216
Identification and Characterization of Common Bacterial Blight Resistance Genes in the Resistant Common Bean (Phaseolus vulgaris) Variety OAC Rex By Denise M. Cooper A Thesis Presented to The University of Guelph In partial fulfilment of requirements for the degree of Master of Science in Plant Agriculture Guelph, Ontario, Canada © Denise Cooper, July, 2015

Identification and Characterization of Common Bacterial

  • Upload
    others

  • View
    22

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Identification and Characterization of Common Bacterial

Identification and Characterization of Common Bacterial Blight

Resistance Genes in the Resistant Common Bean (Phaseolus

vulgaris) Variety OAC Rex

By

Denise M. Cooper

A Thesis

Presented to

The University of Guelph

In partial fulfilment of requirements

for the degree of

Master of Science

in

Plant Agriculture

Guelph, Ontario, Canada

© Denise Cooper, July, 2015

Page 2: Identification and Characterization of Common Bacterial

ABSTRACT

Identification and Characterization of Common Bacterial Blight Resistance

Genes in the Resistant Common Bean (Phaseolus vulgaris) Variety OAC Rex

Denise M. Cooper Advisor: University of Guelph, 2015 Dr. K. Peter Pauls Common bacterial blight (CBB), caused by the bacterium, Xanthomonas

axonopodis pv. phaseoli (Xap), is a major disease of common bean (Phaseolus

vulgaris L.). CBB resistance candidate genes (CGs) were identified in the

genomic library sequence of resistant variety OAC Rex based on pathogen

resistance gene homology. Four CGs (CBB1-4) and other genes in the

surrounding sequence were characterized. The CGs had homology to a putative

leucine-rich-repeat disease resistance gene in Oryza sativa and

retrotransposons. CG expression was not observed in either OAC Rex or OAC

Seaforth (susceptible variety) after Xap or water treatment, indicating they are

unlikely to be involved in CBB resistance. The other genes encoded ascorbate

oxidase, callose synthase and hypothetical proteins. To test Xap susceptibility of

Arabidopsis thaliana, detached leaves were inoculated with bacteria. Chlorosis

and bacterial growth were observed, suggesting it may be a useful system for

testing CGs.

Page 3: Identification and Characterization of Common Bacterial

iii

ACKNOWLEDGEMENTS

Over the long course of this project, I have been blessed with the help of many

individuals. I would like to acknowledge the help of my advisor Dr. Peter Pauls, whose

guidance lead me to get the most of my work and results. Dr. Gregory Perry was a

constant and invaluable source of help in the lab, especially during the bioinformatics

analyses, and my “technical difficulties”. I would also like to thank Drs. Alireza Navabi

and Paul Goodwin for their assistance and feedback on my thesis.

I would like to thank all of the Pauls lab members, who helped me learn new

procedures and perspectives. Our discussions improved my knowledge on many things

inside and out of the lab, and made my time here truly a world class education. I must

mention Jerlene Nessia, Mahbuba Siddiqua, Loo-Sar Chia, Weilong Xie and Jan

Brazolot, who were my troubleshooting pals throughout my lab experiments.

I would like to thank my family and friends for their encouragement and support

throughout my research and personal struggles. My husband Peter was a constant

source of support in many ways during my studies, and I thank you for all your love and

encouragement.

I am grateful to the various forms and sources of financial support provided by

the University of Guelph, making it possible for me to complete this phase of my

education.

Page 4: Identification and Characterization of Common Bacterial

iv

TABLE OF CONTENTS

ACKNOWLEDGEMENTS ……………………………………………………………….. iii

TABLE OF CONTENTS ………………………………………………………………..... iv

LIST OF TABLES ………………………………………………………………………… x

LIST OF FIGURES ……………………………………………………………………….. xii

LIST OF ABBREVIATIONS……………………………………………………………… xiv

1.0 Literature Review ……………………………………….………………………....

1.1 Common Bean ……………………………………………………………....

1.1.1 Nutritional Importance of Dry Bean …………………………....

1.1.2 Centres of Origin …………………………………………………

1.1.2.1 Middle American Landraces ………………………..

1.1.2.2 Andean Cultivated Landraces ………………………

1.1.2.3 Common Bean Diversity and Domestication ……..

1.1.3 Common Bean Genetics ………………………………………..

1.2 Common Bacterial Blight of P. vulgaris …………………………………...

1.2.1 Pathogen Spread, Infection and Symptoms in P. vulgaris …..

1.2.2 CBB Mode of Pathogenicity …………………………………….

1.2.3 Disease Prevention ……………………………………………..

1.3 CBB Resistance in P. vulgaris ……………………………………………..

1.3.1 CBB Resistance Quantitative Trait Loci ……………….………

1.3.1.1 Frequently Used CBB Resistance QTL Markers …

1.3.1.2 Other Significant CBB Resistance Markers ………

1.4 OAC Rex ……………………………………………………………………..

1.4.1 Genetics of CBB Resistance in OAC Rex ………………….....

1.4.2 Major CBB Resistance QTL in OAC Rex ……………………..

1.5 Plant-Pathogen Interactions ………………………………………………..

1.5.1 Mechanisms of Plant Pathogen Resistance …………………..

1.5.1.1 Pathogen Associated Molecular Pattern Triggered

Immunity ……………………………………………...

1

1

1

2

3

4

5

6

6

7

8

9

9

10

11

13

15

15

16

17

17

18

Page 5: Identification and Characterization of Common Bacterial

v

1.5.1.1.1 Recognition of Flagellin: An Example of PTI

1.5.1.2 Effector Triggered Immunity ………………………..

1.5.1.2.1 Downy Mildew Resistance in Arabidopsis:

An Example of ETI …………………………..

1.5.2 Diversity in Plant Pathogen Interactions ………………………

1.5.2.1 Genetic Variability and Pathogen Resistance …….

1.5.2.2 Transposable Elements and Pathogen Resistance

1.5.2.2.1 Classifications of Transposable Elements ..

1.5.2.2.2 Retrotransposons and R Gene Function .....

1.5.2.3 Genetic by Environmental Interaction and

Pathogen Resistance ………………………………..

1.5.2.4 R Gene Expression at Different Developmental

Stages …………………………………………………

1.6 R Genes Classes …………….……………………………………………..

1.6.1 Nucleotide Binding Site-Leucine Rich Repeat Proteins ……..

1.6.1.1 NBS Domain ………………………….………………

1.6.1.2 Role of Amino-Terminal Domain …………………...

1.6.1.3 LRR Domain …………………….……………………

1.6.2 Extracellular LRR Proteins …………………………………..…

1.6.2.1 eLRR Domain ……………………...…………………

1.6.2.2 LRR Receptor-Like Kinases …………………….….

1.6.2.3 Receptor-Like Cytoplasmic Kinases ……………….

1.6.2.4 Receptor-Like Proteins ……………………………...

1.6.2.5 Polygalacturonase Inhibiting Proteins ……………..

1.7 R Gene Specificity Evolution ….……………………………………………

1.8 Disease Resistance Research Genomic Resources…………………….

1.8.1 Common Bean Genomics ………………………………………

1.8.2 OAC Rex Genomic Resources ………………………………...

1.8.3 Other Plant Genomic Resources ………………………………

1.9 Hypothesis and Objectives …………………………………………………

19

20

21

22

23

24

24

25

26

26

27

27

28

28

29

29

30

30

31

31

32

33

34

34

35

36

36

Page 6: Identification and Characterization of Common Bacterial

vi

2.0 Identification and Characterization of Candidate Common Bacterial

Blight Resistance Genes …………………………………………………………

2.1 Abstract ……………………………………………………………………….

2.2 Introduction …………………………………………………………………..

2.2.1 CBB in Common Bean …………………………………………..

2.2.2 OAC Rex Genomic Library ……………………………………..

2.2.3 R Gene Classifications ………………………………………….

2.2.4 R Gene Identification in OAC Rex ……………………………..

2.3 Materials and Methods ……………………………………………………...

2.3.1 R Gene Homology Identification in OAC Rex Library Clones

2.3.2 Confirmation of Contig Sequences …………………………….

2.3.3 Gene Identification and Characterization ……………………..

2.3.4 CG and Surrounding Sequence Genome Location ………….

2.3.4.1 OAC Rex Genome Assembly Location ……………

2.3.4.2 G19833 Genome Assembly Location ……………..

2.3.4.2.1 Repetitive DNA Analysis ……………………

2.3.5 Characterization of Predicted Proteins ………………………..

2.3.5.1 Domain and Secondary Structure Analysis ……….

2.3.5.2 CG-Specific Analysis ………………………………..

2.3.6 Isolation of CGs ...……………….……………………………….

2.3.6.1 Amplification of CGs …………………..…………….

2.3.6.2 Cloning of CGs ..……………………………………..

2.4 Results ……………………………………………………………………….

2.4.1 Gene Identification in Selected OAC Rex Contig Sequences

2.4.1.1 R Gene Homology Identification ……………………

2.4.1.2 Contig Sequence Assembly Confirmation ………...

2.4.1.2.1 Contig1701 Assembly Confirmation ……….

2.4.1.2.2 Contig1455 Assembly Confirmation ……….

2.4.1.3 Contig1701 Gene Annotation ………………………

2.4.1.3.1 Contig1701 CGs ………………..………..…..

2.4.1.4 Contig1455 Gene Annotation ………………………

37

37

38

38

39

39

41

44

44

44

46

46

46

47

47

48

48

49

49

49

50

52

52

52

52

52

54

54

56

56

Page 7: Identification and Characterization of Common Bacterial

vii

2.4.2 OAC Rex Gene Location in OAC Rex Genome Assembly ….

2.4.2.1 G19833 Location for Contig1701 Identified Genes

2.4.2.2 G19833 Location for Contig1455 Identified Genes

2.4.3 Contig1701 Gene and Protein Characterization ……………...

2.4.3.1 Characterization of CGs .……………………………

2.4.3.1.1 CG Similarities …………………..…………...

2.4.3.1.2 CG Domain Analysis ………………………...

2.4.3.1.3 Contig Sequence Retrotransposon Analysis

2.4.3.1.4 CG Secondary Structure Prediction ……….

2.4.3.2 Characterization of Genes with Putative Functions

2.4.3.2.1 Ascorbate Oxidase Gene Homologs ………

2.4.3.2.2 β-Glucan Synthase-Like Homologs ………..

2.4.3.2.3 Hypothetical Gene Annotations …………….

2.4.4 Isolation and Amplification of CGs …………..…………………

2.5 Discussion …………………………………….……………………………..

2.5.1 CG Characteristics ………………………………………………

2.5.1.1 R Gene Characteristics of CGs …………………….

2.5.1.2 Unique/Unusual Characteristics of CGs …………..

2.5.1.3 Retrotransposon Characteristics of CGs ………….

2.5.2 Additional CGs on Contig1701 …………………………………

2.5.2.1 Ascorbate Oxidases and Disease Resistance ……

2.5.3 CG Location in the Bean Genome ……………………………..

2.5.3.1 CBB Resistance Markers and CG G19833

Genome Location ……………………………………

2.6 Conclusion ……………………………………………………………………

59

59

62

67

67

67

67

68

78

78

78

80

90

91

94

94

94

96

97

99

99

100

103

105

3.0 Detached Leaf Xanthomonas axonopodis pv. phaseoli Susceptibility

Assay in Arabidopsis thaliana …………………………………………………..

3.1 Abstract ……………………………………………………………………….

3.2 Introduction …………………………………………………………………..

3.2.1 Common Bean Improvement …………………………………..

3.2.2 CBB in A. thaliana ……………………………………………….

106

106

107

107

107

Page 8: Identification and Characterization of Common Bacterial

viii

3.3 Materials and Methods ……………………………………………………..

3.3.1 Bacterial and Plant Growth …………..…………………………

3.3.2 Plant Wounding and Inoculation ……………………………….

3.3.3 Assessment of Disease Symptoms ……………………………

3.3.3.1 Image Collection and Analysis ……………………..

3.3.3.2 CFU Analysis …………………………………………

3.3.4 Statistical Analyses ………………………………………………

3.4 Results ………………………………………………………………………..

3.4.1 Analysis of Variance ……………………………………………..

3.4.2 Visual Observations and IA ……………………………………..

3.4.2.1 Common Bean Visible Leaf Symptoms ……………

3.4.2.2 Arabidopsis Visible Leaf Symptoms ……………….

3.4.3 Leaf Sample CFUs ………………………………………………

3.4.3.1 Common Bean CFUs …………..……………………

3.4.3.2 Arabidopsis CFUs ……………………………………

3.5 Discussion ……………………………………………………………………

3.5.1 Symptom Development …………………………………………

3.5.1.1 Common Bean Symptom Evaluation by IA ............

3.5.1.2 Arabidopsis Symptom Evaluation by IA …………...

3.5.1.3 Symptom Evaluation by CFUs ..……………………

3.5.2 Suitability of Detached Arabidopsis Leaf CBB Assay ………..

3.5.2.1 Variability in CFU Observations ……………………

3.5.2.2 Detached Leaf vs. Attached Leaf Symptom

Progression …………………………………………..

