII hrpS IS-like hrpY

WT HrpY[D57N] +994 -317 214±26 16 ±4 gusA pMM50 +629 -231 317 ±9 98 ±2 gusA pMM400 +629 -165 gusA pMM396 1±0 1±0.5 +629 -90 70±5 68±2 gusA pMM393

+629 -51 99±24 79±6 gusA pMM403

+629 -5 48±7 39±1 gusA pMM392

Figure 4.1. Expression of plasmid-borne hrpS gene fusions in P. stewartii.

Expression of GUS reporter plasmids shown below the map (the coordinates of the

sequence ends are shown on the map relative to the transcription start site) was

analyzed in strains DC283 (wild-type) and MM254 (DC283 hrpY[D57N]) after growth

in IM5.5 for 16 h. Data are means from two experiments with two replicates each.

-1 –1 70 GUS activities are reported as pmol MU min OD600 ± standard deviation. The σ promoter is indicated by a yellow vertical arrow. RBS = ribosomal binding site.

227

+1 σ54 σ70 σL IHF UAS GUS activity hrpL hrpJ Control +hrpY +220 -148

76 ±9 5 ±1 gusA pDM2785 +220 -95 156±10 15±8 gusA pMM191 -12 -148 4±3 53±13 gusA pMM428

Figure 4.2. Expression of plasmid-borne hrpL gene fusions in E. coli SΦ200.

Expression of GUS reporter plasmids shown below the map (the coordinates of the

sequence ends are shown on the map relative to the transcription start site) was

analyzed in the E. coli strain SΦ200, with and without the regulatory plasmid pMM46

+ P lac-hrpY , after growth in LB broth for 16 h. Data are means from two experiments

-1 –1 with two replicates each. GUS activities are reported as pmol MU min OD600 ± standard deviation. The positions of σ 70, σ 54, and σ L promoters in the hrpJ-hrpL intergenic region are shown by yellow arrows. IHF = integration host factor; UAS = upstream activating sequence.

228

Figure 4.3. Nucleotide sequence alignment of the hrpY-hrpS intergenic regions

from several erwinias and salient features of this region in P. stewartii.

The DNA sequence of the P. stewartii (Pns) hrpS regulatory region minus the 483 bp

IS-like element was aligned with the corresponding regions from P. agglomerans pv.

gypsophylae (Pag), Erwinia amylovora (Ea) and Pectobacterium chrysanthemi (Pch)

using ClustalX. Primers for probes and plasmid fusions are indicated with a name

followed by a small arrow; the coordinates in the primer names refer to P. stewartii

GenBank accession AF282857. Regions protected in footprinting experiments

described in Fig. 4.6 are shown here in bold italics, and the thick arrows indicate the

putative HrpY consensus binding site (MAAYYY). Peaks present in the intrinsic

curvature profiles shown in Fig. 4.4 are marked by asterixes. RBS = ribosomal

binding site; UTR = untranslated region; CUR = Conserved upstream region; VUR =

variable upstream region.

229 SF3509B →

CUR

Pns GATTGCTCTTA-ATTTACAAATTAATGCTACCCACCGTTATTACTTTATCATTTCAACAG Pag GATTGCTCTTATATTTTTTTATTAATGCTACCCATCGTTATTATTTAATTATTTCAATGC Ea GAATGCTCTTATATTTGTCT-CTCGCCCTTCCCTCCTTGAGCAGATATTTATTATCACGC Pch ------CGGTA-ACTTGTACGC---CGCTCCTCTCCGTTATGCGTCATCGCCTGTCGTGA

HindIII VUR

SF3575B → ****** ******* IHF site

Pns GTTACAAAAAAAAGCT-TTTTAAACAGCACGTTATGCCTAAAAATTTTATGTAACCGTTT Pag TTTA-ATAAAAACCCT-TTATTGACAACCTGTTATGTTCATCG-CGATGTGTAAACGCTT Ea CAGATTTAAAAACACCATTAAAAACAATTGGATAAAATGGTTG-TGGAGTGTAACCGCTT Pch CGGAGAGGAGATCCTCGTTTCCCCCTATCCGGAGAGCC------TGCGTAGCGGCG-

Direct Repeats

←SR3650B SF3650 →

Pns TCATTTCACACAGGGTTAGTCAGGAAATCCTTACAATC--CTCCATCAATTCCTTACCCG Pag CCACTCTAATGATGGTGGGTAGGAATAGCCTTACCCTTATCTCAGTGAATTCCTTACTCG Ea CCATTTTATCCCCCGAATGTAGGGTAATCCCTACATTGCCCCCTGAAGATTCCTTACTTA Pch ------ATTTCCGCGGGGTAAGATTATCCCTACTTTTTGCTCTG--AAATCCCTACCTC

σ70 promoter

SF3689B → -35 box -10 box ← SR3735 SF3735B→

Pns ATAT-CAGCA-TATGACGATGGTTTATCT-GTGCTGACATGG-TATTTCTTA-TGCCCCT Pag TGGTTCAGCA-TAAGACGATGGTTAATCT-GAGCTGGCATGA-TATTTTTTA-ATTCATT Ea ATGTTCAGCA-TAAGACGATGGTTTCTGG-ACGTTAGTATGGGTATTTCTTAGTGGCAGC Pch AACTTCAGCGGTTGGCTGATGGTTTGTTTTGCGCCAG-GCAG-TAATGTTTACGCATACC

mRNA start +1 5' UTR

Pns GTAAGCACTTCGTAAACATAAGAATAGTGATTGTAATTAAATGTAATTCAGCTATTGGCT Pag GTAAGCGAAGCGTAATCACGTGTGCAGTCAGTGTAATTAATTGTAAGTGATGAGGTGGCA Ea GTAAACTCAGAGTAAATATTACAGGCATAATTGTAAATAGTTGTGTCTTGTATAACGGCT Pch ATCAATACGTCGTTAAAACGATCAGATTTATCGCAGTTGATCGTTAAGAATACATCAGGG

Site of 483-bp IS-like insertion in Pns ←SR3845

Pns CTAAGTATGTTGTT--AACTGCGATGATGTCAGTTGAAA----GGATAAGTG------Pag ATTAGATTGCTGTG--TACCGCAATGCCGTTAAATAAAACA--GGATAGGAA------Ea GTAACATGACTGTGC-AAGTGAAGCAATGTTAAAAAAAATAATGGATAGGGC------Pch TGATGATGATGGTTGTCAGATCGACGCTGTCGCCGACATGTCGGCACGGCATGGAAATGA ** * ** * * *

RBS Start codon Pns ------AGGGGCAACTTTTTATG Pag ------AGGGGTAACTTCTTATG Ea ------AGGGGTAATTTTTTATG Pch TCCGACAGTCTGATAATGGAATACAAAAAAAGTACAGGAGCAGTC----ATG

230 -231 -154 +1 +95

Figure 4.4. Computer modeling of the intrinsic curvature of the hrpS promoter

region.

The DNA sequence of the hrpS promoter region corresponding to fragment A in Fig.

4.10 was analyzed with the Bend.It algorithm (www.icgeb.trieste.it) to estimate the

intrinsic curvature of the DNA molecule due to its base composition. The graph shows

the predicted curvature in degrees per helical turn and the G+C content in % using a

nucleotide window of 31. Coordinates in bold are relative to the transcription start.

The double-headed arrow indicates the region integrated to calculate the total

bending.

231

Figure 4.5. Primer extension analysis and sequence analysis of the hrpL promoter.

(A) Total RNA from DC283 cells grown in IM5.5 or LB broth was annealed with the

32P-end-labeled primer PEL and cDNA was reverse transcribed in vitro. Lanes A, T,

G, and C show the dideoxy sequencing ladders for the hrpL gene generated with the same primer. The transcriptional start sites are indicated by solid arrows. The transcription starts and the σ 54 promoter region are highlighted. (B) Sequence analysis

of the hrpJ-hrpL intergenic region. Coordinates are relative to Genbank accession

AF282857. Promoter boxes are in bold, underlined characters. RBS = ribosomal

binding site.

232

A

LB IM5.5 A C G T

…AAACCG AA TT A CCGCTATCGTTTCAGACACGGTCGCTAAAATAAAACAA TTTGT AAACTTTAAA +1 σ54 -58 (+1’) B

421 CTCATACAGGGAAGCTGATGCAGATATGTGGCGTGTTTTATGCTCCGGTTCCCAATAAAT 480 R M ← hrpJ UAS

481 TTTCCCAGGCAACAGAGTTAATTACAGAAGATTGCAATGAGTTGCAGATTTAGCTGGAGC 540

sigma 70 box -58

-35 -10 +1’ 541 TTACGGCAAAAAGTTGCGAAAAAATGAAGGGAAATAAAATGTAAATCGTTAAATTTCAAA 600

IHF site sigma 54 box

-24 -12 +1 601 TGTTTAACAAAATAAAATCGCTGGCACAGACTTTGCTATCGCCATTAAGCCAAACGGCAA 660

RBS__ 661 AAAATGGATTAAGATACGGAGCGTATTATGTCAGAAATGAACCTGCACCCTACTGAGTCA 720 M S E M N L H P T E S hrpL →

233

Figure 4.6. Overexpression and purification of HrpY.

Plasmid pMM221 was expressed in E. coli BL21(DE3)/pLysS upon IPTG induction.

Proteins were purified by native IMAC on Ni2+-IDT columns. Various protein

fractions were separated by SDS-PAGE and stained with Coomassie Blue. (A): lane 1,

total cell proteins of uninduced cells; lane 2, total cell proteins of IPTG-induced

cells; lane 3, flow-through from the Ni2+-IDT columns loaded with a French press

lysate; lanes 4 and 5, fractions eluted by using washing buffers with 20 to 60 mM

imidazole; lane 6, 250 mM imidazole fraction; lane 7, 500 mM imidazole fraction.

(B): Step gradient purification fractions with elution buffers containing imidazole at

150 mM (lanes 1 to 3), 250 mM (lanes 4 to 6) and 250 mM (lanes 7 to 9). Fractions 1-

2 and 3-9 were separately pooled, desalted by G25 chromatography and purified by

ion exchange chromatography. (C), fractions purified by anion exchange

chromatography on Q-sepharose columns. Fractions from lanes 5 to 8 were pooled and

desalted/concentrated by ultrafiltration. Proteins purified by ion-exchange

chromatography were used for all the phosphorylation experiments and for most of the

DNA binding assays.

234 123 45 6 7 A 61.3 kDa 49 36.4 24.7 19.2 13.1

9.3

1 2 3 4 5 6 7 8 B 61.3 kDa 49 36.4 24.7 19.2

13.1

9.3

1 2 3 4 5 6 7 8 9 10 11 12 13 14 kDa 61.3 C 49 36.4 24.7 19.2

13.1

235

A B

C D

Figure 4.7. Mass spectrometry analysis of recombinant HrpY and HrpY[D57N].

Protein molecular weight measurements on 300 pmol of desalted HrpY and

HrpY[D57N] were performed by electrospray ionization mass spectrometry using a

Micromass Q-TOFTM II (Micromass, Wythenshawe, UK) mass spectrometer at the OSU

CCIC-MS Facility. m/z spectral data for HrpY (A) and HrpY[D57N] (C) were deconvoluted in the 22 to 30 kDa region using MaxEnt software (Micromass) to

produce distributions of molecular weights for HrpY (B) and HrpY[D57N] (D).

236

Figure 4.8. Homology model of the crystal structure of P. stewartii HrpY

A structural homology model of the HrpY protein was constructed with the Swiss-

Model server using the X-ray diffraction and NMR solution structures of E. coli NarL and Bacillus subtilis Spo0F. Modeling was successful from residue 3 to 206 and

AcelrysView software was used to render the structure. The modular structure of the

protein is shown in panel A. Details of the catalytic site with the conserved

aspartates, lysine and threonine are shown in panel B. Panel C shows the LuxR-like

helix-turn-helix (HTH) motif in the output domain of HrpY. The complete structural

coordinates are reported in the Appendix.

237

Receiver domain Linker Catalytic triad

Output domain

A

D57 D11

D12 K105

T85

DNA binding Helix-turn-helix motif

B Catalytic site residues C

238

HrpY - + - + - - + - + - HrpY[D57N] + - - - + + - - - + BarA - - + + + - - + + +

BarA198~P

26 kDa HrpY~P

Coomassie Autoradiogram

32 32 Figure 4.9. Phospho-transfer of P i from BarA198~ P to HrpY.

Reactions contained 25 µM HrpY, 75 µM Salmonella enterica BarA198, 40 µM

[γ 32P]ATP and 0.1 mM cold ATP in a total volume of 15 µl. Mixtures were incubated

at 25°C for 90 min. Reactions were terminated with 5 µl of 4X stop buffer, separated

by SDS-PAGE (T=12%), and visualized by Coomassie Brilliant Blue staining.

Radiolabeled proteins were visualized by Phosphor-Imager scanning before gel

staining. Negative controls consisted of HrpY and HrpY[D57N] incubated in presence

of [γ 32P]ATP.

239

70 +1 σ

hrpS IS-like hrpY

+95 -231 A 327 bp

-165 -231 B 76 bp +95 -165 C 261 bp -5 -90 C2 86 bp +95 -5 C3 101 bp -90 -165 C1 76 bp

Figure 4.10. Map of the hrpS regulatory region and the promoter fragments used as probes in gel shift experiments

The hrpS promoter region was amplified by PCR from plasmid pMM58 to produce fragments A, B, C, C1, C2 and C3. Coordinates are relative to the transcription start site. Fragments were used in electrophoretic mobility shift experiments as either radiolabeled probes or cold specific competitors.

240

hrpS probe (25 fmol) : A

HrpY (nM): 0 10 40 160 625 625

Competitor (A fragment): - - - - - +

261 bp

Figure 4.11. Electrophoretic mobility shift analysis of the binding of HrpY to the full length hrpS promoter.

Binding conditions are described in Materials and Methods. Increasing concentrations of HrpY were incubated with 25 fmol of 32P-labeled fragment A DNA (Fig. 4.10) for

35-40 min at room temperature. Aliquots were loaded onto a 5% acrylamide gel in

1XTBE buffer and electrophoresed for 2.5 h at room temperature. The positions of the

free probes and complexes were determined by autoradiography of the dried gel at –

80°C. The molar excess of the specific cold competitor in the last lane over the 32P- probe was 100-fold. The size of the free probe in bp is shown.

241

hrpS probe B C1 C2 C3 C (10-20 fmol) : HrpY(4.7 µM): - + + - + + - + + - + + - + + + Competitor (C fragment): - - +200X - - +62X - - +114X - - +46X - - +5X +50X

261 bp

103 bp 86 bp 76 bp

Figure 4.12. Electrophoretic mobility shift analysis of the binding of HrpY to portions of the hrpS promoter.

Binding conditions are described in Materials and Methods. Assay mixtures contained

4.7 µM HrpY and 10-20 fmol of 32P-labeled probe DNA (10,000 cpm). Fragments B,

C, C1, C2 and C3 (whose coordinates are reported in Fig. 4.10) were used as probes.

Mixtures were incubated for 35-40 min at room temperature. Cold competitor DNA

(probe C) was added before HrpY at the concentrations indicated by superscripts

(expressed as molar fold over the 32P-probe). Aliquots were electrophoresed in a 5%

acrylamide gel in 1X TBE buffer for 2.5 h at room temperature. The positions of the

free probes and complexes were determined by exposure of the dried gel to an X-ray

film at –80°C. Sizes of the free probes in bp are shown.

242

hrpS probe: C

HrpY (12.5 µM): - + + + + + + + + + + + + + + + +

Competitor -- BC1C2C3C (25,100,250X):

Figure 4.13. Competition analysis of the binding of HrpY to hrpS fragment C

Binding conditions are described in Materials and Methods. HrpY at 12.5 µM was

incubated with 25 fmol of 32P-labeled DNA fragment C (coordinates reported in Fig.

4.10) with and without competing unlabeled DNA fragments (B, C, C1, C2, C3; see

Fig. 4.10 for the coordinates) at 20, 100 and 250 molar excess for 35-40 min at room temperature. Aliquots were electrophoresed in a 5% acrylamide gel in 1XTBE buffer for 2.5 h at room temperature. The positions of the free probes and complexes were determined by exposure of the dried gel to an X-ray film at –80°C. Sizes of the free probes in bp are shown

243

54 70 L + σ σ σ UAS hrpL hrpJ

+221 -148 L389bp

+221 -94 I 315 bp

+221 -71 H 292 bp

+221 -29 G 250 bp +221 -11 F 232 bp -12 -148 M 156 bp

Figure 4.14. Map of the hrpL regulatory region and of the promoter fragments used as probes in gel shift experiments.

The hrpS promoter region was amplified by PCR from plasmid pMM58 to produce fragment L with the complete promoter region fragments I, H, G and F with 5’ deletions, and fragment M with a 3’ deletion. These were used in electrophoretic mobility shift experiments. Sizes are in bp and coordinates are relative to the transcription start site. UAS = upstream activating sequence.

244

hrpL probe (10-25 fmol) : F G H I L M

HrpY (20 µM): - + - + - + - + - + - +

389 bp 292 bp 315 bp 250 bp 232 bp

156 bp

Figure 4.15. Electrophoretic mobility shift analysis of the binding of HrpY to portions of the hrpL promoter.

Binding conditions are described in Materials and Methods. HrpY at 20 µM was

incubated with 10 to 25 fmol of 32P-labeled DNA (5,000 cpm) corresponding to fragments F, G, H, I, M and L (whose coordinates are reported in Fig. 4.14) for 35-40

min at room temperature. Aliquots were electrophoresed in a 5% acrylamide gel in 1X

TBE buffer for 2.5 h at room temperature. The positions of the free probes and

complexes were determined by exposure of the dried gel to an X-ray film at –80°C.

Sizes of the free probes in bp are shown.

245

hrpL probe: L

Aspecific Specific Competitor(fold):

-- - - -1x 7X 25X 150X 1X 7X 30X 155X

HrpY(µM): 0 1 10 20 50 5050 50 50 50 50 50

389 bp

Figure 4.16. Electrophoretic mobility shift analysis of the binding of HrpY to the full length hrpL promoter fragment.

Binding conditions are described in Materials and Methods. Increasing concentrations of HrpY were incubated with 5 fmol of 32P-labeled DNA corresponding to fragment A

(6000 cpm) in Fig. 4.10 for 35-40 min at room temperature. Specific or aspecific competitor DNAs consisted of unlabeled DNA fragment L and poly(dI:dC), respectively. Aliquots were electrophoresed in a 5% acrylamide gel in 1X TBE buffer for 2.5 h at room temperature. The positions of the free probes and complexes were determined by exposure of the dried gel to an X-ray film at –80°C. Sizes of the free probes in bp are shown.

246

hrpS probe C hrpL probe 1E+08 8000 A B 9E+07 7000 8E+07 HrpY Kd=20 µM HrpY Kd=20 nM 6000 HrpY~P Kd=21 µM 7E+07 HrpY~P Kd=1.7 nM

6E+07 5000

5E+07 4000 ν/P HrpY~P ν/P HrpY~P 4E+07 3000 3E+07 2000 2E+07 HrpY 1000 HrpY 1E+07

0 0 0 0.5 1 0 0.5 1 ν ν HrpY HrpY~P HrpY HrpY~P

Figure 4.17. Effect of phosphorylation on HrpY binding to the hrpS and hrpL promoters

Increasing concentrations of HrpY or HrpY~P (ranging from 5 nM to 80 µM with concentrations increased 4-fold each step; first lane in each series is the “no protein” control) were incubated with an average of 40 fmol of 32P-labeled probe DNA (10,000

cpm) for 35-40 min at room temperature in binding buffer (see Materials and

Methods). Aliquots were loaded onto a 5% acrylamide gel in 1X TBE buffer and

electrophoresed for 2.5 h at room temperature. The positions of the free probes and

complexes and their intensities were determined by exposing the dried gel to a

phosphor storage screen (lower panels). (A): Binding of HrpY to hrpS fragment C

(similar results were obtained for fragment A). (B): Binding of HrpY to hrpL fragment L (coordinates are reported in Fig. 4.10 and Fig. 4.14).

247

DM733/pMM118 D57N

DM733/pMM99 D57E

DM733/pMM74 D57A

DM733/pMM52 D57

DM733

0 100 200 300 400 500 600 β-galactosidase activity (Miller Units)

Fig. 4.18. β-galactosidase activity of a chromosomal hrpS-lacZ fusion in a P. stewartii hrpY null mutant carrying various plasmid-borne Plac-hrpY regulatory

genes

Cells were grown in IM pH 5.5 broth to late log phase and β-galactosidase specific

-1 -1 -1 activity was measured (1 unit = 1 pmol MU min OD600 ml of culture at 25°C).

+ Regulatory gene clones: pMM52 Plac-hrpY , pMM74 Plac-hrpY[D57A], pMM99 Plac-

hrpY[D57E], and pMM118 Plac-hrpY[D57N]. Strain DM733 is DC283 hrpY::Tn5

hrpS::Tn3HoHoI. Vector control not run in this particular experiment. Previous data

suggested no effect of pRK415 on P. stewartii hrp genes and no effect of recombinant

plasmids on DC283 β-galactosidase background activity (not shown).

248

Fig. 4.19. DNase I footprinting assay of HrpY complexes at the hrpS promoter.

Binding reactions used HrpY at the concentrations shown above the lanes and 35 fmol

of a DNA fragment spanning probe A. The equilibrated complexes were partially

digested with DNase I and the digests were amplified by linear PCR with 32P-labeled primer SF3509. The reactions were electrophoresed in a 6% acrylamide/8M urea sequencing gel alongside sequencing ladders for A and G. Solid arrows indicate hypersensitive sites, solid black boxes represent protected sequences. Lane “0” was loaded with 4 times the normal amount of DNA in order to visualize the ladder. The control ladders were visualized by autoradiography for 24 h and the footprint ladders required several days of exposure at -80°C. The coordinates are relative to the +1 transcription start. The bars on the right represent portions of the probe A.

249

HrpY A G0 1.5 15 µM

C3

+1 -10 box -35 box

-50 -50 C2 -71 -80 -78

-100

-102

-118 IHF -120 -125 C1 -132

-148

-165 -170

-174 B (not shown in -182 full)

250

A

R =IHF?=H-NS? RNApol IS-like C3 σ70 C2 C1 B +1 IHF

B

Hierarchical cooperative binding?

RNApol Y Y Y Y Y IS-like C3 σ70 C2 C1 B +1 IHF ?

Figure 4.20. Model for hrpS activation.

(A) Repression may involve a factor ( R ) binding at the C1 fragment. An IHF site is contained in this region. Whether IHF is a repressor of this system, as shown in other

E. coli promoters controlled by cooperative binding of FixJ-like regulators, or a helper for HrpY is not known. H-NS was also shown to bind to C fragments in preliminary experiments (not shown) (B) Cooperative, perhaps hierarchical, binding of HrpY (Y) at multiple sites activates the hrpS promoter. Given that binding to C1 and B is weak or not visible in EMS assays, but observable in the preliminary footprinting experiments, initial binding to C2 sites may be a prerequisite. This is consistent with the stronger binding of HrpY to the A probe compared to binding to its subfragments.

251

Activation loop

IS-like C3 C2

HrpY HrpY C1 B

Repression loop C3

IS-like R?

C2 R?

B C1

Figure 4.21. Model for hrpS activation.

The activation/repression model shown in Fig. 4.20 may involve looping of the hrpS promoter region, supported by the presence of an IHF site and by the intrinsic bending of the DNA.

252

A Activation loop

σ54 σ70

hrpL HrpS IHF

UAS

B Repression loop

hrpL

? -12 HrpY 54 σ

-29 σ70

Figure 4.22. Models for hrpL regulation.

(A) Activation of the σ54 promoter of hrpL may involve binding of HrpS to the UAS

element (probably as simple dimers because of the presence of a single dyad in the

site), DNA looping, perhaps helped by IHF, and isomerization of closed complex. (B)

DNA binding and repression by HrpY requires sites downstream and upstream of the –

12 base. This implies that a looping mechanism may be involved as well. The loop

may prevent proper expression of both σ 54 and σ 70 promoters.

253

CHAPTER 5

GENETIC ANALYSIS OF THE HRPX PAS-KINASE AND PARTIAL

CHARACTERIZATION OF RECOMBINANT HRPX PROTEINS AND

PROTEIN FRAGMENTS

5.1 INTRODUCTION

In bacteria, most adaptive responses to changes in the environment are, directly or indirectly, mediated by two-component signal transduction systems (Stock et al., 2000). These regulatory systems are composed of a sensor kinase and a response regulator, which is often a transcription factor. Most sensor kinases are transmembrane proteins, presumably perceiving extracellular signals via the periplasmic portions of the input domain. Some, however, are soluble proteins, in particular when the signal perceived is intracellular. Examples of cytoplasmic sensors are E. coli NtrB and Bradyrhizobium japonicum FixL, which sense nitrogen starvation conditions and cellular oxygen levels respectively (McFarland et al., 1981; Taylor and Zhulin, 1999; Gilles-Gonzalez et al., 1994). In other cases, the sensor may be a transmembrane protein with its sensory

254 input domain located in the cytoplasm. Examples of this are the E. coli

ArcB redox sensor and the Sinorhizobium meliloti FixL (Gilles-Gonzalez

et al., 1995; Gong et al., 1998; Gilles-Gonzalez, 2001). Many two- component systems are involved in the regulation of virulence genes

(Hoch and Silhavy, 1995; Finlay and Falkow, 1997). In only a few cases,

however, is the stimulus perceived by the sensor kinase known. For

example, in Bordetella pertussis, expression of hemoagglutinin and toxins

is under the control of BvgS/BvgA, which transduce signals such as

2- temperature, SO4 ions, and nicotinic acid (Arico et al., 1989). In

Shigella flexneri, osmolarity modulates virulence gene expression via the

EnvZ/OmpR system (Bernardini et al., 1990). In Salmonella enterica, the

PhoQ/PhoP, PmrB/PmrA and RcsC/YojN/RcsB pairs form a multicomponent, interconnected cascade responding to extracellular Mg2+,

Fe3+ concentrations, pH and cell envelope stresses (Fig. 1.7) (Soncini and

Groisman, 1996; Kox et al., 2000; Bijlsma and Groisman, 2003;

Chamnongpol et al., 2003; Mouslim and Groisman, 2003). In

Agrobacterium tumefaciens, phenolic compounds and monosaccharides are perceived by the VirA/VirG system that regulates the vir type IV

secretion apparatus leading to the development of crown gall tumors

(Shaw et al., 1988; Lee et al., 1995).

P. stewartii uses the hrpXY two-component system to activate the hrp/hrc type III secretion system. The HrpX sensor PAS-kinase is required for full activation of hrp regulatory and secretion genes in vitro

255 and for consistent infection of corn seedlings (Merighi et al., 2003). In

contrast, the ortholog protein in Erwinia amylovora and Pantoea agglomerans subsp. gypsophiliae is not required for HR elicitation and virulence (Wei et al., 2000b; Mor et al., 2001).

The input region of HrpX contains two tandem PAS sensory domains.

These are cytoplasmic modules of sensor kinases that are often implicated in perceiving a wide range of intracellular stimuli in bacteria, including oxygen levels (via heme prosthetic groups; reviewed in Perutz et al.,

1999), redox conditions (via FAD, FMN, and possibly MK and Qo, binding; Repik et al., 2000; Georgellis et al., 2001; Bock and Gross,

2002), energy levels (via ATP/ADP binding) (Stephenson and Hoch,

2001), and light (Cheng et al., 2002). In sensors with multiple PAS domains (Stephenson and Hoch, 2001; Wang et al., 2001), these motifs are also involved in intermolecular protein-protein interactions and probably the two functions are often overlapping. From previous studies, we know that the expression of genes directly or indirectly controlled by hrpXY respond to environmental variables, such as pH, osmolarity, nitrogen starvation and nicotinic acid (Merighi et al., 2003), but whether

HrpX is actually perceiving any of these stimuli is not known. Similarly, the role of the two PAS domains in the input region of HrpX has not been determined in any of the erwinias.

In this chapter, we describe initial experiments to define a role for the

HrpX PAS domains using genetic and biochemical approaches. Plasmid

256 clones with in-frame deletions of hrpX were constructed for expression in

P. stewartii. The phenotypes conferred by these plasmids in ∆hrpX strains

or the phenotypes of mutants constructed by allelic exchange of these mutations are described. Preliminary results for the purification of the sensor kinase protein and of the tandem PAS domains are presented as

well as the construction of several recombinant plasmids that can be used

to purify HrpX and its domains in future studies. We also used the effect

of HrpY[D57N] overexpression, which overrides the need for

phosphorylation (as described in Chapter 2), to define which classes of

signals may be perceived by HrpX. This strategy was chosen because we

do not have a constitutively active, mutant form of HrpX.

5.2 MATERIALS AND METHODS

5.2.1 Media and Strains and general microbial genetic techniques

E. coli and P. stewartii strains and plasmids used in this study are listed

in Table 5.1. Luria-Bertani (LB) broth and agar (Ausubel et al. 1987) were used for strain maintenance and for growth of P. stewartii under

Hrp-repressing conditions. Fusaric acid-containing tetracycline sensitive selection (TSS) medium was prepared as described by Metcalf et al.

(1996). To induce hrp genes in P. stewartii strains, overnight LB broth

cultures were washed twice in inducing medium (IM5.5; Frederick et al.

2001) prepared without casaminoacids and with only 10 mM sucrose

(IM5.5a), inoculated into 2 to 5 ml of the same medium at A600nm ~ 0.05,

257 grown for 16 h and adjusted to a final A600nm of ca. 0.5. Standard IM5.5,

as described in Chapter 2, was used for GUS assays with the hrpX

mutants. In the experiments where chemicals were tested for their activity

on hrp expression, utmost care was taken to assure that the various media

had comparable osmolarity. Carbon and nitrogen sources were used at the

concentrations reported in the respective figure legends. When practical,

the van ’t Hoff law was used to calculate the approximate osmotic

pressure of the various media. HEPES or Tris-HCl buffers were used in

place of MES for the neutral and basic pH ranges, respectively. Liquid

cultures were grown in flasks or tubes shaken at 200 rpm at 37°C for E.

coli or 29°C for P. stewartii. When appropriate, antibiotics were supplied

at the following concentrations: ampicillin, 200 µg ml-1; kanamycin, 50

µg ml-1; and tetracycline, 20 µg ml-1 (or 15 µg ml-1 for selection of single copy Tetr cointegrates). In IM liquid media the concentration of

antibiotics was reduced by half.

5.2.2 Protein modeling

Modeling of the secondary structure of HrpY was performed with the

PSI-PRED modeling server (McGuffin et al., 2000). Threading of portions of the HrpX amino acid sequence into known crystal structures was performed with mGenTHREADER. Both servers are available at http://bioinf.cs.ucl.ac.uk/. The conserved residues in Bradyrhizobium

japonicum FixL crystal structure (PDB accession 1DRM) were identified 258 with mGenTHREADER and rendered as ball-and-stick atoms in secondary

structure schematics. Molecular models were visualized and rendered with

Acelrys ViewerLight v. 5.0 (Acelrys). Multiple sequence alignments were

performed with Clustal X. Phylogenetic trees were designed with

TreeView and the branches were resampled by 1000 cycles of

bootstrapping.

5.2.3 Construction of strains with in-frame deletions in various hrpX

domains

Unmarked in-frame deletions spanning various portions of the input domain of hrpX were constructed in vitro by PCR as previously described

(Merighi et al., 2003). The PCR fragments and positions of the primers are shown in Fig. 5.1. For the PCR reactions, plasmid pMM58 was used as the template DNA and Pfu as the DNA polymerase (Clontech). The

deletion spanning amino acids 15 to 198 (eliminating both PAS1 and only

part of the PAS2 core) was constructed by amplifying flanking DNA

sequences with primers XdX4AB+BdX5B (fragment C) and primers

BdX1+EdX1 (fragment B) (Table 5.2 and Fig. 5.1). A second ∆hrpX

allele had a deletion between residues 15 and 270, spanning PAS1, PAS2

and part of the linker and this was constructed by amplifying flanking

DNA (fragments D and B) with primers XdX4+BdX5A and primers

BdX1+EdX1, respectively. PCR fragments were digested with the

259 restriction enzymes shown in Fig. 5.1 and ligated into pBluescriptKS

(Stratagene) as contiguous EcoRI-BamHI and XbaI-BamHI inserts to produce plasmids pMM243 (∆hrpX15-198) and pMM242 (∆hrpX15-270), respectively. Deletions of the individual PAS domains were constructed by SOE-PCR. To construct the ∆hrpX15-130 allele, primers

dXPASR2+dXPAS1F2 and primers dXPAS1R1+dXPASF1 were used to

amplify fragments E and F (Fig. 5.1). The two PCR products were gel

purified, mixed in a 1:1 molar ratio, melted, annealed and then used as

templates for a second round of PCR with primers dXPASR2+dXPASF1.

The recombinant PCR fragment “E+F” was gel purified and cloned into

pBluescriptSK as BamHI+EcoRI insert to produce pMM283. A similar

strategy was used to construct the ∆hrpX15-130 allele, with the difference that primers dPAS2F2 and dPAS2R1 were used to delimit the deletion

(Fig. 5.1). The recombinant fragment G+H was cloned in pBluescriptSK

(Stratagene) to produce pMM287. All pBluescript recombinant plasmids were verified by sequencing both strands before any further manipulation.

To generate suicide plasmids for allelic exchange, the inserts in pMM242, pMM243, pMM283 and pMM287 were subcloned as XhoI+SacI fragments and ligated into pLD55, to create plasmids pMM249, pMM247, pMM324 and pMM327, respectively. The suicide plasmids were used to generate unmarked mutations in the P. stewartii DC283 chromosome following the

“Buchner selection” protocol described in Chapter 2. The mutant strains

260 were named MM284 (DC283 ∆hrpX15-270), MM266 (DC283 ∆hrpX15-198),

MM360 (DC283 ∆hrpX15-130), and MM348 (DC283 ∆hrpX143-259) (Fig.

5.2). All mutants were verified by PCR and Southern blot analysis.