3.6 Conclusion ……………………………………………………………………

109

109

109

111

111

111

112

113

113

113

113

115

118

118

120

127

127

127

129

130

131

131

133

135

4.0 Candidate Common Bacterial Blight Resistance Gene Expression in a

Resistant (OAC Rex) and Susceptible (Seaforth) Variety of Phaseolus

vulgaris ………………………………………………………………………………

4.1 Abstract ……………………………………………………………………….

4.2 Introduction …………………………………………………………………..

4.2.1 Significance of Common Bean …………………………………

136

136

137

137

Page 9: Identification and Characterization of Common Bacterial

ix

4.2.2 CBB in Common Bean …………………………………………..

4.2.3 CG Testing in Common Bean …………………………………..

4.3 Materials and Methods ……………………………………………………..

4.3.1 Plant and Inoculum Preparation ……………………………….

4.3.1.1 Plant Growth ………………………………………….

4.3.1.2 Inoculum Preparation ………………………………..

4.3.1.3 Plant Inoculation ……………………………………..

4.3.2 Sample Collection and Analysis ………………………………..

4.3.2.1 Sample Collection and Preparation ………………..

4.3.2.2 RNA Isolation …………………………………………

4.3.2.3 cDNA Transcription ………………………………….

4.3.2.4 Primer Design and PCR Amplification …………….

4.4 Results ………………………………………………………………………..

4.4.1 CG Expression …………………………………………………..

4.4.1.1 CBB Susceptible Genotype (Seaforth) …………….

4.4.1.2 CBB Resistant Genotype (OAC Rex) ……………...

4.5 Discussion ……………………………………………………………………

4.5.1 Observed CG Expression ………………………….……………

4.5.2 Expected CG Expression Patterns …………………………….

4.6 Conclusion ……………………………………………………………………

137

138

139

139

139

139

139

140

140

140

142

142

146

146

146

150

153

153

154

156

5.0 Future Directions and Recommendations ..…………………………………..

5.1 OAC Rex Genome Sequence Assembly …………………………………

5.2 Candidate CBB R Gene Analysis ……………………………………..…..

5.2.1 Identification and Characterization of New CGs ……………...

5.2.2 CG Expression Analysis ……………………………….……….

5.3 Model Plant CBB Inoculation Procedure ………………………………….

157

157

157

157

159

159

6.0 Summary ……………………………………………………………………………. 161

7.0 References .…………………………………………………………………………. 164

8.0 Appendices ...………………………………………………………………………. 193

Page 10: Identification and Characterization of Common Bacterial

x

LIST OF TABLES

Table 2.1 DNA sequences and annealing temperatures for primers used to

identify of OAC Rex genomic library clones associated with the PV-ctt001 QTL region of Pv04 ………………………………….…………...

40

Table 2.2 PV-ctt001 associated OAC Rex genomic library clones ……………..

40

Table 2.3 NCBI accession numbers for the NBS-LRR genes used to search for candidate common bacterial blight resistance genes …………….

45

Table 2.4 Locations within the contig1701 sequence assembly and genetic features of contig1701 gene annotations .…………………………......

58

Table 2.5 Locations within the contig1455 sequence assembly and genetic features of contig 1455 gene annotations .………………………….....

60

Table 2.6 Summary of candidate common bacterial blight resistance gene characteristics .………………………………………………………..…..

61

Table 2.7 Predicted locations in G19833 and expression of annotated genes in OAC Rex contig1701 .…………………………………………….…..

64

Table 2.8 Predicted gene locations on OAC Rex contig 1455 in bp with associated G19833 chromosome number, location (in bp), gene or element and possible function ..…………………………………….…..

66

Table 2.9 Similarity between candidate common bacterial blight resistance genes ....……………………………………………………………….…..

69

Table 2.10 Phyre2 secondary structure modelling for candidate gene proteins CBB1, CBB2, CBB3 and CBB4 ……..….………………………….…..

77

Table 2.11 Similarity between contig1701 genes of interest and identified homologues ……………………………………..……….…………..…...

79

Table 2.12 Phyre2 secondary structure modelling for gene 5, Phvul.006G011700, Glyma.20G051900 and 1ASP_A .……………...

87

Table 2.13 Phyre2 secondary structure modelling for gene 7, G2, and Glyma.08G361500 .………………………………………………………

89

Page 11: Identification and Characterization of Common Bacterial

xi

Table 2.14 Primers used to amplify candidate genes CBB2 and CBB3 from

OAC Rex DNA ..………………………………………………………….

93

Table 3.1 LSmeans for common bacterial blight symptoms on OAC Rex, Nautica and Arabidopsis leaves at 144 hours post inoculation (HPI).

117

Table 3.2 LSmeans for common bacterial blight symptoms on OAC Rex, Nautica and Arabidopsis leaves at 192 hours post inoculation (HPI).

117

Table 3.3 LSmeans for colony forming units for X. axonopodis pv. phaseoli (Xap) bacteria at 48 hours post inoculation (HPI) .……………………

122

Table 3.4 LSmeans for colony forming units for X. axonopodis pv. phaseoli (Xap) bacteria at 96 hours post inoculation (HPI) .……………………

122

Table 3.5 LSmeans for colony forming units for X. axonopodis pv. phaseoli (Xap) bacteria at 144 hours post inoculation (HPI) ..………………….

123

Table 3.6 LSmeans for colony forming units for X. axonopodis pv. phaseoli (Xap) bacteria at 192 hours post inoculation (HPI) ..………………….

123

Table 3.7 Inoculated A. thaliana leaf symptom progression .……………………

124

Table 3.8 Inoculated common bacterial blight susceptible P. vulgaris variety leaf symptom progression .………………………………………………

125

Table 3.9 Inoculated common bacterial blight resistant P. vulgaris variety leaf symptom progression .…………………………………………………...

126

Table 4.1 Candidate common bacterial blight resistance gene primer sequences for real time quantitative RT-PCR .………………………..

145

Page 12: Identification and Characterization of Common Bacterial

xii

LIST OF FIGURES

Figure 2.1 Alignment of selected OAC Rex genomic library clones around CBB resistance marker PV-ctt001 …………………………………..

43

Figure 2.2 OAC Rex contig1701 sequence confirmation, with alignments between OAC Rex sequence elements and the contig1701 and contig1455 sequence assemblies …………………………………..

55

Figure 2.3 Locations, coding sequences and potential functions of predicted genes on contig1701 ………………………………………………….

57

Figure 2.4 Location, coding sequence and potential function of predicted genes on contig1455 ………………………………………………….

60

Figure 2.5 Candidate common bacterial blight (CBB) resistance genes …….

61

Figure 2.6 Sequence comparisons between OAC Rex contig1701 and syntenic regions in G19833 …………………………………………..

63

Figure 2.7 Alignments of contig1455 and associated match locations within G19833 …………………………………………………………………

65

Figure 2.8 Alignment of candidate common bacterial blight resistance gene coding sequences ……………………………………………………..

72

Figure 2.9 Alignment of candidate gene protein sequences …………………..

73

Figure 2.10 LRR sequence of candidate genes .………………………………....

74

Figure 2.11 Secondary structure models of predicted protein for candidate CBB resistance genes ………………………………………………....

76

Figure 2.12 Alignment and annotation of ascorbate oxidase coding sequence .

83

Figure 2.13 Alignment and annotation of L-ascorbate oxidase amino acid sequences ……………………………………………………………...

85

Figure 2.14 Secondary structure model of OAC Rex and G19833 L-ascorbate oxidase protreins ………………………………………………………

86

Page 13: Identification and Characterization of Common Bacterial

xiii

Figure 2.15 Secondary structure model for predicted β glucan synthase-like genes …………………………………………………………………...

89

Figure 2.16 Candidate gene PCR amplification from OAC Rex genomic DNA and BiBAC2 library DNA sources ……………………………………

93

Figure 2.17 Common bacterial blight resistance candidate gene clones ……..

94

Figure 2.18 Location of clone sequence match within contig1455 …………….

94

Figure 3.1 Symptom progression in inoculated P. vulgaris leaves …………...

114

Figure 3.2 Symptom progression in inoculated A. thaliana rosette leaves …..

116

Figure 3.3 Pathogen growth in inoculated P. vulgaris leaf samples ………….

119

Figure 3.4 Pathogen growth in inoculated A. thaliana leaf samples ………….

121

Figure 4.1 Locations of RNA expression primers within candidate genes …..

144

Figure 4.2 Initial candidate gene expression in inoculated Phaseolus vulgaris leaves …………………………………………………………………..

147

Figure 4.3 Subsequent candidate gene expression in second experiment replication of Phaseolus vulgaris leaf inoculation ………………….

148

Page 14: Identification and Characterization of Common Bacterial

xiv

LIST OF ABBREVIATIONS

AA – Ascorbic Acid

AO – Ascorbate Oxidase

Avr – Avirulence

BAC – Bacterial Artificial Chromosome

BiBAC – Binary Bacterial Artificial Chromosome

BLAST – Basic Local Alignment Search Tool

CBB – Common Bacterial Blight

CBB1-4 – Common Bacterial Blight Resistance Candidate Genes 1-4

CC – Coiled Coil

CFU – Colony Forming Unit

CG – Candidate Gene

CGM – Candidate Gene Markers

DIRS – Dictyostelium Intermediate Repeat Sequence

EST – Expressed Sequence Transcript

ETI – Effector Triggered Immunity

eLRR – Extracellular Leucine Rich Repeat

HPI – Hours Post Innoculation

HR – Hypersensitive Response

hrp – Hypersensitive Response and Pathogenicity

IA – Image Analysis

Line – Long Interspersed Nuclear Element

LRR – Leucine-Rich Repeat

LTR – Long Terminal Repeat

MAMP – Moleular Associated Molecular Pattern

MAPK – Mitogen-Activated Protein Kinase

MAS – Marker Assisted Selection

NBS – Nucleotide Binding Site

PAMP – Pathogen Associated Molecular Pattern

PCD – Programmed Cell Death

PCR – Polymerase Chain Reaction

PG – Polygalacturonase

PGIP – Polygalacturonase Inhibiting Protein

PLE – Penelope-Like Element

PRR – Pattern Recognition Receptor

PTI – Pathogen Associated Molecular Pattern Triggered Immunity

QTL – Quantitative Trait Locus

Page 15: Identification and Characterization of Common Bacterial

xv

R – Resistance

RAPD – Random Amplified Polymorphic DNA

RLCK – Receptor-Like Cytoplasmic Kinase

RLK – Receptor-Like Kinase

RLP – Receptor-Like Protein

ROS – Reactive Oxygen Species

RT – Reverse Transcriptase

SCAR – Sequence Characterized Amplified Region

SINE – Short Interspersed Nuclear Element

SNP – Single Nucleotide Polymorphism

SSR – Simple Sequence Repeat

T3SS –Type Three Secretion System

TIR – Toll-Interleukin Receptor

Page 16: Identification and Characterization of Common Bacterial

1

Chapter 1: Literature Review

1.1 Common Bean

Common bean (Phaseolus vulgaris L.) is a member of the Fabacaea family, and

is an important grain legume (pulse) crop species, grown in Canada and globally. It is a

highly produced food legume crop in the world, with the majority of production occurring

in developing countries (Gepts et al. 2008). There are many coloured and white

varieties of edible dry bean in a range of different seed sizes. The most important bean

market classes grown in Canada are navy, black, pinto, dark- and light-red kidney, great

northern and cranberry beans (Bewley et al. 2006; Pulse Canada 2007). Dry beans

produced in Canada are mainly exported to the US and Europe (Agriculture and Agri-

Food Canada 2011), accounting for four times the amount consumed domestically

(Goodwin 2005). The main production areas in Canada are Manitoba, Ontario and

Alberta (Goodwin 2005; Pulse Canada 2015).

1.1.1 Nutritional Importance of Dry Bean

Pulses are recommended as a meat alternative in the human diet by Health

Canada (Health Canada, 2014) . Dry beans contain high levels of high-quality proteins

(Yañez et al. 1995), no cholesterol, and are an excellent source of complex

carbohydrates, such as dietary fibre, starch and oligosaccharides, as well as important

vitamins and minerals such as folate (Tosh and Yada, 2010). A large portion of the

protein component of common bean is made up of the phaseolin protein (Ma and Bliss,

1978), which has been extensively studied (Bollini and Chrispeels, 1978; Bollini et al.

Page 17: Identification and Characterization of Common Bacterial

2

1982; Lawrence et al. 1990; Lawrence et al. 1994) and has been shown to be rich in

lysine but poor in methionine (Bressani, 1983; Gepts and Bliss, 1984).

Despite the many beneficial nutritional aspects of incorporating dry beans into

the human diet, there are several factors that contribute to reduced consumption in

developed countries. The first is the presence of anti-nutritional factors such as

protease and amylase inhibitors, lectins, polyphenols and oligosaccharides, whose

effects may be decreased by different cooking methods, but still contribute to low

digestibility (Gupta, 1987). The second is the time required to cook beans, especially

when packaged dry. The long preparation time is a reason they are passed over for

faster prepared foods in developed countries. However, beans are still a major protein

source in some developing countries, especially in South America, and provide up to

40% of the daily protein intake in less developed nations in Africa (Gepts et al. 2008).

1.1.2 Centres of Origin

Common bean originates from Middle- and South-America, with wild common

bean growing from the north of Mexico to the north of Argentina (Gepts and Debouck,

1991). Two major gene pools exist, termed Middle American and Andean, with an

intermediate gene pool from the region between Peru and Ecuador, (Gepts and Bliss,

1986; Gepts, 1988; Debouck et al. 1993). Variation in the sequences of five genes from

Middle American, Andean and Peru-Ecuador wild common bean accessions indicated

that the centre of origin is Mexico, and that subsequent spread and domestication

established the Andean and Peru-Ecuador gene pools (Bitocchi et al. 2012). This was

further supported by Schmutz et al. (2014), using pooled resequencing of 30 individuals

from Middle American and Andean wild genepools, using ~663,000 polymorphic sites at

Page 18: Identification and Characterization of Common Bacterial

3

least 5 kb away from genes and not in repeat sequences. This analysis, which was

facilitated by the genome sequence for common bean variety G19833 (reported in the

same paper), confirmed that the wild Andean gene pool originated from the wild Middle

American population, indicating that it was the centre of origin for common bean. Due

to further domestication and geographic separation, the two groups are distinguished

not only by the location, from which they originate, but also by their agronomic and

molecular traits, which are used to sub-classify them (Singh et al. 1991).

1.1.2.1 Middle American Landraces

Within the Middle American group, which is found in wild populations located

from Mexico to Venezuela, there are three races; Mesoamerica, Durango and Jalisco as

was described by Singh et al. (1991). The race Mesoamerica is from the lowlands and

encompasses the black, navy and small red market classes. Varieties of this race grow

either type II (indeterminate growth with continued vegetative growth after flowering,

vegetative terminal buds and strong upright architecture) or type III (indeterminate

growth with weak branches and twining and pods located primarily at the base of the

plant) growth habits (Singh, 1982). Each pod produces 6 to 8 small, cylindrical, kidney

or oval shaped seeds (<25g/100 seed), which can be all possible seed colours in pods

that are slender and 8-15 cm long. The leaves and internode length can be small,

intermediate or large, with cordate, hastate or ovate terminal leaf shape. Flowers are

striped with broad cordate or lanceolate bracteoles, petals can be white, white with pink

stripes or purple, and inflorescences are multi-noded. The Durango race is from the

semi-arid highlands of Mexico and encompasses pinto, great northern, medium red and

pink market classes. These varieties predominantly have indeterminate, prostrate, type

Page 19: Identification and Characterization of Common Bacterial

4

III growth habits. They have small to medium ovate or cordate leaflets and thin stems

and branches with short internodes. Fruiting mainly occurs from the basal nodes, and

the flowers produce 5 to 8 cm long flattened pods containing 4 to 5 medium sized seeds

(25-40g/100seeds). Seed colour ranges from tan to yellow, cream, gray, black, white,

red or pink and may have stripes or spots. Finally, the Jalisco race is located in the

southern part of the Durango range and encompasses black, small red and yellow

market classes. These varieties have indeterminate type IV growth habit and can reach

total plant height of over 3 m in native environments. The leaves are hastate, ovate or

rhombohedric, stems and branches are weak, and internodes are medium to long.

Fruiting occurs along the entire length of the plant, but is concentrated in the upper part.

The flowers produce 5-8 small to medium sized round or oval seeds in pods that are 8-

15 cm long.

1.1.2.2 Andean Cultivated Landraces

Singh et al. (1991) also described three races that fall within the Andean centre

of origin, and which are found in Peru, Chile, Bolivia and Argentina; these are called

Nueva Granada, Chile and Peru. The Nueva Granada race, which encompasses the

majority of large seeded and horticultural snap bean market classes, is found in a

variety of altitudes and climates as well as continents. These varieties have type I

(determinate growth with reproductive terminal buds and limited or no continued

vegetative growth after flowering) (Singh, 1982), II and III growth habits, with medium to

long internodes and large hastate, ovate or rhombohedric leaves with dense hairs. The

flowers have small to medium bracteoles that are ovate, lanceolate or triangular, and

produce large pods (10-20 cm) with 4-6 seeds. The seeds are medium (25-40g/100

Page 20: Identification and Characterization of Common Bacterial

5

seeds) to large (>40g/100 seeds) and come in a variety of seed colours. Varieties within

the Chile race are found in the southern Andes and have indeterminate type III growth

habits, with small or medium sized hastate, rhombohedric or ovate leaves and small

stem internodes. The flowers are light pink or white with small to medium sized

triangular, spatulate or ovate bracteoles, and produce medium sized pods (5-8 cm) with

3-5 round-to-oval shaped seeds, which are predominantly harvested as green,

immature seeds. The final Andean race, Peru, is grown in the Andean highlands, and

encompasses varieties which have either indeterminate or determinate type IV climbing

growth habits (Debouck et al. 1988). These varieties have large hastate or lanceolate

leaves, and long, weak internodes, which corresponds with the fact that they are

generally grown in association with maize and other crops in their natural habitat (Singh

et al. 1991). Fruiting occurs along the length of the plant or concentrated at the top, and

produces pods that are 10-20 cm long, with large round or oval seeds.

1.1.2.3 Common Bean Diversity and Domestication

As can be seen in the summary above of the more detailed observations

presented by Singh et al. (1991), P. vulgaris is a very diverse species, both between

and within centres of origin and races. This is not only reserved to phenotype, but also

extends to nutritional contents within the seeds and pods (Singh et al. 1991). Much of

the diversity of observed traits can be attributed to the domestication process which

occurred between ~8500-~8200 years ago for the Mesoamerica genepool and ~7000-

~6300 years ago for the Andean genepool (Mamidi et al. 2011). Domestication genes

that have been identified previously are involved with plant inflorescence structure, seed

colour number and size, and harvestability (reviewed by Glémin and Bataillon, 2009).

Page 21: Identification and Characterization of Common Bacterial

6

Genome sequence analysis by Schmutz et al. (2014) identified 1,835 Mesoamerican

and 748 Andean candidate domestication genes, with the number and type being

variable across the subpopulation groups discussed above. Examples of the types of

candidate genes identified include flowering, vernalization, photoperiod regulation,

increased plant size, nitrogen metabolism and increased seed size and weight.

1.1.3 Common Bean Genetics

Common bean is a diploid species with 2n = 2x = 22 chromosomes. The genome is

estimated to be 600 Mb (Gepts et al.2008). P. vulgaris is closely related to Glycine max

(soybean), and both are Phaseoleae legumes. There is evidence that the soybean

genome is the product of a duplication event and further fractionation and reassembly of

the common ancestor for it and dry bean (McClean et al. 2010; Schmutz et al. 2014).

This relationship makes the soybean genome an excellent resource for genetic

comparisons and gene discovery. Even though comparisons between the two genomes

have shown that common bean has evolved more rapidly than G. max (Schmutz et al.

2014), it can still be used as a reference for gene number, function and organization,

and has proven to be useful for ongoing P. vulgaris sequencing efforts as described

below.

1.2 Common Bacterial Blight of P. vulgaris

Common bacterial blight (CBB) is a significant seed borne disease of common

bean, caused by the Gram-negative bacterial pathogen Xanthomonas axonopodis pv.

phaseoli (Xap) and its fuscans variant Xanthomonas fuscans subsp. fuscans (Xff)

(Gabriel et al. 1989; Vauterin et al. 1995; Schaad et al. 2006). Common bean is the

primary host of CBB, although alternative hosts can be found among other members of

Page 22: Identification and Characterization of Common Bacterial

7

the Fabacaea family (Hayward, 1993), and Arabidopsis ecotype Columbia has been

shown to be susceptible to manual inoculation (Perry, 2010). The disease is seen

wherever beans are produced, and can reduce yield from 10-45%, depending on the

environmental conditions and genotype (Saettler 1989; Gillard et al. 2009). The effect of

the disease is most severe on non-resistant varieties grown in warm, humid growing

conditions.

1.2.1 Pathogen Spread, Infection and Symptoms in P. vulgaris

The primary source of Xap inoculum is infected seeds, but the pathogen may

also be introduced into established fields by infected crop debris or secondary hosts

growing alongside fields (Vidaver 1993). Infection by Xap occurs through wounds,

stomata, or from infected cotyledons (Zaumeyer and Thomas, 1957). The initial

bacterial introduction is followed by cell proliferation and accumulation of extracellular

polysaccharide in the intercellular spaces (Rudolph, 1993). Leaf symptoms begin as

water-soaked areas appearing anywhere on the leaf between 4 to10 days after

inoculation, corresponding to the regions of bacterial cell proliferation. The bacteria will

also enter degrading plant cells adjacent to colonies and begin to multiply, resulting in

the expansion of water-soaked lesions on the leaves (Rudolph, 1993). As the spots

grow, they may merge, becoming necrotic and sometimes developing a surrounding

chlorotic zone (Vidaver, 1993).

With the progression of the disease, bacterial cells may enter the vascular tissue

and spread throughout the plant (Rudolph, 1993). Although stem symptoms are less

frequently seen, they have a similar look and progression to leaf symptoms (Vidaver,

1993). Pod symptoms develop differently however, beginning as water-soaked lesions

Page 23: Identification and Characterization of Common Bacterial

8

and becoming slightly sunken and brown or red-brown over time. Seed symptoms are

only visible on light coloured seed coats, and appear as yellow to brown irregular spots.

Seeds may also rot or shrivel, which greatly decreases the economic value of diseased

yield (Vidaver, 1993).

1.2.2 CBB Mode of Pathogenicity

The ability of bacteria to infect a given plant, pathogenicity, involves large

numbers of genes. An excellent example of this are the hypersensitive response and

pathogenicity (hrp) genes, which are a conserved group of bacterial genes involved in

the interaction between gram negative bacteria, such as Xanthomonas, and their hosts

(Kamoun and Kado 1990; Bonas et al. 1990). These genes tend to be associated with

the Type III secretion system (T3SS) in the bacterial cell wall (Van Gijsefem et al. 1995).

However, they have also been implicated in the regulation of expression and secretion

of proteins involved in pathogenicity directly from the bacterial cell to the plant host cell,

as well as bacterial cell motility (Van Gijsefem et al. 1995; Agrios, 2005).

These secreted proteins, called effectors, have a variety of functions related to

pathogenicity. Examples include the formation of motility structures such as the flagella

in Pseudomonas solanacearum (Van Gijsefem et al. 1995); proteases such as prt1 and

prt2, and pectin degrading polygalacturonases (PGs) which are involved in

Xanthomonas campestris pv. campestris (Xcc) infection of crucifers, causing black rot

(Dow et al. 1990; Wang et al. 2008a). The complement of proteins that are employed in

pathogenicity are specific to the bacteria injecting them. In certain cases, host species

have evolved mechanisms to recognize these effectors and use their presence to

initiate resistance reactions (Agrios, 2005), as will be discussed below.

Page 24: Identification and Characterization of Common Bacterial

9

1.2.3 Disease Prevention

One of the most common practises for prevention of Xap infection in affected

areas is the use of disease free seed. This seed is produced in arid areas where a low

frequency of disease is found and where preventative management to limit introduction

of the pathogen is performed. However, in a study performed by Darrasse et al. (2007),

it was found that inoculated common bean plants grown in environments that were

suboptimal for Xap proliferation were still colonized by the bacteria, although symptoms

were not observed. This is a major concern to bean growers, since the use of disease

free seed is so important for disease prevention in a production setting. Given that the

presence of common bacterial blight in a common bean crop can reduce yield

significantly (Saettler, 1989; Tar’an et al. 2001; Gillard et al. 2009), and that chemical

application control measures provide no significant reduction in infection severity and

effect (Weller and Saettler, 1976), the development and use of CBB resistant varieties

of common bean has become a more practical and attractive alternative in the efforts to

reduce yield loss to the disease.

1.3 CBB Resistance in P. vulgaris

P. vulgaris has little natural resistance to CBB, however, there are currently a few

registered varieties of common bean that are resistant to infection by Xap. Resistance

has been introduced to common bean via inter-specific crosses to close relatives of P.

vulgaris such as tepary bean (Phaseolus acutifolius) (Parker, 1985) and scarlet runner

bean (Phaseolus coccineus) (Miklas et al. 1994). P. accutifolius shows differential

levels of resistance to different Xap isolates, which Opio et al. (1996) suggested

indicates a gene-for-gene interaction between host and pathogen, with 5 gene pair

Page 25: Identification and Characterization of Common Bacterial

10

interactions providing the simplest explanation. However, inoculation of the progeny of

four resistant by susceptible crosses with a single isolate (484a) indicated that

resistance was controlled by one or two dominant genes for three resistant parents, and

quantitative resistance for the fourth resistant parent (Urrea et al. 1999).

This wide range in the resistance responses depending on the Xap isolate used

has also been observed in different resistant common bean genotypes (Opio et al.