5.2.4 Construction of plasmids for genetic complementation tests

Plasmids containing hrpX alleles expressed from their own promoters

were constructed by cloning SOE-PCR fragments amplified from plasmids

pMM283 and pMM287 or by subcloning restriction fragments from

pMM58. To construct a plasmid expressing hrpX from its own promoter,

pMM58, which contains the entire hrp regulatory region, was digested with PshAI and FspI and a resulting 2.1 kb kb fragment was ligated into pBKS at the EcoRV site to produce plasmid pDM2890. This insert in this plasmid was then excised by digestion with EcoRI and HindIIII and recloned into pRK415 such that hrpX was downstream of the Plac promoter in the vector; the resulting plasmid was named pMM389 (or pDM2893). To construct plasmids expressing hrpX alleles missing just one PAS domain (i.e. ∆hrpX15-130 and ∆hrpX143-259), inserts from plasmids pMM283 and pMM287 were excised with EcoRI and BamHI and then ligated into pRK415 to produce pMM378 and pMM380, respectively.

261 5.2.5. Construction of plasmids for expression and purification of

HrpX

The entire hrpX ORF and various portions of it were cloned into several protein expression vectors to produce recombinant proteins with tags to facilitate their purification. Vectors pET28b was used to create C-terminal poly-histidine fusions, while pET15b and pQE30 were used for N-terminal poly-histidine fusions. N-terminal fusions to MBP were created with plasmid pMALc2X , while fusions to GST were constructed in plasmids pGEX-2T and pET41b. C-terminal Chitin Binding-Intein domain fusions were made with vector pTYB1. All constructs were produced by 15-20 cycles of PCR using Pfu DNA polymerase and pMM58 as DNA template.

The cloned inserts were verified by sequencing both strands. The full length ORF was amplified by PCR with primers ETXF1+ETXR (for vector pET15b), primers XBAMF(+1)+ETXR (for vectors pQE30 and pGEX-2T) or primers XpMALF+XpMALR (for vector pMALc2X) (Table 5.2).

Products were gel purified, digested with BamHI and ligated in to plasmids pET15b, pQE30, pMALc2X and pGEX-2T to produce plasmids pMM228, pMM377, pMM315 and pMM371, respectively (Fig. 5.3).

Primers XcHISF2+XcHISR2 (Table 5.2) were used to amplify the hrpX

ORF for cloning into pET28b; the products were gel purified and cloned

into the vector as XhoI/NcoI restriction fragments to generate plasmid pMM367 (Fig. 5.3). Finally, to construct a plasmid expressing a HrpX-

Sce-CB fusion, the hrpX ORF was amplified with primers

262 XNHEF+XNHER (Table 5.2) and ligated as a XhoI/NdeI restriction

fragment into vector pTYB1 to produce plasmid pMM362 (Fig. 5.3). The

insert from pMM228 was also cloned into pEXT20 as an XbaI-HindIII

fragment to produce plasmid pMM319, which expresses hrpX from a Ptac promoter.

The transmitter, input and individual PAS domains were cloned in pET15b as BamHI PCR fragments for expression as poly-His tagged proteins (Fig. 5.3). Primers ETXF1 and ETPAS2R (Table 5.2) were used to amplify the input domain-encoding DNA from amino acids 1 to 485 and produce plasmid pMM257, primers ETXF1 and ETPAS1R amplified the

PAS1 domain (amino acids 1 to 142) to produce pMM251, primers

ETPAS2F and ETPAS2R were employed to produce a fragment expressing the PAS2 domain (amino acids 136 to 276) in plasmid pMM280, and finally primers ETXR1 and ETXF2 amplified the transmitter domain from amino acids 277 to 485 in plasmid pMM233. All inserts were sequenced from both directions to check for polymerase errors. The inserts from pMM257 and pMM251 were also subcloned into the BamHI site of pET41b in order to express N-terminal GST-His6 -S-tag protein fusions

from plasmids pMM270 and pMM265, respectively. The input domain was

also cloned into pTYB1, as an NheI/XhoI PCR fragment amplified with

primers XNHEF and PAS2-Xho-R, to produce plasmid pMM363, and into

pGEX-2T to produce a simple GST-PAS1-2 fusion expressed from plasmid

263 pMM374. For the latter, primers XBAMF(1+) and ET-PAS2R were used to

amplify a PCR product, which was ligated into the vector as a BamHI

restriction fragment.

2+ 5.2.6 Overexpression and purification of His6 -HrpX protein by Ni -

NTA affinity chromatography

Plasmid pMM228 was transformed by electroporation into E. coli

BL21(DE3) carrying pLysS. Several independent Apr Cmr isolates were

stored at –20 and -80°C in 40% glycerol, 10 mM MgSO4 , 50 mM Tris-HCl pH 7. For protein expression, bacteria were grown overnight in 5 ml LB broth with antibiotic selection. Two 2-l flasks containing 700 ml LB broth were inoculated with 1.5 ml of overnight culture. The flask was shaken (200 rpm) at 37°C until the culture reached A600=0.57 (measured

in a B&L Spectronic 20 colorimeter with 13 mm x 100 mm tubes). IPTG

was added to a final concentration of 1 mM and the culture was grown at

15°C for 15.5 h to a final A600=1.7. Cells were harvested by

centrifugation in a Sorval GSA rotor at 7000 rpm for 20 min at 4°C.

About 12 g (wet weight) of cells were resuspended in 46 ml ice-cold lysis

buffer (10 mM imidazole, 500 mM NaCl, 5 mM MgCl2 , 10 mM β-

mercaptoethanol, 0.5 mM PMSF, 1 µg ml-1 leupeptin 0.1% NP-40, 5%

glycerol, 20 mM Tris-HCl pH 7.9 at 4°C). The cell suspension was

incubated in 0.5 mg ml-1 lysozyme for 60 min on ice and then passed three

times through a 40-K French pressure cell (Thermo IEC) at 10,000 psi 264 adding 1 mM PMSF after the first passage. The lysate was centrifuged

twice at 17,500 x g for 45 min at 4°C. The supernatant was incubated with

4 µg ml-1 pancreatic DNase I (Sigma) for 15 min at room temperature and

then it was centrifuged at 17,000 x g for 10 min. IMAC column

chromatography of the supernatant was performed at 4°C using a 1.5 ml

glass column (I.D. = 1 cm) packed with 3 ml of Ni2+-NTA agarose slurry

(Novagen) and equilibrated with binding buffer (10 mM imidazole, 500

mM NaCl, 5 mM MgCl2 , 5 mM β-mercaptoethanol, 20 mM Tris-HCl, pH

8.0 at 4°C). The average flow rate was 0.3 ml min-1. The column was

washed with eight column-volumes of binding buffer, ten column-volumes

of washing buffer (40 mM imidazole, 500 mM NaCl, 5 mM MgCl2 , 20 mM

Tris-HCl pH 8.0 at 4°C) and five volumes of elution buffer A (80 mM imidazole, 500 mM NaCl, 5 mM MgCl2 , 20 mM Tris-HCl pH 8.0 at 4°C).

The target His6 -HrpX proteins were eluted with 5 ml 500 mM imidazole

buffer (500 mM imidazole, 500 mM NaCl, 5 mM MgCl2 , 20 mM Tris-HCl

pH 8.0 at 4°C) collecting 1-ml fractions. The target protein migrated in

SDS-PAGE as a 63 kDa band instead of 58 kDa as expected. Fractions

containing higher concentrations of the 63Kda protein were pooled and

either dialyzed overnight against 2 l of storage buffer (100 mM KCl, 1

mM MgCl2, 1 mM DTT, 0.1 mM EDTA, 10% glycerol, 20 mM Tris-HCl

pH7.9) or buffer-exchanged by repeated ultrafiltration in a MicroCon-10

spin filter (Amicon). The semipurified protein preparation was stored at –

265 80°C. The degree of purification was estimated by Coomassie Brilliant

Blue staining (Laemmli, 1970) and the fractions were analyzed by

Western blot using anti-His6 monoclonal antibodies (Novagen) following

the manufacturer’s instructions.

5.2.7 Overexpression and purification of MBP-HrpX protein by

amylose affinity chromatography

Plasmid pMM315 was transformed by electroporation into E. coli strains

DH10B, TB1 and ER2507. For protein expression, bacteria were grown

overnight in 10 ml LB broth with antibiotic selection and 0.2% glucose.

One 2.8-l flask containing 1000 ml LB broth, 200 µg ml-1 ampicillin,

0.2% glucose was inoculated with 1.5 ml of overnight culture. The flask

was shaken (200 rpm) at 37°C until the culture reached A600=0.65. IPTG

was added to a final concentration of 0.3 mM and the culture was grown

at 15°C with shaking for 16 hr. Cells were harvested by centrifugation in

a Sorval GSA rotor at 8000 rpm for 20 min at 4°C. About 12 g (wet

weight) cells were then resuspended in 25 ml of ice-cold column buffer

(CB; 2 mM NaCl, 10 mM β-mercaptoethanol, 1 mM PMSF, 1 µg ml-1 pepstatin, 20 mM Tris-HCl, pH 7.4 at 4°C). The cell suspension was incubated in 0.5 mg ml-1 lysozyme for 20 min on ice and then passed two

times through a 40-K French pressure cell (Thermo IEC) at 10,000 psi

adding 0.5 mM PMSF after the first passage.The lysate was centrifuged

266 twice at 20,000 x g for 45 min at 4°C. The supernatant was sonicated to reduce the viscosity and diluted by adding 100 ml of CB. Column chromatography was performed at 4°C. The supernatant was then applied to a glass column (I.D.= 1 cm) packed using 6 ml of an 80% amylose slurry (Novagen) and pre-equilibrated with 40 ml of column buffer. The average flow rate was 0.3 ml min-1. The column was washed with 12 column-volumes of CB buffer. The MBP-HrpX protein was eluted with four column volumes of elution buffer (10 mM maltose in column buffer) collecting 1-ml fractions. Fractions containing the expected 100-kDa band were pooled and concentrated by ultrafiltration in a Vivaspin 15R device.

The protein preparation was proteolyzed with 0.25% factor Xa for 24 h at

4°C followed by 24 h incubation at room temperature. The protein mixture was dialyzed in buffer A (20 mM Tris-HCl pH8, 25 mM NaCl, 10 mM β- mercaptoethanol) for 24 h and fractionated by ion-exchange chromatography. A linear gradient starting with 100% buffer A and 0% buffer B (20 mM Tris-HCl pH8, 750 mM NaCl, 10 mM β- mercaptoethanol) and ending with 100% buffer B was used. Fractions containing the 56-kDa target protein were pooled and concentrated by ultrafiltration. The degree of purification was estimated by Coomassie

Brilliant Blue staining (Laemmli, 1970).

267 5.2.8 Overexpression and purification of GST-PAS1-2HrpX protein by

glutathione affinity chromatography

For expression of the GST-PAS1-2HrpX protein, E. coli DH10B

(pMM374) was grown overnight in 10 ml LB broth with antibiotic

selection. One 1-l flask containing 100 ml LB broth, 200 µg ml–1 ampicillin, 0.2% glucose was inoculated with 1 ml of spun and washed overnight culture. The flask was shaken (200 rpm) at 37°C until the culture reached A600=0.6. IPTG was added to a final concentration of 1

mM and the culture was grown with shaking at 37°C for 4 hr. Cells were

harvested by centrifugation in a Sorval GSA rotor at 8000 rpm for 20 min

at 4°C (wet weight 12 g) and then resuspended in 5 ml of ice-cold PBS

buffer (1X phosphate buffer saline pH 7.4). The cell suspension was

incubated in 0.5 mg ml-1 lysozyme for 20 min on ice and sonicated by 10

bursts of 10 s each at 20% duty, adding 1 mM PMSF after the first burst.

The lysate was centrifuged twice at 20,000 x g for 45 min at 4°C. Column

chromatography was performed at 4°C. The supernatant was applied to a glass column (I.D. = 1 cm) packed with 3 ml of a 50% GST-BindTM resin slurry (Novagen) and pre-equilibrated with 15 ml of PBS. The column was washed with 12 column-volumes of PBS. The target GST-HrpX protein was eluted with three column-volumes of elution buffer (50 mM Tris-HCl pH 8.0, 10 mM reduced glutathione) collecting 0.5-ml fractions. Fractions containing the expected ca. 80-kDa protein were pooled and concentrated by ultrafiltration in a Vivaspin 15R device. 268 5.2.9 Mass spectrometry protein fingerprinting

Protein samples were analyzed by nano-liquid chromatography tandem

mass spectrometry at the OSU CCIC Proteomics and Mass Spectrometry

Facility. Proteins were separated by SDS-PAGE (10% T; mono to bis-

acrylamide ratio = 30:0.8 ), fixed in 5% acetic acid in ethanol followed by

standard Coomassie staining and destaining. Protein bands were excised

and stored in 5% acetic acid at 4°C. An excised gel fragment without

visible proteins was used for the blank control run. The samples were

cleaved by in-gel trypsin digestion. Tryptic fragments were characterized

by MS/MS fragmentation. Data were analyzed using the Mascot server by

comparison to the Genbank protein database. The sequence of 46% of

HrpX was determined followed this procedure.

5.2.10 In vitro autophosphorylation of HrpX and phosphotransfer

reaction

For autophosphorylation of HrpX, 15 µl of a mixture containing 30 µM

HrpX, 0.22 µM [γ32P]ATP (3000 Ci/mmol; NEN) and 0.1 mM cold ATP in

phosphorylation buffer (125 mM potassium glutamate, 125 mM HEPES pH

8.0, 5 mM MgCl2 , 0.1 mM DTT, 5% glycerol) was incubated at 25°C for up to 60 min. Reactions were terminated with 5 µl of 4X stop/sample buffer (125 mM Tris-HCl pH 6.8, 8 mM EDTA, 4% SDS, 8% β-mercapto- ethanol, 20% glycerol, 0.02% bromophenol blue). For transphosphorylation reactions, HrpX was allowed to autophosphorylate 269 for 20 min as described above and then 40 µM HrpY was added. Reactions

were terminated at intervals using 4X stop buffer. Proteins were separated

by SDS-PAGE (12% monomer gel), and visualized by Coomassie Brilliant

Blue staining (Laemmli, 1970). Gels were dried under vacuum at 60°C for

30 min. The radiolabeled proteins were visualized by autoradiography at –

80°C using Kodak MS film.

5.2.11 UV-crosslinking of the HrpX input domain with [γ 32P]ATP

Direct photo-affinity labeling of the PAS domains with [γ32P]ATP was performed following the protocol of Stephenson et al. (2001). In brief, 50 ng of recombinant protein was incubated in 50 µl reaction mixtures containing 50 mM HEPES pH 7.5, 5 mM MgCl2 , 0.5% acetone and 2 µM

[γ32P]ATP (80 µCi) for 30 min on ice in the dark. A 40 µl aliquot was

placed in an open Eppendorf tube incubated on ice and exposed to 254 nm

light from a UV cross-linker device (Stratagene) for 15, 30, 45 and 60 min

at 125 J s-1 at a distance of 5 cm from the light. The reactions were

stopped by boiling 10 µl of each mixture for 3 min in sample buffer and

then separating the proteins in Laemmli gels followed by autoradiography.

5.2.12 Pathogenicity tests

Pathogenicity tests were performed by inoculating the whorls of 5 to 6-

day-old sweet corn seedlings (Zea mays var. saccharata cv. “Seneca

Horizon”) as previously described (Coplin et al. 1986). The plants were 270 held in growth chambers at 29°C (photoperiod 16 hr, 15000 lux, relative humidity 99%). Each corn whorl was inoculated with 50 µl of ca. A600nm=

0.1 suspensions of P. stewartii cells in 10 mM potassium phosphate buffer

pH6.8, 0.2% Tween 40. After 3-4 days, disease severity was rated using a

0 to 3 scale (0= no symptoms, 1=scattered small lesions, 2 = numerous lesions and 3 = extensive lesions that remained water-soaked with ooze forming on leaf surfaces).

5.3 RESULTS

5.3.1 The input region of P. stewartii HrpX is formed by two PAS

domains with low similarity between them

In order to rationally create a subdomain deletion in hrpX, we decided

to study in more detail the primary and secondary structure of HrpX. The

amino acid sequence of HrpX was initially analyzed by PSI-BLAST and

CDD searches at the NCBI web site to define potential orthologs and

identify functional domains. The closest orthologs in the erwinias and two

paralogs (hypothesized to be orthologs) in the genus Pseudomonas are

described in Chapter 2. Here their sequences are compared by ClustalX.

Branches of the derived unrooted trees were bootstrapped using the

neighbor-joining method (1000 resamplings). The resulting tree (Fig. 5.5)

shows phylogenetic relationships consistent with the taxonomic position

of the species.

271 Two PAS domains comprised 90% of the input region of the sensor kinase. The first PAS domain (PAS1) spanned residues 16 to 130 and the second domain (PAS2) was located between residues 142 to 259. The C- terminal region of PAS domains, which contains the β-scaffold, is usually poorly conserved, so these coordinates have a higher margin of error in that position. The two domains were 21.2% identical to each other (28% similarity). Based on the CDD search results, the PAS1 core was 100% identical to the consensus motif Pfam00989 (E=1x10-5), whereas PAS2

core was only 89.2% identical to it (E=0.002). Pfam00989 contains only

the highly conserved PAS core, not the helical connector and the β-

scaffold, but similar results were obtained from comparisons with

consensus motif Cd00130.1, which also contains the C-terminal portion of

the PAS domain. The secondary structure of HrpX was predicted using

PSIPRED. The PAS1 domain has a typical α/β fold (topology:

AβBβCαDαEαFαGβHβIβ). Two independent PSI-PRED analyses

predicted that the short linker between PAS1 and PAS2 may fold either as

a β-strand/loop (Fig. 5.6) or as a helix (not shown), but the statistical

confidence was only moderate in both analyses. PAS2 appears to be

missing the Eα helix and contains a slightly longer Dα helix. The pocket

between the Eα and Fα helices and the Gβ, Hβ and Iβ strands in FixL and

Phy3 PAS domains were shown by NMR and X-ray crystallography to

form the ligand binding site (Amezcua et al., 2002). The differences at

this site between FixL and HrpX could reflect the fact that different 272 ligands are bound by the kinase. The HrpX PAS core domains were compared to other known PAS domains in an attempt to predict what ligand(s) might interact with HrpX. A CDD search produced a list of 50 domains most similar to PAS1 out of the many thousands found in

GenBank. It included, in order of decreasing similarity, Rhizobium and

Bradyrhizobium FixL (1EW0 and 1DRM), Rhizobium NodV (gi 128494),

E. coli NtrB (gi 128594), Bacillus subtilis KinC and KinA (gi 729901, gi

3041695), the Neurospora crassa protein WC1 (gi 2494692), and E. coli

NtrY (gi 585587) and ArcB (gi 1168485). The corresponding list using

PAS2 as a query included TraJ (gi 136225), PhoR (gi 1172485), NtrB (gi

128597) and ArcB (gi 1168485). The associated ligands are heme and oxygen for FixL and ATP for KinA; ArcB is believed to bind the respiratory chain cofactors MK3 and Q0 . The phylogenetic trees in Fig.

5.7 show the relative genetic distances between these PAS domains.

Unfortunately, this tree again demonstrates that, based on similarity alone, it is impossible to predict what ligand is bound by any given PAS domain, since WCI, FixL and KinA PAS modules cluster together but bind three very different molecules (FAD, heme, and ATP, respectively).

The threading of the HrpX PAS1 into the FixL PAS backbone crystal structure (1DRM) was performed with mGenTHREADER (Fig. 5.8) and the corresponding three-dimensional model is shown in Fig. 5.9.

Conserved residues are shown as ball-and-sticks. Automated homology modeling with the ProMod algorithm at the SwissModel server failed due

273 to a low level of identity (<25%) to the reference structure, and structural alignments with high RSM (~19Å) were obtained (not shown). This information was helpful in designing in frame deletions of HrpX that did not alter any secondary structure elements that might cause misfolding the mutant protein. These secondary and tertiary structural models may be a useful reference for future studies aimed at engineering mutations in the active site of the PAS structure that will help to identify the ligands.

5.3.2 The two PAS domains of HrpX play different roles in modulating its activity

In frame deletions in the hrpX gene were engineered in vitro to produce plasmids for complementation experiments and constructing chromosomal mutations in P. stewartii. The deletions were designed to preserve the secondary structural elements identified in Fig. 5.6 and minimize the risk of producing locally misfolded proteins. Deletion of the entire hrpX gene in strain DM2790 ∆hrpX17-414 is described in Chapter 2. When inoculated into corn whorls, the virulence of this mutant was only slightly reduced and its watersoaking severity rating was 2.0-2.5 as compared to 3.0 for the wild-type strain (Fig. 5.10). Deletion of both PAS domains in strains

MM266 ∆hrpX15-198 and MM284 ∆hrpX15-270 or just PAS2 in strain

MM360 ∆hrpX143-259 reduced the severity rating to about 2.0, which corresponds to numerous lesions but not collapse of entire leaves. In contrast, deletion of just the PAS1 domain in strain MM348 ∆hrpX15-130 274 strongly reduced virulence to a level below 1.0 (i.e. only a few small,

scattered lesions) (Fig. 5.10). All of the mutants were complemented to

full virulence by pMM389 hrpX+. Interestingly, strain MM360 (∆PAS2)

was fully complemented by pMM380 (∆PAS1), and, vice versa, strain

MM348 (∆PAS1) was complemented by pMM378 (∆PAS2). Intracistronic

complementation suggests that the HrpX protein may dimerize as shown

for other prokaryotic sensor kinases. These data indicate the critical

importance of the PAS1 domain for proper signal transduction in vivo. On

the other hand, PAS2, although required for full virulence, appears only

to cooperate with PAS1. In other words, PAS1 is necessary but not

sufficient for full HrpX activity. Expression of pMM380 in the wild-type

background did not affect virulence (not shown), ruling out negative

trans-dominance of the ∆PAS1 allele. This is somewhat unexpected, because the same allele appears to reduce the virulence of the strain

DM2790 ∆hrpX17-414. Such results imply an active suppression by the

HrpX∆PAS1 protein of the cellular cross-talk to HrpY occurring in this

genetic background (Merighi et al., 2003), while the same mutant protein

appears unable to compete with wild type HrpX.

The expression of a representative hrp secretion gene (hrpJ-uidA in

pDM2531) in the above hrpX mutants was not always correlated with their

virulence (Fig. 5.11). We observed that while the deletion of the entire

hrpX gene in DM2790 and the ∆PAS1-PAS2 deletion in mutant MM284

reduced hrp expression by 90% from the wild-type levels, these mutants

275 still produced 70% as much watersoaking as the wild-type strain. These relative levels of expression were comparable to those observed in planta by epifluorescence microscopy of the same strains carrying hrpJ-gfp fusions (not shown). This implies that the transmitter domain in the

∆(PAS1-PAS2) mutants is not functioning as a kinase, and that, both in planta and in vitro, cross-talk between HrpY and other kinases or small phosphodonors is controlling hrp gene expression in these genetic backgrounds (Chapter 2 and Fig. 5.11). On the other hand, both hrpJ

expression and virulence was drastically reduced in the ∆PAS1 mutant

MM348 (1.8% of wild-type). (Fig. 5.11). In contrast, the ∆PAS2 strain

(MM360) expressed hrp genes in IM5.5 at about 20% of the wild-type level but was only slightly affected in virulence (watersoking score ~2).

In summary, PAS1 seems to be absolutely required for both stimuli perception in vitro and virulence in planta. Therefore, it is possible that this truncated protein is locked in a conformation that suppresses cross- talk (as shown in strain DM2790/pMM380; Fig. 5.10) or/and the functioning of the PAS2 domain. For example, this could happen if it is locked in a phosphatase mode. On the other hand, PAS2 is still required for full activation of the sensor protein. One possibility is that PAS1 and

PAS2 sense separate signals and then act cooperatively to control the kinase domain. Another possibility is that PAS2 might function as an helper of PAS1, by being involved in dimerization of the sensor molecules, as shown for Bacillus KinA.

276 Mutant MM266 does not contain a full deletion of the PAS2 domain and

a small portion of PAC2 motif carrying the regulatory Fα helix-F-G loop is preserved. hrp expression in MM266 is only 1.2% of the wild-type, similarly to the ∆PAS1 strain MM348. However, differently from MM348, virulence was not affected as severely. The reason for these differences is unclear.

5.3.3 HrpX does not transduce pH, osmolarity or many catabolic metabolite-associated signals but it appears to sense stimuli related to

nitrogen metabolism and the Krebs cycle

As shown in Chapter 2, many different environmental signals and/or

growth conditions affect the expression of hrp genes. To start dissecting

which of these stimuli are perceived by the HrpX/HrpY signal

transduction system, we determined which repressing conditions could be

overridden by overexpression of HrpY and ergo might be affecting HrpY

activation. This strategy assumes that other conditions operating directly

on hrpS would not affect HrpY overexpression. To do this we exploited

the observation that ectopic expression of the hrpY[D57N] allele in

pMM118 activates hrpS expression independently of any phosphorylation

event catalyzed by HrpX. The use of the mutant hrpY[D57N] gene instead

of wild-type hrpY eliminates the risk that phosphorylation or

dephosphorylation of HrpY might confuse (dampen or stimulate) its

effects on hrpS. P. stewartii strain DM733 hrpY::Tn5 hrpS-lacZ was used 277 as a recipient for pMM118 (Fig. 5.12). DM733 has very low basal β- galactosidase activity and was used as negative control in all experiments

(data not shown). Strain MEX1 hrpS-lacZ was used a positive control and enzyme activity measurements were corrected for the basal β- galactosidase activity in DC283. Bacteria were grown in IM5.5A with one or two chemical variables changed at a time, β-galactosidase activity was measured and data were compared to the expression of each strain in

IM5.5A (data in Fig. 5.12 expressed as a percentage of activity in

IM5.5A).

Elevated osmolarity due to potassium phosphate or sodium chloride caused repression of hrpS in the MEX1 hrpS-lacZ control strain and overexpression of hrpY[D57N] did not override this effect (Fig. 5.12).

Likewise, the reduced expression of hrpS in media with high pH,

asparagine as a sole source of C and N, nicotinic acid at 1 mM or peptone

was not increased by hrpY overexpression. Curiously, at higher population

density, repression of hrpS in cells grown in IM5.5 with asparagine as the

sole C/N source was overcome by pMM118 in one experiment. The

expression hrpS-lacZ fusion was strongly reduced by growth in IM5.5

with tryptone, alone and in combination with sucrose (4.5-fold repression)

and overexpression of hrpY[D57N] restored hrpS transcription. In

contrast, the low hrpS-lacZ expression observed in LB broth, which also

contains tryptone, was not increased by pMM118. Presumably, this is

because LB broth also contains high salt concentrations (0.5%). The 278 downregulation of hrpS by growth in IM with citric acid or succinate as sole C source was variable and ranged from 40-60%, compared to the

IM5.5 control, but hrpS transcription was always fully restored by

pMM118. Other respiratory catabolites, such as pyruvate, aspartate and

glutamate, did not have any significant effect on hrpS expression (not

shown).

5.3.4 Expression and Purification of recombinant HrpX proteins

Several types of vectors and polypeptide fusions were used to engineer

versions of HrpX that could be purified by affinity chromatography. This

was necessary because initial experiments with a His6 -HrpX fusion

indicated that we might encounter solubility problems. Consequently,

various combinations of tags, promoters of different strength and fusions

to various locations in HrpX were constructed so that we could be ready

to switch expression construct/purification chemistry if required.

Initially, His6 -HrpX was overexpressed from pMM228, which uses the

T7lac promoter in a pET-15b vector. A PCR insert was engineered to

maintain the correct translation frame with a vector-encoded N-terminal

polyhistidine tag (ca. 2 kDa in size). The his6 -hrpX gene was also

subcloned from pMM228 into pEXT20 as a HindIII+XbaI fragment for

expression from Ptac (plasmid pMM319). Two additional His6 -tagged versions of this gene were constructed in pET28b and pQE30 (Fig. 5.3) to express proteins with minimal size tags in C- and N-terminal positions

279 respectively. Other tags were also engineered in order to increase the recombinant protein solubility and yield. In particular, N-terminal genes expressing the maltose binding protein, encoded by malE (MBP), and the

glutathione S-transferase polypeptide (GST) were translationally fused to

hrpX and expressed from pMM315 and pMM371, respectively. Finally, a

C-terminal fusion of hrpX to the Intein-Chitin binding domain encoded in

plasmid pTYB1 was engineered in plasmid pMM362. This last construct is

expected to have increased solubility, comparable to the pMM315 MBP-

HrpX fusions, but the removal of the tag is performed in vitro using an

autocatalytic mechanism triggered by the intein domain in presence of

high concentrations of reductants, such as DTT. Most of these constructs

were prepared recently as alternatives to pMM228 and pMM315, which

gave mixed results, as shown below.

Induction of the T7lac promoter of plasmid pMM228 in E. coli

BL21(DE3)(pLysS) by IPTG resulted in the production of a ca. 63-kDa

polypeptide visible in total cell protein lysates, but not present in

uninduced cells (Fig. 5.13). Fig. 5.14 shows the purification steps by immobilized metal affinity chromatography (IMAC) of the poly(His) tagged HrpX proteins using step-elution gradients. Highly purified His6 -

HrpX was concentrated with a VivaSpin 15R ultrafiltration device and the

purified fraction dialyzed against storage buffer is shown in Fig. 5.14 -B.

The purity of the samples was assessed by densitometric analysis of

Coomassie-stained SDS-PAGE gels and by nano-LC-MS/MS spectrometry.

280 In many cases, the purity indicated by SDS-PAGE was >80%, but a

second band of about 30 kDa always copurified with the 63-kDa band,

which is close to the expected size of 58 kDa. Small alterations of

electrophoretic mobility in Laemmli gels are observed for proteins with

hydrophobic domains or with a high proportion of positively charged

amino acids. Western blot analysis using anti-His monoclonal antibodies

confirmed both the 63-kDa and the 30-kDa bands were His-tagged

proteins (Fig. 5.14, C). This suggests that the 30 kDa protein may be a

proteolytic fragment from HrpX (probably corresponding to the PAS1-2

input domain; amino acids 1 to 260, ca. 30 kDa) released during the

purification procedure. This may be due to the fact that metalloprotease

inhibitors were not added because they would interfere with the

immobilized metal affinity chromatography. The average yield was of

0.25 mg of purified soluble protein per liter of culture.

To determine the sequence identity of the 63-kDa polypeptide, the band was excised from an SDS-PAGE gel and subjected to tryptic digestion.

The carbamidomethyl-modified peptides were separated by liquid chromatography and individually fragmented in a MS/MS detector (Fig.

5.15). Thirty peptides, corresponding to 46% of the sequence of the protein, were isolated and characterized. Peptide sequences were matched by Mascott searches to the GeBank accession gi 9885631 corresponding to

HrpX. This His6 -HrpX preparation was used for the autophosphorylation and transphosphorylation experiments described below. The major

281 problem with the recombinant HrpX expressed from pMM228 was its low solubility and yield and the difficulty of obtaining reproducible results between purification batches. This limited the number and complexity of experiments we conducted with His6 -HrpX from pMM228 and forced us to use different His6 -tagged constructs. We attempted various strategies,

including the reduction of the expression level by using weaker Ptac

promoters (plasmid pMM319), the repositioning of the polyHis tag in the

C-terminal position (plasmid pMM367), and the engineering of minimal

polyHis tags (plasmid pMM377). We also tried the expression and

purification of HrpX in P. stewartii DM2790(pMM319), hoping that the

presence of the cognate response regulator might stabilize the

overexpressed protein. None of these approaches helped in solving the

solubility problems.

We then switched to other purification chemistries involving the use of

larger, well structured and soluble recombinant tags. The production of

highly soluble MBP-HrpX recombinant proteins was achieved by

expressing pMM315 in E. coli DH10, ER2507 or TB1. The expected 100-

kDa fusion protein was observed in IPTG-induced samples (Fig. 5.16) and

amylose affinity chromatography was used to purify it. A large amount of

recombinant protein was obtained (ca. 10 mg l-1 of culture). Proteolytic

fragments of ca. 70 kDa were present in spite of the use of a wide-

spectrum protease inhibitor cocktail. We did not attempt to use E. coli

backgrounds with mutations in housekeeping protease systems. The

282 protein preparations eluted from the amylose column were hydrolyzed by factor Xa to release the MBP tag, which was subsequently removed by ion-exchange chromatography (Fig. 5.16). The final yield of soluble tag- less HrpX was 5.6 mg l-1. Removal of the tag was necessary because the

MBP-HrpX fusion protein had extremely weak autophosphorylation activity (not shown), probably due to interference by the N-terminal MBP polypeptide. The caveat was that the removal of the tag resulted in totally inactive kinase preparations

Plasmids pMM371 and pMM362 were only verified for the production of the expected recombinant protein and not tested exhaustively for the purification of soluble recombinant proteins. These are available for future experimentation and should facilitate the preparation of larger amounts of soluble, active HrpX.

5.3.5 Autophosphorylation kinetics and phosphotransfer ability of

HrpX

This section describes functional experiments performed with the only active HrpX preparation obtained. The ability of His6 -HrpX to autophosphorylate was assessed in vitro. The purpose of this experiment was to optimize the conditions for maximal phosphorylation of the sensor kinase so that we could use it to phosphorylate HrpY. In addition, this

283 information will be needed for future studies to test if potential, known

ligand molecules can interact with the PAS domains and affect

autophosphorylation rates in end-point experiments.

Clear autophosphorylation was demonstrated with one preparation of

His6 -HrpX (Fig. 5.17). This was maximal after 60 min of incubation with

0.22 µM of [γ32P]ATP. We were able to use this enzyme preparation to perform a preliminary experiment to test the ability of several known PAS ligands to modulate the rate of autophosphorylation of HrpX. These results indicated that ADP inhibits the reaction, while AMP and menaquinone weakly stimulate it (not shown). A kinetic analysis to determine their effect on the reaction rate was not done for lack of enzyme.

Multiple subsequent attempts to re-purify His6 -HrpX for further

experimentation yielded very little protein and the preparations had no

measurable autokinase activity. Furthermore, the larger amounts of

soluble HrpX obtained from MBP-tagged proteins also showed little or no

autophosphorylation activity (Fig. 5.18). When either form of the protein

was used in transphosphorylation reactions with HrpY, we never observed any clear signals corresponding to HrpY~32P, except in one experiment

using the His6 -HrpX preparation where we observed a radioactive signal

corresponding to a 26-kDa band after 5 to 10 min incubation with

suboptimal concentrations of HrpX (Fig. 5.18). Salmonella BarA/HrpY

reaction mixtures were used as positive controls. The difficulties in

284 obtaining larger amounts of soluble His6 -HrpX halted this line of experiments. Our difficulties in repurifying His6 -HrpX were quite

frustrating. Similar problems of irreproducible purification of soluble

proteins with large hydrophobic domains in general, and PAS-kinases in

particular have been described. These problems are especially exacerbated

by the absence of sufficient amounts of the cognate ligand in cells

overexpressing the recombinant protein (Dr. K. Gardner, pers. comm.).

PAS domains in absence of the cognate ligand tend to compact in an

astructured conformation. To this respect, in two instances we attempted

without success to add ATP and FAD to cleared cell lysates hoping to

increase the protein solubility. On the other hand, we suspect that the

absence of activity for the tag-less MBP-HrpX derivative is due to the

digestion of the preparations at room temperature for 24 h to efficiently release the MBP tag. The kinase may have lost activity during this time.

In the future, it might be better to use highly soluble tags that can be

released more quickly at low temperature (e.g. pMM362).

5.3.6 GST-PAS1-2 does not bind ATP

Some PAS kinases, such as Bacillus KinA (Stephenson et al., 2000), use

ATP to sense internal energy levels in the cell. To test if HrpX may be

sensing ATP levels, we tried to purify subdomains corresponding to the

input region of HrpX. His-tagged recombinant proteins (Fig. 5.13) were

insoluble, and instead of trying to repurify them from the inclusion

285 bodies, we decided to engineer more soluble tags in the C-terminus.

Several other expression plasmids were constructed (Fig. 5.4) and small- scale purifications were attempted. Plasmids pMM270 and pMM265 did not show any upregulated polypeptide upon IPTG induction, possibly because of the instability of the recombinant protein. A soluble recombinant GST-PAS1-2 protein was successfully purified from a small- scale batch of E. coli TB1 (pMM374) (Fig. 5.19A). The recombinant protein spans the whole input domain of HrpX. The protein was soluble, with only a little localizing in inclusion bodies. For the purification, we used a simple affinity column without downstream purification and the protein appear >90% pure in a Coomassie-stained gel. The protein was incubated with a large molar excess of [γ32P]ATP. A clear signal corresponding to the size of the GST-PAS1-2 fragment was not observed

(Fig. 5.19A), so it is unlikely that HrpX binds ATP. As a positive control, we showed that exposure of the reaction mixture to UV light resulted in cross-linking of radiolabeled ATP to background proteins that were not visible in Coomassie-stained gels of the GST-PAS1-2 preparation (Fig.

5.19A). These experiments should be repeated with higher concentrations of more purified protein and with batches of GST-PAS1-2 with the GST tag removed. It is possible that the tag could be interfering with the binding, although this is unlikely due to the small size of the ligand involved.

286

5.4 DISCUSSION

P. stewartii HrpX is a sensor kinase that is important for the

transduction of signals generated by growth in IM5.5, an apoplast

mimicking medium, and for full and consistent virulence on corn

seedlings (Merighi et al., 2003). The input domain of HrpX is composed

of two tandem PAS domains, designated PAS1 and PAS2. In this study,

we found that HrpX PAS1 and PAS2 are not redundant. Their sequence

similarity is low (28% similarity; 21.2% identity) and the effects of

deleting PAS1 or PAS2 are not identical. PAS1 appears to be essential for

signal sensing in IM5.5 (Fig. 5.10) and it is very important for virulence

on corn. In contrast, PAS2 appears to play a secondary role in virulence.

Specifically, the virulence of the ∆PAS1 (∆hrpX15-130) strain was about 1

(“small scattered lesions”), while that of the ∆PAS2 mutant (∆hrpX143-259)

was about 2 (“many lesions”). In this regard, the ∆PAS2 mutant is roughly

comparable to the ∆hrpX17-414 strain (DM2790) (Merighi et al., 2003)

(Fig. 5.10). Thus, PAS2 does appear to be involved in signal sensing in

IM5.5 (Fig. 5.10) but to a much lesser extent than PAS1. It is important to

note that the ∆hrpX15-130 allele does not seem to be negatively dominant,

because it did not reduce virulence in the wild-type background. On the

other hand, its expression in the ∆hrpX17-414 strain seemed to eliminate cross-talk with HrpY, as inferred by a reduction in the virulence of this strain (Fig. 5.10). If the expression of hrp genes in a wild-type strain is 287 affected by the trans-expression of the ∆hrpX15-130 allele is not known.

Finally, intragenic complementation suggests that HrpX molecules can oligomerize.

In other sensor proteins, PAS domains have been shown to control the autophosphorylation activity of the kinase core. Ligand binding to the

PAS pocket leads to structural rearrangements of the Fα-FG-Hβ interaction surface in the PAS fold and modulation of the kinase activity of the catalytic domain (Amezcua et al., 2002). This happens even when purified PAS domain fragments are added to in vitro reactions containing the kinase domain (Rutter et al., 2001). Our genetic data seem to suggest that the two PAS domains in HrpX cooperatively interact to control HrpX autophosphorylation, but that ligand interactions at PAS1 may be dominant over events at PAS2. Consistent with this model is the possibility that PAS2 may work also as a secondary dimerization domain, similar to what has been described for KinA (Wang et al., 2001). We have started to analyze the protein-protein interactions of the individual PAS domains with each other and with the kinase core using a yeast two-hybrid system (Merighi, Ham and Coplin, unpublished), but so far we have not been able to show interactions between the input domain and full-length

HrpX protein fused to CytoTrap tags. The mapping of these interactions may provide insights into the mechanism of regulation of the kinase core by the PAS modules.

288 The nature of the ligand/signal sensed by the PAS domains is of

particular theoretical and practical interest. Very few PAS domain ligands

have been identified under natural conditions. Among those known are 4-

hydroxy-cinnamic acid, FAD, heme, ATP and FMN. A recent NMR screen of 750 organic compounds showed that the hydrophobic core of the hPASK PAS-A domain can specifically bind small molecules (Amezcua

et al., 2002). In this system and several others, ligand binding is

important for the transition between the unstructured state of the

apoprotein and the folded state of the conjugated protein. The tendency

apo-PAS domains to precipitate in the absence of ligand may explain why

we could not obtain soluble His6 -tagged HrpX and PAS proteins, while

addition of highly structured tags, such as MBP or GST, allowed the

purification of reasonable amounts of HrpX or PAS1-2 protein. The low

activity of the purified proteins (especially the MBP-HrpX version), in

addition to thermal inactivation for the reason discussed above, may be

due to loss of the cofactor during purification, especially during dialysis.

Using the only batch of soluble His6 -HrpX we were able to obtain, we

demonstrated rapid autophosphorylation of HrpX, which reached

saturation around 60 minutes from the start of the reaction (Fig. 5.19).

This is comparable to what has been observed for other sensors (Loh et

al., 1997). We do not know if any ligand was copurified with the His6 -

HrpX protein. Absorbance spectral data did not reveal any signals

consistent with the binding of flavin cofactors (not shown). An attempted

289 transphosphorylation experiment with the same protein preparation resulted in a weak signal consistent with the production of HrpY~P, but the negative control was ambiguous due to the presence of background kinases (Fig. 5.18). In future experiments, these problems might be overcome by the use of pMM362 as expression plasmid, which should allow for the expression of a larger amount of soluble protein and easy removal of the tag at low temperature (Fig. 5.3). More work should also be done on refolding denatured HrpX, which is present in large and highly purified amounts in inclusion bodies.

At this point, we do not know the nature of the ligand(s) that interacts with HrpX. Sequence analysis alone is unable to predict its nature, because of the generally low similarity (around 20-28%) among various

PAS domains and the small number of ligands known so far. Only five crystal or solution structures of PAS proteins are known and the residues important for ligand binding have been identified in only a few of these systems (Repik et al., 2000; Dunham et al., 2003; Gerharz et al., 2003).

The PAS domains most similar to HrpX are found in NodV PAS-A (for

PAS1) and TraJ (for PAS2) and the ligands for these have been described.

NodV senses genistein signals during the interaction with legume plants

(Loh et al., 1997) and its loss leads to restriction of symbiotic host range

(Gottfert et al., 1990). NodV has five tandem PAS domains, but there is no proof that any of them are involved in genistein-sensing. TraJ PAS

290 domains represent an interesting hit. TraJ is a nucleotide-binding outer

membrane protein, unrelated to two-component system kinases, that regulates conjugal transfer of F factors.

In our attempts to identify the growth conditions and signals transduced by HrpX/HrpY, we discovered that HrpY[D57N], when expressed from

multicopy plasmids, could overcome the repression of hrp genes by either

citrate or succinate were supplied as the sole C-source and, to a lesser

extent, by 0.125% tryptone. In contrast, the other environmental

conditions that modulate hrpS expression, such as pH, osmolarity,

nicotinic acid and peptone, were independent of HrpY regulation. The significance of the effect of citrate, succinate and tryptone is still under investigation. Citrate and succinate are important intermediates of the

tricarboxylic acid and glyoxylate cycles. Tryptone, the major component

of LB broth, contains large amounts of free amino acids and small

peptides. Aerobic growth in media containing tryptone as sole carbon and

nitrogen source should activate gluconeogenic and anaplerotic pathways

feeding into the Krebs cycle. Whether the particular amino acid

composition of tryptone leads to a build up of Kreb’s cycle intermediates,

or else, whether the growth in minimal media with succinate or citrate as

sole C-source alter particular amino acid synthesis pathways is not

known. Intriguingly, the periplasmic PAS domain of the E. coli CitA

sensor, which regulates the transport and metabolism of citrate, has

recently been shown to cocrystalize with citric acid molecules bound to

291 the PAS core pocket (Gerharz et al., 2003; Reinelt et al., 2003). By analogy, it is tempting to hypothesize that either citrate or succinate may function as a ligand for one of the HrpX PAS domains, but it is also likely that they may indirectly affect it. In this model the HrpX-citrate conjugate would be locked in a HrpY-phosphatase conformation. At this point, we have only shown modest repression using these two Kreb’s cycle intermediates (only 60 to 40% downregulation compared to the positive control) and it may not be significant from a biological point of view. In our suppression experiments, we tried to control most of the variables, osmotic pressure in particular, by balancing chemical concentrations of ions in the various media tested. Since we could not perform direct osmotic measurements, we used the van t’ Hoff law to approximate their values. When using complex organic substrates such as tryptone, these calculations are impossible. Therefore, in the IM5.5 tryptone treatments some of the repression may be due to the osmotic pressure. This could explain why hrpY[D57N] did not fully overcome the repression (70% suppression) (Fig. 5.12) caused by peptone, because we used 2% peptone versus 0.1% tryptone.

A biochemical approach is probably the best avenue for identifying the molecular ligands for HrpX. If soluble HrpX PAS domains are eventually purified in large amounts, the determination of their solution structure by

292 NMR and the analysis of their chemical shifts in presence of putative cofactors could quickly identify what ligands they bind, as shown for other PAS domains (Amezcua et al., 2002).

From a practical point of view, the study of the biochemistry of HrpX and the identification of its ligands may open the way to large throughput screenings for structurally analogous compounds that could inhibit kinase activation. PAS domains are found in sensors of many other plant and animal pathogens, and, if the molecular basis for their ligand specificity is eventually understood, they could become novel and universal targets for therapeutic drugs and plant disease control chemicals.

5.5 SUMMARY

P. stewartii HrpX/HrpY two-component system activates the hrp/hrc type III secretion system through a regulatory cascade involving the HrpS enhancer-binding protein and the HrpL alternative sigma factor. The HrpX sensor kinase is required for full activation of hrp regulatory and secretion genes in vitro and for consistent infection of corn seedlings. The signal(s) controlling HrpX are not known. The input domain of the sensor is composed of two tandem PAS domains showing little similarity to each others. The presence of PAS domains and the predicted cytoplasmic localization of the sensor suggest that it could be sensing cellular redox levels, the energy charge or other internal metabolic signals.

293 In this study we begun testing the role of the PAS domains in HrpX

activity. We first engineered in frame deletions spanning both individual

motifs and the whole input domain. The effects of the new alleles on

virulence and hrp gene expression were analyzed by corn whorl

inoculations and reporter gene assays. The results suggest that the two

PAS domains are non-redundant, are both required to different extent for

full virulence and a ∆PAS1 deletion had negative effect on virulence when

expressed in a ∆hrpX or ∆(PAS1-PAS2)

We then tried to determine which of the many environmental signals

controlling hrp gene transcriptions are perceived by HrpX. Given that we

were unable to isolate a constitutive HrpX* mutant, we used

overexpression of hrpY[D57N] to attempt to suppress the downregulatory

effects of high pH, high osmolarity, nicotinic acid, complex nitrogen

sources, and some organic acids on a hrpS::lacZ reporter fusion. We

indirectly found that the HrpX/HrpY mediates the effect of tryptone, citrate and succinate but none of the other growth conditions used.

Experiments aimed to establish reagents for the biochemical analysis of

HrpX and the two PAS domains are also presented in this chapter.

294 Strain, or plasmid Relevant phenotype(s) and/or genotype(s) Source or reference

Escherichia coli

DH10B F- mcrA ∆(mrr-hsdRMS-mcrBC) φ80dlacZ∆M15 ∆lacX74 endA1 recA1 deoR

∆(ara,leu)7697 araD139 galU galK nupG rpsL λ- Invitrogen

- - - r HB101 F thi-1 hsd20 (r Bm B) sup E44 recA13 ara-14 leuB6 proA2 lacY1 rpsL20 (Sm )

xyl-5 mtl-1 Boyer and Roulland-Dussoix (1969)

BL21(DE3)/pLysS F- ompT hsdSB (rB- mB-) gal dcm (DE3) carrying pLysS Novagen

- R TB1 F ara ∆(lac-proAB) [φ80 dlac ∆(lacZ)M15] rpsL (Str ) thi hsdR Novagen

- R ER2507 F ara-14 leuB6 fhuA2 ∆(argF-lac)U169 lacY1 glnV44 gal44 galK2 rpsL-20 (Str )

R xyl-5 mtl-5 ∆(malB) zjc::Tn5 (Kan ) ∆(mcrC-mrr)HB101 Novagen

P. stewarti subsp. stewartiii

SS104 Wild type, Wts+, HR+ ICPPBa

DC283 SS104 Nalr Coplin et al. (1986)

DM064 DC283 hrpY1296::Tn5 (Kanr) Frederick et al. (2001)

DM733 DC283 hrpY64::Tn5 hrpS1::Tn3HoHoI (Kanr, Ampr) Merighi et al. (2003)

DM2790 DC283 ∆hrpX17-414 Merighi et al. (2003)

MM254 DC283 hrpY[D57N] Merighi et al. (2003)

MM266 DC283 ∆hrpX15-198 This study

MM284 DC283 ∆hrpX15-270 This study

MM348 DC283 ∆hrpX15-130 This study

MM360 DC283 ∆hrpX143-259 This study

Plasmids pBluescript

KS and SK (+) ColE1 αlacZ (ApR) cloning vector Stratagene pPL6GUSC pLAFR6 derivative carrying a promoterless uidA gene (TcR) Knoop et al. (1991)

R Q pQE30 ColE1 (Ap ) Ptac lacI N-terminal His6-tag expression plasmid Qiagen

pRK415 IncP αlacZ (TcR) Keen et al. (1988)

R Q pTYB1 ColE1 (Ap ), PT7 lacI C-terminal SceVMW intein-CBD tagexpression plasmid NEB

Table 5.1 (continued)

295 Table 5.1 (continued)

r pET15b ColE1 (Ap ) PT7lac lacI N-terminal His6-tag expression plasmid Novagen

r pET28b ColE1 (Km ) PT7lac lacI C-terminal His6-tag expression plasmid Novagen

r pET41b ColE1 PT7lac lacI (Km ) N-terminal GST-S-His-tag expression plasmid Novagen

R Q pEXT20 ColE1 (Ap ) Ptac lacI expression plasmid Wilson et al., 1996

r Q pGEX-2T ColE1 (Ap ) Ptac lacI N-terminal GST- -tag expression plasmid Pharmacia

r Q pMALc2X ColE1 (Ap ) Ptac lacI N-terminal malE fusion expression plasmid NEB

pMM58 pBSK carrying hrpL, hrpXY and hrpS Merighi et al., 2003

+ pMM228 pET15b with a 1.5 kb BamHI hrpX PCR fragment to express a His6-HrpX[1-485] protein fusion This study pMM233 pET15b with a 0.6 kb BamHI ∆hrpX PCR fragment to express His6-HrpX[277-485] protein fusion This study

pMM242 pBSK with a 1.3 kb EcoRI-XbaI fragment carrying a ∆hrpX15-270 allele This study pMM243 pBSK with a 1.3 kb EcoRI-XbaI fragment carrying a ∆hrpX15-198 allele This study

pMM247 pLD55 with a XhoI-SstI insert from pMM243 This study

pMM249 pLD55 with a XhoI-SstI insert from pMM242 This study

pMM251 pET15b with a 0.4 kb BamHI ∆hrpX PCR fragment to express a His6-HrpX[1-142] protein fusion This study pMM257 pET15b with a 0.8 kb BamHI ∆hrpX PCR fragment to express a His6-HrpX[1-276] protein fusion This study pMM265 pET41b with a 0.4 kb BamHI ∆hrpX PCR fragment to express a GST-HrpX[1-142] protein fusion This study pMM270 pET41b with a 0.8 kb BamHI ∆hrpX PCR fragment to express a GST-HrpX[1-276] protein fusion This study pMM280 pET15b with a 0.4 kb BamHI ∆hrpX PCR fragment to express a His6-HrpX[136-276] protein fusion This study pMM283 pBSK with a BamHI-EcoRI insert carrying a SOE-PCR ∆hrpX15-130 allele This study

pMM287 pBSK with a BamHI-EcoRI insert carrying a SOE-PCR ∆hrpX143-259 allele This study

pMM315 pMALc2x with a 1.5 kb BamHI hrpX+ PCR fragment to express a MBP-HrpX[1-485] protein fusion This study

+ pMM319 pEXT20 with the HindIII-XbaI His6-hrpX fragment of pMM228 This study pMM324 pLD55 with a XhoI-SstI insert from pMM283 This study pMM327 pLD55 with a XhoI-SstI insert from pMM287 This study pMM362 pTYB1 with a 1.5 kb BamHI hrpX+ PCR fragment to express a HrpX[1-485]-SceCBD fusion This study pMM363 pTYB1 with a 0.8 kb NheI-XhoI hrpX PCR fragment to express a HrpX[1-276]- SceCBD fusion This study pMM367 pET28b with a 1.5 kb BamHI hrpX+ PCR fragment to express a HrpX[1-485]- His6 protein fusion This study

pMM371 pGEX-2T with a 1.5 kb BamHI hrpX+ PCR fragment to express a GST-HrpX[1-485] protein fusion This study

pMM374 pGEX-2T with a 0.8 kb BamHI hrpX PCR fragment to express a GST-HrpX[1-276] protein fusion This study

pMM377 pQE30 with a 1.5 kb BamHI hrpX+ PCR fragment to express a His6-HrpX[1-485] protein fusion This study pMM378 pRK415 with a BamHI-EcoRI insert from pMM287 This study

Table 5.1 (continued)

296 Table 5.1 (continued) pMM380 pRK415 with a BamHI-EcoRI insert from pMM283 This study

pMM389 pRK415 with a HindIII-EcoRI insert carrying hrpX+ driven by the native promooter This study

Table 5.1 Bacterial strains and plasmids.

297 Sequence coordinates

Primer Sequence (5’Æ 3’) (5’Æ3’)a Source

EdX1 CGCGAATTCTTCGGTATTGCCCTGAACCT 919-938 This study

BdX1 CGCGGATCCTTTGATTGGCGTATCACTGCT 1427-1407 This study

BdX5A CGCGGATCCATGCGCTTGCGTTTGTTAACA 2190-2210 This study

BdX5B CGCGGATCCCATATTGAACGCTTTCTTAAA 974-1994 This study

BdX4AB TCTAGAACGGCCGTTATCACTCACTG 2671-2687 This study

XdX3 CGCTCTAGATTTGCTCGGCGAGCATTTGG 3272-3253 This study

ETXF1 CGCGGATCCGATGCAGATTTTGGACAGAC 1379-1398 This study

ETXF2 CGCGGATCCGACAAGGCAGCGAGAAATATC 2207-2227 This study

ETXR GGATCCTCACTCACTCAGGGTATTAAAGGAATGGTC 2818-2837 This study

ETPAS1R CGCGGATCCCTATCATCTACTGTTCCCGAGCATAAGC 1788-1805 This study

ETPAS2F CGCGGATCCGGGTGTAGCTTATGCTCGGGAACAG 1785-1805 This study

ETPAS2R CGCGGATCCCTATCATCTACAAACGCAAGCGCATAT 2188-2205 This study

XCHISF2 CATGCCATGGAGATGCAGTATTTGGACAGAC 1378-1398 This study

XCHISR2 CCGCTCGAGGGGTATTAAAGGAATGGTC 2816-2834 This study

XBAMF(+1) CGCGGATCCATGCAGTATTTGGACAGAC 1380-1398 This study

XNHEF ACTAGCTAGCATGCAGTATTTGGACAGAC 1380-1398 This study

PAS2XHOR ATGCCTCGAGATGAATGGCTGTGATATC 2142-2159 This study

XXHOIR ATGCCTCGAGGGGTATTAAAGGAATGGTC 2816-2834 This study

dXPASF1 CGGAATTCAGAGGAAGACGGCAGTTATC 1122-1141 This study dXPAS1R1 ACCTGCTCTAGTTGCTTGATTGGCGTATCACTGCTATG 1404-1424 This study dXPAS1F2 AGTGATACGCCAATCAAGCAACTAGAGCAGGTAG 1770-1788 This study dXPASR2 CGGGATCCTCAGTCAGGGTATTAAAGGAATGGTCA 2815-2838 This study

dPAS2R1 CTTCTTTTCAGCGTAATGCTGTTCCCGAGCATAAGCTAC 1805-1785 This study

dPAS2F2 TATGCTCGGGAACAGCATTACGCTGAAAAGAAGTTAG 2157-2178 This study

XPMALR CGGAATTCTCACTCAGGGTATTAAAGGAATGGTC 2816-2837 This study

XPMALF GGGAATTCATGCAGTATTTGGACAGAC 1380-1398 This study

Table 5.2 Oligonucleotide primers.

298

hrpY 2837 hrpX 1380 hrpL

PAS2 PAS1

XdX3 BdX4 BdX1 EdX1 XbaI A BamHI BamHI B EcoRI pDM2842, ∆ pDM2848 1421 919

XdX4AB BdX5B XbaI CDBamHI BamHI EcoRI pMM243, pMM247 ∆ 2690 1974 1421 919

BdX5A XbaI EFBamHI BamHI EcoRI

pMM242, pMM249 ∆

2690 2190

dXPAS1F2 dXPAS1R1 GH BamHI EcoRI pMM283, pMM324 ∆

2837 1770 1425 1122

dXPASR2 dXPAS2F2 dXPAS2R1 dXPASF1 IL BamHI EcoRI pMM287, pMM327 ∆ 2837 2157 1805 1122

PshAI M FspI pMM389, pDM2890, pDM2893 2945 816

HindIII HindIII BamHI

5363 3579 1 pMM58

Figure 5.1. Map of primers and PCR fragments used to construct hrpX mutations

(plasmids and chromosomal mutants).

Plasmids carrying the various alleles are listed on the left and PCR primers are shown by arrowheads. Sequences of the primers are in Table 5.2. Nucleotide coordinates are relative to Genbank accession AF282857

299

DM2790 ∆17−414

MM284 ∆15−270

MM266 ∆15−198

PshAI FspI hrpX H hrpY PAS2 PAS1 hrpL 2837 1380

MM36 MM34 ∆143−259 ∆15−130

PshAI FspI pRK415 Plac pMM389

BamHI EcoRI

pRK415 Plac ∆ pMM378 BamHI EcoRI pRK415 Plac ∆ pMM380

Figure 5.2. Physical maps of the hrpX mutations and plasmids used for

complementation studies.

The coordinates of the deletions refer to the amino acid sequence of HrpX.

300

Figure 5.3. Plasmids used for expression and purification of HrpX.

Lines below the genetic map depict inserts cloned into various expression vectors.

Relevant restriction sites and promoters (P) are indicated. Plasmid names are listed on the right. The nature of the tag used for affinity column chromatography and the plasmid vector name is given in parentheses. Triangles indicate the proteolytic cleavage site for releasing the tag from the recombinant protein. Nucleotide coordinates are relative to Genbank accession AF282857. CBD/Sce = chitin binding/intein domains.

301

2837 hrpX 1380 PAS2 PAS1

ETXR ETXF BamHI BamHI (His tag pET15b) PT7 pMM228 6

ETXR XBAMF(+1) BamHI BamHI

Plac pMM377 (His6 tag pQE30) SG

XcHISR2 XcHISF2 XhoI NcoI

PT7 pMM367 (His6 tag pET28b)

XpMALR XpMALF EcoRI EcoRI

Ptac

Xa pMM315 (MBP tag pMALc2x)

ETXR XBAMF(+1) BamHI BamHI

Ptac

pMM371 (GST tag pGEX-2T)

XNHER XNheF XhoI NheI

CBD Sce PT7

SSGEL pMM362 (Intein-CB tag pTYB1)

302

Figure 5.4. Plasmids used for expression and purification of HrpX domains.

Lines below the genetic map depict inserts cloned into various expression vectors.

Relevant restriction sites and promoters (P) are indicated. Plasmid names are listed on the right. The nature of the tag used for affinity column chromatography and the plasmid vector name is given in parentheses. Triangles indicate the proteolytic cleavage site for releasing the tag from the recombinant protein. Nucleotide coordinates are relative to Genbank accession AF282857. CBD/Sce = chitin binding/intein domains.

303 2837 hrpX 1380 PAS2 PAS1

BamHI BamHI

PT7 pMM257 (His6 tag pET15b)

BamHI BamHI

PT7 pMM251 (His6 tag pET15b) BamHI BamHI

PT7 pMM280 (His6 tag pET15b) BamHI BamHI

PT7 pMM233 (His6 tag pET15b)

BamHI BamHI

Ptac

pMM270 (GST tag pET41b)

BamHI BamHI

Ptac

pMM265 (GST tag pET41b)

XhoI NheI

CBD Sce PT7 pMM363 (Intein-CB tag pTYB1) SSGEL MSAM

BamHI BamHI

Ptac SGR pMM374 (GST tag pGEX-2T)

304

P. agglomerans pv. gypsophilae P. stewartii E. amylovora

P. chrysanthemi

P. syringae*

P. aeruginosa* X. campestris

0.1

Figure 5.5. Phylogenetic tree of HrpX homologs.

HrpX orthologs, or presumed orthologs (*) found in the genome of P. stewartii, P.

agglomerans pv. gypsophiliae, P. chrysanthemum, E. amylovora, X. campestris, P.

syringae (PSPTO559; gi28851022) and P. aeruginosa (PAO600; gi9946473) were

aligned with ClustalX. Other accession numbers are shown in Table 2.3. Large gaps,

if any, were eliminated and a phylogenetic tree was constructed with the neighbor-

joining methods and 1000 bootstraps cycles. The unrooted tree is shown. Each branch

was supported by a bootstrap value >980. The bar indicates a genetic distance of 0.1.

305

Figure 5.6. PSI-PRED Secondary structure analysis of HrpX.

The amino acid sequence of HrpX was analyzed with PSI_PRED at http://bioinf.cs.ucl.ac.uk/ to identify the secondary structure elements. PAS domains are labeled following the convention of Amezcua et al., (2002). Conserved motifs in the transmitter domain were identified according to Hoch et al., (1995) and Stock et al., (1989). α-helices are shown as green cylinders, β-sheets are represented by yellow arrows and loops are shown as black lines.

306

PAS1

Aβ Bβ

Cα Dα Eα Fα

Gβ Hβ Iβ

PAS2

Aβ Bβ

Cα D/Eα? Fα

Gβ Hβ

linker

(continued)

307 Figure 5.6. (continued)

H~P H box

H-ATPase domain

N box D box

G box

α-helix CONF: CONFIDENCE OF PREDICTION β-strand Pred: prediction of secondary structure coil

Fig.B. PSI-PRED Secondary structure analysis of HrpX

308

PHYLIP_1 consensus

128494 NodV PAS1 PAS1

gi|2494692 WC1 gi|3041695 KinA gi|120207 FixL

6980436

8569635 gi|2145515 FixL

128594 NtrB gi|128596 NtrB 1171792 NtrB

gi|729901 KinC gi|1168485 ArcB gi|585587 NtrY

5107751 0.1

Figure 5.7. Phylogenetic tree of PAS1-like domains in other two component systems

The 14 PAS domains most similar to the P. stewartii PAS1 core domain were identified by a CDD search at NCBI and aligned with ClustalX. Accession numbers are shown next to the protein names. Large gaps were eliminated and a phylogenetic tree was constructed with the neighbor-joining method and 1000 bootstrap cycles.

TreeView was used to construct the cladogram above. The bar indicates a genetic distance of 0.1.

309

10 20 30 ------CCCEEEEEECCCCEEEECHHHHHHHCCCHHHHCCCC 1drmA0 ------IPDAMIVIDGHGIIQLFSTAAERLFGWSELEAIGQN * || || | |* |*| | * HrpX MQYLDRRLHSSDTPIKLLEFAMNMLRDAVYLIDSDTCFFYVNDEACRMLGYSNAELLNMR CHHHHHHHHHHHHHHHHHHHHHHHCCCCEEEECCCCCEEECCHHHHHHHCCCHHHHHCCC 10 20 30 40 50 60

40 50 60 70 80 90 HHHHCCCCHHHHHHHHHHHHHHHCCCCCCCCCEEEEEECCCCCEEEEEEEEEEEEECCEE 1drmA0 VNILMPEPDRSRHDSYISRYRTTSDPHIIGIGRIVTGKRRDGTTFPMHLSIGEMQSGGEP | * || | | | | | *| HrpX VSDIDPEWDFDDVMDMWKRVREQNGGR--PFTFETFHKSSQGEMIAVEISANPFVYEEKS HHHCCCCCCHHHHHHHHHHHHHHCCCC--EEEEEEEEECCCCCEEEEEEEEECCCCCCEE 70 80 90 100 110

100 110 EEEEEEEECHHHHHHHHHHHHCC------1drmA0 YFTGFVRDLTEHQQTQARLQELQ------| |*|* | | * * HrpX YSMCVVKDIRERKQLEQVAYAREQEFRVLVENSPDMVVRFSPDLKCQYANPAALRHLRLS EEEEEECCCCCCEEEEEEECCCCCCEEEEECCCCEEEEEECCCCEEEECCHHHHHHHCCC 130 140 150 160 170

Figure 5.8. Threading of the PAS1 amino acid sequence into the FixL PAS secondary structure.

The PAS1 domain of HrpX was threaded into the B. japonicum FixL PAS domain

using the mGenTREADER algorithm at http://bioinf.cs.ucl.ac.uk/. α-helical regions

are indicated by the letter H. β-sheets are labeled with the letter E. Random coil loops

are indicated by Cs. Identical residues are indicated with (|) and similar residues are

marked by (*).

310

Helical connector (Fα)

PAS core

β-scaffold

Ligand binding pocket

E67 P66 G100 K96 R81 V61

R79 E55 Eα Fα S52 Dα

Gβ G50 Bβ Cα V104 D33 Aβ Eβ Hβ

Y97 V102 I32 D29 A45 R47 S108 D28

Figure 5.9. Structural model of the P. stewartii PAS1 threaded into the backbone of B. japonicum FixL PAS.

311

Figure 5.10. Complementation analysis and virulence phenotypes of P. stewartii hrpX deletion mutants.

Inocula were prepared and plants were inoculated as described in Material and

Methods. Bacterial suspensions were adjusted to A600nm = 0.1. Whorls of 6-day-old corn seedlings were inoculated and watersoaking symptoms were scored at 4 days using a 0 to 3 scale. The cartoons on the left illustrate which part of the HrpX protein was deleted in the various mutants. Strains: DC283 is wild-type; DM2790 is

∆hrpX17−414, MM266 is ∆hrpX 15−198, MM284 is ∆hrpX 15−270, MM348 is ∆hrpX

15−130, MM360 is ∆hrpX 143−259. Plasmids: in all plasmids the expression of the gene is driven by the native hrpX promoter; pMM389 carries hrpX+ , pMM380 carries a ∆hrpX

15−130 allele and pMM378 carries ∆hrpX 143−259.

312 .

1 2 ~ DC283 MM360/pMM378 MM360/pMM380 1 ∆ ~ MM360/pMM389 MM360 MM348/pMM378 MM348/pMM380 ∆ 2 ~ MM348/pMM389 MM348

MM284/pMM378 A MM284/pMM380 ∆ ~ MM284/pMM389 MM284 MM266/pMM378 MM266/pMM380 ∆ ~ MM266/pMM389 MM266 DM2790/pMM378 DM2790/pMM380 ∆~ DM2790/pMM389 DM2790

00.511.522.53 Watersoaking

313

100 90 80 70 60 GUS activity % 50 Wts%

Percent 40 30 20 10 0

/2531 M2531 M2531 M2531 254 66/pD 48/pD 2 3 MM DC283/pDM2531M2790/pD D MM MM284/pDM2531MM MM360/pDM2531

Figure 5.11. The effect of various hrpX mutations on virulence and the expression of hrpJ. a typical hrp secretion gene.

Expression of the hrpJ-lacZ reporter gene fusion in pDM2531 was measured in P.

stewartii strains resuspended in IM5.5 to an A600=0.05-0.1 and grown for 16 h at

28°C. Data are from two sets of experiments with at least two replicates per strain.