1996; Mutlu et al. 2008), and multiple genomic regions have been shown to control the

reaction (Shi et al. 2011). Resistance to CBB in common bean is characterized by a

high to moderate slowing of disease symptom progression, and a lack of internal seed

infection (Aggour et al. 1989). This partial resistance suggests that not all of the

multiple CBB resistance-related genes were transferred to P. vulgaris through the

interspecific crosses, or that the transferred genes don’t function at full capacity in the

common bean background.

Current advances in developing CBB resistant common bean varieties include

combining multiple sources of resistance (pyramiding). This approach may lead to

higher levels of resistance and may reduce the chances that the resistance may be

overcome by the pathogen. This may ultimately result in the production of varieties

containing a resistance system which is more robust and would reduce the potential for

development of resistance in the pathogen to plant defences.

1.3.1 CBB Resistance Quantitative Trait Loci

Resistance to common bacterial blight in P. vulgaris is a quantitative trait,

meaning it segregates in a non-Mendelian fashion and involves multiple genes (St.

Clair, 2010). Genes involved in the trait are affected by other genes as well as the

Page 26: Identification and Characterization of Common Bacterial

11

environment. Quantitative trait loci (QTLs) are regions in the genome containing

gene(s) for the trait of interest. To map QTLs, individuals in a population are analyzed

for genetic markers to determine the relationship to a particular phenotype, for example,

resistance to a disease. The phenotypic and genotypic data obtained from the

population as a whole is then used to determine the location of the QTL (St Clair, 2010),

and markers shown to be associated with particular QTL are then used to determine its

presence or absence in a plant for the purpose of trait selection.

Since the introduction of CBB resistance through interspecific crosses with P.

acutifolius, many linkage maps have been cdeveloped to determine the locations of

QTLs for CBB resistance (i.e. Miklas et al. 2000; Tar’an et al. 2001; Blair et al. 2003;

Cordoba et al. 2010; Shi et al. 2011), which identified 22 minor and major effect CBB

resistance QTLs across the 11 linkage groups. Although these maps do not identify

specific genes associated with resistance to CBB, markers linked to QTL can be used

for selection, as they identify the regions in the genome which likely contain genes for

resistance to the disease. Markers also provide the opportunity to identify genes that

may be related to a particular QTL when the linkage map is compared to a physical map

(i.e. Perry et al. 2013). Thus, QTL, linkage maps and markers can, and have been

excellent tools, not only for the selection of traits related to CBB resistance in common

bean, but to study the genetic mechanisms of resistance.

1.3.1.1 Frequently Used CBB Resistance QTL Markers

Markers which have been frequently used for CBB resistance selection, and

account for the most variability in resistance phenotypes across multiple genetic

backgrounds are SU91 and BC420 (Pedraza et al. 1997; Yu et al. 2000b). The

Page 27: Identification and Characterization of Common Bacterial

12

resistance associated with both sequence characterized amplified region (SCAR)

markers originated from the dry bean line XAN159, which resulted from an interspecific

cross with P. acutifolius accession PI31944.

Marker SU91 has been reported to be on Pv08 (Miklas et al. 2006; Shi et al

2011; Perry et al. 2013) and BC420 has been reported to be on Pv06 (Yu et al. 2000b;

Shi et al. 2011). In the study by Vandemark et al. (2008), it was shown that both of

these dominant markers are part of a recessive epistatic interaction. In this interaction,

the recessive su91//su91genotype supresses the expression of heterozygous or

dominant homozygous BC420. The most resistant phenotypes were seen when at least

one dominant allele was present for both BC420 and SU91. However, since both

markers are dominant, it is difficult to determine when a given genotype is homozygous

for the dominant allele of either or both markers. This makes selection for genotypes

that are homozygous dominant for both BC420 and SU91 difficult.

BC420 and SU91 are the most often used markers for determining the presence

of resistance QTL in common bean accounting for 64% and 17% of phenotypic variation

respectively (Miklas et al. 2006). BC420 is estimated to be 7.1 cM away from the

associated resistance QTL (Yu et al. 2004), and the distance between SU91 and its

associated resistance gene is estimated to be 1-2 cM (Xie et al., personal

communication). These two markers are not perfect indicators of resistance, and may

result in the identification of false positives or negatives, but they are still useful tools

when breeding for CBB resistance.

To improve prediction capabilities for the QTL which SU91and BC420 are

associated with, Shi et al. (2012) developed new markers through the use of a genomic

Page 28: Identification and Characterization of Common Bacterial

13

library available for resistant common bean variety HR45 (Liu et al. 2010). Candidate

gene markers (CGMs) were developed for genes present in bacterial artificial

chromosome (BAC) library clones associated with both SU91 and BC420 (Shi et al.

2012).

Although common bean is recalcitrant to transformation, Shi et al. (2012) were

able to determine that several of the CGMs accounted for more of the variability in

resistance phenotypes than the original markers did. This was shown by testing for the

presence of the CGM expression in resistant (Heterozygous dominant BC420 and

SU91) and susceptible (heterozygous recessive BC420 and SU91) near isogenic lines

(NILs) under disease pressure. Since the CGMs were developed by identifying genes

shown to be involved in disease resistance within the SU91 and BC420 QTL regions, it

was expected that they would provide a more specific method of selection for the trait

during breeding. This assumption was validated for SU91 in a recombinant inbred line

(Shi et al. 2012), which showed that CGMs accounted for greater phenotypic variation

both in the growth room (R2 value up to 0.36) and in the field (R2 value up to 0.42). One

of the CGMs was also found to be co-dominant, making identification of the

homozygous or heterozygous states easier.

1.3.1.2 Other Significant CBB Resistance Markers

Two other significant markers used for identification of CBB resistance QTL

include simple sequence repeat (SSR) marker PV-ctt001 and SCAR marker SAP6. The

PV-ctt001 marker was identified in the resistant common bean variety OAC Rex,

accounts for 42.2% of variation (Tar’an et al. 2001), and is located at the end of the

Pv04 (Yu et al. 2000a; Perry, 2010). However, subsequent studies using OAC Rex as a

Page 29: Identification and Characterization of Common Bacterial

14

resistant parent have not shown this marker to be involved in CBB resistance (Larsen,

2005; Durham et al. 2013). This suggests that PV-ctt001 may have been a false positive

detection in the original study, and that other abiotic (ie temperature and humidity) and

biotic (ie. common bean disease Colletitrichum lindemuthianum) factors may have been

responsible for the increased variability accounted for by PV-ctt001 (personal

communication; Alireza Navabi). The effect of the common bean disease anthracnose

(C. lindemuthianum) is a likely possibility, since the same region of the Pv04 has been

the focus of studies on a cluster of multiple resistance (R) genes and major effect QTLs

for resistance to anthracnose (Geffroy et al 1999; Geffroy et al 2000; Ferrier-Cana et al.

2003; Campa et al. 2011). However, multiple resistance type genes are predicted to be

present in a large cluster in the same region of Pv04 (Schmutz et al. 2014), suggesting

other potential resistance interactions

SAP6 is a dominant marker for CBB resistance that was identified in great

northern cultivar Montana No. 5 on Pv10 (Miklas et al. 2000) and accounts for 35% of

resistance variation in an F2 population (Miklas et al. 2003). Although this marker was

shown to be associated with a major effect QTL for CBB resistance, when Vandemark

et al. (2009) attempted to combine SAP6 and SU91 in a single population, it was found

that of the two, SU91 was primarily responsible for resistance. They discussed the

possibility that the linkage between SAP6 and its associated QTL might not be strong,

and that recombination was the cause of the lack of resistance conferred by the

presence of the marker, as was evidenced by the presence of the marker in susceptible

bean cultivar Matterhorn (Miklas et al. 2003). Although markers provide a good

indication of the desired trait, the SAP6 marker highlights the need for identifying the

Page 30: Identification and Characterization of Common Bacterial

15

gene(s) that are responsible for CBB resistance. With gene-based markers there is an

improved ability to select for the desired phenotype.

1.4 OAC Rex

OAC Rex was the first CBB-resistant variety of common bean registered in

Canada, and is the result of a cross between the HR20-728 and MBE 7 bean lines

(Michaels et al. 2006). Resistance to CBB originated from an inter-specific cross with

naturally resistant tepary bean (P1440795) (Parker, 1985). OAC Rex produces white

navy bean seeds and displays resistance to CBB in leaves and pods (Michaels et al.

2006). Although resistance to infection by Xap in P. acutifolius is characterized as a

hypersensitive response (HR), and likely involves gene-for-gene resistance (Opio et al.

1996), OAC Rex lacks HR when infected. Instead, a decrease in disease symptom

severity is observed, allowing for increased yield in resistant varieties under disease

pressure when compared to non-resistant varieties under the same conditions (Tar’an

et al. 2001; Gillard et al. 2009).

1.4.1 Genetics of CBB Resistance in OAC Rex

In the case of OAC-Rex and other varieties that are resistant to Xap but do not

exhibit HR in the presence of the pathogen, it is likely that some, but not all of the genes

for resistance were transferred from the resistant donor. Assuming this is the case, key

signalling or recognition genes may be missing, which slows the resistance reaction.

This is supported by the observation that disease progression and symptom

development is often delayed and less severe in these resistant varieties as opposed to

the small contained necrotic region associated with HR. Although complete resistance

isn’t present, the slowed disease progression is sufficient to protect yield compared to

Page 31: Identification and Characterization of Common Bacterial

16

susceptible varieties under disease pressure (Tar’an et al. 2001, Gillard et al. 2009). It

must also be noted that a strategy to use multiple partial resistance genes may be a

more robust system of resistance, since it may be more difficult for the pathogen to

overcome that kind of resistance.

1.4.2 Major CBB Resistance QTL in OAC Rex

QTL analysis of CBB resistance in OAC Rex performed by Tar’an et al. (2001)

indicated the involvement of at least one major R gene and two minor QTLs. The major

R gene, associated with the PV-ctt001 marker, was originally placed on the Pv05

linkage group (Tar’an et al. 2001) but the original SSR mapping study by Yu et al.

(2000a) had reported the position of Pvctt001 on Pv04. Also, subsequent studies

localized the marker on Pv04 (Perry, 2010; Perry et al. 2013). The location of this

marker at the end of the linkage group in both cases is most likely the reason for

differing reports. Since the most recent report (Perry, 2010) suggested the marker was

present on the Pv04 linkage group, the region associated with the PV-ctt001 marker on

this linkage group has been used for a preliminary candidate gene search as described

below.

Another disease resistance marker, SU91, was not initially identified in OAC Rex

as being associated with a resistance QTL (Tar’an et al. 2001). However, the SU91

sequence is present in the OAC Rex genome as confirmed by Perry et al. (2013), and

further testing has shown the marker to be associated with a major effect QTL in

progeny of OAC Rex (Durham et al. 2013). Association mapping performed by Shi et

al. (2011), which used 469 bean cultivars and breeding lines and 132 bean single

nucleotide polymorphisms (SNPs) has indicated marker SU91 is located on the Pv08

Page 32: Identification and Characterization of Common Bacterial

17

linkage group, and comparisons between the genome sequence of Andean common

bean line G19833 and OAC Rex by Perry et al. (2013) provide support for this location

in OAC Rex.

1.5 Plant-Pathogen Interactions

The majority of microbes are non-pathogenic towards most plant species, but in

some cases, a parasitic, or disease interaction may evolve. Due to their stationary

nature and lack of a circulatory immune system such as those found in mammals,

plants have developed complex systems for pathogen recognition and signalling

defense responses. This involves a diverse complement of genes related to pathogen

resistance, which encode proteins specific to the role they are needed for. These genes

may be major effect resistance genes, which have a high level of specificity and efficacy

on their own, or they may be minor effect resistance genes, which act in conjunction

with other minor and major effect genes to provide resistance to a particular disease,

race or strain. In either case, the expression of these genes can be induced in

response to an invading pathogen or constitutively expressed in the plant as a whole, or

in specific locations.

1.5.1 Mechanisms of Plant Pathogen Resistance

One of the most important steps in resisting a pathogen for any organism is the

recognition that there is an attacking entity present. The absence or retardation of

pathogen recognition can mean the difference between disease resistance and

susceptibility, as it can mean responses may not be started or may be too slow to be

effective. Plants are no exception, and just as there are a variety of pathogen types

with diverse strategies for plant infection, there are a variety of mechanisms for

Page 33: Identification and Characterization of Common Bacterial

18

recognizing them. These mechanisms fall into two broad categories called pathogen

associated molecular pattern (PAMP) or microbe associated molecular pattern (MAMP)

triggered immunity (PTI), or effector triggered immunity (ETI), and which may overlap,

or complement each other (reviewed by Jones and Dangl 2006; Block et al. 2008;

Padmanabhan et al. 2009; Rafiqi et al. 2009).

1.5.1.1 Pathogen Associated Molecular Pattern Triggered Immunity

The initial, more general form of recognition and defense is called PTI (Jones

and Dangl 2006). This line of defense involves the recognition of molecules which are

conserved among many classes of microbes (PAMPs/MAMPs), but are not necessarily

involved in pathogenicity. Examples of PAMPs/MAMPs include the flagellin in bacterial

flagella, or chitin which is a cell wall component specific to fungal cells (Zipfel and Felix,

2005). These molecules are recognized via pattern recognition receptors (PRRs),

which are generally membrane bound proteins with recognition domains that are

located extra-cellularly, and which may have an intracellular kinase domain for further

signalling (Zipfel, 2008). Their role is to monitor for the presence of microbes, by

recognizing the conserved PAMPs/MAMPs, whether or not it is generally pathogenic

toward that species.

Once a pathogen is recognized, different intracellular responses ensue, such as

signalling via MAP kinase phosphorylation cascades (Asai et al. 2002). Ultimately, a

defense response occurs via the activation of defense genes. The effect of PTI on the

plant may range from increased cell wall thickness, closing of stomata to prevent

bacterial access, increases in hormone levels, production of reactive oxygen species

production and defense gene expression (Taguchi et al. 2003; Desaki et al. 2006;

Page 34: Identification and Characterization of Common Bacterial

19

Melotto et al. 2006). PTI is an effective defense strategy for the majority of microbes

that plants come into contact with, on a daily basis. However, the initial recognition of a

microbe can be overcome by a virulent species through mutation of the conserved

molecule, which allows it to avoid PTI and become pathogenic (Zipfel, 2008).

Resistance is then dependent on the recognition of other molecules produced by the

pathogen, whether it is another PAMP or a protein directly involved in pathogenicity, as

will be discussed below.

1.5.1.1.1 Recognition of Flagellin: An Example of PTI

A well-studied example of PTI is the recognition and response to flagellin. This

molecule is the building block molecule of the bacterial flagella, the membrane

protrusion responsible for cellular motility, and is an important factor for many bacteria,

whether plant pathogenic or otherwise. Although the internal parts of flagellin can be

very different between bacterial species, there are well conserved regions at the ends

(Murthy et al. 2003), which are the targets of recognition of PRRs (Gómez-Gómez et al.

1999; Asai et al. 2002). The N-terminal domain is the most highly conserved part of

flagellin, and a synthesized peptide (flg22) from this region has been shown to induce

typical defense responses, such as increased ethylene and reactive oxygen species

production and induction of defence related genes (Felix et al. 1999; Gómez-Gómez et

al. 1999; Asai et al. 2002).

Recognition of the flagellin molecule occurs via the PRR flagellin sensing 2

(FLS2), which is a LRR-receptor like kinase (LRR-RLK) that is expressed in flowers,

stems, leaves and roots, even in the absence of flg22 (Gómez-Gómez and Boller,

2000). The FLS2 protein has an LRR domain for PAMP/MAMP recognition, a

Page 35: Identification and Characterization of Common Bacterial

20

membrane spanning domain, and a serine/threonine protein kinase domain (Gómez-

Gómez and Boller, 2002). It acts in the presence of flg22 with another LRR-RLK called

BRI1-associated receptor kinase 1 (BAK1) due to its association with the

brassinosteroid receptor BRI1 (Li et al. 2002; Chinchilla et al. 2007). Defense

responses are induced by the FLS2-BRI1 association by initiating signalling cascades

via mitogen-activated protein kinases (MAPKs), resulting in the activation of defense-

related genes via WRKY transcription factors (Asai et al. 2002). In Arabidopsis,

downstream responses such as oxidative burst, callose deposition, ethylene production

and the expression of defense-related genes FRK1, GST1, PR1 and PR5 were

observed at different time points post-treatment (Asai et al. 2002)

1.5.1.2 Effector Triggered Immunity

Although PTI may be effective in responding to the presence of PAMPs/MAMPs,

it can also be overcome by a successful pathogen, by either altering the conserved

molecule, or by producing molecules that restrict host resistance responses such as

kinase signalling and protein transport (Reviewed by Jones and Dangl, 2006; Block et

al. 2008; Padmanabhan et al. 2009). These proteins that are involved with initiating and

maintaining pathogenesis are called effectors. They are secreted by the pathogen into

the host cell or surrounding extracellular region, and may trigger the form of resistance

called ETI. Recognition of effectors generally occurs within the cell via the protein

products of R genes. R genes are generally major effect genes, and are most

commonly of the NBS-LRR class. ETI can involve direct or indirect recognition of the

molecule. Direct recognition of the pathogen involves the interaction of the LRR domain

of the R protein with the pathogen effector. Indirect recognition, which is the more

Page 36: Identification and Characterization of Common Bacterial

21

common form of pathogen recognition, is mediated by NBS-LRR proteins, and involves

monitoring a host molecule for changes made by a pathogen effector protein (Jones

and Dangl, 2006; Padmanabhan et al. 2009).

As with PTI, recognition of effectors leads to a series of responses resulting in

ETI. This includes an oxidative burst, changes in hormone ratios, and signalling

cascades via kinase activation and transcription factors, and eventually leads to a

resistant response (Ryals et al. 1996; Dorey et al. 1997; Torres et al. 2005).

Alternatively to PTI, ETI may lead to the more severe hypersensitive response (HR), in

which infected and surrounding cells undergo apoptosis to restrict expansion of the

pathogen. This method of resistance is only effective on obligate biotrophs, such as

bacteria, because they require a living host cell to obtain nutrients and survive. By

sacrificing the infected cell, the plant stops the progression of the infection into the

surrounding healthy tissue. Conversely, if the pathogen were necrotrophic, killing cells

to obtain nutrients from them, the HR would not decrease pathogen expansion since it

would still be able to obtain nutrients from the dead cells (reviewed by Glazebrook,

2005).

1.5.1.2.1 Downy Mildew Resistance in Arabidopsis: An Example of ETI

An example of ETI are the Recognition of Peronospora parasitica genes (RPP),

which confer resistance to downy mildew (Hyaloperonospora arabidopsidis) (Hpa) in

Arabidopsis (Parker et al. 1997; Botella et al. 1998; van der Biezen et al. 2002). The

RPP genes are present in two clusters, the RPP5 and RPP1 clusters, which are located

on chromosomes 4 and 3 respectively (Parker et al. 1997; Botella et al. 1998; Noël et al.

1999). The functional members of these large gene clusters encode Toll interleukin

Page 37: Identification and Characterization of Common Bacterial

22

receptor (TIR) nucleotide binding site NBS-LRR class of resistance proteins (discussed

below), and recognize the presence of the Hpa effector proteins Arabidopsis Thaliana

Recognized (ATR) (Parker et al. 1997; Botella et al. 1998; van der Biezen et al. 2002;

Gunn et al. 2002; Krasileva et al. 2011; Goritschnig et al. 2012). There is variability

seen in the LRR domain sequence, however, the NBS region for RPP gene family

members are highly conserved (Noël et al. 1999). The variability which is seen in the

different RPP gene family members corresponds to the range in specificities required to

recognize the different Hpa effectors in a gene-for-gene fashion (Botella et al. 1998;

Rehmany et al. 2005; Fabro et al. 2011; Krasileva et al. 2011). In the case of RPP1,

recognition of ATR1 occurs through the association of the LRR domain of RPP1 with

the effector protein, and the TIR domain facilitates signalling and induction of defense

genes (ie PAL, PBS2/3 and SID1/2), and hormone (ie. salicylic acid) production, which

ultimately results in HR (van der Biezen et al. 2002; Krasileva et al. 2010).

1.5.2. Diversity in Plant-Pathogen Interactions

Throughout the interactions between the different plant species and the

pathogens that affect them, similarities and trends can be seen, and classifications of

the related genes and proteins can be made. However, the mechanisms for individual

plant-pathogen interactions are complex, and may also be affected by several

interchangeable factors. These factors include the genetic background of the plant

and/or the pathogen, the environment in which the interaction is occurring, and the

developmental stages of the plant and pathogen.

Page 38: Identification and Characterization of Common Bacterial

23

1.5.2.1 Genetic Variability and Pathogen Resistance

The genetic background of the plant is an important consideration, especially

when dealing with highly cultivated crops. Some pathogen resistance mechanisms

require the presence of multiple genes for recognition of a pathogen to produce a

complete resistance reaction. Examples include the FLS2/BAK1 complex required for

complete downstream resistance reaction to flagellin in Arabidopsis (Chinchilla et al.

2007), the complex formed between Pi5-1 and Pi5-2 that confer resistance to rice blast

(Magnaporthe oryzae) in rice (Lee et al. 2009), and the complexes formed between Prf,

Pto and Pto-homologues which confer resistance to bacterial speck disease

(Pseudomonas syringae pv. tomato) in tomato (Chang et al. 2002; Mucyn et al. 2006;

Gutierrez et al. 2010). In these cases, the absence of regulatory genes or genes that

would normally work together to form a complex translates into a reduction in, or non-

reaction to the pathogen.

There is also a complex interplay between the genetics of the plant host and the

pathogen. The interaction between resistance genes in the host plant and avirulence

(Avr) genes in the pathogen, combine to produce the resistance reaction on the host

that we see. In general terms, when the pathogen has superior Avr expression, disease

symptoms are seen, and when the plant expresses superior resistance genes, disease

symptoms are either not present, or significantly reduced. Pathogens with different

genetic backgrounds that are of the same species are called races, to make a

distinction between the differences in pathogenic abilities. This interplay between host

and pathogen genetic variability is called race-specific resistance. An excellent

example of race-specific resistance is the interaction between races of anthracnose

Page 39: Identification and Characterization of Common Bacterial

24

(Coletotrichum lindemuthianum) and common bean. There are multiple races of bean

anthracnose, and multiple resistance genes, called Co, have been identified, and they

provide resistance to different anthracnose races when present in different

combinations (reviewed by Kelly and Vallejo, 2004; Ferreira et al. 2013).

1.5.2.2 Transposable Elements and Pathogen Resistance

In most cases, the genetic background of a plant species refers to which genes

and alleles are present, and how their presence and combination affects the reaction to

the presence of a pathogen. However, sometimes other genetic elements, such as

transposable elements, can have an effect on gene function.

1.5.2.2.1 Classifications of Transposable Elements

Transposable elements are repetitive sequences within a genome that are able

to move within the genome. There are two classifications of transposable element,

which is based on the intermediate used for their movement (transposition) within the

genome: Type I retrotransposons and Type II DNA transposons, which use RNA and

DNA as their transposition intermediate, respectively (Finnegan, 1989). These

classifications have been further divided into orders and super-families, to reflect the

differences in insertion mechanisms and enzyme characteristics (Wicker et al. 2007).

Type I retrotransposons are divided into the orders long terminal repeat (LTR),

Dictyostelium intermediate repeat sequence (DIRS), Penelope-like elements (PLE),

long interspersed nuclear element (LINE) and short interspersed nuclear element

(SINE), which are found in plants, mammals, fungi and other eukaryotes. Despite

differences in their size, organization, and the number and type of encoded proteins, the

main distinguishing features of retrotransposons are the presence of a reverse

Page 40: Identification and Characterization of Common Bacterial

25

transcriptase, and the production of a new copy with each replication cycle (for

complete review see Wicker et al. 2007).

Type II DNA transposons are classified into two subclasses based on the number

of DNA strands cut during transposition: subclass I encompasses orders TIR and

crypton, which mainly use transposase to “cut and paste” from and into double stranded

sites; subclass II is comprised of the helitron and maverick orders, which use a range of

proteins to “copy and paste” from and into single stranded sites. Unlike the Type I

retrotransposons, the main unifying feature of Type II DNA transposon subclasses are

the use of the DNA intermediate for transposition (for complete review of mechanisms

and proteins see Wicker et al. 2007).

1.5.2.2.2 Retrotransposons and R Gene Function

In addition to their role as a major source of highly repetitive regions,

retrotransposons also have been shown to affect the function of disease resistance

genes in plants. In many species, retrotransposons are found in areas where gene

clusters are rapidly evolving, such as disease resistance related genes in rice (Miyao et

al. 2003), lettuce (Meyers et al. 1998), barley (Marcel et al. 2007), common bean

(Vallejos et al. 2006; David et al. 2009), poplar (Lescot et al.2004), sugar beet

(Kuykendall et al. 2009) and Arabidopsis (Yi and Richards, 2007). Retrotransposons

have also been shown to cause transcription of neighboring disease resistance genes

(Hernández-Pinzón et al. 2009), cause chimeric proteins with resistance genes

(Kashkush et al. 2003; Wang and Warren, 2010), and re-functionalize inactive

resistance genes (Hayashi and Yoshida, 2009).

Page 41: Identification and Characterization of Common Bacterial

26

1.5.2.3 Genetic by Environmental Interaction and Pathogen Resistance

In addition, the genetic background is affected by the environment, which can

make the evaluation of resistance gene effectiveness difficult. This genetic (G) by

environmental (E) interaction (G x E) is not unique to pathogen resistance, but the effect

is two-fold in that the conditions present during plant growth and development affect

host plant physiology and host gene expression as well as the ability of the pathogen to

infect the host. Plant resistance genes have been shown to be affected by temperature

(Fu et al. 2009; Jorgensen, 2012), soil nutrient composition (Jorgensen, 2012) humidity

(Xiao et al. 2003; Zhou et al. 2004a) and light levels (Xiao et al. 2003), and may be

affected at either high or low extremes. The typical response measured in these

experiments is the lack of HR, which indicates that resistance genes with high resource

and energy requirement are the primary target to conserve energy in increased abiotic

stress. However, there are likely other less visibly affected resistance genes that have

yet to be studied.