Error bars indicate standard deviations (when not visible is because the SD is too

small for the scale of the graph). All data are given as a percentage of wild-type

levels. DM2790 is ∆hrpX17−414, MM266 is ∆hrpX15−198, MM284 is ∆hrpX15−270, MM348 is

∆hrpX15−130, MM360 is ∆hrpX143−259, MM254 is hrpY[D57N]. MM254(pDM2531) had

no GUS activity or virulence

314

Figure 5.12. The effect of environmental stimuli on hrpS-lacZ expression and suppression of these effects by ectopic overexpression of hrpY[D57N]

P. stewartii strains MEX1 hrpS-lacZ and its derivative DM733 hrpY::Tn5 hrpS-lacZ carrying pMM118 Plac-hrpY[D57N] were resuspended in IM5.5+supplement (initial

A 600nm = ~0.2-0.3) and grown at 28C for 12-16 h. Expression of hrpS-lacZ was measured and the data are shown as a percentage of the respective control strain grown in IM. Media: IM5.5A was standard IM inducing media lacking casaminoacids and with 10 mM sucrose as the C-source; “high Pi ” was IM5.5A with 250 mM

KH2 PO4 ; “High NaCl” was IM5.5A with 250 mM NaCl; “pH8.0” was IM buffered with

HEPES to pH 8.0; “citrate” was IM5.5A with 10 mM trisodium citrate as a sole C-

source; “succinate” was IM5.5A with 13 mM succinic acid as a sole C-source;

“asparagine” was IM5.5A with 56 mM L–asparagine as sole source for C and N;

“peptone” was IM5.5A with 2.0% peptone as a sole source for C and N; “tryptone”

was IM5.5A with 0.125% tryptone as sole source of C and N; “tryptone-sucrose” was

IM5.5A with 0.125% tryptone as sole N source. The pH of the media was adjusted

with NaOH and the milliequivalents of base were included in calculations of osmotic

pressures. Error bar = SD.

315 120 IM5.5 High Pi 100 High NaCl pH8.0 80 citrate succinate 60 asparagine

% IM5.5 control peptone 40 tryptone tryptone-sucrose 20 LB Nicotinic acid 0 MEX1 DM733/pMM118

316

Figure 5.13. Pilot experiment to overexpress polyHis-tagged HrpX proteins

E. coli BL21(DE3) pLysS cells carrying various expression plasmids were grown in

LB to A600=0.6 and induced with 1 mM IPTG for 3 h. Cells were concentrated 10- fold in PBS and lysed with reducing sample buffer. Total cell proteins (TCP) were run in a Laemmli gel. Lane 1, prestained molecular weight marker (Invitrogen), l, lane 2-

3, uninduced-IPTG induced cells carrying pMM223 (His6 -HrpS); lane 4-5, uninduced-

IPTG induced cells carrying pMM228 (His6 -HrpX); lane 6-7, uninduced-IPTG induced

cells carrying pMM233 (His6 -HKHrpX); lane 8-9, uninduced-IPTG induced cells

carrying pMM251 (His6 -PAS1); lane 10-11, uninduced-IPTG induced cells carrying pMM257 (His6 -PAS1-2). Subsequent analysis has shown that in all cases the proteins

precipitate in inclusion bodies.

317

1 2 3 4 5 6 7 8 9 10 11

78 kDa His -HrpX (lane 5) 61 kDa 6

His6-HrpS (lane 3)

His6-PAS1-2 (lane 11)

His6-HK (lane 11) 25 kDa

His6-PAS1 (lane 9) 13 kDa

318

Figure 5.14. Overexpression and purification of His6 -HrpX from E. coli

BL21(DE3) (pMM228).

(A) Coomassie-stained gel showing various steps in the purification of His6 -HrpX:

Lane 1, molecular weight markers (Invitrogen); lane 2, total cell protein (TCP) from uninduced cells; lane 3, TCP from IPTG-induced cells; lanes 4 to 10, eluted fractions at 40 mM, 60mM, 80 mM, 100 mM, 150 mM, 250 mM, and 500 mM imidazole. (B)

Anti-poly(His)6 Western blot analysis of purified fractions: lane 1, TCP from lane 2, panel A; lanes 2 to 4, fractions eluted at 40 mM, 100 mM and 150 mM imidazole. (C)

Fractions from a separate purification (used for MS/MS protein fingerprinting) eluted at 500 mM imidazole after extensive washing in 80 mM imidazole: Lane 1, molecular weight markers; lane 2, HrpX fraction in 500 mM imidazole. (D) Anti-poly(His)6

Western blot analysis of fractions from the purification of the HrpX preparation used for MS/MS analysis: Lane 1, TCP from uninduced cells (10X); lane 2, TCP from

IPTG-induced cells (10X); lane 3, cleared lysate from IPTG-induced cells loaded onto the Ni2+-NTA columns (30X); lane 4, inclusion bodies (30X); lane 5, column flow- through; lane 6, 500 mM imidazole eluted fraction (same as lane 2, panel C).

319 A 1 2 3 4 5 6 7 8 9 10

61 kDa His6-HrpX

B

α-Poly(His)6 1 2 3 4

His6-HrpX 61 kDa

C D 1 2

1 2 3 4 5 6 7 α-Poly(His)6

61 kDa

36 kDa 19 kDa

320

Figure 5.15. Example of the results from nano LC-MS/MS protein fingerprinting of the tryptic fragment LHSSDTPIK from gel-excised His6 -HrpX.

321

Figure 5.16. Overexpression and purification of MBP-HrpX by amylose affinity and ion exchange chromatography.

(A) Purification by amylose affinity chromatography from strain DH10B (pMM315):

Lane 1, prestained molecular weight marker (Invitrogen); lane 2, TCP from uninduced cells; lane 3, TCP from IPTG-induced cells; lane 4, amylose columns flow-through; lanes 5-6, wash fractions; lanes 7-15, eluted fractions using 10 mM maltose. (B)

Factor Xa proteolysis: Lane 1, prestained molecular weight marker (Invitrogen); lanes

2-4, various 10 mM maltose fractions from Panel A after pooling, ultrafiltration and storage at 4°C for 16 h; lane 5, undigested MBP-HrpX preparation (note again the presence of degradation fragments); lane 6, MBP-HrpX preparation from lane 5 digested for 6 h at 22°C with factor Xa to release the 56 kDa HrpX protein. (C) Ion exchange chromatography: lane 1, prestained molecular weight marker (Invitrogen); lanes 2-13, protein preparations digested with factor Xa for 24 h and purified by ion- exchange chromatography in a Q-sepharose column using a linear gradient of NaCl from 20 mM to 750 mM. (fractions from 20 to 320 mM are shown in the gel). (D).

Summary of MBP-HrpX purification steps: Lane 1, prestained molecular weight marker (Invitrogen); lane 2, TCP from uninduced cells; lane 3, TCP from IPTG- induced cells; lane 4, soluble total proteins; lane 5, flow-though from the amylose column; lane 6, concentrated 10 mM amylose eluted fraction; lane 7, Factor Xa digested preparation; lane 8, anion echange chromatography purified concentrated fraction.

322 A 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 113 kDa

81 kDa

1 2 3 4 5 6 7 B

63 kDa 56 kDa

C 1 2 3 4 5 6 7 8 9 10 11 12 13

63 kDa

D 1 2 3 4 5 6 7 8

63 kDa 56 kDa

323

A 0 2 5 10 15 30 60 min

HrpX~32P

B 160000

140000

120000

100000

80000

60000 Sum of pixels

40000

20000

0 0 20406080 minutes

32 Figure 5.17. Kinetics of autophosphorylation of His6 -HrpX by [γ P]ATP.

(A) Reaction mixtures containing 30 µM His6 -HrpX were incubated with 0.22 µM µCi

[γ 32P]ATP for increasing times as indicated in the figure. Reactions were terminated with stop sample buffer and separated in a Laemmli gel for subsequent autoradiography of the radiolabeled proteins. (B) The radioactive signals were quantified after exposure to a Phosphor storage cassette. The signals are reported in the plot as arbitrary units (sum of pixels in a defined constant sampling area).

324

Incubation time: 60 60 0 2 5 10 15 30 60 60 min

1 2 3 4 5 6 7 8 9 10 A 89 kDa

26.1 kDa

B 89 kDa 58 kDa BarA~32P HrpX~32P Background kinase

26.1 kDa HrpY~32P

32 32 Figure 5.18. Transphosphorylation of HrpY by BarA~ P and His6 -HrpX~ P.

Coomassie-stained PAGE gel (A) and autoradiography (B) of reaction mixtures containing [γ 32P]ATP and the following semipurified proteins: lane 1, BarA; lane 2,

25 pmol His6 -HrpX; lane 3, 380 pmol His6-HrpY; lanes 4 to 9, His6 -HrpX and His6 -

HrpY (1:15 molar ratio) incubated for 0, 2, 5, 10, 30 and 60 min; lane 10, BarA and

His6 -HrpY (1:1 molar ratio). Note the presence of a copurifying ~35-kDa uncharacterized background kinase in all preparations. The His6 -HrpY preparation also showed a moderate intensity signal at 26.1 kDa in the absence of any kinase. This observation prompted further purification of the His6 -HrpY protein by ion-exchange

chromatography as described in Chapter 4. The higher intensity signal at 26.1 kDa in

lanes 6 and 7 suggest a low level of phosphorylation of HrpY by HrpX~P.

325 A 1 2 3 4

66 kDa

B 1 2 3 4

Background kinase 66 kDa ? Background kinase

Figure 5.19. Purification of the GST-PAS1-2 protein from E. coli DH10B

(pMM374) and attempted direct photoaffinity labeling to [γ 32P]ATP.

(A) Overexpression and purification of the GST-PAS1-2 peptide containing both HrpX

PAS domains by glutathione affinity chromatography: lane 1, TCP from uninduced cells; lane 2, TCP from IPTG-induced cells, lane 3, cleared lysate; lane 4, fraction eluted from the GST-BindTM column with 10 mM reduced glutathione. (B) Direct

photoaffinity labeling of GST-PAS1-2 with [γ 32P]ATP. Lanes 1 to 4, GST-PAS1-2

preparation shown in Panel A, lane 4 incubated with [γ 32P]ATP under UV light for 5,

15, 30 and 45 min, respectively. The BSA negative control produced a smear (not

shown). A GST control was not available for this experiment. Note the presence of

background kinases not visible in the Coumassie gel of panel A, lane 4. The arrow

with a question mark indicates a possible PAS1-2~P signal, but the presence of

background kinases will require further purification to confirm it.

326

CHAPTER 6

N-ACYLHOMOSERINE LACTONE-DEPENDENT QUORUM-

SENSING IS REQUIRED FOR FULL EXPRESSION OF THE HRP-

TYPE III SECRETION SYSTEM OF PANTOEA STEWARTII SUBSP.

STEWARTII

6.1 INTRODUCTION

Quorum sensing is the phenomenon by which bacterial populations

coordinate the expression of complex phenotypes in response to their population density (Fuqua et al., 1994). In more than 30 gram-negative

species, quorum sensing is mediated by N-acyl-homoserine lactone (AHL)

signals that are perceived by receptor proteins acting as transcriptional

factors. AHLs are constitutively synthesized and can usually move freely

across biological membranes. The AHL concentration outside of the cell

is proportional to the population density in the milieu. When critical

concentrations are reached (>>Kd at equilibrium), AHL binds

quantitatively to the receptor/transcription factor thereby modulating its

affinity for DNA regulatory sites. Many review articles are available on

this subject (Fuqua et al., 1996; Fuqua and Greenberg, 1998; Dunny and 327 Winans, 1999; Parsek and Greenberg, 2000; Bassler and Schauder, 2001;

Miller and Bassler, 2001; Loh et al., 2002; Von Bodman et al., 2003).

AHL-independent quorum sensing mechanisms in gram-negative bacteria have also been described (Surette et al., 1999; Schauder et al., 2001).

Among the bacterial functions regulated by AHLs are: biofilm formation and light production in Vibrio spp. (Nealson et al., 1970; Camara et al.,

2002), toxins and other virulence factors in Pseudomonas aeruginosa

(Passador et al., 1993; Wagner et al., 2003), conjugal transfer of virulence plasmids in Agrobacterium tumefaciens (Piper et al., 1993), secondary metabolites and antibiotic production in Pseudomonas aureofaciens and Pectobacterium carotovorum (McGowan et al., 1995;

Pierson et al., 1998), and exoenzymes and harpinEcc in Pectobacterium

carotovorum (Pirhonen et al., 1993; Cui et al., 1996; Chatterjee et al.,

2002). Often, more than one quorum sensing system operates in the same cell (Bassler and Schauder, 2001). These regulatory networks may be organized into parallel-converging (Mok et al., 2003), parallel- independent (Zhang and Pierson, 2001) or hierarchical pathways, such those described in P. aeruginosa (Pearson et al., 1995). Quorum sensing systems interact with various housekeeping global regulators, such as

RpoN (Lilley and Bassler, 2000), RpoS (Whiteley et al., 2000) and rsmA/rsmB (=csrA/csrB) (Cui et al., 1995; Koiv and Mae, 2001; Pessi et al., 2001), or with two-component systems (Sperandio et al., 2002;

Ledgham et al., 2003; Ledgham et al., 2003). 328 In the plant macerogenic bacterium P. carotovorum subsp. carotovorum,

a complex regulatory network, involving quorum sensing, coregulates Pel,

Peh, Cel and Prt exoenzyme production, carbapenem antibiotic synthesis and the hrp genes.

In Pantoea stewartii subsp. stewartii, which causes Stewart’s wilt and leaf blight of sweet corn and maize, pathogenicity depends on a type III secretion (TTS) system encoded by hrp genes (Alfano and Collmer, 1997).

The Hrp phenotype in this species is defined as the lack of virulence in corn and the inability to elicit an HR upon infiltration of high doses of

IM-induced cells into tobacco leaves (Frederick et al., 2001). In addition, extracellular polysaccharide (EPS) is also a primary virulence factor, being required for wilt symptoms, but not for water-soaked lesion formation on young leaves (reviewed in Leigh and Coplin, 1992). The

EsaI/R quorum sensing system of P. stewartii controls synthesis of EPS

(von Bodman and Farrand, 1995; reviewed in Von Bodman et al., 2003).

The esaI gene encoding the synthase for the production of AHLs and the esaR gene that encodes its cognate AHL-receptor/transcription regulator have been well characterized. Mutations in esaI have pleiotropic effects, including elimination of AHL production, non-mucoidy, reduced growth rates and decreased virulence (von Bodman and Farrand, 1995). The major

AHL produced by P. stewartii is N-3-oxohexanoyl-L-homoserine lactone

(OHHL); subnanomolar concentrations of N-3-oxooctanoyl-L-HSL are also made (von Bodman and Farrand, 1995; Von Bodman et al., 1998). In 329 contrast to the paradigm for LuxR-like receptors (Fuqua et al., 1994),

EsaR negatively regulates EPS production at low population-density (Von

Bodman et al., 1998 by acting as a repressor in the apoprotein

conformation. Binding of the cognate AHL relieves the repression by

decreasing the affinity of EsaR for its DNA binding sites at the esaR promoter (Minogue et al., 2002), the only target DNA site for EsaR identified so far. On the other hand, the regulatory cascade leading to the esaR/esaI-dependent activation/derepression of the EPS-producing cps gene cluster is not known, and experiments in S. von Bodman laboratory have not shown in vitro binding of semipurified EsaR to cps promoter fragments. A putative EsaR-binding site in front of esaR

(GCCTGTCTA:TAGTGCAGGT) partially resembles the luxR-box

C G T T T consensus ( / ANCT / ANCNNNTAN / NG / NCAG / NG) (S. von Bodman,

pers. commun.).

The virulence of the esaI mutant was originally tested using a stem-

inoculation technique. Under these conditions, the esaI strain was unable

to cause the typical water-soaked lesions and wilt observed with wild-type

strains (Von Bodman et al., 1998). Lack of EPS usually leads only to loss

of the ability to move systemically in the plant and cause severe wilting;

most non-mucoid cps mutants can still produce water-soaked lesions on

the pseudostem (Dolph et al., 1988; Von Bodman et al., 1998). More

relevant assays such whorl inoculation (which test specifically the ability

to elicit watersoaking) had not been done with the esaI strain at the start 330 of this study. The apparent lack of watersoaking prompted us to examine

the effect of the Esa quorum sensing system on the expression of hrp

genes.

As shown in previous work (Coplin et al., 1986; Coplin et al., 1992;

Coplin et al., 1992; Frederick et al., 2001) and in previous chapters of

this dissertation, P. stewartii needs to fully express the Hrp TTS system

and effector genes to cause watersoaking in corn.

We therefore hypothesized that hrp expression was repressed in the

esaI mutant and that the regulatory effect could be mediated by the

cognate EsaR receptor. Given that the regulatory Hrp cascade is extremely

complex (Frederick et al., 1993; Merighi et al., 2003), quorum sensing

regulation could enter the pathway at any of many levels: hrpXY, hrpS or

hrpL transcription, HrpX/HrpY post-translational activation, or secretion

and effector gene transcription.

The results of this study showed that P. stewartii esaI mutants have a

Hrp- phenotype at low inoculum doses and that this mutation significantly reduces expression of hrp genes. We also found that transcriptional regulation of hrpS is the first step in the hrp regulatory cascade that is

affected by the esaI mutation. Our findings are consistent with the

hypothesis that EsaR is mediating the effects of esaI. The presence of a

putative esaR box in the 5’-regulatory region of hrpS may indicate direct

regulation. The relevance of quorum sensing to the biology of P. stewartii

will be discussed. 331

6.2 MATERIALS AND METHODS

6.2.1 Bacterial strains and plasmids

The Escherichia coli and P. stewartii strains and plasmids used in this study are listed in Table 6.1. Luria-Bertani (LB) broth and agar (Ausubel

et al. 1987) were used for strain maintenance and for growth of P.

stewartii under Hrp-repressing conditions. To induce hrp genes in P.

stewartii strains, overnight LB broth cultures were washed twice in pH

5.5 inducing medium (IM5.5; Frederick et al. 2001), inoculated into 2 ml

of the same medium at A600nm = 0.01 and grown for 12-16 h, unless

otherwise specified. The final optical density ranged from A600nm = 0.6 to

0.9. For time-course experiments, bacteria were grown in 250-ml flasks with side arms and optical density readings were taken with a Spectronic

20 spectrophotometer (Thermo Spectronic). Liquid cultures were grown in flasks or tubes shaken at 200 rpm at 37°C for E. coli or 28°C for P.

stewartii. When appropriate, antibiotics were supplied at the following

concentrations: ampicillin, 200 µg ml-1; kanamycin, 50 µg ml-1; and tetracycline, 10 µg ml-1. Synthetic N-3-oxo-hexanoyl-homoserine lactone

(OHHL; Sigma) was resuspended in methanol at 13 mM and stored at –20

°C. OHHL was used at a working concentration of 4 or 10 µM.

332

6.2.2 Genetic and molecular biology techniques

DNA isolation, agarose gel electrophoresis, restriction enzyme

digestion and ligation were done according to standard procedures

(Ausubel et al. 1987). PCRs used either native Taq (for colony-PCR

screens and construct mapping) or cloned Pfu (for molecular cloning)

DNA polymerases (Invitrogen and Clontech respectively) and standard amplification protocols (Ausubel et al., 1987). For colony-PCR, we used

cells lysed in 50 mM NaOH for 10 min at 100°C as templates. Plasmids

were introduced into E. coli and P. stewartii strains by electroporation or

mobilized by triparental mating using pRK2013::Tn7 as a helper plasmid

(Ditta et al. 1980).

6.2.3 Plasmid construction

For complementation and over-expression experiments, the hrpY and

hrpL ORFs, including their ribosomal binding sites, were amplified by

PCR and cloned into high and low copy number cloning vectors, such that

the inserts were expressed from the Plac promoter in the vector (Fig. 6.1;

the construction of these plasmids is described in Chapter 2).

Construction of pPL6GUSC plasmids containing uidA reporter genes fused

to the hrpX, hrpS, hrpL, hrpJ and wtsE promoter regions is described in

Chapter 2.

333 6.2.4 Enzyme assays

β-glucuronidase (GUS) activity was assayed fluorometrically using 4-

methyl-umbelliferyl-β-D-glucuronide as described by Jefferson (1987).

However, the assays were scaled down to fit microtiter plates and

analyzed using a Victor 14202 multilabel reader (PE-Applied Biosystems,

Wellesley, MA, U.S.A.) as described in Chapter 2. Net GUS activity of

each strain was corrected for the basal fluorescence of P. stewartii DC283

carrying pL6GUSC without an insert. Specific activity was expressed in

-1 -1 -1 GUS units (pmol MU min OD600nm ml of culture at 37°C).

6.2.5 Plant assays

To test for HR elicitation in tobacco (Nicotiana tabacum L., cv. “White

Burley”) plants, P. stewartii strains were grown overnight in IM5.5 liquid

9 medium, washed and resuspended in water at OD600nm = 0.52 (ca. 1x10

CFU ml-1). Cells were infiltrated into tobacco leaves using two replicates

per strain as previously described (Frederick et al. 2001) and necrosis was

rated 24-48 h after infiltration. Pathogenicity tests were performed by

inoculating the whorls of 6-day-old sweet corn seedlings (Zea mays var.

saccharata, cv. “Seneca Horizon”) as previously described (Coplin et al.

1986). This assay tests specifically for water-soaking ability. The plants

were held in growth chambers at 29°C (photoperiod 16 h, 15,000 lux,

relative humidity 99%). After 3 days, disease severity was rated using a 0

334 to 3 scale (0= no symptoms, 1=scattered small lesions, 2 = numerous

lesions, and 3 = extensive lesions that remained water-soaked with ooze

forming on leaf surfaces).

6.3 RESULTS

6.3.1 An esaI mutant, containing a mutation in the AHL synthase gene,

shows a Hrp- phenotype.

The esaI mutant used in this study, ESN51, is defective in N-3-oxo-C6-

HSL synthesis (OHHL) and EPS production and is unable to cause wilting

of corn (von Bodman and Farrand, 1995). We were initially interested in confirming whether this mutant was indeed unable to produce water- soaked lesions as previous stem-inoculations had suggested (Von Bodman

et al., 1998). To this purpose, we performed repeated corn whorl

inoculations using low inoculum doses (about 5x107 CFU per whorl) of various quorum sensing mutants. The ES∆R ∆esaR::cat and ES∆IR

∆esaIR::cat mutants produced water-soaked lesions that were basically undistinguishable from the wild-type strain after 3 to 4 days-post- inoculation (DPI) (Table 6.2). ESN51, the esaI strain, produced from none to a few scattered small lesions (Table 6.2) and some sporadic leaf flecking. To test if the esa mutations affected the ability to elicit a non- host HR, the mutants were also infiltrated into tobacco leaves. The ability to cause an HR was positively correlated with virulence (Table 6.2), confirming the Hrp phenotype of the esaI strain. 335

6.3.2 In epistasis experiments, constitutive expression of hrpS, but not hrpY, suppresses the Hrp phenotype of esaI strains

To verify the epistatic interaction between hrp and esa genes, we tested the ability of several known hrp regulatory genes to restore virulence to the esaI mutant. Clones containing hrpY, hrpS and hrpL expressed from a constitutive promoter (Plac) (Fig.6.1) were introduced into strains ESN51,

ES∆R and ES∆IR. As shown in Table 6.2, ectopic expression of hrpS and hrpL, but not hrpY, suppressed the Hrp- phenotype of the esaI strain.

Considering that one known role of HrpY is to activate the transcriptional expression of hrpS (Merighi et al., 2003), these results suggest that esaI is epistatic to hrpY. In other words, this implies the existence of regulatory interactions between the hrp genes and quorum sensing.

6.3.3 Expression of most hrp/wts genes is dramatically reduced in the

N-acyl-homoserine synthase mutant

To determine if reduced expression of hrp genes is responsible for the

Hrp- phenotype of the esaI strain and to identify which hrp promoter responds to quorum sensing signals, several plasmid borne hrp-uidA fusions were introduced into strain ESN51 and GUS activity was measured after growth in IM5.5 to late exponential/early stationary phase

(Fig. 6.2). Fusions carried by pMM50 hrpS-uidA, pDM2785 hrpL-uidA, pDM2531 hrpJ-uidA and pDM2791 wtsE-uidA were all downregulated 336 from 10 to 200 fold. However, the hrpX-uidA fusion (pMM25) exhibited only a two-fold difference. The modest effect of quorum sensing on PhrpX may be due to the decreased autoregulation of hrpXY by HrpY, which would in turn be a consequence of the autoregulatory loops described in

Chapter 3. Given this assumption, our results indicate that quorum sensing-dependent regulation of the hrp regulon probably enters the cascade at the level of hrpS regulation, since it is the earliest gene in the pathway to be significantly affected (about 12-fold).

To verify that the effect of esaI mutations on hrp gene expression was due to the defect in OHHL production and not to second-site mutations, we added synthetic OHHL, at a concentration comparable to that observed previously in P. stewartii cultures by Von Bodman et al., 1998, to ESN51 carrying the hrpJ-uidA reporter. hrpJ expression was restored to wild- type levels indicating that the esaI mutation can be complemented by synthetic OHHL (Fig. 6.3).

6.3.4 Expression of hrp secretion genes is population density- dependent and EsaR mediates the effects of esaI

Quorum-sensing regulated genes are expected to differ widely in their expression at various cell densities. This may complicate comparisons of expression data between samples with different cell density. We decided to more carefully analyze the population density dependence of hrpJ, which has been used throughout this dissertation as a representative 337 secretion gene. This was done in wild-type, esaI, esaR and esaIR strains

(Fig. 6.4). Cells were first grown in LB and then diluted 1:200 into fresh

medium. The GUS activity of the hrpJ-uidA reporter (in plasmid

pDM2531) was measured at different cell densities and data were

normalized per A600nm. (The relationship between A600nm and viable cells was the same for all strains used.) As shown in Fig. 6.4, in the wild-type background (strain DC283) the response curve follows a logistic trend, with a quorum for hrp expression estimated around 1 x 108 CFU ml-1 (A quorum was defined as the population density where the second derivative of the logistic curve equal zero; not shown) On the other hand, hrp gene expression in ESN51 was suppressed and basically constant at different cell densities, with only a slight rise at very high populations (Fig. 6.4).

This confirmed that esaI is absolutely required for full hrp gene expression over a wide range of cell densities. Addition of synthetic

OHHL (Sigma) at 10 µM to the growth medium did not up-regulate hrp gene expression in the wild-type strain at low cell densities (strain

DC283(pDM2531), Fig 6.5), which would be expected if quorum sensing was the only mechanism controlling hrp genes. This is in contrast to what was previously observed for EPS production, which was upregulated by exogenous AHL at low cell densities (von Bodman and Farrand, 1995).

EsaR is the only known cognate receptor protein for OHHL in P. stewartii. To test whether esaR mediates the regulatory effect of esaI on hrp genes, we evaluated the expression of a plasmid-borne hrpJ-uidA 338 fusion in a ∆esaRI::cat mutant. In this double esa mutant, hrp gene

expression was restored to levels comparable to wild-type or ∆esaR::cat

strains. For reason we do not fully understand, the ∆esaRI::cat strain showed somewhat lower hrp transcription levels than the ∆esaR::cat mutant, which was mildly upregulated at higher cell densities compared to the wild-type (Fig. 6.4). Collectively, these and the previous results suggest that esaR is either directly or indirectly involved in the repression of hrp genes and that additional regulators may be required. These additional regulators must be expressed only at later growth stages either downstream of or, more likely, in parallel with EsaR. Supporting this model, expression of hrpJ was not upregulated at low cell density in the

∆esaR::cat mutant (Fig. 6.4).

6.4 DISCUSSION

The results of this study show that the esaI/esaR quorum sensing system plays an important role in modulating P. stewartii virulence, irrespective of its effects on EPS synthesis. In this study we showed that a deletion mutation in the OHHL-synthase gene caused a significant reduction in water-soaked lesion formation, HR elicitation and in vitro hrp gene expression. The Hrp phenotype and the regulatory effects of the esaI mutation were suppressed by a second mutation in esaR, suggesting the involvement of the EsaR receptor as a repressor of hrp genes (Fig. 6.4).

Transcriptional control of the hrpS promoter seems to be the step directly 339 or indirectly modulated by the esaI/esaR system and this is supported by

both epistasis analysis (Table 6.2) and gene expression experiments (Fig.

6.2). In particular, post-translational activation of HrpY does not appear

to be involved in mediating the quorum sensing effects because ectopic expression of HrpY does not suppress the Hrp phenotype of the esaI strain

(Table 6.2). Addition of synthetic OHHL complemented the esaI mutation

(Fig. 6.3) but did not upregulate hrp expression of the wild-type strain at low cell densities (Fig. 6.5). Similarly, hrp expression was not upregulated at low population densities in the ∆esaR::cat strain, but it is moderately higher than the wild-type at high cell densities (Fig. 6.4).

These last observations may imply that a growth-phase-dependent regulator acts in parallel to, or downstream of, the esaR/esaI system. A similar scenario has been described for P. aeruginosa (Diggle et al.,

2002). In that species, genetic screening showed the parallel involvement of RpoS and MvaT as growth-phase regulators acting in concert with the

LasR/I and RhlR/I systems.

The observation that hrp expression in the ∆esaRI::cat strain is intermediate between the wild type and the ∆esaR::cat strain is somewhat puzzling. Assuming it is significant from a regulatory point of view, it may suggest that the OHHL produced by ∆esaI is required by other unknown AHL-receptors that also have a positive regulatory effect on hrp expression.

340 P. stewartii can multiply in plants up to ca. 105 -106 CFU g-1 tissue without causing any apparent host response (Coplin et al., 1992). After this population level is reached, a functional type III secretion system is essential for the bacteria to grow to much higher cell densities (109 CFU

g-1 and above) and cause disease symptoms (Coplin et al., 1992).

Activation of a hrp-gfpmut3[AAV] gene fusion in P. stewartii was observed

around 20-24 h post-inoculation when using low inoculum doses (M.

Merighi, unpublished observation). Other necrogenic phytobacteria of the

genera Erwinia, Pseudomonas and Xanthomonas cannot grow in plant tissue when their Hrp secretion systems are defective (Willis et al., 1991;

Bonas, 1994; Alfano and Collmer, 1997), and induction of hrp genes occurs at as little as two hours after inoculation (Boureau et al., 2003).

The fact that the Hrp pathogenicity system in P. stewartii is not only environmentally regulated (Merighi et al., 2003), but also under quorum sensing control, is consistent with the “stealth” model for growth in plant tissues. Activation of the Hrp system of P. stewartii causes the translocation of several effector proteins that are cytotoxic (Ahmad et al.,

2001; J-H, Ham and D. L. Coplin, unpublished). In other diseases, it has been shown that the host responds to this aggression with the induction of defense genes in the attempt to restrict pathogen growth (Yalpani and

Raskin, 1993; Shirasu et al., 1996; Fritig et al., 1998; Sandermann, 2000;

341 Conrath et al., 2002; Klement et al., 2003) and some pathogens even deploy active mechanisms to counteract host defenses (Abramovitch et al., 2003).

Perhaps P. stewartii delays hrp gene induction up to the point during infection when the host can no longer respond effectively to invasion. In this regard, induced defense responses have been described in corn

(Morris et al., 1998; Tiffin and Gaut, 2001), but it is not known if any of them are induced by P. stewartii. Alternatively, this delay may reflect the need for biofilm formation as a prerequisite for the interaction of the Hrp

secretion apparatus with host cells. Work in the von Bodman lab suggests

that quorum sensing controls biofilm formation on glass surfaces in P.

stewartii (Koutsoudis et al., 2002). Consequently hrp genes might be expressed only after microcolonies have already formed. This hypothesis would agree with early observations on the development of P. stewartii

populations in corn leaves (Braun, 1982). Based on Braun’s electron

microscopy studies, synthesis of EPS is completed around 12 h post-

inoculation, while the cells are still isolated or in small groups. The first

microcolonies appear only after 30 h (Braun, 1982). This timing roughly

corresponds to when hrp genes are turned on, if the plants receive low

inoculum doses. An additional advantage of expressing hrp genes in

biofilms is that the bacteria might be more resistant to hrp-induced host

oxidative responses.

342 The only other Hrp TTS system known to be regulated by quorum

sensing is P. carotovorum. This species is of particular interest because it

is evolutionarily related to P. stewartii. The RNA-binding protein RsmA

is the central regulator in this network (Fig. 6.6) and it acts by reducing

the half-life of target mRNAs. rsmA null mutants are upregulated for

exoenzyme synthesis and tissue maceration, hrpL mRNA steady-state

levels and harpinEcc production, whereas overexpression of RsmA reduces

antibiotic production and HR elicitation (Cui et al., 1995; Cui et al.,

1996; Mukherjee et al., 1996; Chatterjee et al., 2002). rsmA is controlled at the transcriptional and post-translational levels by several regulators.

The list includes the stationary phase sigma factor RpoS (Mukherjee et al., 1998) (activating rsmA), the regulatory RNA rsmB (Liu et al., 1998)

(titrating RsmA and therefore repressing its effects), the transcriptional adapter RsmC (Cui et al., 1999) (which acts negatively on rsmB and positively on rsmA expression), the two-component system GacA/GacS

(ExpA/ExpS) (activating rsmB and repressing rsmA) (Cui et al., 2001;

Hyytiainen et al., 2001), and finally by the quorum sensing system expR/expI (Andersson et al., 2000; Koiv and Mae, 2001). This means that quorum sensing controls harpin production and hrp gene expression indirectly via its effect on rsmA.

In particular, data from (Koiv and Mae, 2001) shows that an expI N- acyl-homoserine synthase knock-out mutant has increased expression of rsmA and rsmB RNAs and a net negative effect on exoenzyme production 343 and virulence. Recent work in Chatterjee’s laboratory (Chatterjee et al.,

2002) is in partial agreement with that conclusion, i.e. the absence of N-

(3-oxohexanoyl)-L-homoserine lactone (OHHL) in P. carotovorum

apparently activates transcription of rsmA but not of rsmB. The net effect of the increased concentration of cellular RsmA is the reduction of virulence and of rsmB RNA. Consequently, the effect of quorum sensing signal molecules on virulence in Pectobacterium appears to be channeled via the rsmA/B system. Work from E. Palva’s laboratory tried to define

the role for ExpR, the cognate LuxR-like receptor of OHHLs (Andersson

et al., 2000). Null expR mutations did not decrease virulence or

production of exoenzymes in vitro, but resulted in a slight increase in the

maceration capacity of the mutant strain in vivo. At the same time,

ectopic expression of ExpR reduced exoenzyme production and virulence.