1.5.2.4 R Gene Expression at Different Developmental Stages

Some resistance genes are differentially expressed depending on the

developmental state of the plant (Reviewed by Develey-Rivière and Galiana, 2007).

Resistance has been shown to be affected by transitions between developmental

stages such as Zea mays resistance to Puccinia sorghi which is dependent on the

expression of adult characteristics (Abedon and Tracy, 1996). It may also be affected

by the maturity of tissues or organs, such as the increased resistance to Venturia

inaequalis with increased leaf maturity in apple (Li and Xu, 2002), or tissue rank, such

as the increased resistance in rice from juvenile leaf ranks to adult ones (Koch and

Page 42: Identification and Characterization of Common Bacterial

27

Mew, 1991). Resistance has also been shown to increase over several stages once it

begins to have an effect, such as resistance to Xanthomonas oryzae pv. oryzae in rice,

which starts at the transition from the juvenile to adult stage, and increases to maximize

at the full adult state (Mazzola et al. 1994; Century et al. 1999).

1.6 R Gene Classes

Just as resistance genes have been shown to be involved in a variety of

pathogen recognition roles, there are several different classes of proteins which these

genes encode. This includes the nucleotide binding site – leucine-rich repeat (NBS-

LRR), extracellular leucine rich repeat (eLRR) containing proteins such as LRR-receptor

like kinases (RLKs), receptor like proteins (RLPs), polygalacturonase inhibiting proteins

(PGIPs), and receptor like cytoplasmic kinases (RLCKs). Many genes are also involved

in peripheral roles, such as downstream signalling (i.e. MAPKs) or physical prevention

of pathogenesis (i.e. callose production). Although there are multiple classes of genes

and resultant proteins involved in pathogen resistance pathways in plants, the most

commonly found, and most well studied group of resistance genes encode NBS-LRR

proteins (Mindrinos et al. 1994; Ori et al.1997; Wang et al. 1999; Meyers et al. 2003).

1.6.1 Nucleotide Binding Site-Leucine Rich Repeat Proteins

The basic NBS-LRR protein contains two main domains: the nucleotide binding

site domain, and the leucine rich repeat domain. These domains have different motifs,

and play different roles in the recognition and defense pathway. The NBS domain of R

proteins is involved in binding nucleotides for induction of downstream resistance

processes (McHale et al. 2006). The LRR region is involved in recognition of a

pathogen via protein-protein interaction (Jones and Jones, 1997).

Page 43: Identification and Characterization of Common Bacterial

28

1.6.1.1 NBS Domain

The NBS region of R proteins is made up of several conserved motifs, including

the P-loop, kinase-2, resistance nucleotide binding site (RNBS)-B, Gly-Leu-Pro-Leu

(GLPL) and MHDV subdomains (Meyers et al. 2002; Liu et al. 2007). These motifs are

important for anchoring and hydrolyzing ADP molecules, which is the beginning of the

signal transduction cascade for the resistance response (Tameling et al. 2002; McHale

et al. 2006; Mestre and Baulcombe, 2006). The NBS site is kept inactive by other

domains of the protein, or with host proteins, and is activated by the presence of a

specific pathogen effector molecule either by direct interaction or by conformational

changes in the host protein it is monitoring (Reviewed by Muthamilarasan and Prasad,

2013).

1.6.1.2 Role of Amino-Terminal Domain

In addition to the nucleotide binding region, an NBS-LRR protein may have other

domains on the N terminal end. These extra domains are either the coiled coil (CC) or

toll-interleukin receptor (TIR) motifs, and are the basis for two subfamilies of NBS-LRR

proteins having different sequences and downstream pathways they affect. Both the

CC and TIR domains are comprised of ~ 200 amino acids (Meyers et al. 2003; Bernoux

et al. 2011). Crystallography has shown the CC domain structure to be a rod-shaped

dimer of α-helices (Maekawa et al. 2011), and the TIR structure is comprised of a β-

sheet connected by α-helices (Chan et al. 2010; Bernoux et al. 2011). It has been

shown for both N terminal domains that self-association occurs, and is an important

component for initiation of downstream signalling and ultimately the induction of HR

(Moffett et al. 2002; Rairdan et al. 2008; Bernoux et al. 2011; Maekawa et al. 2011).

Page 44: Identification and Characterization of Common Bacterial

29

Bernoux et al. (2011) hypothesized that dimerization of the TIR region after effector

binding, might provide a target for binding proteins that would cause downstream

signalling, however, it is unclear whether this is a unique or more common action .

1.6.1.3 LRR Domain

The LRR region is involved in protein-protein interactions (Kobe and Kajava,

2001). The amino acid sequence is hyper-variable, yet maintains a basic consensus

sequence of LxxLxxLxLxxc/nxx, where L represents the location of leucine or other

allopathic amino acids, and x represents the location of any amino acid (Kobe and

Kajava, 2001; Kajava and Kobe, 2002; McHale et al. 2006). The leucine rich consensus

sequence forms a β-strand followed by a β-turn structure, and makes up the backbone

of the LRR domain (Kobe and Deisenhofer, 1993; Kobe and Kajava, 2001). The other

variable amino acids comprise the exposed region, and changes in these amino acids

may alter the specificity of the protein (Parniske et al. 1997; Botella et al. 1998; Hwang

et al. 2000; Krasileva et al. 2010) as well as the structure (for review see Kobe and

Kajava, 2001).

1.6.2 Extracellular LRR Proteins

Another class of genes that have been shown to be involved in pathogen

resistance is the extracellular leucine rich repeat protein encoding genes. There are

several classifications of these proteins based on their domains, whether they are

membrane spanning, and if they contain a cytoplasmic domain. Their unifying feature is

mainly the presence of an eLRR, as well as their localization to the cell membrane.

Page 45: Identification and Characterization of Common Bacterial

30

1.6.2.1 eLRR Domain

The extracellular leucine rich repeat domain is comprised of 20-29 repeats of the

leucine rich repeat motif found in cytoplasmic genes with a variation in the leucine rich

repeat motif consensus sequence (Zhang et al. 2013). The eLRR consensus sequence

is xxLxxLxxLxxLxLxxNxLt/sGxIP where x may be any amino acid and L represents

either leucine or other allopathic residues (Kobe and Kajava, 2001; Zhang et al. 2013).

Like the LRR motif of cytoplasmic proteins, the eLRR has been shown to be involved in

pathogen ligand binding, such as recognition of flg22 by the eLRR domain of the

receptor like kinase FLS2 (Chinchilla et al. 2006).

1.6.2.2 LRR Receptor-Like Kinases

RLKs are comprised of an extracellular leucine-rich repeat domain, a membrane

spanning domain, an intracellular serine/threonine kinase domain and a signal peptide

(Sun et al. 2004; Zipfel et al. 2006). Two variations of RLK domain content have also

been identified in various plants. The variants RLCK and RLP have also been shown to

be involved in disease resistance reactions (discussed below). While the eLRR domain

of RLKs have been shown to bind pathogen ligands as was discussed above, the

kinase domain is involved in the direct and indirect initiation of signalling for downstream

resistance responses by phosphorylation (Ellis et al. 2000; Lu et al. 2010; Zhang et al.

2013).

RLKs have been shown to be involved in resistance to fungal and bacterial

pathogens both by positive and negative regulation (Reviewed by Yang et al. 2012).

Some examples of isolated and characterized RLK pathogen resistance genes are

Page 46: Identification and Characterization of Common Bacterial

31

Xa26 and OsBISERK1 in rice (Sun et al. 2004; Song et al. 2008), EFR, FLS2 and BAK1

in Arabidopsis (Zipfel et al. 2006; Yang et al. 2012).

1.6.2.3 Receptor-Like Cytoplasmic Kinases

Receptor-like cytoplasmic kinases have similar domain content to RLKs, but are

lacking the eLRR domain. Although there is no LRR ligand-binding domain, RLCKs

have been implicated in signalling roles in disease resistance. For example, resistance

to bacterial pathogens was demonstrated by the BIK1 gene which encodes an RLCK

that mediates signalling post-recognition of flagellin by FLS2/BAK1 complex in

Arabidopsis (Lu et al. 2010). Interestingly, the BIK1 protein was first reported to be

involved in Arabidopsis resistance to fungal pathogens Botrytis cinerea and Alternaria

brassicicola, and decrease resistance to the bacterial pathogen Pseudomonas syringae

pv. tomato (Veronese et al. 2006). This indicates the existence of more diverse roles

for proteins involved in disease resistance signalling than those involved in pathogen

recognition.

1.6.2.4 Receptor-Like Proteins

RLPs have a similar structure to RLKs as they contain an eLRR and

transmembrane domain and lack only the cytoplasmic kinase domain (Wang et al.

2008b). Within these two major domains exist several smaller conserved domains

including a putative signal peptide, a cysteine-rich domain, the eLRR domain which

consists of two LRR sequences interrupted by a non-LRR region, a spacer, an acidic

domain, a single transmembrane domain and a short cytoplasmic region (Reviewed by

Kruijt et al. 2005). RLPs have been implicated in resistance to fungal pathogens such

as the tomato Cf-9 gene that confers resistance to Cladosporum fulvum (Jones et al.

Page 47: Identification and Characterization of Common Bacterial

32

1994) and LepR3 gene involved in resistance to Leptosphaeria maculans in Brassica

napus (Larkan et al. 2013), as well as bacterial pathogens such as AtRLP30, which has

been shown to be involved in non-host resistance in Arabidopsis to Pseudomonas

syringae pv. phaseolicola (Ellendorff et al. 2008; Wang et al. 2008b).

1.6.2.5 Polygalacturonase Inhibiting Proteins

Polygalacturonase inhibiting proteins (PGIPs) are extracellular proteins that are

involved in both the general PTI and specific ETI forms of pathogen resistance in a plant

(reviewed by Dangl and Jones, 2001; Nüremberger et al. 2004; Federici et al. 2006).

These proteins have been well studied for their role in inhibiting fungal pathogen

infection in many dicot and monocot plants (reviewed by De Lorenzo et al. 2001; Ferrari

et al. 2002; Federici et al. 2006), but have also been reported to be involved in

resistance to bacterial pathogens (Huang and Allen, 2000; Hwang et al. 2010). PGIPs

are also implicated in stress-induced and developmental roles, due to their inhibition of

the pectinase polygalacturonase (PG) (for review see Protsenko et al. 2008).

Pectinases are used by pathogenic microbes in the initial stages of infection to

degrade the cell wall either for nutrients, or to gain access to the cytoplasm (Collmer

and Keen, 1986; Alghisi and Favaron, 1995). PG in particular has been shown to be

the first protein secreted by pathogenic microorganisms during the initial stages of

penetration (Jones et al. 1972). PGIPs inhibit PG activity in two ways. The first method

of inhibition is by physically occupying the PG cleavage site on the homogalacturonan

molecules (Protsenko et al. 2008). The second method is by binding the PG enzyme

and regulating its action (Leckie et al. 1999; Manfredini et al. 2005). This leads to the

release of degraded plant cell wall oligogalacturide fragments in the apoplast (Cook et

Page 48: Identification and Characterization of Common Bacterial

33

al. 1999), which elicit further resistance reactions in surrounding unaffected cells (De

Lorenzo and Ferrari, 2002).

There are multiple plants from which PGIPs have been isolated, such as

common bean, soybean, Arabidopsis, tomato and pear (reviewed by De Lorenzo et al.

2001). These proteins have different triggers that induce their expression as well as

different specificities for both the pathogen and the range of PGs they employ

(Desiderio et al. 1997). For example, in P. vulgaris, four PGIPs have been identified:

PvPGIP1-4. PvPGIP1 is involved with resistance to Fusarium moniliforme via binding of

its PG, whereasPvPGIP2 is able to inhibit the of PGs secreted by both F. moniliforme

and Aspergillus niger, but has also been shown to inhibit Botrytis cinerea PG1 (Sicilia et

al. 2005).

1.7 R Gene Specificity Evolution

Resistance gene specificity for a particular pathogen is an important component

for the resistance reaction since loss of specificity in a given gene may cause a resistant

plant to become susceptible to the associated disease. New specificities may be

achieved through different means which have been observed in multiple plants for both

NBS-LRR and kinase genes (Meyers et al. 1998; Meyers et al. 2003; Kim et al. 2009;

Schmutz et al. 2010; Huard-Chauveau et al. 2013). Events that lead to new specificities

include inter-allelic recombination, unequal pairing and mispairing, gene duplication of

small and large DNA fragments and single genes or gene clusters, diversifying

selection, intergenic sequence diversion and rearrangements due to transposable

element activity (reviewed by Michelmore and Meyers, 1998; Meyers et al. 2005; Liu et

al. 2007; Joshi and Nayak, 2013).