These results are consistent with a model where apo-ExpR is an activator

of rsmA transcription at low OHHL concentration and increased OHHL levels may sequester ExpR, making it incompetent for DNA binding. No one has yet tested this hypothesis using biochemical approaches.

RsmA and gacA/S have been shown to be present in P. stewartii as well

(Mukherjee et al., 1998), but nothing is known about their role in quorum

sensing-mediated phenotypes and/or virulence. In spite of the similarities

in the genetic complement of the two species, substantial differences exist

between P. carotovorum and P. stewartii, both in the role of the Hrp

system in pathogenesis and the manner in which quorum sensing 344 interfaces with the Hrp regulatory cascade. The P. carotovorum Hrp

system is not required for growth in compatible hosts; tissue maceration caused by quorum-sensing controlled pectic enzymes is the primary virulence mechanism. Moreover, the effect of quorum sensing on hrp

genes is not direct. Instead, it is mediated by the global regulator RsmA,

which post-transcriptionally controls hrpL mRNA stability (Chatterjee et

al., 2002).

In P. stewartii, a necrogenic pathogen like P. syringae and Erwinia

amylovora, hrp genes are central to its virulence. In this species a direct

effect of EsaR on hrpS expression appears more likely considering that a

putative esaR box has been identified in the long 5’-UTR of hrpS

(GCCTGTCTAGTGAGGGT) (Fig. 6.7). It is interesting that the putative

esaR box is located in the IS-like element found in front of the hrpS

ribosomal binding site and that this element is absent in all the other

erwinias. Direct quorum sensing regulation of hrp genes may therefore be

a unique adaptation of P. stewartii. Other possible regulatory loops may

involve various global regulators, perhaps those described in P.

carotovorum (RsmA, rsmB, RpoS) (Fig. 6.6), but acting on hrpS. This

model is attractive because of the involvement of a known stationary-

phase regulator (RpoS) acting in parallel to the ExpI/ExpR system for

RsmA regulation (Mukherjee et al., 1998). Multiple regulators acting in

concert could be involved as well, as recently shown for P. aeruginosa

(Diggle et al., 2002). 345 6.5 SUMMARY

N-acylhomoserine lactone (HSL)-dependent quorum sensing is involved

in coordinate regulation of many bacterial phenotypes in response to

population density. In the corn pathogen P. stewartii subsp. stewartii, the

quorum sensing system encoded by the esaI OHHL synthase and esaR

receptor-regulator genes controls expression of exopolysaccharide

synthesis. In this study, we show that expression of P. stewartii hrp genes

is also under control of esaI and esaR. The expression of a hrp secretion

gene was shown to have a typical quorum sensing response curve with

time and had a threshold set at 1 x 108 CFU per ml in IM pH5.5 growth

medium. When assayed in corn whorls, an esaI mutant caused much less water-soaking than the wild-type and was unable to cause HR in tobacco.

This phenotype was suppressed by ectopic overexpression of hrpS and

hrpL, but not of hrpY. Expression of hrp secretion and effector genes was

reduced one to two orders of magnitude in an esaI mutant background and

restored in an esaRI strain. Collectively our results indicate that EsaR

mediates the effect of the esaI mutation and acts, directly or indirectly, as

a repressor of hrpS. The observations that hrp gene expression was not

upregulated significantly at low cell density in esaR mutants and that

addition of exogenous OHHL could not upregulate hrp expression in the

wild-type at low cell density suggest that additional regulation, perhaps

346 involving late exponential-stationary phase signals, may be involved. This is the first report of a Hrp type III secretion system of a stealth plant pathogen that is under direct quorum sensing control.

347

Strain or plasmid Relevant characteristics Reference

P. stewartii

SS104 Wild type, Wts+, HR+ ICPPBa

DC283 SS104 Nalr Coplin et al. (1986)

ESN51 esaI::Tn5 von Bodman et. al, 1998

ES∆R ∆esaR::cat von Bodman et al., 1998

ES∆IR ∆esaIR::cat von Bodman et al., 1998

Escherichia coli

DH10B F- mcrA ∆(mrr-hsdRMS-mcrBC) φ80dlacZ∆M15 ∆lacX74 endA1 recA1 deoR

∆(ara,leu)7697 araD139 galU galK nupG rpsL λ- Gibco BRL

Plasmids pDM2531 pPL6GUSC with a 0.9 kb BamHI fragment to create a hrpJ-uidA fusion Merighi et al., 2003

pDM2785 pPL6GUSC with a 389 bp BamHI PCR fragment to create a hrpL-uidA fusion Merighi et al., 2003

pDM2791 pPL6GUSC with a 1.1 kb HindIII-BamHI fragment from pES411 to create

a wtsE-uidA fusion Merighi et al., 2003

pMM25 pPL6GUSC with a 244 bp HindIII-BamHI PCR fragment to create a

hrpX-uidA fusion Merighi et al., 2003

pMM50 pPL6GUSC with a 1.3 kb BamHI PCR fragment to create a hrpS-uidA fusion Merighi et al., 2003

+ pMM6 pRK415 with a 0.6 kb BamHI-EcoRI hrpL PCR fragment, transcribed from Plac Merighi et al., 2003

pMM52 pRK415 with the 0.7 kb BamHI-EcoRI hrpY+ fragment from pMM46, transcribed

from Plac Merighi et al., 2003

pRF205 pVK100 with the 1.7 kb HindIII fragment containing hrpS from pES1044 Frederick et al. (1993)

+ - r r r pRK2013::Tn7 ColE1 mob traRK2 ∆ repRK2 repE kan::Tn7 (Tp , Sm , Sp ) Ditta et al. (1980)

Table 6.1. Bacterial strains and plasmids.

348

Strain Genotype Plasmid-borne hrp regulator

None hrpY hrpS hrpL

DC283 esaRI+ 2.9±0.1 a (+)b ND ND ND

ESN51 esaI- 0.8±0.4 (-) 0.6±0.25(-) 2.6±0.8(+) 2.5±0.3(+)

ES∆R ∆esaR::cat - 1.9±0.4 (+) 2.5 (+) 2.6 (+) 2.7 (+)

ES∆IR esaRI- 2.8±0.2 (+) 3.0 (+) 2.9 (+) 3.0 (+)

a Water-soaking severity in sweet corn whorls was rated using a 0 to 3 scale at 3-4

days post-inoculation. Data are the means up to 6 independent experiments with 6

replicates each ± the standard error. b HR elicitation is indicated between parenthesis.

Table 6.2. Watersoaking symptom severity ratings and hypersensitive response elicitation caused by P. stewartii esaI-esaR mutants and epistatic effects of other

Hrp regulatory genes.

349

hrpS hrpY hrpX hrpL hrpJ

GUS pMM50

GUS pMM25

pDM2531 GUS

GUS pDM2785

wtsF wtsE hrpN

GUS pDM2791 1 kb

Figure 6.1. Maps of plasmid-borne reporter gene fusions used in this study.

(Top) Map of the hrpL-hrpXY-hrpS regulatory region, including a portion of the secretion operon hrpJ . (Bottom) Map of the Hrp-secreted WtsE effector gene. The inserts cloned into the promoter probe plasmid pPL6GUSC and their orientations are shown below the maps as bars.

350

10000

1000 DC283 100 ESN51

10

GUS activity (pmol MU min-1 OD-1) 1

85 31 91 MM25 MM50 M27 M25 M27 p p D D D p p p hrp-uidA fusions

Figure 6.2. Effect of an esaI mutation on the expression of several hrp regulatory

and structural genes.

DC283 (wild-type) and ESN51 (esaI) strains carrying hrp-uidA reporter gene fusions

in plasmid pPL6GUS were grown in IM5.5 to an A600 >0.4 and GUS assays were

performed to assess hrp transcriptional activity. Data shown are the mean of two

independent experiments with two replicates each. Pair-wise averages were analyzed

using a t-test and they all were significant at P<0.05. Error bars indicate the standard

error of the mean. pMM25 is PhrpX-uidA; pMM50 is PhrpS-uidA; pDM2785 is PhrpL-

uidA; pDM2531 is PhrpJ-uidA; pDM2791 is PwtsE-uidA.

351

A600: 0.75 0.4 1.0 0.9

10000 9000 8000 7000 6000 -OHHL 5000 +10 uM OHHL 4000 3000 2000 Specific GUS activity

(pmol MU min-1 OD-1) 1000 0 DC283/pDM2531 ESN51/pDM2531

Figure 6.3. Suppression of esaI regulatory effects on hrpJ by addition of synthetic N-3-oxo-hexanoyl-homoserine lactone (OHHL).

Strains ESN51 (esaI) and DC283 (wild-type) carrying the hrpJ-uidA gene fusion in plasmid pDM2531 were grown in IM5.5 with and without the addition of 10 µM

OHHL. After 24 and 48 h of growth at 28°C, for DC283 and ESN51 respectively, the

cultures were assayed for GUS activity. Data are from one experiment with two

replicates. Optical densities (A600) of the cultures at the time of the assay are shown above the bars.

352

WT 12000 esaI esaR 10000 esaRI

8000 R2 = 0.8663 esaR 6000 R2 = 0.9403 WT 4000 R2 = 0.8341 esaIR 2000 R2 = 0.8793 GUS activity (pmol MU min-1 OD-1) esaI 0 0.01 0.1 1

OD600

Figure 6.4. Transcriptional expression of hrpJ as a function of cell density in esaI, esaR and esaIR mutants.

Cells containing a hrpJ-gusA fusion in plasmid pDM2531 were grown in LB broth

overnight, diluted 1:200 in fresh medium and grown at 28°C in flasks with shaking.

Aliquots were removed at different times for cell density and GUS activity measurements. The plot shows data from a representative experiment with 2-4 replicates per strain. Curves were interpolated by third-order polynomial regression

and the R2 of the fitting is shown. WT is strain DC283; esaI is strain ESN51 [DC283 esaI::Tn5]; esaR is strain ES∆R [DC283 ∆esaR::cat]; and esaRI is strain ES∆IR

[DC283 ∆esaIR::cat].

353

1600 DC283/pDM2531 1400 1200 1000 800 - OHHL 600 + 4 uM 400 OHHL Specific GUS activity (pmol MU min-1 OD-1) 200 0 0 0.2 0.4 0.6 OD600

Figure 6.5. Effect of addition of synthetic N-3-oxo-hexanoyl-homoserine lactone

(OHHL) on expression of hrpJ.

DC283 [wild-type] containing a hrpJ-gusA fusion in plasmid pDM2531 was grown in

LB overnight. Cells were washed in 100 mM MES pH 5.5, diluted 1:200 in IM pH5.5

medium with or without 4 µM OHHL and grown at 28°C in flasks with shaking.

Aliquots were removed at different times for measurement of cell density and GUS

activity. Data are from a representative experiment.

354

GacS/Exp expI ExpM HexA ~P RpoS

GacA/ExpA ExpR/OHHL?

rsmA rsmB RsmC

rsmB KdgR RsmA RNA

hrpL mRNA L exoenzymes σ hrp genes

Figure 6.6 Model describing the RsmA regulatory network in P. carotovorum

Refer to the Discussion section for details on the model.

355

Apo-EsaR OHHL (active)

EsaR

(inactive)

SAM +acyl- carrier protein EsaI

esa I esaR

esaR-Box ?

Stationary-phase signal?

hrpS

esaR-like box HrpY~P

HrpS BOX: GCCTGT-CTA:--GTG-AGGGT

LuxR BOX: MNCTRN-CNN:WANNGCAGNG-

EsaR BOX: GCCTGTACTA:TAGTGCA-GGT

Figure 6.7. Model for quorum sensing regulation of P. stewartii hrpS and

alignment of the putative esaR-like box (“HrpS box”).

356

CHAPTER 7

CONCLUSIONS

The proper expression of virulence factors in time and space is essential for the success of bacteria as pathogens. What molecular signals and which transduction mechanisms control plant pathogen type III secretion systems is for the most part unknown. We have seen that in P. stewartii as in most other phytobacteria, growth conditions mimicking the apoplast

(low pH, low concentration of inorganic nitrogen and phosphate sources) can activate hrp gene expression, whereas media containing complex nitrogen sources suppress it. Specific Hrp-inducing plant factors have not been characterized for any plant-microbe pathosystem. Regulation by contact has been proposed for a Hrp group II pathogen, but this is probably not a common paradigm among plant pathogens.

In the first part of this dissertation, I delineated the Hrp-activating pathway in P. stewartii (Fig. 7.1). All hrp secretion and effector genes identified in P. stewartii have “hrp boxes” in the –10/-35 region of their promoters. Genetic experiments using transcriptional fusions in both P. stewartii and E. coli demonstrated that the alternative sigma factor HrpL

357 is necessary and sufficient for recognition of “hrp box“-containing

promoters by the RNA polymerase. The expression of hrpL is controlled

by the response regulator HrpS, a receiver-less member of the NtrC family

of enhancer-binding transcription factors. A regulatory element upstream

of P. stewartii hrpS was identified by deletion analysis. These regulatory

steps are common to all group I hrp clusters in erwinias and

pseudomonads.

The most novel research presented in this dissertation is the analysis of

the two-component system HrpX/HrpY, which is also present in all

erwinias. In particular, we studied its roles in controlling hrpS and hrpL

and in autoregulation of its own operon. By using gene fusions and

ectopically expressed regulatory genes, we demonstrated that HrpY positively controls hrpS transcription and that it does not stimulate

expression of hrpL. These findings contradict a previous model published

for E. amylovora (Wei et al., 2000b), but are fully supported by the data

for P. agglomerans pv. gypsophilae, published concurrently with ours by

I. Barash’s lab (Merighi et al., 2003; Nizan-Koren et al., 2003). Sensor

and response regulators are located in the same operon, which is also

autoregulated by HrpY. We further showed that a conserved Asp residue

is required for the activity of HrpY. The response regulator was purified

as a recombinant protein for in vitro studies. We demonstrated that HrpY

is phosphorylated in vitro at the D57 residue and we confirmed its direct

interaction with its target promoters by gel shift assays and preliminary 358 DNA footprinting analysis. In these studies, phosphorylation appeared to increase the affinity of HrpY for the hrpS promoter (Fig. 7.1). However,

phosphorylation may not be needed for activation of the RNA

polymerase/DNA complex, but simply for increased oligomerization and

cooperative binding at multiple regulatory sites within the hrpS promoter

region.

Regulation of the hrpS promoter probably involves more factors than

just HrpY. This was suggested by the following observations: (a) Crude

extracts from a P. stewartii hrpY null mutant still retard hrpS promoter

fragments in gel shift experiments. (b) An IHF-like site has been

identified several hundred bp upstream of the promoter of hrpS. (c)

Deletion analysis of hrpS indicates the existence of an operator for an

unknown repressor (and in this respect, HrpY may function as an

antirepressor). (d) The quorum-sensing regulator EsaR was shown to

mediate the repression of hrpS in an OHHL-synthase mutant and a

potential EsaR-like box was found downstream of the hrpS promoter. (f)

hrpS expression is upregulated by a mutation in a nikR-like gene (Ham

and Coplin, unpublished).

The critical role of regulation at the hrpS promoter in integrating

multiple regulatory pathways is also highlighted by the results described

in Chapter 5. The environmental and metabolic signals generated upon

growth in media with high pH, high osmolarity, nicotinic acid, complex

nitrogen sources and some organic acids as sole C sources all repress 359 hrpS. Moreover, only a few of these (nitrogen sources, citrate and

succinate) are mediated by HrpX /HrpY. Therefore, HrpX may function

primarily as a metabolic starvation sensor. Its predicted cytoplasmic

location, the presence of PAS domains in its input region and the essential

role of these domains in hrp gene activation all support this hypothesis.

The remaining external signals must be transduced via alternative

regulatory pathways acting on PhrpS. For example, in enteric bacteria, osmolarity and pH are usually sensed by two-component systems or by direct changes in the topology of the promoter DNA (Finlay and Falkow,

1997). Given that HrpY is absolutely required for virulence and expression of hrpS, most of these postulated parallel regulatory mechanisms probably involve repression (Fig. 7.1).

The presence of so many independent mechanisms controlling a single master regulator may reflect the need to tightly regulate hrp gene transcription, allowing full expression only upon the fulfillment of multiple prerequisites in addition to simple starvation. Our discovery that the expression of hrpS is indirectly autoregulated by a novel transcriptional strategy further stresses the importance of maintaining a stable level of HrpS once the Hrp-pathway has been activated by external clues. The need for positive autoregulation perhaps also reflects high turnover rates for HrpS. This possibility is supported by the observation

360 that the Lon protease is a negative regulator of hrp gene expression in P.

stewartii, similar to what was described for P. syringae (Bretz et al.,

2002).

In addition to the primary role of HrpY in activating hrpS, we found

that, when it is overexpressed, HrpY has a negative effect on the basal

expression of hrpL from its tandem σ 70 promoter. This has not been characterized or described in any other erwinias. We still need to show that this repression happens in P. stewartii but it could represent a novel means to directly downregulate a critical regulator of hrp/hrc genes in order to dampen the autoregulatory loop triggered by HrpS via the hrpL promoter.

Perhaps my most surprising finding was the discovery that hrp genes are also regulated by quorum sensing, again at the level of hrpS transcription

(Fig. 7.1). Regulation by quorum sensing implies that hrp genes are only fully expressed when a local quorum of cells is reached (i.e. after microcolonies are formed). However, it is believed that the early multiplication in planta of most necrogenic phytobacteria is dependent on early induction of hrp genes. This tautology may be explained by the ability of P. stewartii to grow in plant tissues to considerable population densities as endophyte independently of hrp genes. In other words, hrp genes are used in this species only for a short, but critical, final attack.

The reasons for this strategy are unknown, but it could be that a quorum

361 sensing-dependent protective biofilm must be formed before the pathogen starts secreting effectors that could trigger host defenses.

Alternatively, or concurrently, P. stewartii may use a “brute-force” approach, delaying the critical attack of the host to when a sufficient population of invading cells is reached, and therefore overwhelming host defenses.

362

High pH, High osmolarity Nicotinic acid Nitrogen starvation? EsaR

HrpY~P σL ~P HrpX Hrp regulon

High KD Low KD High KD

UAS

(+/-)?

IHF σ54

Lon HrpS

Figure 7.1. Global regulation of hrp genes in P. stewartii.

363

LIST OF REFERENCES

Abramovitch, R.B., Kim, Y.J., Chen, S., Dickman, M.B., and Martin, G.B. (2003) Pseudomonas type III effector AvrPtoB induces plant disease susceptibility by inhibition of host programmed cell death. EMBO J 22: 60-69.

Ahmad, M., Majerczak, D.R., Pike, S., Hoyos, M.E., Novacky, A., and Coplin, D.L. (2001) Biological activity of harpin produced by Pantoea stewartii subsp. stewartii. Mol Plant Microbe Interact 14:1223-1234.

Ahmer, B.M., van Reeuwijk, J., Watson, P.R., Wallis, T.S., and Heffron, F. (1999) Salmonella SirA is a global regulator of genes mediating enteropathogenesis. Mol Microbiol 31: 971-982.

Aiba, H., Kato, N., Tsuzuki, M., and Mizuno, T. (1994) Mechanism of gene activation by the Escherichia coli positive regulator, OmpR: a mutant defective in transcriptional activation. FEBS Lett 351: 303-307.

Aiyar, A., and Leis, J. (1993) Modification of the megaprimer method of PCR mutagenesis: improved amplification of the final product. Biotechniques 14: 366-369.

Aldon, D., Brito, B., Boucher, C., and Genin, S. (2000) A bacterial sensor of plant cell contact controls the transcriptional induction of Ralstonia solanacearum pathogenicity genes. EMBO J 19: 2304-2314.

Alfano, J.R., and Collmer, A. (1997) The type III (Hrp) secretion pathway of plant pathogenic bacteria: trafficking harpins, Avr proteins, and death. J Bacteriol 179: 5655-5662.

364 Alfano, J.R., Bauer, D.W., Milos, T.M., and Collmer, A. (1996) Analysis of the role of the Pseudomonas syringae pv. syringae HrpZ harpin in elicitation of the hypersensitive response in tobacco using functionally non-polar hrpZ deletion mutations, truncated HrpZ fragments, and hrmA mutations. Mol Microbiol 19: 715-728.

Alfano, J.R., Charkowski, A.O., Deng, W.L., Badel, J.L., Petnicki- Ocwieja, T., van Dijk, K., and Collmer, A. (2000) The Pseudomonas syringae Hrp pathogenicity island has a tripartite mosaic structure composed of a cluster of type III secretion genes bounded by exchangeable effector and conserved effector loci that contribute to parasitic fitness and pathogenicity in plants. Proc Natl Acad Sci U S A 97: 4856-4861.

Allaoui, A., Woestyn, S., Sluiters, C., and Cornelis, G.R. (1994) YscU, a Yersinia enterocolitica inner membrane protein involved in Yop secretion. J Bacteriol 176: 4534-4542.

Altschul, S.F., Madden, T.L., Schaffer, A.A., Zhang, J., Zhang, Z., Miller, W. et al. (1997) Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res 25: 3389-3402.

Amezcua, C.A., Harper, S.M., Rutter, J., and Gardner, K.H. (2002) Structure and interactions of PAS kinase N-terminal PAS domain: model for intramolecular kinase regulation. Structure 10: 1349-1361.

Anderson, D.M., and Schneewind, O. (1997) A mRNA signal for the type III secretion of Yop proteins by Yersinia enterocolitica. Science 278: 1140-1143.

Anderson, D.M., and Schneewind, O. (1999) Yersinia enterocolitica type III secretion: an mRNA signal that couples translation and secretion of YopQ. Mol Microbiol 31: 1139-1148.

Anderson, D.M., Fouts, D.E., Collmer, A., and Schneewind, O. (1999) Reciprocal secretion of proteins by the bacterial type III machines of plant and animal pathogens suggests universal recognition of mRNA targeting signals. Proc Natl Acad Sci U S A 96: 12839-12843.

365 Andersson, R.A., Eriksson, A.R.B., Heikinheimo, R., Mae, A., Pirhonen, M., Koiv, V. et al. (2000) Quorum sensing in the plant pathogen Erwinia carotovora subsp. carotovora: the role of expREcc. Mol Plant Microbe Interact 13: 384-393.

Anonymous (1983) Reappearance of Erwinia stewartii in the Po Valley. FAO Plant Protection Bulletin 31: 96.

Appleman, J.A., and Stewart, V. (2003) Mutational analysis of a conserved signal-transducing element: the HAMP linker of the Escherichia coli nitrate sensor NarX. J Bacteriol 185: 89-97.

Appleman, J.A., Chen, L.L., and Stewart, V. (2003) Probing conservation of HAMP linker structure and signal transduction mechanism through analysis of hybrid sensor kinases. J Bacteriol 185: 4872-4882.

Arico, B., Miller, J.F., Roy, C., Stibitz, S., Monack, D., Falkow, S. et al. (1989) Sequences required for expression of Bordetella pertussis virulence factors share homology with prokaryotic signal transduction proteins. Proc Natl Acad Sci U S A 86: 6671-6675.

Arlat, M., Gough, C.L., Barber, C.E., Boucher, C., and Daniels, M.J. (1991) Xanthomonas campestris contains a cluster of hrp genes related to the larger hrp cluster of Pseudomonas solanacearum. Mol Plant Microbe Interact 4: 593-601.

Arlat, M., Gough, C.L., Zischek, C., Barberis, P.A., Trigalet, A., and Boucher, C.A. (1992) Transcriptional organization and expression of the large hrp gene cluster of Pseudomonas solanacearum. Mol Plant Microbe Interact 5: 187-193.

Arlat, M., Van Gijsegem, F., Huet, J.C., Pernollet, J.C., and Boucher, C.A. (1994) PopA1, a protein which induces a hypersensitivity-like response on specific Petunia genotypes, is secreted via the Hrp pathway of Pseudomonas solanacearum. EMBO J 13: 543-553.

Ausubel, F.M., Brent, R., Kingston, R.E., Moore, D.D., Seidman, J.G., Smith, J.A. and Struhl, K. (eds) (1987) Current Protocols in Molecular Biology, Vol. 1. New York: John Wiley & Sons.

366 Ayers, J.E., MacNab, A.A., Tetrault, R.C., and Yocum, J.O. (1979) The influence of selected insecticides on yield and the incidence of Stewart's wilt in sweet corn. Plant Dis Rep 63: 634-638.

Backert, S., Churin, Y., and Meyer, T.F. (2002) Helicobacter pylori type IV secretion, host cell signalling and vaccine development. Keio J Med 51 Suppl 2: 6-14.

Badel, J.L., Charkowski, A.O., Deng, W.L., and Collmer, A. (2002) A gene in the Pseudomonas syringae pv. tomato Hrp pathogenicity island conserved effector locus, hopPtoA1, contributes to efficient formation of bacterial colonies in planta and is duplicated elsewhere in the genome. Mol Plant Microbe Interact 15: 1014-1024.

Bajaj, V., Hwang, C., and Lee, C.A. (1995) hilA is a novel ompR/toxR family member that activates the expression of Salmonella typhimurium invasion genes. Mol Microbiol 18: 715-727.

Barny, M.A. (1995) Erwinia amylovora hrpN mutants, blocked in harpin synthesis, express a reduced virulence on host plants and elicit variable hypersensitive reactions on tobacco. European Journal of Plant Pathology 101: 333-340.

Barny, M.A., Guinebretiere, M.H., Marcais, B., Coissac, E., Paulin, J.P., and Laurent, J. (1990) Cloning of a large gene cluster involved in Erwinia amylovora CFBP1430 virulence. Mol Microbiol 4: 777-786.

Barta, T.M., Kinscherf, T.G., and Willis, D.K. (1992) Regulation of tabtoxin production by the lemA gene in Pseudomonas syringae. J Bacteriol 174: 3021-3029.

Bartolome, B. Y., Jubete, Y., Martinez, E. and de la Cruz, F. (1991) Construction and properties of a famility of pACYC184-derived cloning vectors compatible with pBR322 and its derivatives. Gene 102:75-78.

367 Bassler, B.L., and Schauder, S. (2001) The languages of bacteria. Annu Review Microbiol 55: 165-199.

Bauer, D.W., Bogdanove, A.J., Beer, S.V., and Collmer, A. (1994) Erwinia chrysanthemi hrp genes and their involvement in soft rot pathogenesis and elicitation of the hypersensitive response. Mol Plant Microbe Interact 7: 573-581.

Bauer, D.W., Wei, Z.M., Beer, S.V., and Collmer, A. (1995) Erwinia chrysanthemi harpinEch: an elicitor of the hypersensitive response that contributes to soft-rot pathogenesis. Mol Plant Microbe Interact 8: 484- 491.

Beer, S.V., Bauer, D.W., Jiang, X.H., Laby, R.J., Sneath, B.J., Wei, Z.M. et al. (1991) The Hrp gene cluster of Erwinia amylovora. Curr Plant Sci Biotechnol Agric 10: 53-60.

Beer, S.V., Wei, Z.M., Laby, R.J., He, S.Y., Bauer, W., Collmer, A., and Zumoff, C. (1993) Are harpins universal elicitors of the hypersensitive response of phytopathogenic bacteria? . Curr Plant Sci Biotechnol Agric 14:281-286.

Behlau, I., and Miller, S.I. (1993) A PhoP-repressed gene promotes Salmonella typhimurium invasion of epithelial cells. J Bacteriol 175: 4475-4484.

Bell, K.S., Avrova, A.O., Holeva, M.C., Cardle, L., Morris, W., De Jong, W. et al. (2002) Sample sequencing of a selected region of the genome of Erwinia carotovora subsp. atroseptica reveals candidate phytopathogenicity genes and allows comparison with Escherichia coli. Microbiology 148: 1367-1378.

Bergey, D.H. et al. (1923) Bergey's Manual of Determinative Bacteriology. Williams and Wilkins Company, Baltimore, MD.

Bergstrom, L.C., Qin, L., Harlocker, S.L., Egger, L.A., and Inouye, M. (1998) Hierarchical and co-operative binding of OmpR to a fusion construct containing the ompC and ompF upstream regulatory sequences of Escherichia coli. Genes Cells 3: 777-788.

368 Bernardini, M.L., Fontaine, A., and Sansonetti, P.J. (1990) Two- component regulatory system ompR-envZ controls the virulence of Shigella flexneri. J Bacteriol 172: 6274-6281.

Bernhard, F., Coplin, D.L., and Geider, K. (1993) A gene cluster for amylovoran synthesis in Erwinia amylovora: characterization and relationship to cps genes in Erwinia stewartii. Mol Gen Genetics 239: 158-168.

Bernhard, F., Schullerus, D., Bellemann, P., Nimtz, M., Coplin, D.L., and Geider, K. (1996) Genetic transfer of amylovoran and stewartan synthesis between Erwinia amylovora and Erwinia stewartii. Microbiology 142: 1087-1096.

Bijlsma, J.J., and Groisman, E.A. (2003) Making informed decisions: regulatory interactions between two-component systems. Trends Microbiol 11: 359-366.

Blakemore, E.J.A., Law, J.R., and and Reeves, J.C. (1999) PCR identification of Erwinia stewartii and its comparison with other two methods. Seed Sci Technol 23: 385-396.

Blakemore, E.J.A., Reeves, J.C., and Ball, S.F.L. (1992) Polymerase chain reaction used in the development of a DNA probe to identify Erwinia stewartii, a bacterial pathogen of maize. Seed Sci Technol 20: 331-335.

Boch, J., Joardar, V., Gao, L., Robertson, T.L., Lim, M., and Kunkel, B.N. (2002) Identification of Pseudomonas syringae pv. tomato genes induced during infection of Arabidopsis thaliana. Mol Microbiol 44: 73- 88.

Bock, A., and Gross, R. (2002) The unorthodox histidine kinases BvgS and EvgS are responsive to the oxidation status of a quinone electron carrier. Eur J Biochem 269: 3479-3484.

Bogdanove, A.J., Bauer, D.W., and Beer, S.V. (1998) Erwinia amylovora secretes DspE, a pathogenicity factor and functional AvrE homolog, through the Hrp (type III secretion) pathway. J Bacteriol 180: 2244-2247. 369 Bogdanove, A.J., Beer, S.V., Bonas, U., Boucher, C.A., Collmer, A., Coplin, D.L. et al. (1996) Unified nomenclature for broadly conserved hrp genes of phytopathogenic bacteria. Mol Microbiol 20: 681-683.

Bogdanove, A.J., Kim, J.F., Wei, Z., Kolchinsky, P., Charkowski, A.O., Conlin, A.K. et al. (1998) Homology and functional similarity of an hrp-linked pathogenicity locus, dspEF, of Erwinia amylovora and the avirulence locus avrE of Pseudomonas syringae pathovar tomato. Proc Natl Acad Sci U S A 95: 1325-1330.

Bogdanove, A.J., Wei, Z.M., Zhao, L., and Beer, S.V. (1996) Erwinia amylovora secretes harpin via a type III pathway and contains a homolog of yopN of Yersinia spp. J Bacteriol 178: 1720-1730.

Bonas, U. (1994) hrp genes of phytopathogenic bacteria. Curr Topics Microbiol and Immunol 192: 79-98.

Bonas, U., and Van den Ackerveken, G. (1999) Gene-for-gene interactions: bacterial avirulence proteins specify plant disease resistance. Curr Opin Microbiol 2: 94-98.

Boucher, C.A., Van Gijsegem, F., Barberis, P.A., Arlat, M., and Zischek, C. (1987) Pseudomonas solanacearum genes controlling both pathogenicity on tomato and hypersensitivity on tobacco are clustered. J Bacteriol 169: 5626-5632.

Boureau, T., Routtu, J., Roine, E., Taira, S., and Romantschuk, M. (2003) Localization of hrpA-induced Pseudomonas syringae pv. tomato DC3000 in infected tomato leaves. Mol Plant Pathol 3: 451-460.

Bourret, R.B., Hess, J.F., Borkovich, K.A., Pakula, A.A., and Simon, M.I. (1989) Protein phosphorylation in chemotaxis and two-component regulatory systems of bacteria. J Biol Chem 264: 7085-7088.

Boyer, H.W., and Roulland-Dussoix, D. (1969) A complementation analysis of the restriction and modification of DNA in Escherichia coli. J Mol Biol 41: 459-472.

370 Bradshaw-Rouse, J.J., Whatley, M.H., Coplin, D.L., Woods, A., Sequeira, L., and Kelman, A. (1981) Agglutination of Erwinia stewartii strains with a corn agglutinin: correlation with extracellular polysaccharide production and pathogenicity. Appl Environ Microbiol 42: 344-350.

Braun, E.J. (1982) Ultrastructure investigation of resistant and susceptible maize infected with Erwinia stewartii. Phytopathology 72: 159-166.

Braun, E.J. (1990) Colonization of resistant and susceptible maize plants by Erwinia stewartii strains differing in exopolysaccharide production. Physiological and Mol Plant Pathol 36: 363-379.

Breed, R.S., and al., e. (1948) Bergey's Manual of Determinative Bacteriology, 6th ed. Williams and Wilkins Company, Baltimore, MD.

Bretz, J., Losada, L., Lisboa, K., and Hutcheson, S.W. (2002) Lon protease functions as a negative regulator of type III protein secretion in Pseudomonas syringae. Mol Microbiol 45: 397-409.

Brito, B., Aldon, D., Barberis, P., Boucher, C., and Genin, S. (2002) A signal transfer system through three compartments transduces the plant cell contact-dependent signal controlling Ralstonia solanacearum hrp genes. Mol Plant Microbe Interact 15: 109-119.

Brito, B., Marenda, M., Barberis, P., Boucher, C., and Genin, S. (1999) prhJ and hrpG, two new components of the plant signal-dependent regulatory cascade controlled by PrhA in Ralstonia solanacearum. Mol Microbiol 31: 237-251.

Brown, A.F., Juvik, J.A., and Pataky, J.K. (2001) Quantitative trait loci in sweet corn associated with partial resistance to Stewart's wilt, northern corn leaf blight, and common rust. Phytopathology 91: 293-300.

Browning, D.F., Cole, J.A., and Busby, S.J. (2000) Suppression of FNR- dependent transcription activation at the Escherichia coli nir promoter by Fis, IHF and H-NS: modulation of transcription initiation by a complex nucleo-protein assembly. Mol Microbiol 37: 1258-1269.

371 Buell, C.R., Joardar, V., Lindeberg, M., Selengut, J., Paulsen, I.T., Gwinn, M.L. et al. (2003) The complete genome sequence of the Arabidopsis and tomato pathogen Pseudomonas syringae pv. tomato DC3000. Proc Natl Acad Sci U S A 100: 10181-10186.

Bull, C.T., Duffy, B., Voisard, C., Defago, G., Keel, C., and Haas, D. (2001) Characterization of spontaneous gacS and gacA regulatory mutants of Pseudomonas fluorescens biocontrol strain CHAO. Antonie Van Leeuwenhoek 79: 327-336.

Burkholder, W.H. (1957) In Bergey's Manual of Determinative Bacteriology. Baltimore: Williams and Wilkins Company, pp. 1094.

Burkholder, W.H., and Starr, M.P. (1948) The generic and specific characters of phytopathogenic species of Pseudomonas and Xanthomonas. Phytopathology 38: 494-502.

Buttner, D., and Bonas, U. (2002) Getting across--bacterial type III effector proteins on their way to the plant cell. EMBO J 21: 5313-5322.

Buttner, D., Nennstiel, D., Klusener, B., and Bonas, U. (2002) Functional analysis of HrpF, a putative type III translocon protein from Xanthomonas campestris pv. vesicatoria. J Bacteriol 184: 2389-2398.

Camara, M., Hardman, A., Williams, P., and Milton, D. (2002) Quorum sensing in Vibrio cholerae. Nat Genet 32: 217-218.

Casper-Lindley, C., Dahlbeck, D., Clark, E.T., and Staskawicz, B.J. (2002) Direct biochemical evidence for type III secretion-dependent translocation of the AvrBs2 effector protein into plant cells. Proc Natl Acad Sci U S A 99: 8336-8341.

Castaneda, M., Guzman, J., Moreno, S., and Espin, G. (2000) The GacS sensor kinase regulates alginate and poly-beta-hydroxybutyrate production in Azotobacter vinelandii. J Bacteriol 182: 2624-2628.

Castor, L.L., Ayers, J.E., McNab, A.A., and Krause, R.A. (1975) Computerized forecasting system for Stewart's bacterial disease of corn. Plant Dis Rep 59: 533-536. 372 Chamnongpol, S., Cromie, M., and Groisman, E.A. (2003) Mg2+ sensing by the Mg2+ sensor PhoQ of Salmonella enterica. J Mol Biol 325: 795-807.

Chancey, S.T., Wood, D.W., and Pierson, L.S., 3rd (1999) Two- component transcriptional regulation of N-acyl-homoserine lactone production in Pseudomonas aureofaciens. Appl Environ Microbiol 65: 2294-2299.

Chang, J.H., Rathjen, J.P., Bernal, A.J., Staskawicz, B.J., and Michelmore, R.W. (2000) avrPto enhances growth and necrosis caused by Pseudomonas syrinage pv. tomato in tomato lines lacking either Pto or Prf. Mol Plant-Microbe Interact 13: 568-571.

Charkowski, A.O., Alfano, J.R., Preston, G., Yuan, J., He, S.Y., and Collmer, A. (1998) The Pseudomonas syringae pv. tomato HrpW protein has domains similar to harpins and pectate lyases and can elicit the plant hypersensitive response and bind to pectate. J Bacteriol 180: 5211-5217.

Charkowski, A.O., Huang, H.C., and Collmer, A. (1997) Altered localization of HrpZ in Pseudomonas syringae pv. syringae hrp mutants suggests that different components of the type III secretion pathway control protein translocation across the inner and outer membranes of gram-negative bacteria. J Bacteriol 179: 3866-3874.

Chatterjee, A., Cui, Y., and Chatterjee, A.K. (2002)a Regulation of Erwinia carotovora hrpL(Ecc) (sigma-L(Ecc)), which encodes an extracytoplasmic function subfamily of sigma factor required for expression of the HRP regulon. Mol Plant Microbe Interact 15: 971-980.

Chatterjee, A., Cui, Y., Chaudhuri, S., and Chatterjee, A.K. (2002)b Identification of regulators of hrp/hop genes of Erwinia carotovora ssp carotovora and characterization of HrpL(Ecc) (SigmaL(Ecc)), an alternative sigma factor. Mol Plant Pathol 3: 359-370.

Chatterjee, A., Cui, Y.Y., and Chatterjee, A.K. (2002)c RsmA and the quorum-sensing signal, N-[3-oxohexanoyl]-L- homoserine lactone, control the levels of rsmB RNA in Erwinia carotovora subsp carotovora by affecting its stability. J Bacteriol 184: 4089-4095.

373 Cheng, L.W., and Schneewind, O. (1999) Yersinia enterocolitica type III secretion. On the role of SycE in targeting YopE into HeLa cells. J Biol Chem 274: 22102-22108.

Cheng, P., Yang, Y., Gardner, K.H., and Liu, Y. (2002) PAS domain- mediated WC-1/WC-2 interaction is essential for maintaining the steady- state level of WC-1 and the function of both proteins in circadian clock and light responses of Neurospora. Mol Cell Biol 22: 517-524.

Christie, P.J. (2001) Type IV secretion: intercellular transfer of macromolecules by systems ancestrally related to conjugation machines. Mol Microbiol 40: 294-305.

Claflin, L.E. (1999) Stewart's bacterial wilt. p 3-4 in: Compendium of Corn Diseases. White, D.G. (ed). American Phytopathological Society Press, St. Paul, MN, U.S.A.

Collmer, A., Badel, J.L., Charkowski, A.O., Deng, W.L., Fouts, D.E., Ramos, A.R. et al. (2000) Pseudomonas syringae Hrp type III secretion system and effector proteins. Proc Natl Acad Sci U S A 97: 8770-8777.

Collmer, A., Lindeberg, M., Petnicki-Ocwieja, T., Schneider, D.J., and Alfano, J.R. (2002) Genomic mining type III secretion system effectors in Pseudomonas syringae yields new picks for all TTSS prospectors. Trends in Microbiology 10: 462-469.

Conrath, U., Pieterse, C.M., and Mauch-Mani, B. (2002) Priming in plant-pathogen interactions. Trends Plant Sci 7: 210-216.

Cooksey, D.A. (1986) Galls of Gypsophila paniculata caused by Erwinia herbicola. Plant Dis 70: 464-468.

Coplin, D.L. (1978) Properties of F and P group plasmids in Erwinia stewartii. Phytopathology 68: 1637-1643.

Coplin, D.L. (1979) Introduction of bacteriophage Mu into Erwinia stewartii by use of an RK2::Mu hybrid plasmid. J Gen Microbiol 113: 181-184.

374 Coplin, D.L., and Majerczak, D.R. (1990) Extracellular polysaccharide genes in Erwinia stewartii: directed mutagenesis and complementation analysis. Mol Plant Microbe Interact 3: 286-292.

Coplin, D.L., Fredercik, R.D., Majerczak, D.R., and Tuttle, L.D. (1992) Characterization of a gene cluster that specifies pathogenicity in Erwinia stewartii. Mol Plant-Microbe Interact 5: 81-88.

Coplin, D.L., Frederick, R.D., and Majerczak, D.R. (1992) New pathogenicity loci in Erwinia stewartii identified by random Tn5 mutagenesis and molecular cloning. Mol Plant Microbe Interact 5: 266- 268.

Coplin, D.L., Frederick, R.D., Majerczak, D., and Haas, E. (1987) Complementation of avirulent mutants of Erwinia stewartii by recombinant plasmids. Plant Pathog Bact, Proc Int Conf, 6th: 413-418.

Coplin, D.L., Frederick, R.D., Majerczak, D.R., and Haas, E.S. (1986) Molecular cloning of virulence genes from Erwinia stewartii. J Bacteriol 168: 619-623.

Coplin, D.L., Majerczak, D.R., Zhang, Y., Kim, W.-S., Jock, S., and Geider, K. (2002) Identification of Pantoea stewartii subsp. stewartii by PCR and strain differentiation by PFGE. Plant Dis 86: 304-311.

Coplin, D.L., Rowan, R.G., Chisholm, D.A., and Whitmoyer, R.E. (1981) Characterization of plasmids in Erwinia stewartii. Appl Environ Microbiol 42:599-604.

Corbell, N., and Loper, J.E. (1995) A global regulator of secondary metabolite production in Pseudomonas fluorescens Pf-5. J Bacteriol 177: 6230-6236.

Cornelis, G., Sluiters, C., Lambert de Rouvroit, C., and Michiels, T. (1989) Homology between VirF, the transcriptional activator of Yersinia virulence regulon, and AraC, the Escherichia coli arabinose operon regulator. J Bacteriol 171: 254-262.

375 Cornelis, G., Vanooteghem, J.-C., and Sluiters, C. (1987) Transcription of the yop regulon from Y. enterocolitica requires trans-acting pYV and chromosomal genes. Microb Pathog 2: 367-379.

Cornelis, G.R. (1998) The Yersinia deadly kiss. J Bacteriol 180: 5495- 5504.

Cornelis, G.R. (2002) The Yersinia Ysc-Yop 'type III' weaponry. Nat Rev Mol Cell Biol 3: 742-752.

Cornelis, G.R., and Van Gijsegem, F. (2000) Assembly and function of type III secretory systems. Annu Rev Microbiol 54: 735-774.

Cornelis, G.R., and Wolf-Watz, H. (1997) The Yersinia Yop virulon: a bacterial system for subverting eukaryotic cells. Mol Microbiol 23: 861- 867.

Cornelis, G.R., Sluiters, C., Delor, I., Geib, D., Kaniga, K., Lambert de Rouvroit, C., Sory, M.P. et al. (1991) ymoA, a Yersinia enterocolitica chromosomal gene modulating the expression of virulence functions. Mol Microbiol 5: 1023-1034.

Cossart, P., Bochet, P., Normark, S., and Rappuoli, R. (ed.) (2000) Cellular Microbiology. ASM Press, Washington, D. C.

Cui, Y., Chatterjee, A., and Chatterjee, A.K. (2001) Effects of the two- component system comprising GacA and GacS of Erwinia carotovora subsp. carotovora on the production of global regulatory rsmB RNA, extracellular enzymes, and harpinEcc. Mol Plant Microbe Interact 14: 516-526.

Cui, Y., Chatterjee, A., Liu, Y., Dumenyo, C.K., and Chatterjee, A.K. (1995) Identification of a global repressor gene, rsmA, of Erwinia carotovora subsp. carotovora that controls extracellular enzymes, N-(3- oxohexanoyl)-L-homoserine lactone, and pathogenicity in soft-rotting Erwinia spp. J Bacteriol 177: 5108-5115.

Cui, Y., Madi, L., Mukherjee, A., Dumenyo, C.K., and Chatterjee, A.K. (1996) The RsmA- mutants of Erwinia carotovora subsp. carotovora 376 strain Ecc71 overexpress hrpNEcc and elicit a hypersensitive reaction-like response in tobacco leaves. Mol Plant Microbe Interact 9: 565-573.

Cui, Y., Mukherjee, A., Dumenyo, C.K., Liu, Y., and Chatterjee, A.K. (1999) rsmC of the soft-rotting bacterium Erwinia carotovora subsp. carotovora negatively controls extracellular enzyme and harpin (Ecc) production and virulence by modulating levels of regulatory RNA (rsmB) and RNA-binding protein (RsmA). J Bacteriol 181: 6042-6052.

da Silva, A.C.R., Ferro, J.A., Reinach, F.C., Farah, C.S., Furlan, L.R., Quaggio, R.B. et al. (2002) Comparison of the genomes of two Xanthomonas pathogens with differing host specificities. Nature 417: 459-463.

Dahl, J.L., Wei, B.-Y., and Kdner, R.J. (1997) Protein phosphorylation affects binding of the Escherichia coli transcription activator UhpA to the uhpT promoter. J Biol Chemistry 272: 1910-1919.

Dale, C., Young, S.A., Haydon, D.T., and Welburn, S.C. (2001) The insect endosymbiont Sodalis glossinidius utilizes a type III secretion system for cell invasion. Proc Natl Acad Sci U S A 98: 1883-1888.

Dangl, J.L. (1994) The enigmatic avirulence genes of phytopathogenic bacteria. Curr Top Microbiol Immunol 192: 99-118.

Dangl, J.L., and Jones, J.D. (2001) Plant pathogens and integrated defence responses to infection. Nature 411: 826-833.

Darwin, A.J., Li, J., and Stewart, V. (1996) Analysis of nitrate regulatory protein NarL-binding sites in the fdnG and narG operon control regions of Escherichia coli K-12. Mol Microbiol 20: 621-632.

Darwin, K.H., and Miller, V.L. (2001) Type III secretion chaperone- dependent regulation: activation of virulence genes by SicA and InvF in Salmonella typhimurium. EMBO J 20: 1850-1862.

377 de Lorenzo, V., Herrero, M., Jakubzik, U., and Timmis, K.N. (1990) Mini-Tn5 transposon derivatives for insertion mutagenesis, promoter probing, and chromosomal insertion of cloned DNA in gram-negative eubacteria. J Bacteriol 172: 6568-6572.

Delgado, J., Forst, S., Harlocker, S., and Inouye, M. (1993) Identification of a phosphorylation site and functional analysis of conserved aspartic acid residues of OmpR, a transcriptional activator for ompF and ompC in Escherichia coli. Mol Microbiol 10: 1037-1047.

Deng, W.L., Preston, G., Collmer, A., Chang, C.J., and Huang, H.C. (1998) Characterization of the hrpC and hrpRS operons of Pseudomonas syringae pathovars syringae, tomato, and glycinea and analysis of the ability of hrpF, hrpG, hrcC, hrpT, and hrpV mutants to elicit the hypersensitive response and disease in plants. J Bacteriol 180: 4523- 4531.

Deng, W.L., Rehm, A.H., Charkowski, A.O., Rojas, C.M., and Collmer, A. (2003) Pseudomonas syringae exchangeable effector loci: sequence diversity in representative pathovars and virulence function in P. syringae pv. syringae B728a. J Bacteriol 185: 2592-2602.

Deslandes, L., Olivier, J., Peeters, N., Feng, D.X., Khounlotham, M., Boucher, C. et al. (2003) Physical interaction between RRS1-R, a protein conferring resistance to bacterial wilt, and PopP2, a type III effector targeted to the plant nucleus. Proc Natl Acad Sci U S A 100: 8024-8029.

Diggle, S.P., Winzer, K., Lazdunski, A., Williams, P., and Camara, M. (2002) Advancing the quorum in Pseudomonas aeruginosa: MvaT and the regulation of N-acylhomoserine lactone production and virulence gene expression. J Bacteriol 184: 2576-2586.

Ditta, G., Stanfield, S., Corbin, D., and Helinski, D.R. (1980) Broad host range DNA cloning system for gram-negative bacteria: Construction of a gene bank of Rhizobium meliloti. Proc Natl Acad Sci USA 77: 7347- 7351.

Dolph, P.J., and Coplin, D.L. (1987) Characterization of a gene cluster for polysaccharide biosynthesis in Erwinia stewartii, p. 419-424, Plant Pathog Bact, Proc Int Conf, 6th.

378 Dolph, P.J., Majerczak, D.R., and Coplin, D.L. (1988) Characterization of a gene cluster for exopolysaccharide biosynthesis and virulence in Erwinia stewartii. J Bacteriol 170: 865-871.

Dong, H., Delaney, T.P., Bauer, D.W., and Beer, S.V. (1999) Harpin induces disease resistance in Arabidopsis through the systemic acquired resistance pathway mediated by salicylic acid and the NIM1 gene. Plant J 20: 207-215.

Dong, X.R., Li, S.F., and DeMoss, J.A. (1992) Upstream sequence elements required for NarL-mediated activation of transcription from the narGHJI promoter of Escherichia coli. J Biol Chem 267: 14122-14128.

Dowson, W.J. (1939) On the systematic position and generic names of the Gram negative bacterial plant pathogens. Zentr Bakt 100: 177-193.

Duan, Y.P., Castaneda, A., Zhao, G., Erdos, G., and Gabriel, D.W. (1999) Expression of a single, host-specific, bacterial pathogenicity gene in plant cells elicits division, enlargement, and cell death. Mol Plant- Microbe Interact 12: 556-560.

Dunham, C.M., Dioum, E.M., Tuckerman, J.R., Gonzalez, G., Scott, W.G., and Gilles-Gonzalez, M.A. (2003) A distal arginine in oxygen- sensing heme-PAS domains is essential to ligand binding, signal transduction, and structure. Biochemistry 42: 7701-7708.

Dunny, G.M., and Winans, S.C. (ed.) (1999) Cell-cell signaling in bacteria. ASM Press, Washington, D.C.

Dye, D.W. (1963) The taxonomic position of Xanthomonas stewartii (Erw. Smith 1914) Dowson 1939. New Zealand J Sci 6: 495-506.

Elliott, C. (1937) The genus Phytomonas. Phytopathology 27: 1181-1182.

Elliott, C., and Poos, F.W. (1934) Overwintering of Aplanobacter stewartii. Science 80: 289-290.

379 Elliott, C., and Poos, F.W. (1940) Seasonal development, insect vectors, and host range of bacterial wilt of sweet corn. J Agric Res 60: 645-686.

EPPO. (1999) EPPO PQR Database v. 3.8. EPPO, Paris, France.

Eriksson, A.R., Andersson, R.A., Pirhonen, M., and Palva, E.T. (1998) Two-component regulators involved in the global control of virulence in Erwinia carotovora subsp. carotovora. Mol Plant Microbe Interact 11: 743-752.

Espinosa, A., Guo, M., Tam, V.C., Fu, Z.Q., and Alfano, J.R. (2003) The Pseudomonas syringae type III-secreted protein HopPtoD2 possesses protein tyrosine phosphatase activity and suppresses programmed cell death in plants. Mol Microbiol49: 377-387.

Ezra, D., Barash, I., Valinsky, L., and Manulis, S. (2000) The dual function in virulence and host range restriction of a gene isolated from the pPATH (Ehg) plasmid of Erwinia herbicola pv. gypsophilae. Mol Plant Microbe Interact 13: 683-692.

Fellay, G., Rahme, L.G., Mindrinos, M.N., Frederick, R.D., Pisi, A., and Panopoulos, N.J. (1991) Genes and signals controlling the Pseudomonas syringae pv. phaseolicola-plant interaction. In Advances in Molecular Genetics of Plant-Microbe Interactions. Dordrech, The Netherlands: Kluwer Academic.

Fenselau, S., Balbo, I., and Bonas, U. (1992) Determinants of pathogenicity in Xanthomonas campestris pv. vesicatoria are related to proteins involved in secretion in bacterial pathogens of animals. Mol Plant Microbe Interact 5: 390-396.

Fessehaie, A., De Boer, S.H., and Levesque, C.A. (2002) Molecular characterization of DNA encoding 16S-23S rRNA intergenic spacer regions and 16S rRNA of pectolytic Erwinia species. Can J Microbiol 48: 387-398.

380 Ffrench-Constant, R.H., Waterfield, N., Burland, V., Perna, N.T., Daborn, P.J., Bowen, D., and Blattner, F.R. (2000) A genomic sample sequence of the entomopathogenic bacterium Photorhabdus luminescens W14: potential implications for virulence. Appl Environ Microbiol 66: 3310-3329.

Fiedler, U., and Weiss, V. (1995) A common switch in activation of the response regulators NtrC and PhoB: phosphorylation induces dimerization of the receiver modules. EMBO J 14: 3696-3705.

Finlay, B.B., and Falkow, S. (1997) Common themes in microbial pathogenicity revisited. Microbiol Mol Biol Rev 61: 136-169.

Fischer, W., Haas, R., and Odenbreit, S. (2002) Type IV secretion systems in pathogenic bacteria. Int J Med Microbiol 292: 159-168.

Flor, H. (1971) Current status of the gene-for-gene concept. Annu Rev Phytopathol 9: 275- 285.

Forsberg, A., and Wolf-Watz, H. (1988) The virulence protein Yop5 of Yersinia pseudotubercolosis is regulated at transcriptional level by plasmid-pIB1-encoded trans-acting elements controlled by temperature and calcium. Mol Microbiol 2: 121-133.

Fouts, D.E., Abramovitch, R.B., Alfano, J.R., Baldo, A.M., Buell, C.R., Cartinhour, S. et al. (2002) Genomewide identification of Pseudomonas syringae pv. tomato DC3000 promoters controlled by the HrpL alternative sigma factor. Proc Natl Acad Sci U S A 99: 2275-2280.

Francis, M.S., Wolf-Watz, H., and Forsberg, A. (2002) Regulation of type III secretion systems. Curr Opin Microbiol 5: 166-172.

Frank, D.W., and Iglewski, B.H. (1991) Cloning and sequence analysis of trans-regulatory locus required for exoenzyme S synthesis in Pseudomonas aeruginosa. J Bacteriol 173: 6460-6468.

Frankel, G., Phillips, A.D., Rosenshine, I., Dougan, G., Kaper, J.B., and Knutton, S. (1998) Enteropathogenic and enterohaemorrhagic Escherichia coli: more subversive elements. Mol Microbiol 30: 911-921. 381 Frederick, R.D., Ahmad, M., Majerczak, D.R., Arroyo-Rodriguez, A.S., Manulis, S., and Coplin, D.L. (2001) Genetic organization of the Pantoea stewartii subsp. stewartii hrp gene cluster and sequence analysis of the hrpA, hrpC, hrpN, and wtsE operons. Mol Plant Microbe Interact 14: 1213-1222.

Frederick, R.D., Chiu, J., Bennetzen, J.L., and Handa, A.K. (1997) Identification of a pathogenicity locus, rpfA, in Erwinia carotovora subsp. carotovora subsp. carotovora that encodes a two-component sensor- regulator protein. Mol Plant Microbe Interact 10: 407-415.

Frederick, R.D., Majerczak, D.R., and Coplin, D.L. (1993) Erwinia stewartii WtsA, a positive regulator of pathogenicity gene expression, is similar to Pseudomonas syringae pv. phaseolicola HrpS. Mol Microbiol9: 477-485.

Frey, J., and Krisch, H.M. (1985) Omega mutagenesis in gram-negative bacteria: a selectable interposon which is strongly polar in a wide range of bacterial species. Gene 36: 143-150.

Friedman, A.M., Long, S., Brown, S., Buikema, W., and Ausubel, F. (1982) Construction of a broad host range cosmid cloning vector and its use in the genetic analysis of Rhizobium mutants. Gene 18: 289-296.

Fritig, B., Heitz, T., and Legrand, M. (1998) Antimicrobial proteins in induced plant defense. Curr Opin Immunol 10: 16-22.

Fu, J.-F., Chang, H.-C., Chen, Y.-M., Chang, Y.-S., and Liu, S.-T. (1995) Sequence analysis of an Erwinia stewartii plasmid, pSW100. Plasmid 34: 75-84.

Fu, J.-F., Chang, H.-C., Chen, Y.-M., Chang, Y.-S., and Liu, S.-T. (1996) Characterization of the replicon of plasmid pSW500 of Erwinia stewartii. Mol Gen Genetics 250:699-704.

Fu, J.-F., Ying, S.W., and Liu, S.-T. (1997) Cloning and characterization of the ori region of pSW1200 of Erwinia stewartii: similarity with plasmid P1. Plasmid 38: 141-147.

382 Fuqua, C., and Greenberg, E.P. (1998) Self perception in bacteria: quorum sensing with acylated homoserine lactones. Curr Opin Microbiol 1: 183-189.

Fuqua, C., Winans, S.C., and Greenberg, E.P. (1996) Census and consensus in bacterial ecosystems: The LuxR-LuxI family of quorum- sensing transcriptional regulators. Annual Review of Microbiology 50: 727-751.

Fuqua, W.C., Winans, S.C., and Greenberg, E.P. (1994) Quorum sensing in bacteria: the LuxR-LuxI family of cell density-responsive transcriptional regulators. J Bacteriol 176: 269-275.

Gabriel, D.W., Burges, A., and Lazo, G.R. (1986) Gene-for-gene interaction of five cloned avirulence genes from Xanthomonas campestris pv. malvacearum with specific resistance genes in cotton. Proc Natl Acad Sci U S A 83: 6415-.

Galan, J.E., and Bliska, J.B. (1996) Cross-talk between bacterial pathogens and their host cells. Annu Rev Cell Dev Biol 12: 221-255.

Galan, J.E., and Collmer, A. (1999) Type III secretion machines: bacterial devices for protein delivery into host cells. Science 284: 1322- 1328.

Galinier, A., Garnerone, A.M., Reyrat, J.M., Kahn, D., Batut, J., and Boistard, P. (1994) Phosphorylation of the Rhizobium meliloti FixJ protein induces its binding to a compound regulatory region at the fixK promoter. J Biol Chem 269: 23784-23789.

Galperin, M.Y., Nikolskaya, A.N., and Koonin, E.V. (2001) Novel domains of the prokaryotic two-component signal transduction systems. FEMS Microbiol Lett 203: 11-21.

Garibaldi, A., and Gibbins, L.N. (1975) Induction of avirulent variants in Erwinia stewartii by incubation at supraoptimal temperatures. Canad J Microbiol 21:1282-1287.

383 Gaudriault, S., Brisset, M.N., and Barny, M.A. (1998) HrpW of Erwinia amylovora, a new Hrp-secreted protein. FEBS Lett 428: 224-228.

Gaudriault, S., Malandrin, L., Paulin, J.P., and Barny, M.A. (1997) DspA, an essential pathogenicity factor of Erwinia amylovora showing homology with AvrE of Pseudomonas syringae, is secreted via the Hrp secretion pathway in a DspB-dependent way. Mol Microbiol 26: 1057- 1069.

Genin, S., Gough, C.L., Zischek, C., and Boucher, C.A. (1992) Evidence that the hrpB gene encodes a positive regulator of pathogenicity genes from Pseudomonas solanacearum. Mol Microbiol 6: 3065-3076.

Georgellis, D., Kwon, O., and Lin, E.C. (2001) Quinones as the redox signal for the arc two-component system of bacteria. Science 292: 2314- 2316.

Gerharz, T., Reinelt, S., Kaspar, S., Scapozza, L., and Bott, M. (2003) Identification of basic amino acid residues important for citrate binding by the periplasmic receptor domain of the sensor kinase CitA. Biochemistry 42: 5917-5924.

Gibbins, L.N., Bennett, P.M., Saunders, J.R., Grinsted, J., and Connolly, J.C. (1976) Acceptance and transfer of R-factor RP1 by members of the "herbicola" group of the genus Erwinia. J Bacteriol 128:309-316.

Gilles-Gonzalez, M.A. (2001) Oxygen signal transduction. IUBMB Life 51: 165-173.

Gilles-Gonzalez, M.A., Gonzalez, G., and Perutz, M.F. (1995) Kinase activity of oxygen sensor FixL depends on the spin state of its heme iron. Biochemistry 34: 232-236.

Gilles-Gonzalez, M.A., Gonzalez, G., Perutz, M.F., Kiger, L., Marden, M.C., and Poyart, C. (1994) Heme-based sensors, exemplified by the kinase FixL, are a new class of heme protein with distinctive ligand binding and autoxidation. Biochemistry 33: 8067-8073.

384 Gomez-Duarte, O.G., and Kaper, J.B. (1995) A plasmid-encoded regulatory region activates chromosomal eaeA expression in the enteropathogenic Escherichia coli. Infect Immun 63: 1767-1776.

Glass, L., and Kauffman, S.A. (1972) Co-operative components, spatial localization and oscillatory cellular dynamics. J Theor Biol 34: 219-237.

Gottfert, M., Grob, P., and Hennecke, H. (1990) Proposed regulatory pathway encoded by the nodV and nodW genes, determinants of host specificity in Bradyrhizobium japonicum. Proc Natl Acad Sci U S A 87: 2680-2684.

Gong, W., Hao, B., Mansy, S.S., Gonzalez, G., Gilles-Gonzalez, M.A., and Chan, M.K. (1998) Structure of a biological oxygen sensor: a new mechanism for heme-driven signal transduction. Proc Natl Acad Sci U S A 95: 15177-15182.

Goodier, R.I., and Ahmer, B.M. (2001) SirA orthologs affect both motility and virulence. J Bacteriol 183: 2249-2258.

Goodrich, J.A., Schwartz, M.L., and McClure, W.R. (1990) Searching for and predicting the activity of sites for DNA binding proteins: compilation and analysis of the binding sites for Escherichia coli integration host factor (IHF). Nucleic Acids Res 18: 4993-5000.

Gopalan, S., Bauer, D.W., Alfano, J.R., Loniello, A.O., He, S.Y., and Collmer, A. (1996) Expression of the Pseudomonas syringae avirulence protein AvrB in plant cells alleviates its dependence on the hypersensitive response and pathogenicity (Hrp) secretion system in eliciting genotype- specific hypersensitive cell death. Plant Cell 8: 1095-1105.

Gough, C.L., Genin, S., Zischek, C., and Boucher, C.A. (1992) hrp genes of Pseudomonas solanacearum are homologous to pathogenicity determinants of animal pathogenic bacteria and are conserved among plant pathogenic bacteria. Mol Plant Microbe Interact 5: 384-389.

Gowen, J.W. (1941) Mutation in Drosophila, bacteria and viruses. Cold Spring Harbor Sympos Quant Biol 9: 187-193.

385 Grimm, C., and Panopoulos, N.J. (1989) The predicted protein product of a pathogenicity locus from Pseudomonas syringae pv. phaseolicola is homologous to a highly conserved domain of several procaryotic regulatory proteins. J Bacteriol 171: 5031-5038.

Grimm, C., Aufsatz, W., and Panopoulos, N.J. (1995) The hrpRS locus of Pseudomonas syringae pv. phaseolicola constitutes a complex regulatory unit. Mol Microbiol 15: 155-165.

Gueneron, M., Timmers, A.C., Boucher, C., and Arlat, M. (2000) Two novel proteins, PopB, which has functional nuclear localization signals, and PopC, which has a large leucine-rich repeat domain, are secreted through the hrp-secretion apparatus of Ralstonia solanacearum. Mol Microbiol 36: 261-277.

Han, D.C., and Winans, S.C. (1994) A mutation in the receiver domain of the Agrobacterium tumefaciens transcriptional regulator VirG increases its affinity for operator DNA. Mol Microbiol 12: 23-30.

Harlocker, S.L., Bergstrom, L., and Inouye, M. (1995) Tandem binding of six OmpR proteins to the ompF upstream regulatory sequence of Escherichia coli. J Biol Chem 270: 26849-26856.

Harrison-McMonagle, P., Denissova, N., Martinez-Hackert, E., Ebright, R.H., and Stock, A.M. (1999) Orientation of OmpR monomers within an OmpR:DNA complex determined by DNA affinity cleaving. J Mol Biol 285: 555-566.

Hauben, L., Moore, E.R., Vauterin, L., Steenackers, M., Mergaert, J., Verdonck, L., and Swings, J. (1998) Phylogenetic position of phytopathogens within the Enterobacteriaceae. Syst Appl Microbiol 21: 384-397.

Hauck, P., Thilmony, R., and He, S.Y. (2003) A Pseudomonas syringae type III effector suppresses cell wall-based extracellular defense in susceptible Arabidopsis plants. Proc Natl Acad Sci U S A 100: 8577-8582.

He, S.Y. (1997) Hrp-controlled interkingdom protein transport: learning from flagellar assembly? Trends Microbiol 5: 489-495.

386 He, S.Y. (1998) Type III protein secretion systems in plant and animal pathogenic bacteria. Annu Rev Phytopathol 36: 363-392.

He, S.Y., and Jin, Q. (2003) The Hrp pilus: learning from flagella. Curr Opin Microbiol 6: 15-19.

He, S.Y., Huang, H.C., and Collmer, A. (1993) Pseudomonas syringae pv. syringae harpin Pss: a protein that is secreted via the Hrp pathway and elicits the hypersensitive response in plants. Cell 73: 1255-1266.

Heeb, S., and Haas, D. (2001) Regulatory roles of the GacS/GacA two- component system in plant-associated and other gram-negative bacteria. Mol Plant Microbe Interact 14: 1351-1363.

Heichel, G.H., Sands, D.C., and Kring, J.B. (1977) Seasonal patterns and reduction by carbofuran of Stewart's bacterial wilt of sweet corn. Plant Dis Rep 61:149-153.

Helmann, J.D. (2002) The extracytoplasmic function (ECF) sigma factors. Advances in Microbial Physiology 46:47-110.

Hendrickson, E.L., Guevera, P., and Ausubel, F.M. (2000) The alternative sigma factor RpoN is required for hrp activity in Pseudomonas syringae pv. maculicola and acts at the level of hrpL transcription. J Bacteriol 182: 3508-3516.

Hendrickson, E.L., Guevera, P., Penaloza-Vazquez, A., Shao, J., Bender, C., and Ausubel, F.M. (2000) Virulence of the phytopathogen Pseudomonas syringae pv. maculicola is rpoN dependent. J Bacteriol 182: 3498-3507.

Henikoff, S., Wallace, J.C., and Brown, J.P. (1990) Finding protein similarities with nucleotide sequence databases. Methods Enzymol 183: 111-132.

Hienonen, E., Roine, E., Romantschuk, M., and Taira, S. (2002) mRNA stability and the secretion signal of HrpA, a pilin secreted by the type III system in Pseudomonas syringae. Mol Genet Genomics 266: 973-978.

387 Hirschman, J., Wong, P.K., Sei, K., Keener, J., and Kustu, S. (1985) Products of nitrogen regulatory genes ntrA and ntrC of enteric bacteria activate glna transcription in vitro - evidence that the ntrA product is a sigma-factor. Proc Natl Acad Sci USA 82: 7525-7529.

Hobbs, M., and Mattick, J.S. (1993) Common components in the assembly of type IV fimbriae, DNA trasnfer systems, filamentous phage and protein secretion apparatus: a general system for the formation of surface-associated protein complexes. Mol Microbiol 10: 233-243.

Hoch, J.A. (2000) Two-component and phosphorelay signal transduction. Curr Opin Microbiol 3: 165-170.