Page 49: Identification and Characterization of Common Bacterial

34

These events cause shuffling and duplication of large and small chunks of DNA,

which cause changes in amino acid sequences of the resistance proteins, which are

seen primarily as high sequence variability and changes in the specificity area of the

LRR region (Michelmore and Meyers, 1998). This not only promotes the development

of increased, or new pathogen specificities, but also the occurrence of clusters of R

genes and pseudogenes. The genes in these clusters are typically related, but have

highly divergent sequences indicating duplication events and followed by divergence

caused by chromosomal rearrangements (Hulbert et al. 2001; David et al. 2009).

Selection of these changes can occur either through positive or balancing selection,

depending on the plant and the genes observed (Bakker et al. 2006; Chen et al. 2010;

Gos et al. 2012).

1.8 Disease Resistance Research Genomic Resources

1.8.1 Common Bean Genomics

Recently, the genomic sequence of an Andean line G18933, primarily used as a

parent of the principal CIAT (Centro Internacional de Agricultura Tropical, Cali

Columbia) mapping population was released by a group in the US (available at

www.phytozome.net, Schmutz et al. 2014). The sequencing of the G19833 BAC genetic

library was performed using the Roche 454 platform with additional information from

BAC and fosmid end sequences. Assembly was performed using the Arachne2 system

(Jaffe et al. 2003) in combination with genetic markers and the G. max genomic

sequence (Schmutz et al. 2010) as a reference. The released assembly is

approximately 521.1 Mb, with 27,197 genes which encode 31,638 transcripts, and

Page 50: Identification and Characterization of Common Bacterial

35

transposons accounting for 41% of the genome (www.phytozome.net, Schmutz et al.

2014).

Although the sequencing was performed on an Andean variety of common bean

and some rearrangements can be expected between it and Mesoamerican varieties, it

is still an excellent tool for characterization and comparison of gene organization,

structure and function between the gene pools. It also serves as a useful reference for

other P. vulgaris sequencing efforts. Current projects include the South American-

Spanish “PhaseIbeAm” consortium focused on sequencing the Mesoamerican variety

BAT93, as well as the Canadian “Applied Bean Genomics & Bioproducts” effort,

focused on sequencing CBB resistant varieties HR67 and OAC Rex, which will be

discussed below.

1.8.2 OAC Rex Genomic Resources

A binary bacterial artificial chromosome 2 (BiBAC2) genomic library has been

created for the OAC-Rex genome (Perry, 2010), providing a resource for genetic

analysis of CBB resistance in OAC-Rex. With the use of this library, the OAC Rex has

been sequenced using several sequencing methods. Initial sequencing was performed

on 16 BiBAC2 clones selected by markers surrounding the PV-ctt001 CBB disease

resistance marker using Roche 454 sequencing. The obtained sequence was then

assembled into 1,309 contigs based on homology to the G. max chromosome 19.

Further sequencing of the entire OAC Rex genome was performed using the Ilumina

(Illumina Inc, California) and PacBio (Pacific Biosciences Inc., California) platforms,

which provided a more complete assembly based on homology to the P. vulgaris variety

G19833 genome (Schmutz et al. 2014).

Page 51: Identification and Characterization of Common Bacterial

36

1.8.3 Other Plant Genomic Resources

Multiple other plant genomes have been sequenced, including A. thaliana, G.

max, Oryza sativa, and annotations of genes involved in plant-pathogen interactions are

available through the National Centre for Biotechnology Information (NCBI -

www.http://www.ncbi.nlm.nih.gov). These available gene sequences provide an

excellent resource for homologue comparison to locate genes of interest within the OAC

Rex genome. The collection of gene and protein sequences in the NCBI database, as

well as their related studies, also make it possible to determine putative gene function,

characteristics and protein structure, which can be further tested in the plant.

1.9 Hypothesis and Objectives

It was hypothesized that candidate CBB resistance genes could be identified in

the genomic sequence of OAC Rex genomic library clones selected from the PV-ctt001

region of Pv04 using homology to previously identified disease resistance genes with an

LRR domain. The first of four objectives relating to the exploration of this hypothesis

was to identify and characterize candidate common bacterial blight resistance genes

associated with disease resistance marker Pv-ctt001 on Pv04. The second objective

was to isolate the predicted candidate genes for future disease resistance testing in the

model plant Arabidopsis. The third objective was to develop a detached Arabidopsis leaf

tissue culture assay to be used in the testing of candidate genes in the model plant.

The last objective was to measure candidate gene expression in susceptible and

resistant varieties of common bean, to observe differences in gene expression.

Page 52: Identification and Characterization of Common Bacterial

37

Chapter 2: Identification and Characterization of Candidate Common Bacterial

Blight Resistance Genes

2.1 Abstract

Common bacterial blight (Xanthomonas axonopodis pv. phaseoli) is a serious

bacterial disease of common bean (Phaseolus vulgaris L.) which causes lesions on the

leaves, stems, pods and seeds of the plant. Since the disease reduces the quantity and

quality of seed and fresh pod harvests, efforts to breed for resistance have been

ongoing for several decades. To improve the understanding of the genetic mechanisms

underlying CBB resistant variety OAC Rex, resistance gene homologs were identified

within selected BiBAC2 clones from the OAC Rex genomic library. Characterization of

the gene homologs identified four candidate genes, as well as six genes in the

surrounding sequence that have properties associated with genes involved in disease

resistance. Identified genomic features include retrotransposons, a predicted ascorbate

oxidase, a callose/glucan synthase and a hypothetical gene. Genetic and protein

analyses of these genes are described.

Page 53: Identification and Characterization of Common Bacterial

38

2.2 Introduction

2.2.1 CBB in Common Bean

CBB is a major disease of common bean, caused by the bacterial pathogen Xap

and its fuscans variant Xff. Symptoms produced by the disease include the

development of lesions on the leaves, stems, pods and seeds of the plant (Vidaver

1993). The disease affects seed quality and can reduce yield by up to 45% (Saettler

1989; Gillard et al. 2009). There is little natural resistance found in wild or domesticated

common bean varieties. However, sources of resistance within the same genus have

been used to introduce improved CBB resistance into breeding populations; including

Phaseolus acutifolius (tepary bean) (Parker 1985) and Phaseolus coccineus (scarlet

runner bean) (Miklas et al. 1994).

The first CBB resistant variety of common bean, OAC Rex, was registered in

Canada in 2002. It was the result of a cross between HR20-728 and MBE7 (Michaels

et al. 2006), and resistance originated from an inter-specific cross with P. acutifolius

accession P1440795 (Parker 1985). Two major resistance QTLs were found in OAC

Rex, associated with PV-ctt001 and SU91, as well as two minor affect QTL, associated

with BNG71DraI and BNG21EcoRV (Tar’an et al. 2001; Perry et al. 2013). PV-ctt001 was

first reported in OAC Rex on linkage group Pv05 by Tar’an et al. (2001) and was

mapped 21.6 cM away from a major CBB gene locus, and accounted for 42.2% of the

phenotypic variation. Subsequently, a physical location for PV-ctt001 was identified on

chromosome 4 , through the creation of a genomic library for OAC Rex (Perry 2010;

Perry et al. 2013), which is in agreement with a previous study performed by Yu et al.

(2000a) in a different mapping population. The difference in the location of the marker

Page 54: Identification and Characterization of Common Bacterial

39

between reports was attributed to the location of the marker at the end of the linkage

group, which is difficult to map accurately in the absence of a physical map (Tar’an et al.

2001; Perry 2010).

2.2.2 OAC Rex Genomic Library

Resistance to CBB in P. vulgaris has been a main focus of bean improvement for

decades. However, the genes involved have yet to be discovered. To assist in CBB R

gene identification, the OAC Rex genomic library created by Perry (2010) was used to

select clones associated with the PV-ctt001 region of Pv04. Selection was performed

by Southern hybridization of library membranes, using primers for PV-ctt001, as well as

P. vulgaris variety G19833 library clone end sequences PV_GBa00015K06 5’,

PV_GBa000015K06 3’ and PV_GBa00100K10 3’ (Schuleter et al. 2008) as probes

(Table 2.1). Sixteen positive clones were confirmed to have either 1 or 2 of the probe

sequences by PCR (Table 2.2), and the locations of the probes, as well as clone end

sequences that were used to orient the sequences into three contigs (Figure 2.1).

These 16 clones were sequenced using 454 sequencing, yielding 433,233 reads which

were pooled and assembled into 1,550 contigs. The current work was designed to

identify candidate CBB resistance genes associated with the PV-ctt001 marker in the

contig sequences.

2.2.3 R Gene Classifications

One of the main classifications of pathogen R genes are the NBS-LRR protein

encoding genes (Jones and Dangl 2006; Padmanabhan et al. 2009). Two basic

components of NBS-LRR proteins are NBS domains, which bind and hydrolyze ADP for

downstream signalling (Tameling et al. 2002), and LRR domains, which are involved in

Page 55: Identification and Characterization of Common Bacterial

40

Table 2.2 PV-ctt001 associated OAC Rex genomic library clones. Presence (+) or absence (-) of markers is indicated for each Binary Bacterial Artificial Chromosome 2 (BiBAC2) library clone. (Adapted from Perry 2010)

Clone Marker

PV_GBa00015K06 5’ PV_GBa000015K06 PV_GBa00100K10 3’ PV-ctt001

Rex001A06 + + - +

Rex012B03 - + + -

Rex020F10 - + + +

Rex021F10 + - - +

Rex028A07 - - - +

Rex045C01 - - - +

Rex056D11 - + - -

Rex057D11 - + - +

Rex063G02 - + + -

Rex075F07 - + + -

Rex076C04 - - + -

Rex090F06 - + + -

Rex091D05 - + + -

Rex092G02 + - - -

Rex213E05 + - - +

Rex281F03 + - - -

Table 2.1. DNA sequences and annealing temperatures for primers used to identify of OAC Rex genomic library clones associated with the PV-ctt001 QTL region of Pv04 (Adapted from Perry 2010)

Primer Sequence (5’-3’) Annealing Temperature

PV-ctt001 Forward GAGGGTGTTTCACTATTGTCACTGC 48oC

PV-ctt001 Reverse TTCATGGATGGTGGAGGAACAG 48oC

PV_GBa00015K06 5’ Forward CACCTAGTACTTCTGCATATAAAG 50oC

PV_GBa00015K06 5’ Reverse TCTTGGTAATTGCCTATTTAGG 50oC

PV_GBa00015K06 3’ Forward GGATATGACCCTAATTGCC 50oC

PV_GBa00015K06 3’ Reverse TCCCTTCCCTTCTCTCAG 50oC

PV_GBa00100K10 3’ Forward GCCGGAGTATAATTCAACCAGC 48oC

PV_GBa00100K10 3’ Reverse ATTCAGACTTTCATGATGTGTG 48oC

Page 56: Identification and Characterization of Common Bacterial

41

protein-protein interactions and pathogen specificity (Jones and Jones 1997; Ellis et al.

1999; Dodds et al. 2001). Additional small motifs may also be present at the amino

terminal ends, including the CC or TIR motifs, which help mediate signalling and the

induction of HR (Meyers et al. 2002; Moffett et al. 2002; Rairdan et al. 2008; Bernoux et

al. 2011; Maekawa et al. 2011).

Another type of gene involved in plant pathogen resistance are the eLRR genes.

This includes RLKs, which contain an eLRR domain, a transmembrane domain and a

kinase domain (Sun et al. 2004; Zipfel et al. 2006), RLPs which contain an eLRR

domain, transmembrane domain and small cytoplasmic tail (Wang et al. 2008b), and

PGIP which is located in the apoplast and contains only an eLRR (Di Matteo et al. 2003;

Sakamoto et al. 2003). The secondary structure has been resolved by X-ray

crystallography for PGIP2, which was isolated in P. vulgaris (Di Matteo et al. 2003), and

may be used as a reference to model the structure of LRR domain containing proteins.

RLCKs are similar to RLKs, but do not contain an eLRR and are anchored to the cell

membrane and have a kinase domain (Veronese et al. 2006; Lu et al. 2010).

2.2.4 R Gene Identification in OAC Rex

Both NBS-LRR and the eLRR genes have been studied in many different

pathogen systems. However, they have yet to be linked to resistance to CBB in the

common bean. The aim of this research was to identify and characterize CGs in the

region of the Pv04 linkage group associated with the PV-ctt001 marker in the OAC Rex

genome. It was hypothesized that the gene associated with this CBB resistance QTL

encoded an LRR protein, since it is the most commonly reported pathogen R gene

domain. It was further hypothesized that the conserved sequence of LRR genes could

Page 57: Identification and Characterization of Common Bacterial

42

be used to identify candidate CBB R genes in OAC Rex that could be tested for their

ability to reduce the disease.

Page 58: Identification and Characterization of Common Bacterial

43

Figure 2.1 Alignment of selected OAC Rex genomic library clones around CBB resistance marker PV-ctt001. The ~75 kb region of BiBAC2 library clone 21 (Rex021F10; ~150 kb) that represents the likely location for contig1701 is highlighted in blue to indicate the region of focus for this chapter. (Adapted from Perry 2010; Perry et al. 2013).