Hoch, J.A., and Silhavy, T.J., eds. (1995) Two-Component Signal Transduction. American Society for Microbiology, Washington, D.C., U.S.A.

Hoch, J.A., and Varughese, K.I. (2001) Keeping signals straight in phosphorelay signal transduction. J Bacteriol 183: 4941-4949.

Holland, D.F. (1920) Generic index of the common forms of bacteria. In The families and genera of the bactera. J Bacteriol 5: 191-229.

Holt, B.F., 3rd, Hubert, D.A., and Dangl, J.L. (2003) Resistance gene signaling in plants--complex similarities to animal innate immunity. Curr Opin Immunol 15: 20-25.

Holt, J.G. (ed.) (1984) Bergey's Manual of Determinative Bacteriology, 9th ed, Vol. 1. Williams & Wilkins, Baltimore.

Hoyos, M.E., and al., e. (1996) The interaction of Harpin Pss with plant cell walls. Mol Plant-Microbe Interact 9: 608-616.

Hrabak, E.M., and Willis, D.K. (1992) The lemA gene required for pathogenicity of Pseudomonas syringae pv. syringae on bean is a member of a family of two-component regulators. J Bacteriol 174: 3011-3020.

388 Hromockyj, A.E., Tucker, S.C., and Maurelli, A.T. (1992) Temperature regulation of Shigella virulence: identification of the repressor gene virR, an analogue of hns, and partial complementation by tyrosyl transfer RNA (tRNA1(Tyr)). Mol Microb 6: 2113-2124.

Hu, W., Yuan, J., Jin, Q.L., Hart, P., and He, S.Y. (2001) Immunogold labeling of Hrp pili of Pseudomonas syringae pv. tomato assembled in minimal medium and in planta. Mol Plant Microbe Interact 14: 234-241.

Huang, H.C., Xiao, Y., Lin, R.H., Lu, Y., Hutcheson, S.W., and Collmer, A. (1993) Characterization of the Pseudomonas syringae pv. syringae 61 hrpJ and hrpI genes: homology of HrpI to a superfamily of proteins associated with protein translocation. Mol Plant Microbe Interact 6: 515-520.

Huang, K.J., and Igo, M.M. (1996) Identification of the bases in the ompF regulatory region, which interact with the transcription factor OmpR. J Mol Biol 262: 615-628.

Huang, K.J., Lan, C.Y., and Igo, M.M. (1997) Phosphorylation stimulates the cooperative DNA-binding properties of the transcription factor OmpR. Proc Natl Acad Sci U S A 94: 2828-2832.

Hueck, C.J. (1998) Type III protein secretion systems in bacterial pathogens of animals and plants. Microbiol Mol Biol Rev 62: 379-433.

Hughes, K.T., and Mathee, K. (1998) The anti-sigma factors. Annual Review of Microbiology 52: 231-286.

Huguet, E., and Bonas, U. (1997) hrpF of Xanthomonas campestris pv. vesicatoria encodes an 87-kDa protein with homology to NoIX of Rhizobium fredii. Mol Plant Microbe Interact 10: 488-498.

Hutcheson, S.W. (1999) The hrp cluster of Pseudomonas syrinagae: a pathogenicity island encoding a type III protein translocation complex? In Pathogenicity Islands and Other Mobile Virulence Elements. Washington, D. C.: ASM Press.

389 Hutcheson, S.W., Bretz, J., Sussan, T., Jin, S., and Pak, K. (2001) Enhancer-binding proteins HrpR and HrpS interact to regulate hrp- encoded type III protein secretion in Pseudomonas syringae strains. J Bacteriol 183: 5589-5598.

Huynh, T.V., Dahlbeck, D., and Staskawicz, B.J. (1989) Bacterial blight of soybean: regulation of a pathogen gene determining host cultivar specificity. Science 245: 1374-1377.

Hyytiainen, H., Montesano, M., and Palva, E.T. (2001) Global regulators ExpA (GacA) and KdgR modulate extracellular enzyme gene expression through the RsmA-rsmB system in Erwinia carotovora subsp. carotovora. Mol Plant Microbe Interact 14: 931-938.

Innes, R.W., Bent, A.F., Kunkel, B.N., Bisgrove, S.R., and Staskawicz, B.J. (1993) Molecular analysis of avirulence gene avrRpt2 and identification of a putative regulatory sequence common to all known Pseudomonas syringae avirulence genes. J Bacteriol 175: 4859-4869.

Inoue, H., Nojima, H., and Okayama, H. (1990) High efficiency transformation of Escherichia coli with plasmids. Gene 96: 23-28.

Ivanoff, S.S. (1933) Stewart’s wilt disease of corn, with emphasis on the life history of Phytomonas stewartii in relation to pathogenesis. J Agric Res 47: 749-770.

Ivanoff, S.S., Riker, A.J., and Dettwiler, H.A. (1938) Studies on cultural characteristics, physiology and pathogenicity of strain types of Phytomonas stewarti. J Bact 35: 235-253.

Jefferson, R.A. (1987) Assaying chimeric genes in plants: the GUS gene fusion system. Plant Mol Biol Rep 5: 387-405.

Jin, Q., and He, S.Y. (2001) Role of the Hrp pilus in type III protein secretion in Pseudomonas syringae. Science 294: 2556-2558.

Jin, S.G., Prusti, R.K., Roitsch, T., Ankenbauer, R.G., and Nester, E.W. (1990) Phosphorylation of the VirG protein of Agrobacterium

390 tumefaciens by the autophosphorylated VirA protein: essential role in biological activity of VirG. J Bacteriol 172: 4945-4950.

Johnston, C., Pegues, D.A., Hueck, C.J., Lee, A., and Miller, S.I. (1996) Transcriptional activation of Salmonella typhimurium invasion genes by a member of the phosphorylated response-regulator superfamily. Mol Microbiol 22: 715-727.

Jung, K., and Altendorf, K. (1998) Truncation of amino acids 12-128 causes deregulation of the phosphatase activity of the sensor kinase KdpD of Escherichia coli. J Biol Chem 273: 17406-17410.

Kadner, R. J. (1995) Expression of the Uhp sugar phosphate transport system of Escherichia coli. Pages 263-274 in: Two-Component Signal Transduction. Hoch, J.A., and Silhavy, T.J. (eds.) American Society for Microbiology, Washington, D.C., U.S.A.

Kado, C., and Coplin, D.L. (2001) Laboratory Guide for the Identification of Plant Pathogenic Bacteria. APS Press, St. Paul, MI, USA.

Kaniga, K., Bossio, J.C., and Galan, J.E. (1994) The Salmonella typhimurium invasion genes invF and invG encode homologues of the AraC and PulD family of proteins. Mol Microbiol 13: 555-568.

Kaper, J.B., and Hacker, J. (ed.) (1999) Pathogenicity Islands and Other Mobile Virulence Elements. ASM Press, Washington, D.C., USA.

Kariola, T., Palomaki, T.A., Brader, G., and Palva, E.T. (2003) Erwinia carotovora subsp. carotovora and Erwinia-derived elicitors HrpN and PehA trigger distinct but interacting defense responses and cell death in Arabidopsis. Mol Plant Microbe Interact 16: 179-187.

Keen, N.T., Tamaki, S., Kobayashi, D., and Trollinger, D. (1988) Improved broad-host-range plasmids for DNA cloning in gram-negative bacteria. Gene 70: 191-197.

391 Keener, J., and Kustu, S. (1988) Protein kinase and phosphoprotein phosphatase activities of nitrogen regulatory proteins NtrB and NtrC of enteric bacteria: roles of the conserved amino-terminal domain of NtrC. Proc Natl Acad Sci U S A 85: 4976-4980.

Kim, J.F., and Beer, S.V. (1998) HrpW of Erwinia amylovora, a new harpin that contains a domain homologous to pectate lyases of a distinct class. J Bacteriol 180: 5203-5210.

Kim, J.F., Ham, J.H., Bauer, D.W., Collmer, A., and Beer, S.V. (1998) The hrpC and hrpN operons of Erwinia chrysanthemi EC16 are flanked by plcA and homologs of hemolysin/adhesin genes and accompanying activator/transporter genes. Mol Plant Microbe Interact 11: 563-567.

Kim, J.F., Wei, Z.M., and Beer, S.V. (1997) The hrpA and hrpC operons of Erwinia amylovora encode components of a type III pathway that secretes harpin. J Bacteriol 179: 1690-1697.

Kim, Y.J., Lin, N.C., and Martin, G.B. (2002) Two distinct Pseudomonas effector proteins interact with the Pto kinase and activate plant immunity. Cell 109: 589-598.

Kinscherf, T.G., and Willis, D.K. (1999) Swarming by Pseudomonas syringae B728a requires gacS (lemA) and gacA but not the acyl- homoserine lactone biosynthetic gene ahlI. J Bacteriol 181: 4133-4136.

Klement, Z. (1982) Hypersensitivity. In Phytopathogenic prokaryotes. New York, N.Y.: Academic Press, pp. 149-177.

Klement, Z., Bozso, Z., Kecskes, M.L., Besenyei, E., Arnold, C., and Ott, P.G. (2003) Local early induced resistance of plants as the first line of defence against bacteria. Pest Manag Sci 59: 465-474.

Klose, K.E., North, A.K., Stedman, K.M., and Kustu, S. (1994) The major dimerization determinants of the nitrogen regulatory protein NtrC from enteric bacteria lie in its carboxy-terminal domain. J Mol Biol 241: 233-245.

392 Klose, K.E., Weiss, D.S., and Kustu, S. (1993) glutamate at the site of phosphorylation of nitrogen-regulatory protein NtrC mimics aspartyl- phosphate and activates the protein. J Mol Biol 232: 67-78.

Knauf, V.C., and Nester, E.W. (1982) Wide host range cloning vectors: a cosmid clone bank of an Agrobacterium Ti plasmid. Plasmid 8: 45-54.

Knoop, V., Staskawicz, B., and Bonas, U. (1991) Expression of the Avirulence gene avrbs3 from Xanthomonas- campestris pv vesicatoria is not under the control of hrp genes and is independent of plant factors. J Bacteriol 173: 7142-7150.

Koiv, V., and Mae, A. (2001) Quorum sensing controls the synthesis of virulence factors by modulating rsmA gene expression in Erwinia carotovora subsp. carotovora. Mol Genet Genomics 265: 287-292.

Kolesnikow, T., Schroder, I., and Gunsalus, R.P. (1992) Regulation of narK gene expression in Escherichia coli in response to anaerobiosis, nitrate, iron, and molybdenum. J Bacteriol 174: 7104-7111.

Koster, M., Bitter, W., and Tommassen, J. (2000) Protein secretion mechanisms in Gram-negative bacteria. Int J Med Microbiol 290: 325- 331.

Koutsoudis M., Minogue T.D., von Bodman S.B. (2002) Quorum sensing regulationof capsular polysaccharide synthesis in Pantoea stewartii subsp. stewartii and its relationship to surface attachmentand biofilm formation. Abstr. #64475. Gen. Meet. ASM, Salt Lake City

Kox, L.F., Wosten, M.M., and Groisman, E.A. (2000) A small protein that mediates the activation of a two-component system by another two- component system. EMBO J 19: 1861-1872.

Kramer, G., and Weiss, V. (1999) Functional dissection of the transmitter module of the histidine kinase NtrB in Escherichia coli. Proc Natl Acad Sci U S A 96: 604-609.

393 Krause, A., Doerfel, A., and Gottfert, M. (2002) Mutational and transcriptional analysis of the type III secretion system of Bradyrhizobium japonicum. Mol Plant Microbe Interact 15: 1228-1235.

Kuhar, T.P., Stivers-Young, L.J., Hoffmann, M.P., and Taylor, A.G. (2002) Control of corn flea beetle and Stewart's wilt in sweet corn with imidacloprid and thiamethoxam seed treatments. Crop Protection 21:25- 31.

Kwon, S.-W., Go, S.-J., Kang, H.-W., Ryu, J.-C., and Jo, J.-K. (1997) Phylogenetic analysis of Erwinia species based on 16S rRNA gene sequences. Int J System Bacteriol 47:1061-1067.

Laby, R.J., and Beer, S.V. (1992) Hybridization and functional complementation of the hrp gene cluster from Erwinia amylovora strain Ea321 with DNA of other bacteria. Mol Plant Microbe Interact 5: 412- 419.

Lamka, G.L., Hill, J.H., McGee, D.C., and and Braun, E.J. (1991) Development of an immunosorbert assay for seedborne Erwinia stewartii in corn seeds. Phytopathology 81: 839-846.

Lavie, M., Shillington, E., Eguiluz, C., Grimsley, N., and Boucher, C. (2002) PopP1, a new member of the YopJ/AvrRxv family of type III effector proteins, acts as a host-specificity factor and modulates aggressiveness of Ralstonia solanacearum. Mol Plant Microbe Interact 15: 1058-1068.

Lawhon, S.D., Maurer, R., Suyemoto, M., and Altier, C. (2002) Intestinal short-chain fatty acids alter Salmonella typhimurium invasion gene expression and virulence through BarA/SirA. Mol Microbiol 46: 1451-1464.

Ledgham, F., Soscia, C., Chakrabarty, A., Lazdunski, A., and Foglino, M. (2003) Global regulation in Pseudomonas aeruginosa: the regulatory protein AlgR2 (AlgQ) acts as a modulator of quorum sensing. Res Microbiol 154: 207-213.

394 Ledgham, F., Ventre, I., Soscia, C., Foglino, M., Sturgis, J.N., and Lazdunski, A. (2003) Interactions of the quorum sensing regulator QscR: interaction with itself and the other regulators of Pseudomonas aeruginosa LasR and RhlR. Mol Microbiol 48: 199-210.

Lee, J., Klusener, B., Tsiamis, G., Stevens, C., Neyt, C., Tampakaki, A.P. et al. (2001) HrpZ(Psph) from the plant pathogen Pseudomonas syringae pv. phaseolicola binds to lipid bilayers and forms an ion- conducting pore in vitro. Proc Natl Acad Sci U S A 98: 289-294.

Lee, Y.W., Jin, S., Sim, W.S., and Nester, E.W. (1995) Genetic evidence for direct sensing of phenolic compounds by the VirA protein of Agrobacterium tumefaciens. Proc Natl Acad Sci U S A 92: 12245-12249.

Leigh, J.A., and Coplin, D.L. (1992) Exopolysaccharides in plant- bacterial interactions. Annu Rev Microbiol 46: 307-346.

Li, J., Kustu, S., and Stewart, V. (1994) In vitro interaction of nitrate- responsive regulatory protein NarL with DNA target sequences in the fdnG, narG, narK and frdA operon control regions of Escherichia coli K- 12. J Mol Biol 241: 150-165.

Liao, C.H., McCallus, D.E., Fett, W.F., and Kang, Y. (1997) Identification of gene loci controlling pectate lyase production and soft- rot pathogenicity in Pseudomonas marginalis. Can J Microbiol 43: 425- 431.

Lidell, M.C., and Hutcheson, S.W. (1994) Characterization of the hrpJ and hrpU operons of Pseudomonas syringae pv. syringae Pss61: similarity with components of enteric bacteria involved in flagellar biogenesis and demonstration of their role in HarpinPss secretion. Mol Plant Microbe Interact 7: 488-497.

Lilley, B.N., and Bassler, B.L. (2000) Regulation of quorum sensing in Vibrio harveyi by LuxO and sigma-54. Mol Microbiol 36: 940-954.

Lincoln, R.E. (1939) Host-parasite interactions with bacterial wilt of maize. Science 89: 159-160.

395 Lincoln, R.E. (1947) Mutation and adaptation of Phytomonas stewartii. J Bacteriol 54: 745-757.

Lincoln, R.E., and Gowen, J.W. (1942) Mutation of Phytomonas stewartii by X-ray irradiation. Genetics 27: 441-462.

Lindgren, P.B. (1997) The role of hrp genes during plant-bacterial interactions. Annu Rev Phytopathol 35.

Lindgren, P.B., Peet, R.C., and Panopoulos, N.J. (1986) Gene cluster of Pseudomonas syringae pv. phaseolicola controls pathogenicity on bean plants and hypersensitivity on nonhost plants. J Bacteriol 168: 512-.

Liu, Y., Cui, Y., Mukherjee, A., and Chatterjee, A.K. (1998) Characterization of a novel RNA regulator of Erwinia carotovora ssp. carotovora that controls production of extracellular enzymes and secondary metabolites. Mol Microbiol 29: 219-234.

Lloyd, S.A., Norman, M., Rosqvist, R., and Wolf-Watz, H. (2001) Yersinia YopE is targeted for type III secretion by N-terminal, not mRNA, signals. Mol Microbiol 39: 520-531.

Lloyd, S.A., Sjostrom, M., Andersson, S., and Wolf-Watz, H. (2002) Molecular characterization of type III secretion signals via analysis of synthetic N-terminal amino acid sequences. Mol Microbiol 43: 51-59.

Lockwood, J.L., and Williams, L.E. (1956) Field experiments for control of bacterial wilt of sweet corn by antibiotic and Tween 20 sprays Plant Dis Reptr 40: 622-625.

Loh, J., Garcia, M., and Stacey, G. (1997) NodV and NodW, a second flavonoid recognition system regulating nod gene expression in Bradyrhizobium japonicum. J Bacteriol 179: 3013-3020.

Loh, J., Pierson, E.A., Pierson, L.S., 3rd, Stacey, G., and Chatterjee, A. (2002) Quorum sensing in plant-associated bacteria. Curr Opin Plant Biol 5: 285-290.

396 Lonetto, M.A., Brown, K.L., Rudd, K.E., and Buttner, M.J. (1994) Analysis of the Streptomyces coelicolor sigE gene reveals the existence of a subfamily of eubacterial RNA polymerase sigma factors involved in the regulation of extracytoplasmic functions. Proc Natl Acad Sci U S A 91: 7573-7577.

Lorang, J.M., and Keen, N.T. (1995) Charaterization of avrE from Pseudomonas syringae pv. tomato: a hrp-linked avirulence locus consisting of at least two transcriptional units. Mol Plant-Microbe Interact 8: 49-57.

Lukat, G.S., McCleary, W.R., Stock, A.M., and Stock, J.B. (1992) Phosphorylation of bacterial response regulator proteins by low molecular weight phospho-donors. Proc Natl Acad Sci U S A 89: 718-722.

Maloy, S.R., Stewart, V.J., and Taylor, R.K. (1996) Genetic Analysis of Pathogenic Bacteria. A laboratory Manual. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY.

Manulis, S., Gafni, Y., Clark, E., Zutra, D., Ophir, Y., and Barash, I. (1991) Identification of a plasmid DNA probe for detection of Erwinia herbicola pathogenic on Gypsophila paniculata. Phytopathology 81: 54- 57.

Marchler-Bauer, A., Panchenko, A.R., Shoemaker, B.A., Thiessen, P.A., Geer, L.Y., and Bryant, S.H. (2002) CDD: a database of conserved domain alignments with links to domain three-dimensional structure. Nucleic Acids Res 30: 281-283.

Marenda, M., Brito, B., Callard, D., Genin, S., Barberis, P., Boucher, C., and Arlat, M. (1998) PrhA controls a novel regulatory pathway required for the specific induction of Ralstonia solanacearum hrp genes in the presence of plant cells. Mol Microbiol 27: 437-453.

Maris, A.E., Sawaya, M.R., Kaczor-Grzeskowiak, M., Jarvis, M.R., Bearson, S.M., Kopka, M.L. et al. (2002) Dimerization allows DNA target site recognition by the NarL response regulator. Nat Struct Biol 9: 771-778.

397 Marois, E., Van den Ackerveken, G., and Bonas, U. (2002) The Xanthomonas type III effector protein AvrBs3 modulates plant gene expression and induces cell hypertrophy in the susceptible host. Mol Plant Microbe Interact 15: 637-646.

Mascher, F., Moenne-Loccoz, Y., Schnider-Keel, U., Keel, C., Haas, D., and Defago, G. (2002) Inactivation of the regulatory gene algU or gacA can affect the ability of biocontrol Pseudomonas fluorescens CHA0 to persist as culturable cells in nonsterile soil. Appl Environ Microbiol 68: 2085-2088.

McCammon, S.L., Coplin, D.L., and Rowan, R.G. (1985) Isolation of avirulent mutants of Erwinia stewartii using bacteriophage Mu pf7701. J Gen Microbiol 131: 2993-3000.

McCleary, W.R., Stock, J.B., and Ninfa, A.J. (1993) Is acetyl phosphate a global signal in Escherichia coli? J. Bacteriol. 175: 2793-2798.

McCulloch, L. (1918) Phytopathology 1:440.

McDowell, J.M., and Dangl, J.L. (2000) Signal transduction in the plant immune response. Trends Biochem Sci 25: 79-82.

McFarland, N., McCarter, L., Artz, S., and Kustu, S. (1981) Nitrogen regulatory locus glnR of enteric bacteria is composed of cistrons ntrB and ntrC, Identification of their protein products. Proc Natl Acad Sci U S A 78: 2135-2139.

McGowan, S., Sebaihia, M., Jones, S., Yu, B., Bainton, N., Chan, P.F. et al. (1995) Carbapenem antibiotic production in Erwinia carotovora is regulated by carR, a homolog of the LuxR transcriptional activator. Microbiology 141: 541-550.

McGuffin, L.J., Bryson, K., et al. (2000) The PSIPRED protein structure prediction server. Bioinformatics 16: 404-405.

Menard, R., Sansonetti, P.J., and Parsot, C. (1993) Nonpolar mutagenesis of the ipa genes defines IpaB, IpaC, and IpaD as effectors of Shigella flexneri entry into epithelial cells. J Bacteriol 175: 5899-5906. 398 Mergaert, J., Verdonck, L., and Kersters, K. (1993) Transfer of Erwinia ananas (synonym, Erwinia uredovora) and Erwinia stewartii to the genus Pantoea emend. as Pantoea ananas (Serrano 1928) comb. nov. and Pantoea stewartii (Smith 1898) comb. nov., respectively, and description of Pantoea stewartii subsp. indologenes subsp. nov. In J Syst Bacteriol 43: 162-173.

Merighi, M., Majerczak, D.R., and Coplin, D.L. (2001) The hrp genes of Pantoea stewartii are regulated by a complex system that senses environmental signals. Plant Pathogenic Bacteria, Proceedings of the International Conference on Plant Pathogenic Bacteria, 10th, Charlottetown, PE, Canada, July 23-27, 2000: 201-204.

Merighi, M., Majerczak, D.R., Stover, E.H., and Coplin, D.L. (2003) The HrpX/HrpY two-component system activates hrpS expression, the first step in the regulatory cascade controlling the Hrp regulon in Pantoea stewartii subsp. stewartii. Mol Plant Microbe Interact 16: 238-248.

Metcalf, W.W., Jiang, W., Daniels, L.L., Kim, S.K., Haldimann, A., and Wanner, B.L. (1996) Conditionally replicative and conjugative plasmids carrying lacZα for cloning, mutagenesis, and allele replacement in bacteria. Plasmid 35: 1-13.

Mettke, I., Fiedler, U., and Weiss, V. (1995) Mechanism of activation of a response regulator: interaction of NtrC-P dimers induces ATPase activity. J Bacteriol 177: 5056-5061.

Michiels, T., and Cornelis, G.R. (1991) Secretion of hybrid proteins by the Yersinia Yop export system. J Bacteriol 173: 1677-1685.

Michiels, T., Vanooteghem, J.C., Lambert de Rouvroit, C., China, B., Gustin, A., Boudry, P., and Cornelis, G.R. (1991) Analysis of virC, an operon involved in the secretion of Yop proteins by Yersinia enterocolitica. J Bacteriol 173: 4994-5009.

Miller, J.H. (1992) A short course in bacterial genetics: A laboratory manual and handbook for Escherichia coli and related bacteria. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY, U.S.A.

399 Miller, M.B., and Bassler, B.L. (2001) Quorum sensing in bacteria. Annu Rev Microbiol 55: 165-199.

Miller, W.G., Leveau, J.H., and Lindow, S.E. (2000) Improved gfp and inaZ broad host-range promoter-probe vectors. Mol Plant Microbe Interact 13: 1243-1250.

Ming, R., Brewbaker, J.L., Moon, H.G., Musket, T.A., Holley, R.N., Pataky, J.K., and McMuller, M.D. (1999) Identification of RFLP markers linked to a major gene, swl, conferring resistance to Stewart's wilt in maize. Maydica 44: 319-323.

Minogue, T.D., Wehland-Von Trebra, M., Bernhard, F., and von Bodman, S.B. (2002) The autoregulatory role of EsaR, a quorum-sensing regulator in Pantoea stewartii ssp. stewartii: evidence for a repressor function. Mol Microbiol44: 1625-1635.

Minogue, T.D., Wehland-Von Trebra, M., Bernhard, F., and von Bodman, S.B. (2002) The autoregulatory role of EsaR, a quorum-sensing regulator in Pantoea stewartii ssp. stewartii: evidence for a repressor function. Mol Microbiol 44: 1625-1635.

Mizuno, T. (1997) Compilation of all genes encoding two-component phosphotransfer signal transducers in the genome of Escherichia coli. DNA Res 4: 161-168.

Mok, K.C., Wingreen, N.S., and Bassler, B.L. (2003) Vibrio harveyi quorum sensing: a coincidence detector for two autoinducers controls gene expression. EMBO J 22: 870-881.

Monson, E.K., Weinstein, M., Ditta, G.S., and Helinski, D.R. (1992) The FixL protein of Rhizobium meliloti can be separated into a heme- binding oxygen-sensing domain and a functional C-terminal kinase domain. Proc Natl Acad Sci U S A 89: 4280-4284.

Moore, J.B., Shiau, S.P., and Reitzer, L.J. (1993) Alterations of highly conserved residues in the regulatory domain of nitrogen regulator I (NtrC) of Escherichia coli. J. Bacteriol. 175: 2692-2701.

400 Mor, H., Manulis, S., Zuck, M., Nizan, R., Coplin, D.L., and Barash, I. (2001) Genetic organization of the hrp gene cluster and dspAE/BF operon in Erwinia herbicola pv. gypsophilae. Mol Plant Microbe Interact 14: 431-436.

Morris, S.W., Vernooij, B., Titatarn, S., Starrett, M., Thomas, S., Wiltse, C.C. et al. (1998) Induced resistance responses in maize. Mol Plant Microbe Interact 11: 643-658.

Mouslim, C., and Groisman, E.A. (2003) Control of the Salmonella ugd gene by three two-component regulatory systems. Mol Microbiol 47: 335- 344.

Mudgett, M.B., and Staskawicz, B.J. (1999) Characterization of the Pseudomonas syringae pv. tomato AvrRpt2 protein: demonstration of secretion and processing during bacterial pathogenesis. Mol Microbiol 32: 927-941.

Mukherjee, A., Cui, Y., Liu, Y., and Chatterjee, A.K. (1997) Molecular characterization and expression of the Erwinia carotovora hrpNEcc gene, which encodes an elicitor of the hypersensitive reaction. Mol Plant Microbe Interact 10: 462-471.

Mukherjee, A., Cui, Y., Liu, Y., Dumenyo, C.K., and Chatterjee, A.K. (1996) Global regulation in Erwinia species by Erwinia carotovora rsmA, a homolog of Escherichia coli csrA: repression of secondary metabolites, pathogenicity and hypersensitive reaction. Microbiology 142: 427-434.

Mukherjee, A., Cui, Y., Ma, W., Liu, Y., Ishihama, A., Eisenstark, A., and Chatterjee, A.K. (1998) RpoS (sigma-S) controls expression of rsmA, a global regulator of secondary metabolites, harpin, and extracellular proteins in Erwinia carotovora. J Bacteriol 180: 3629-3634.

Mukhopadhyay, S., Audia, J.P., Roy, R.N., and Schellhorn, H.E. (2000) Transcriptional induction of the conserved alternative sigma factor RpoS in Escherichia coli is dependent on BarA, a probable two- component regulator. Mol Microbiol 37: 371-381.

401 Muller, S., Feldman, M.F., and Cornelis, G.R. (2001) The Type III secretion system of Gram-negative bacteria: a potential therapeutic target? Expert Opin Ther Targets 5: 327-339.

Munkvold, G.P., Iles, A., and McGee, D.C. (1996) Effects of imidacloprid seed treatment of corn on foliar feeding and Erwinia stewartii transmission by the corn flea beetle. Plant Dis 80:747-749 :.

Nakayama, S., and Watanabe, H. (1995) Involvement of cpxA, a sensor of a two-component regulatory system, in the pH-dependent regulation of expression of Shigella sonnei virF gene. J Bacteriol 177: 5062-5069.

Natti, J.J. (1958) Effect of seed treatments with antibiotics on bacterial wilt of corn. Plant Dis Reptr 42: 953-958.

Nealson, K.H., Platt, T., and Hastings, J.W. (1970) Cellular control of the synthesis and activity of the bacterial luminescence system. J Bacteriol 104.

Neidhardt, F.C. (ed.) (1993) Escherichia coli and Salmonella: Cellular and Molecular Biology, 1st Ed., Vol. 1-2. ASM Press, Washington, D. C.

Nixon, B.T., Ronson, C.W., and Ausubel, F.M. (1986) Two-component regulatory systems responsive to environmental stimuli share strongly conserved domains with the nitrogen assimilation regulatory genes ntrB and ntrC. Proc Natl Acad Sci U S A 83: 7850-7854.

Nizan, R., Barash, I., Valinsky, L., Lichter, A., and Manulis, S. (1997) The presence of hrp genes on the pathogenicity-associated plasmid of the tumorigenic bacterium Erwinia herbicola pv. gypsophilae. Mol Plant Microbe Interact 10: 677-682.

Nizan-Koren, R., Manulis, S., Mor, H., Iraki, N.M., and Barash, I. (2003) The regulatory cascade that activates the Hrp regulon in Erwinia herbicola pv. gypsophilae. Mol Plant Microbe Interact 16: 249-260.

Noel, L., Thieme, F., Nennstiel, D., and Bonas, U. (2001) cDNA-AFLP analysis unravels a genome-wide hrpG-regulon in the plant pathogen Xanthomonas campestris pv. vesicatoria. Mol Microbiol 41: 1271-1281. 402 Noel, L., Thieme, F., Nennstiel, D., and Bonas, U. (2002) Two novel type III-secreted proteins of Xanthomonas campestris pv. vesicatoria are encoded within the hrp pathogenicity island. J Bacteriol 184: 1340-1348.

Ochman, H., Soncini, F.C., Solomon, F., and Groisman, E.A. (1996) Identification of the pathogenicity island required for Salmonella survival in host cells. Proc Natl Acad Sci USA 93: 7800-7804.

Olekhnovich, I.N., and Kadner, R.J. (2002) Mutational scanning and affinity cleavage analysis of UhpA-binding sites in the Escherichia coli uhpT promoter. J Bacteriol 184: 2682-2691.

Parkins, M.D., Ceri, H., and Storey, D.G. (2001) Pseudomonas aeruginosa GacA, a factor in multihost virulence, is also essential for biofilm formation. Mol Microbiol 40: 1215-1226.

Parsek, M.R., and Greenberg, E.P. (2000) Acyl-homoserine lactone quorum sensing in gram-negative bacteria: a signaling mechanism involved in associations with higher organisms. Proc Natl Acad Sci U S A 97: 8789-8793.

Pasinetti, L. (1937) La "bacteriosi" del mais in Italia da "Aplanobacter stewartii" Smith. Nota II. Riv Pat Veg 27: 221-229.

Passador, L., Cook, J.M., Gambello, M.J., Rust, L., and Iglewski, B.H. (1993) Expression of Pseudomonas-Aeruginosa Virulence Genes Requires Cell-to-Cell Communication. Science 260: 1127-1130.

Pataky, J.K. (2000) Pest risk analysis:: The risk of introducing Erwinia stewartii in maize seed. Pest risk analysis. International Seed Trade Federation.

Pataky, J.K., du Toit, L.J., and Freeman, N.D. (2000) Stewart's wilt reactions of an international collection of Zea mays germ plasm inoculated with Erwinia stewartii. Plant Dis 84: 901-905.

403 Pataky, J.K., du Toit, L.J., Revilla, P., and Tracy, W.F. (1998) Reactions of open-pollinated sweet conr cultivars to Stewart's wilt, common rust, northern leaf blight, and southern leaf blight. Plant Dis 82: 939-944.

Pataky, J.K., Michener, P.M., Freeman, N.D., Weinzierl, R.A., and Teyker, R.H. (2000) Control of Stewart's wilt in sweet corn with seed treatment insecticides. Plant Dis 84: 1104-1108.

Pataky, J.K. (2003) Stewart's wilt of corn. APSnet Feature Article. URL: www.apsnet.org/online/feature/stewarts.

Pearson, J.P., Passador, L., Iglewski, B.H., Greenberg, E.P., and Seed, P.C. (1995) A second N-acylhomoserine lactone signal produced by Pseudomonas aeruginosa. Activation of the Pseudomonas aeruginosa lasI gene by LasR and the Pseudomonas autoinducer PAI: an autoinduction regulatory hierarchy. Proc Natl Acad Sci USA 92: 1490-1494.

Pelton, J.G., Kustu, S., and Wemmer, D.E. (1999) Solution structure of the DNA-binding domain of NtrC with three alanine substitutions. J Mol Biol 292: 1095-1110.

Pepper, E.H. (1967) Stewart's Bacterial Wilt of Corn. American Phytopathological Society, St. Paul, MN.

Perino, C., Gaudriault, S., Vian, B., and Barny, M.A. (1999) Visualization of harpin secretion in planta during infection of apple seedlings by Erwinia amylovora. Cell Microbiol 1: 131-141.

Pernestig, A.K., Georgellis, D., Romeo, T., Suzuki, K., Tomenius, H., Normark, S., and Melefors, O. (2003) The Escherichia coli BarA-UvrY two-component system is needed for efficient switching between glycolytic and gluconeogenic carbon sources. J Bacteriol 185: 843-853.

Pernestig, A.K., Melefors, O., and Georgellis, D. (2001) Identification of UvrY as the cognate response regulator for the BarA sensor kinase in Escherichia coli. J Biol Chem 276: 225-231.

404 Perombelon, M.C.M. (2002) Potato diseases caused by soft rot Erwinias: an overview of pathogenesis. Plant Pathology 51: 1-12.

Perret, X., Freiberg, C., Rosenthal, A., Broughton, W.J., and Fellay, R. (1999) High-resolution transcriptional analysis of the symbiotic plasmid of Rhizobium sp. NGR234. Mol Microbiol 32: 415-425.

Perutz, M.F., Paoli, M., and Lesk, A.M. (1999) Fix L, a haemoglobin that acts as an oxygen sensor: signalling mechanism and structural basis of its homology with PAS domains. Chem Biol 6: R291-297.

Pessi, G., Williams, F., Hindle, Z., Heurlier, K., Holden, M.T., Camara, M. et al. (2001) The global posttranscriptional regulator RsmA modulates production of virulence determinants and N-acylhomoserine lactones in Pseudomonas aeruginosa. J Bacteriol 183: 6676-6683.

Petnicki-Ocwieja, T., Schneider, D.J., Tam, V.C., Chancey, S.T., Shan, L., Jamir, Y. et al. (2002) Genomewide identification of proteins secreted by the Hrp type III protein secretion system of Pseudomonas syringae pv. tomato DC3000. Proc Natl Acad Sci U S A 99: 7652-7657.

Petterson, J., Nordfelth, R., Dubinina, R., Bergman, T., Gustafsson, M., Magnusson, K.E., and Wolf-Watz, H. (1996) Modulation of virulence factor expression by pathogen-target cell contact. Science 273: 1231-1233.

Pierson, L.S., Wood, D.W., and Pierson, E.A. (1998) Homoserine lactone-mediated gene regulation in plant-associated bacteria. Annual Review of Phytopathology 36: 207-225.

Pike, S.M., Adam, A.L., Pu, X.A., Hoyos, M.E., Laby, R., Beer, S.V., and Novacky, A. (1998) Effects of Erwinia amylovora harpin on tobacco leaf cell membranes are related to leaf necrosis and electrolyte leakage and distinct from perturbations caused by inoculated E- amylovora. Physiological and Mol Plant Pathol 53: 39-60.

Piper, K.R., Beck von Bodman, S., and Farrand, S.K. (1993) Conjugation factor of Agrobacterium tumefaciens regulates Ti plasmid transfer by autoinduction. Nature 362: 570-572.

405 Pirhonen, M., Flego, D., Heikinheimo, R., and Palva, E.T. (1993) A Small diffusible signal molecule is responsible for the global control of virulence and exoenzyme production in the plant pathogen Erwinia- Carotovora. EMBO Journal 12: 2467-2476.

Poetter, K., and Coplin, D.L. (1991) Structural and functional analysis of the rcsA gene from Erwinia stewartii. Mol Gen Genetics 229:155-160.

Poos, F.W. (1939) Host plants harboring Aplanobacter stewartii without showing external symptoms after inoculation by Chaetocnema pulicaria. J Econ Ent 32: 881-882.

Porter, S.C., North, A.K., and Kustu, S. (1995) Mechanism of transcriptional activation by NtrC. Pages 147-158 in: Two-Component Signal Transduction. Hoch, J.A. and Silhavy, T.J. (eds). American Society for Microbiology Press, Washington, D.C., U.S.A.

Porter, S.C., North, A.K., Wedel, A.B., and Kustu, S. (1993) Oligomerization of NtrC at the glnA enhancer is required for transcriptional activation. Genes & Devel 7: 2258-2273.

Preston, G., Deng, W.L., Huang, H.C., and Collmer, A. (1998) Negative regulation of hrp genes in Pseudomonas syringae by HrpV. J Bacteriol 180: 4532-4537.

Preston, G., Huang, H.C., He, S.Y., and Collmer, A. (1995) The HrpZ proteins of Pseudomonas syringae pvs. syringae, glycinea, and tomato are encoded by an operon containing Yersinia Ysc homologs and elicit the hypersensitive response in tomato but not soybean. Mol Plant Microbe Interact 8: 717-732.

Preston, G.M., Bertrand, N., and Rainey, P.B. (2001) Type III secretion in plant growth-promoting Pseudomonas fluorescens SBW25. Mol Microbiol 41: 999-1014.

Prouty, A.M., and Gunn, J.S. (2000) Salmonella enterica serovar typhimurium invasion is repressed in the presence of bile. Infect Immun 68: 6763-6769.

406 Rahme, L.G., Mindrinos, M.N., and Panopoulos, N.J. (1991) Genetic and transcriptional organization of the hrp cluster of Pseudomonas syringae pv phaseolicola. J Bacteriol 173: 575-586.

Rahme, L.G., Mindrinos, M.N., and Panopoulos, N.J. (1992) Plant and environmental sensory signals control the expression of hrp genes in Pseudomonas syringae pv phaseolicola. J Bacteriol 174: 3499-3507.

Raivio, T.L., and Silhavy, T.J. (2001) Periplasmic stress and ECF sigma factors. Annu Rev Microbiol 55: 591-624.

Rampersaud, A., Harlocker, S.L., and Inouye, M. (1994) The OmpR protein of Escherichia coli binds to sites in the ompF promoter region in a hierarchical manner determined by its degree of phosphorylation. J Biol Chem 269: 12559-12566.

Rampersaud, A., Norioka, S., and Inouye, M. (1989) Characterization of OmpR binding sequences in the upstream region of the ompF promoter essential for transcriptional activation. J Biol Chem 264: 18693-18700.

Rand, F.V., and Cash, L.C. (1924) Further evidence of insect dissemination of bacterial wilt of corn. Science 59: 67-69.

Rantakari, A., Virtaharju, O., Vahamiko, S., Taira, S., Palva, E.T., Saarilahti, H.T., and Romantschuk, M. (2001) Type III secretion contributes to the pathogenesis of the soft-rot pathogen Erwinia carotovora: partial characterization of the hrp gene cluster. Mol Plant Microbe Interact 14: 962-968.

Reese, M.G., Harris, N.L., and Eckman, F.H. (1996) Large scale sequencing specific neural networks for promoter and splice site secognition. Biocomputing: Proceedings of the 1996 Pacific Symposium. Hunter, L. and Klein, T.E. (eds).

Reimmann, C., Beyeler, M., Latifi, A., Winteler, H., Foglino, M., Lazdunski, A., and Haas, D. (1997) The global activator GacA of Pseudomonas aeruginosa PAO positively controls the production of the autoinducer N-butyryl-homoserine lactone and the formation of the virulence factors pyocyanin, cyanide, and lipase. Mol Microbiol 24: 309- 319. 407 Reinelt, S., Hofmann, E., Gerharz, T., Bott, M., and Madden, D.R. (2003) The structure of the periplasmic ligand-binding domain of the sensor kinase CitA reveals the first extracellular PAS domain. J Biol Chem.

Repik, A., Rebbapragada, A., Johnson, M.S., Haznedar, J.O., Zhulin, I.B., and Taylor, B.L. (2000) PAS domain residues involved in signal transduction by the Aer redox sensor of Escherichia coli. Mol Microbiol 36: 806-816.

Rich, J.J., and Willis, D.K. (1997) Multiple loci of Pseudomonas syringae pv. syringae are involved in pathogenicity on bean: restoration of one lesion-deficient mutant requires two tRNA genes. J Bacteriol 179: 2247-2258.

Rich, J.J., Hirano, S.S., and Willis, D.K. (1992) Pathovar-specific requirement for the Pseudomonas syringae lemA gene in disease lesion formation. Appl Environ Microbiol 58: 1440-1446.

Rich, J.J., Kinscherf, T.G., Kitten, T., and Willis, D.K. (1994) Genetic evidence that the gacA gene encodes the cognate response regulator for the lemA sensor in Pseudomonas syringae. J Bacteriol 176: 7468-7475.

Ritter, C., and Dangl, J.L. (1995) The avrRpm1 gene of Pseudomonas syringae pv. maculicola is required for virulence on Arabidopsis. Mol Plant Microbe Interact 8: 444-453.

Roine, E., Saarinen, J., Kalkkinen, N., and Romantschuk, M. (1997) Purified HrpA of Pseudomonas syringae pv. tomato DC3000 reassembles into pili. FEMS Microbiol Lett 417: 168-172.

Roine, E., Wei, W., Yuan, J., Nurmiaho-Lassila, E.L., Kalkkinen, N., Romantschuk, M., and He, S.Y. (1997) Hrp pilus: an hrp-dependent bacterial surface appendage produced by Pseudomonas syringae pv. tomato DC3000. Proc Natl Acad Sci U S A 94: 3459-3464.

Rombel, I., Peters-Wendisch, P., Mesecar, A., Thorgeirsson, T., Shin, Y.K., and Kustu, S. (1999) MgATP binding and hydrolysis determinants of NtrC, a bacterial enhancer-binding protein. J Bacteriol 181: 4628- 4638. 408 Rosenberg, C., Casse-Delbart, F., Dusha, I., David, M., and Boucher, C. (1982) Megaplasmids in the plant-associated bacteria Rhizobium meliloti and Pseudomonas solanacearum. J Bacteriol 150: 402-406.

Rossier, O., Van den Ackerveken, G., and Bonas, U. (2000) HrpB2 and HrpF from Xanthomonas are type III-secreted proteins and essential for pathogenicity and recognition by the host plant. Mol Microbiol 38: 828- 838.

Rutter, J., Michnoff, C.H., Harper, S.M., Gardner, K.H., and McKnight, S.L. (2001) PAS kinase: an evolutionarily conserved PAS domain-regulated serine/threonine kinase. Proc Natl Acad Sci U S A 98: 8991-8996.

Ruvkun, G.B., and and Ausubel, F.M. (1981) A general method for site- directed mutagenesis in prokaryotes. Nature 289: 85-88.

Sakai, T., Sasakawa, C., Makino, S., and Yoshikawa, M. (1986) DNA sequence and product analysis of the virF locus responsible for Congo red binding and cell invasion in Shigella flexneri 2a. Infect Immun 54: 395- 402.

Salanoubat, M., Genin, S., Artiguenave, F., Gouzy, J., Mangenot, S., Arlat, M. et al. (2002) Genome sequence of the plant pathogen Ralstonia solanacearum. Nature 415: 497-502.

Salmeron, J.M., and Staskawicz, B.J. (1993) Molecular characterization and hrp dependence of the avirulence gene avrPto from Pseudomonas syringae pv. tomato. Mol Gen Genet 239: 6-16.

Sandermann, H., Jr. (2000) Active oxygen species as mediators of plant immunity: three case studies. Biol Chem 381: 649-653.

Sanders, D.A., Gillece-Castro, B.L., Burlingame, A.L., and Koshland, D.E., Jr. (1992) Phosphorylation site of NtrC, a protein phosphatase whose covalent intermediate activates transcription. J Bacteriol 174: 5117-5122.

409 Sandkvist, M. (2001) Biology of type II secretion. Mol Microbiol 40: 271-283.

Sands, D.C., Heichel, G.H., and Kring, J.B. (1979) Carbofuran reduction of Stewart's disease in corn inoculated with Erwinia stewartii. Plant Dis Rep 63: 631-633.

Schauder, S., Shokat, K., Surette, M.G., and Bassler, B.L. (2001) The LuxS family of bacterial autoinducers: biosynthesis of a novel quorum- sensing signal molecule. Mol Microbiol 41: 463-476.

Scholten, M., and Tommassen, J. (1993) Topology of the PhoR protein of Escherichia coli and functional analysis of internal deletion mutants. Mol Microbiol 8: 269-275.

Schulte, R., and Bonas, U. (1992) Expression of the Xanthomonas campestris pv. vesicatoria hrp gene cluster, which determines pathogenicity and hypersensitivity on pepper and tomato, is plant inducible. J Bacteriol 174: 815-823.

Shao, F., Merritt, P.M., Bao, Z., Innes, R.W., and Dixon, J.E. (2002) A Yersinia effector and a Pseudomonas avirulence protein define a family of cysteine proteases functioning in bacterial pathogenesis. Cell 109: 575- 588.

Shattuck-Eidens, D. M. and Kadner, R. J. (1983) Molecular cloning of the uph region and evidence for a positive activator for expression of the hexose phosphate transport system of Escherichia coli. J. Bacteriol. 155:1060-1070.

Shaw, C.H., Ashby, A.M., Brown, A., Royal, C., and Loake, G.J. (1988) virA and virG are the Ti-plasmid functions required for chemotaxis of Agrobacterium tumefaciens towards acetosyringone. Mol Microbiol 2: 413-417.

Shea, J.E., Hensel, M., Gleeson, C., and Holden, D. (1996) Identification of a virulence locus encoding a second type III secretion system in Salmonella typhimurium. Proc Natl Acad Sci USA 93: 2593- 2597.

410 Shen, H., and Keen, N.T. (1993) Characterization of the promoter of avirulence gene D from Pseudomonas syringae pv. tomato. J Bacteriol 175: 5916-5924.

Shirasu, K., Dixon, R.A., and Lamb, C. (1996) Signal transduction in plant immunity. Curr Opin Immunol 8: 3-7.

Simon, R., Priefer, U., and Puhler, A. (1983) A broad host range mobilization system for in vivo genetic engineering: Transposon mutagenesis in gram-negative bacteria. Bio/Technology 1: 784-791.

Skerman, V.B.D., McGowan, V., and Sneath, P.H.A. (1980) Approved lists of bacteria names. Int J Syst Bacteriol 30: 225-420.

Smith, E.F. (1898) Address presented at the Proc Am Assoc Adv Science, Boston, August 1898, Published in Salem, Dec. 1898.

Smith, E.F. (1901) The cultural characters of Pseudomonas hyacinthi, Ps. campestris, Ps. phaseoli and Ps. stewartii- Four one-flagellate yellow bacteria parasitic on plants. Bulletin Div. of Veg. Physiology and Pathology, 28. U.S.Dept. of Agriculture.

Smith, E.F. (1903) Complete proof that Ps. stewartii is the cause of the sweet-corn disease of Long Island. Science 17: 457.

Smith, E.F. (1909) Seed corn as a means of dissemination of Bacterium stewartii. Science 30: 223-224.

Smith, E.F. (1914) Bacteria in Relation to lant Diseases- Vascular Diseases, Vol. 3. Carnegie Institution of Washington, Washington, D.C.

Soncini, F.C., and Groisman, E.A. (1996) Two-component regulatory systems can interact to process multiple environmental signals. J Bacteriol 178: 6796-6801.

Sory, M.P., and Cornelis, G.R. (1994) Translocation of a hybrid YopE- adenylate cyclase from Yersinia enterocolitica into HeLa cells. Mol Microbiol 14: 583-594.

411 Sperandio, V., Mellies, J.L., Nguyen, W., Shin, S., and Kaper, J.B. (1999) Quorum sensing controls expression of the type III secretion gene transcription and protein secretion in enterohemorrhagic and enteropathogenic Escherichia coli. Proc Natl Acad Sci U S A 96: 15196- 15201.

Sperandio, V., Torres, A.G., and Kaper, J.B. (2002) Quorum sensing Escherichia coli regulators B and C (QseBC): a novel two-component regulatory system involved in the regulation of flagella and motility by quorum sensing in E. coli. Mol Microbiol 43: 809-821.

Staskawicz, B., Dahlbeck, D., Keen, N., and Napoli, C. (1987) Molecular characterization of cloned avirulence genes from race 0 and race 1 of Pseudomonas syringae pv. glycinea. J. Bacteriol. 169: 5789- 5794.

Staskawicz, B.J., Dahlbeck, K.D., and Keen, N.T. (1984) Cloned avirulence gene of Pseudomonas syringae pv glycinea determines race- specific incompatibility on Glycine max (L.) Merr. Proc Natl Acad Sci U S A 81.

Staskawicz, B.J., Mudgett, M.B., Dangl, J.L., and Galan, J.E. (2001) Common and contrasting themes of plant and animal diseases. Science 292: 2285-2289.

Steinberger, E.M., and Beer, S.V. (1988) Creation and complementation of pathogenicity mutants of Erwinia amylovora. Mol Plant-Microbe Interact 1: 135-.

Stephenson, K., and Hoch, J.A. (2001) PAS-A domain of phosphorelay sensor kinase A: a catalytic ATP-binding domain involved in the initiation of development in Bacillus subtilis. Proc Natl Acad Sci U S A 98: 15251- 15256.

Stevens, N.E. (1934) Stewart's disease in relation to winter temperatures. Plant Dis Rep 18: 141-149.

Stevenson, F.J., and Jones, H.A. (1953) Some sources of reistance in crop plants. In U S Dept Agric Yearbook - Plant Diss. U.S.D.A., pp. 192- 216. 412 Stewart, F.C. (1897) A bacterial disease of sweet corn. New York Agric Exp Sta Bull, Geneva NY 130: 423-439.

Stewart, V. (1993) Nitrate regulation of anaerobic respiratory gene expression in Escherichia coli. Mol Microbiol 9: 425-434.

Stock, A.M., Robinson, V.L., and Goudreau, P.N. (2000) Two- component signal transduction. Annu Rev Biochem 69: 183-215.

Stock, J., Surette, M., and Park, P. (1994) Chemosensing and signal transduction in bacteria. Curr Opin Neurobiol 4: 474-480.

Su, W., Porter, S., Kustu, S., and Echols, H. (1990) DNA-Looping and enhancer activity - association between DNA- bound NtrC activator and RNA polymerase at the bacterial glnA promoter. Proc Natl Acad Sci U S A 87: 5504-5508.

Suparyono, and Pataky, J.K. (1989) Influence of host resistance and growth stage at the time of inoculation on Stewart's wilt and Goss' wilt. Plant Dis 73: 339-345.

Surette, M.G., Miller, M.B., and Bassler, B.L. (1999) Quorum sensing in Escherichia coli, Salmonella typhimurium, and Vibrio harveyi: a new family of genes responsible for autoinducer production. Proc Natl Acad Sci U S A 96: 1639-1644.

Suzuki, K., Wang, X., Weilbacher, T., Pernestig, A.K., Melefors, O., Georgellis, D. et al. (2002) Regulatory circuitry of the CsrA/CsrB and BarA/UvrY systems of Escherichia coli. J Bacteriol 184: 5130-5140.

Swanson, R. V., Alex, L. A., Simon, M. I. (1994) Histidine and aspartate phosphorylation: Two-component systems and the limits of homology. Trends Biol Sci 19:485-490.

Swarup, S., De Feyter, R., Brlansky, R.H., and Gabriel, D.W. (1991) A pathogenicity locus from Xanthomonas citri enables strains from several pathovars of X. campestris to elicit canker-like lesions on citrus. Phytopathology 81:802.

413 Swarup, S., Yang, Y., Kingsley, M.T., and Gabriel, D.W. (1992) Xanthomonas citri pathogenicity gene, pthA, pleiotropically encodes gratuitous avirulence on nonhosts. Mol Plant-Microbe Interact 5: 204-.

Swords, K.M., Dahlbeck, D., Kearney, B., Roy, M., and Staskawicz, B.J. (1996) Spontaneous and induced mutations in a single open reading frame alter both virulence and avirulence in Xanthomonas campestris pv. vesicatoria avrBs2. J Bacteriol 178: 4661-4669.

Szurek, B., Marois, E., Bonas, U., and Van den Ackerveken, G. (2001) Eukaryotic features of the Xanthomonas type III effector AvrBs3: protein domains involved in transcriptional activation and the interaction with nuclear import receptors from pepper. Plant J 26: 523-534.

Szurek, B., Rossier, O., Hause, G., and Bonas, U. (2002) Type III- dependent translocation of the Xanthomonas AvrBs3 protein into the plant cell. Mol Microbiol 46: 13-23.

Tan, M.W., Rahme, L.G., Sternberg, J.A., Tompkins, R.G., and Ausubel, F.M. (1999) Pseudomonas aeruginosa killing of Caenorhabditis elegans used to identify P. aeruginosa virulence factors. Proc Natl Acad Sci U S A 96: 2408-2413.

Taylor, B.L., and Zhulin, I.B. (1999) PAS domains: internal sensors of oxygen, redox potential, and light. Microbiol. Mol. Biol. Review. 63: 479- 506.

Teplitski, M., Goodier, R. I., Ahmer, M. M. (2003) Pathways leading from BarA/SirA to motility and virulence gene expression in Salmonella. J Bacteriol 185: (in press)

Thomas, R.C. (1938) Transmissible lysins in water extracts of seeds. Science 88: 56-57.

Thomas, R. (1978) Logical analysis of systems comprising feedback loops. J Theor Biol 73: 631-656.

Thomas, R. (1973) Boolean formalization of genetic control circuits. J Theor Biol 42: 563-585. 414 Tiffin, P., and Gaut, B.S. (2001) Molecular evolution of the wound- induced serine protease inhibitor wip1 in Zea and related genera. Mol Biol Evolut 18: 2092-2101.

Torres, A.G., and Kaper, J.B. (2002) Pathogenicity islands of intestinal E. coli. Curr Top Microbiol Immunol 264: 31-48.

Torres-Cabassa, A., Gottesman, S., Frederick, R.D., Dolph, P.J., and Coplin, D.L. (1987) Control of extracellular polysaccharide synthesis in Erwinia stewartii and Escherichia coli K-12: a common regulatory function. J Bacteriol 169: 4525-4531.

Tsiamis, G., Mansfield, J.W., Hockenhull, R., Jackson, R.W., Sesma, A., Athanassopoulos, E. et al. (2000) Cultivar-specific avirulence and virulence functions assigned to avrPphF in Pseudomonas syringae pv. phaseolicola, the cause of bean halo-blight disease. EMBO J 19: 3204- 3214.

Tsuzuki, M., Ishige, K., and Mizuno, T. (1995) Phosphotransfer circuitry of the putative multi-signal transducer, ArcB, of Escherichia coli: in vitro studies with mutants. Mol Microbiol 18: 953-962.

Tyson, K.L., Cole, J.A., and Busby, S.J. (1994) Nitrite and nitrate regulation at the promoters of two Escherichia coli operons encoding nitrite reductase: identification of common target heptamers for both NarP- and NarL-dependent regulation. Mol Microbiol 13: 1045-1055.

Valinsky, L., Manulis, S., Nizan, R., Ezra, D., and Barash, I. (1998) A pathogenicity gene isolated from the pPATH plasmid of Erwinia herbicola pv. gypsophilae determines host specificity. Mol Plant Microbe Interact 11: 753-762.

Van den Ackerveken, G., Marois, E., and Bonas, U. (1996) Recognition of the bacterial avirulence protein AvrBs3 occurs inside the host plant cell. Cell 87: 1307-1316.

Van der Zwett, T., and Keil, H.L. (1979) Fire blight - A bacterial disease of rosaceous plants. U. S. Department of Agriculture, Washington, D. C., p.

415 Van Gijsegem, F., Genin, S., and Boucher, C. (1995) hrp and avr genes, key determinants controlling the interactions between plants and gram- negative phytopathogenic bacteria. In Prokaryotes: Pathogenesis and Host specificity in plant disease. (ed.^eds.). Pergamon/Elsevier Science LTD Pub., pp.

Van Gijsegem, F., Gough, C., Zischek, C., Niqueux, E., Arlat, M., Genin, S. et al. (1995) The hrp gene locus of Pseudomonas solanacearum, which controls the production of a type III secretion system, encodes eight proteins related to components of the bacterial flagellar biogenesis complex. Mol Microbiol 15: 1095-1114.

Van Gijsegem, F., Vasse, J., Camus, J.C., Marenda, M., and Boucher, C. (2000) Ralstonia solanacearum produces hrp-dependent pili that are required for PopA secretion but not for attachment of bacteria to plant cells. Mol Microbiol 36: 249-260.

Vasse, J., Genin, S., Frey, P., Boucher, C., and Brito, B. (2000) The hrpB and hrpG regulatory genes of Ralstonia solanacearum are required for different stages of the tomato root infection process. Mol Plant Microbe Interact 13: 259-267.

Viprey, V., Del Greco, A., Golinowski, W., Broughton, W.J., and Perret, X. (1998) Symbiotic implications of type III protein secretion machinery in Rhizobium. Mol Microbiol 28: 1381-1389.

Von Bodman, S.B., and Farrand, S.K. (1995) Capsular polysaccharide biosynthesis and pathogenicity in Erwinia stewartii require induction by an N-acylhomoserine lactone autoinducer. J Bacteriol 177: 5000-5008.

Von Bodman, S.B., Bauer, W.D., and Coplin, D.L. (2003) Quorum sensing in plant-pathogenic bacteria. Annu Rev Phytopathol 41:455-482.

Von Bodman, S.B., Majerczak, D.R., and Coplin, D.L. (1998) A negative regulator mediates quorum-sensing control of exopolysaccharide production in Pantoea stewartii subsp. stewartii. Proc Natl Acad Sci USA 95: 7687-7692.

Wagner, R. (2000) Transcription regulation in prokaryotes. Oxford University Press, Oxford, N.Y. 416 Wagner, V.E., Bushnell, D., Passador, L., Brooks, A.I., and Iglewski, B.H. (2003) Microarray analysis of Pseudomonas aeruginosa quorum- sensing regulons: effects of growth phase and environment. J Bacteriol 185: 2080-2095.

Waldee, E.L. (1941) Relationship of some bacterial plant pathogens to the coliform bacteria. Proc Iowa Acad Sci 48: 197.

Waleron, M., Waleron, K., Podhajska, A.J., and Lojkowska, E. (2002) Genotyping of bacteria belonging to the former Erwinia genus by PCR- RFLP analysis of a recA gene fragment. Microbiology 148:583-595.

Wang, L., Fabret, C., Kanamaru, K., Stephenson, K., Dartois, V., Perego, M., and Hoch, J.A. (2001) Dissection of the functional and structural domains of phosphorelay histidine kinase A of Bacillus subtilis. J Bacteriol 183: 2795-2802.

Warren, J.R. (1951) The use of radioisotopes in determining the distribution of Bacterium stewartii Erw. Smith within corn plants. Phytopathology 41: 794-800.

Watson, W.T., Minogue, T.D., Val, D.L., von Bodman, S.B., and Churchill, M.E. (2002) Structural basis and specificity of acyl- homoserine lactone signal production in bacterial quorum sensing. Mol Cell 9: 685-694.

Wedel, A., and Kustu, S. (1995) The bacterial enhancer-binding protein NtrC is a molecular machine - ATP hydrolysis is coupled to transcriptional activation. Genes & Devel 9: 2042-2052.

Wehland, M., Kiecker, C., Coplin, D.L., Kelm, O., Saenger, W., and Bernhard, F. (1999) Identification of an RcsA/RcsB recognition motif in the promoters of exopolysaccharide biosynthetic operons from Erwinia amylovora and Pantoea stewartii subspecies stewartii. J Biol Chem 274: 3300-3307.

417 Wei, W., Plovanich-Jones, A., Deng, W.L., Jin, Q.L., Collmer, A., Huang, H.C., and He, S.Y. (2000)a The gene coding for the Hrp pilus structural protein is required for type III secretion of Hrp and Avr proteins in Pseudomonas syringae pv. tomato. Proc Natl Acad Sci U S A 97: 2247-2252.

Wei, Z.M., and Beer, S.V. (1993) HrpI of Erwinia amylovora functions in secretion of harpin and is a member of a new protein family. J Bacteriol 175: 7958-7967.

Wei, Z.M., and Beer, S.V. (1995) hrpL activates Erwinia amylovora hrp gene transcription and is a member of the ECF subfamily of sigma factors. J Bacteriol 177: 6201-6210.

Wei, Z.M., Kim, J.F., and Beer, S.V. (2000)b Regulation of hrp genes and type III protein secretion in Erwinia amylovora by HrpX/HrpY, a novel two-component system, and HrpS. Mol Plant Microbe Interact 13: 1251-1262.

Wei, Z.M., Laby, R.J., Zumoff, C.H., Bauer, D.W., He, S.Y., Collmer, A., and Beer, S.V. (1992) Harpin, elicitor of the hypersensitive response produced by the plant pathogen Erwinia amylovora. Science 257: 85-88.

Wei, Z.M., Sneath, B.J., and Beer, S.V. (1992) Expression of Erwinia amylovora hrp genes in response to environmental stimuli. J Bacteriol 174: 1875-1882.

Weiss, D.S., Batut, J., Klose, K.E., Keener, J., and Kustu, S. (1991) The phosphorylated form of the enhancer-binding protein NtrC has an ATPase activity that is essential for activation of transcription. Cell 67: 155-167.

Wellhausen, E.J. (1936) Histological changes in resistant and susceptible strains of maize infected with Phytomonas stewartii. Phytopathology 26: 112-113.

Wellhausen, E.J. (1937) Effect of the genetic constitution of the host on the virulence of Phytomonas stewartii. Phytopathology 27: 1070-1089.

418 Wells, J.M., van der Zwet, T., and Butterfield, J.E. (1994) Differentiation of Erwinia species in the "herbicola" group by class analysis of cellular fatty acids. J Phytopathol 140:39-48.

Wengelnik, K., and Bonas, U. (1996) HrpXv, an AraC-type regulator, activates expression of five of the six loci in the hrp cluster of Xanthomonas campestris pv vesicatoria. J Bacteriol 178: 3462-3469.

Wengelnik, K., Rossier, O., and Bonas, U. (1999) Mutations in the regulatory gene hrpG of Xanthomonas campestris pv. vesicatoria result in constitutive expression of all hrp genes. J Bacteriol 181: 6828-6831.

Wengelnik, K., Van den Ackerveken, G., and Bonas, U. (1996) HrpG, a key hrp regulatory protein of Xanthomonas campestris pv. vesicatoria is homologous to two-component response regulators. Mol Plant Microbe Interact 9: 704-712.

Whistler, C.A., and Pierson, L.S., 3rd (2003) Repression of phenazine antibiotic production in Pseudomonas aureofaciens strain 30-84 by RpeA. J Bacteriol 185: 3718-3725.

Whistler, C.A., Corbell, N.A., Sarniguet, A., Ream, W., and Loper, J.E. (1998) The two-component regulators GacS and GacA influence accumulation of the stationary-phase sigma factor sigmaS and the stress response in Pseudomonas fluorescens Pf-5. J Bacteriol 180: 6635-6641.

Whiteley, M., Parsek, M.R., and Greenberg, E.P. (2000) Regulation of quorum sensing by RpoS in Pseudomonas aeruginosa. J Bacteriol 182: 4356-4360.

Whitney, N.J., and Mortimore, C.G. (1961) Effect of 6- methoxybenzoxazolinone on the growth of Xanthomonas stewartii and its presence in sweet corn. Nature 189: 596-597.

Willis, D.K., Rich, J.J., and Hrabak, E.M. (1991) hrp genes of phytopathogenic bacteria. Mol Plant Microbe Interact 4: 132-138.

Wilson, M., McNab, R., and Henderson, B. (2002) Bacterial Disease Mechanisms. Cambridge University Press, Cambridge, UK. 419 Wilson, W.J., Wiedmann, M., Dillard, H.R., and Batt, C.A. (1994) Identification of Erwinia stewartii by a ligase chain reaction assay. Appl Environ Microbiol 60: 278-284.

Woestyn, S., Allaoui, A., Wattiau, P., and Cornelis, G.R. (1994) YscN, the putative energizer of the Yersinia Yop secretion machinery. J Bacteriol 176: 1561-1569.

Wong, S.M., Carroll, P.A., Rahme, L.G., Ausubel, F.M., and Calderwood, S.B. (1998) Modulation of expression of the ToxR regulon in Vibrio cholerae by a member of the two-component family of response regulators. Infect Immun 66: 5854-5861.

Wu, C.-Y., Fu, J.-F., and Liu, S.-T. (2001) The replicon of pSW800 from Pantoea stewartii. Microbiology 147:2757-2767

Wu, H., Tyson, K.L., Cole, J.A., and Busby, S.J. (1998) Regulation of transcription initiation at the Escherichia coli nir operon promoter: a new mechanism to account for co-dependence on two transcription factors. Mol Microbiol 27: 493-505.

Wyman, C., Rombel, I., North, A.K., Bustamante, C., and Kustu, S. (1997) Unusual oligomerization required for activity of NtrC, a bacterial enhancer-binding protein. Science 275: 1658-1661.

Xiao, Y., and Hutcheson, S.W. (1994) A single promoter sequence recognized by a newly identified alternate sigma factor directs expression of pathogenicity and host range determinants in Pseudomonas syringae. J Bacteriol 176: 3089-3091.

Xiao, Y., Heu, S., Yi, J., Lu, Y., and Hutcheson, S.W. (1994) Identification of a putative alternate sigma factor and characterization of a multicomponent regulatory cascade controlling the expression of Pseudomonas syringae pv. syringae Pss61 hrp and hrmA genes. J Bacteriol 176: 1025-1036.

Xiao, Y., Lu, Y., Heu, S., and Hutcheson, S.W. (1992) Organization and environmental regulation of the Pseudomonas syringae pv. syringae 61 hrp cluster. J Bacteriol 174: 1734-1741.

420 Yalpani, N., and Raskin, I. (1993) Salicylic acid: a systemic signal in induced plant disease resistance. Trends Microbiol 1: 88-92.

Yang, C.H., Gavilanes-Ruiz, M., Okinaka, Y., Vedel, R., Berthuy, I., Boccara, M. et al. (2002) hrp genes of Erwinia chrysanthemi 3937 are important virulence factors. Mol Plant Microbe Interact 15: 472-480.

Yother, J., Chamess, T.W., and Goguen, J.D. (1986) Temperature- controlled plasmid regulon associated with the low calcium response in Yersinia pestis. J Bacteriol 165: 443-447.

Zapf, J., Sen, U., Madhusudan, Hoch, J.A., and Varughese, K.I. (2000) A transient interaction between two phosphorelay proteins trapped in a crystal lattice reveals the mechanism of molecular recognition and phosphotransfer in signal transduction. Structure Fold Des 8: 851-862.

Zhang, Z., and Pierson, L.S., 3rd (2001) A second quorum-sensing system regulates cell surface properties but not phenazine antibiotic production in Pseudomonas aureofaciens. Appl Environ Microbiol 67: 4305-4315.

Zuber, S., Carruthers, F., Keel, C., Mattart, A., Blumer, C., Pessi, G. et al. (2003) GacS sensor domains pertinent to the regulation of exoproduct formation and to the biocontrol potential of Pseudomonas fluorescens CHA0. Mol Plant Microbe Interact 16: 634-644.

421