Page 59: Identification and Characterization of Common Bacterial

44

2.3 Materials and Methods

2.3.1 R Gene Homology Identification in OAC Rex Library Clones

Contigs assembled from the sequence raw reads obtained from clones 1, 12, 20,

21, 28, 45, 56, 57, 63, 75, 76, 90, 91, 92, 213 and 281 (Figure 2.1) were analyzed for

homologous regions to known LRR genes (Table 2.3) in other plants using BLAST. The

query was run through the CLC Genomics Workbench, and comparisons were based

on a collection of plant LRR genes available at the National Centre for Biotechnology

Information (NCBI) website (www.ncbi.nlm.nih.gov). The collection included sequences

from A. thaliana, G. max, O. sativa, Populus trichocarpa and Z. mays, which were

deposited in the CLC Genomics Workbench for BLAST against the selected BiBAC

clone contig sequences (Table 2.3). When selecting which regions of LRR to explore

further, the size of contig and the size and number of homologous regions were taken

into account. Preference was given to large contigs containing multiple predicted gene

regions with E values (probability of obtaining the hit by chance) less than 1E-20.

2.3.2 Confirmation of Contig Sequences

Two contigs, contig1455 and contig1701, were analyzed for sequence continuity.

Raw reads from PacBio (Pacific Biosciences, CA) and Illumina sequencing systems

were assembled into a consensus sequence using each contig as a reference to

produce a consensus sequence. The consensus sequences were aligned to their

respective contig sequence (from Roche 454 sequencing). Scaffolds, contigs and raw

sequencing reads were also used to confirm the sequence assembly and associated

gene annotations of contig1701. The sequence for contig1701 from 35,528 – 53,077 bp

was determined to be an assembly error (results presented in section 2.4.1.2), and

Page 60: Identification and Characterization of Common Bacterial

45

Table 2.3 NCBI accession numbers for the NBS-LRR genes used to search for candidate common bacterial blight resistance genes.

Species of Origin NCBI Accession Number

Arabidopsis Thaliana NM_001123924.1

NM_001125847.1

NM_001161067.1

NM_102893.2

NM_112307.2

NM_114955.4

NM_118071.5

NM_123196.1

NM_124542.2

NM_128428.4

NM_180323.3

NM_180841.2

Glycine Max EU839671.1

EU888329.1

EU888330.1

EU888332.1

FJ014886.1

Oryza Sativa NM_001055547.2

NM_197391.1

Populus Trichocarpa XM_002301537.1

Zea Mays NM_001112339.1

Page 61: Identification and Characterization of Common Bacterial

46

analyses performed on the 1-35,527 bp section, only, are reported below. However, the

final consensus sequence was produced after the majority of analyses were performed

on the entire contig1701 sequence (1-53,077 bp). This includes the CG CBB3, which

was identified and characterized using the same procedures as the other CGs, and is

included in the results reported below. The sequence for contig1455 was aligned to the

~7 kb sequence surrounding CBB2 (25,469-32,210 bp) to determine the similarity of the

two sequences.

2.3.3 Gene Identification and Characterization

Genes were identified in contig1701 and contig1455 using Fgenesh

(www.softberry.com) to identify transcription start points, exons and poly A tails. The

coding sequences and the corresponding proteins were predicted for each identified

gene by Fgenesh. Confirmation of gene predictions were performed by identifying the

presence of promoter elements (TAT box or CAAT core) in the putative promoter

region, expected intron start (GT) and stop (AG) (Luehrsen et al. 1994), internal content

of plant introns (Lorković et al. 2000), and poly A tail sequences. Confirmed genes

which also had a corresponding resistance gene homology match hit were designated

CGs.

2.3.4 CG and Surrounding Sequence Genome Location

2.3.4.1 OAC Rex Genome Assembly Location

Sequences of the contigs as well as the individual genes were BLASTed against

the most current assembly and raw sequencing data from the OAC Rex genome. The

most current OAC Rex assembly was comprised of 11 pseudochromosomes, 8,532

scaffolds and 30,023 contigs using Illumina, and 1,550 contigs using Roche 454

Page 62: Identification and Characterization of Common Bacterial

47

sequencing data (Perry 2010). The most current raw sequencing data (March 2015)

was produced using the PacBio (Pacific Biosciences, CA) sequencing system, and

consisted of 992,295 total raw reads.

2.3.4.2 G19833 Genome Assembly Location

The coding sequences and full sequences for genes identified in contig1455 and

contig1701, as well as the contig sequences were BLASTed against the sequence

assembly reported by Schmutz et al. (2014) for P. vulgaris variety G19833, to determine

their locations within the common bean genome

(http://phytozome.jgi.doe.gov/pz/portal.html). An initial indication of putative gene type

and function was obtained when a OAC Rex gene matched to a previous gene

annotation or other features (i.e. EST, transcript, protein match, RNA expression). In

the case where multiple good (>E63) matches within the G19833 sequence assembly

were found, the first match, with the highest E-value, was used for further comparisons

and analyses.

2.3.4.2.1 Repetitive DNA Analysis

The presence of repetitive DNA elements (i.e. transposable elements) was

identified within the regions of the G19833 genome that matched to genes in selected

contigs using the Phytozome RepeatMasker feature. The OAC Rex contigs were

analyzed using LTR_Finder (Xu and Wang 2007) and LTRphyler

(http://tools.bat.infspire.org/ltrphyler/) to identify the presence of long terminal repeats

(LTRs) and LTR retrotransposon domains.

Page 63: Identification and Characterization of Common Bacterial

48

2.3.5 Characterization of Predicted Proteins

The protein sequences were predicted for genes which were associated with

annotations of the G19833 sequence assembly. The annotation types which were used

to select OAC Rex genes for further analysis included previously identified genes

(including hypothetical genes/proteins), locations where ESTs or transcripts matched to,

or where RNASeq data indicated expression was occurring. OAC Rex genes that

matched to locations which contained transposable elements were the only annotations

that were not included, except for CGs.

2.3.5.1 Domain and Secondary Structure Analysis

The protein sequences were analyzed for the presence of signal peptide using

SignalP 4.1 (Petersen et al. 2011), and transmembrane domain prediction using

TMHMM Server v. 2.0 (www.cbs.dtu.dk/services/TMHMM). To identify functional

domains, the protein sequences were analyzed using InterPro

(http://www.ebi.ac.uk/interpro/; Mitchell et al. 2014). The secondary structures for the

peptide sequences were modelled using Phyre2 (Kelley and Sternberg 2009). BLAST

searches at UniProt (www.uniprot.org), NCBI (http://blast.ncbi.nlm.nih.gov/Blast.cgi),

InterPro (http://www.ebi.ac.uk/interpro/; Mitchell et al. 2014) and PvGEA

(http://plantgrn.noble.org/PvGEA/blasttranscript.jsp; O’Rourke et al. 2014) were

performed for each of the proteins to confirm potential functions. Alignments of the

coding and peptide sequences for the genes with putative functions with a previously

predicted gene of the same type from G19833 or other species were performed to

identify whether key features were conserved.

Page 64: Identification and Characterization of Common Bacterial

49

2.3.5.2 CG-Specific Analysis

CG protein sequences were manually screened for leucines and other allopathic

amino acids to determine the locations of leucine rich repeat motifs. Leucine rich

repeats were identified using the consensus LRR motif LxxxLxLxLxxL. This motif was

designed based on the general consensus motif reported in literature (Tornero et al.

1996; Kajava and Kobe 2002; Di Matteo et al. 2003), but modified to fit the recurring

consensus in the predicted proteins encoded by the CGs. Manual analysis of the CG

protein sequences was also performed to identify NBS, CC and TIR motifs within the

sequence, using the motifs described by Meyers et al. (2003) for Arabidopsis TIR- and

CC- NBS-LRR genes. Alignment of CG protein sequences was performed using

ClustalW (http://www.genome.jp/tools-bin/clustalw), and the locations of LRR motifs

were indicated to show protein homology as well as the percent similarity between the

CG proteins. CG protein sequences were also aligned to PvPGIP-2 and Swiss-Model

(Arnold et al. 2006) was used to provide an additional secondary structure prediction for

CGs using PvPGIP-2 as a reference template.

2.3.6 Isolation of CGs

2.3.6.1 Amplification of CGs

Primers were designed for the identified CG regions from contig1701 (Table 2.3)

using CLC Genomics. Polymerase Chain Reaction (PCR) was performed using the

OAC Rex BiBAC library clone 21 DNA as the template with the following conditions.

Each PCR reaction consisted of 2.5 µl of 10x PCR buffer, 2.5 µl of 15 mM MgCl2, 1 µl of

10 mM dNTP mix, 1 µl of reverse primer, 1 µl of forward primer, 0.5 µl of JumpStart Taq

DNA Polymerase (Sigma-Alderich, St. Louis, MO, USA), 1 µl of DNA template and 14.5

Page 65: Identification and Characterization of Common Bacterial

50

µl of sterile distilled water for a total of 25 µl. The PCR parameters included an initial

denaturation cycle at 94oC for 1 minute, followed by 30 cycles of denaturation (94oC for

30 seconds), annealing (30 seconds at the temperature specific to the primers – see

Table 2.3), and extension (72oC for 3 minutes), this step was followed by 10 minutes of

final extension at 72oC, and held at 4oC. The PCR products were separated by

electrophoresis through a 1% agarose gel in 1x TBE buffer (108 g Tris, 55 g Boric Acid,

40 mL EDTA in 10 L of distilled water) stained with 2.5% ethidium bromide, and

visualized by exposure to UV light. To isolate DNA fragments of interest from the

agarose gel the region containing the band of interest was cut from the gel and

extracted using PureLink Quick Gel Extraction Kit (Invitrogen, CA).

2.3.6.2 Cloning of CGs

Isolated fragments were inserted into TOPO cloning vectors using a TOPO TA™

Cloning Kit (Invitrogen), and the ligations were then transformed into TOPO E. coli cells

using heat shock, according to manufacturer instructions. After overnight culture at

37oC on solid LB selective medium (1 µl ampicillin/mL LB), white colonies were selected

and cultured in liquid ampicillin selective medium (1µLampicillin/mL LB). PCR was

performed for cultures that grew in the selective liquid media to determine the size of

the ligated fragment, with the following conditions. Each PCR reaction consisted of 2.5

µl of 10x PCR buffer, 2.5 µl of 15 mM MgCl2, 1 µl of 10 mM dNTP mix, 1 µl of reverse

primer, 1 µl of forward primer, 0.5 µl of JumpStart Taq DNA Polymerase (Sigma-

Alderich, St. Louis, MO, USA), 1 µl of each positive culture as the DNA template and

14.5 µl of sterile distilled water for a total of 25 µl. The PCR parameters included an

initial denaturation cycle at 94oC for 1 minute, followed by 30 cycles of denaturation

Page 66: Identification and Characterization of Common Bacterial

51

(94oC for 30 seconds), annealing (30 seconds at the temperature specific to the primers

– see Table 2.3), and extension (72oC for 3 minutes), this step was followed by 10

minutes of final extension at 72oC, and held at 4oC. The PCR products were separated

by electrophoresis through a 1% agarose gel in 1x TBE buffer (108 g Tris, 55 g Boric

Acid, 40 mL EDTA in 10 L of distilled water) stained with 2.5% ethidium bromide, and

visualized by exposure to UV light.

For cultures containing an insert of the appropriate size, the plasmid was isolated

using PureLink Quick Plasmid Mini Prep kit (Invitrogen). For each isolated plasmid, 1

kb of the 5’ and 3’ ends of the ligated fragment was sequenced at the University of

Guelph Advanced Analysis Centre Genomics Facility (Guelph, ON, CA). The 1 kb

fragments were sequenced using M13 forward and reverse primers which correspond to

locations on the 5’ and 3’ ends of the cloning site, to confirm the correct sequence was

cloned. The end sequences were then compared to CG and contig sequences to

determine the locations the fragments that were isolated.

Page 67: Identification and Characterization of Common Bacterial

52

2.4 Results

2.4.1 Gene Identification in Select OAC Rex Contig Sequences

2.4.1.1 R Gene Homology Identification

Several contigs from OAC Rex BiBAC2 library clone 21 were identified as having

significant homology to previously reported NBS-LRRs (Figure 2.1). Contigs1701 and

1455 were selected for analysis. Contig1701 was selected for further analysis because

of its large size (~50 kb) and the presence of 3 NBS-LRR homologous regions in its

sequence. Contig1455 was added after the initial BLAST and selection of CGs,

although it was not originally analyzed. Interest in this contig arose from cloning and

sequencing work in later experiments (described below).

2.4.1.2 Contig Sequence Assembly Confirmation

2.4.1.2.1 Contig1701 Assembly Confirmation

When the contig1701 sequence was BLASTed against the current OAC Rex

assembly and raw sequence data, 2 scaffolds (scaffold1045 and scaffold296), 4 contigs

(contig9955, contig9956, contig4106 and contig 4107) and 3 PacBio raw reads (PBR2,

PBR4 and PBR11) matched the best (between 87-99% similarity) to 1-35,527 bp, with

some overlapping sequence (Figure 2.2). When the contig1701 sequence was

BLASTed to the OAC Rex pseudochromosomes there were no locations with high

similarity observed. Within the contig1701 sequence, the1-19,586bp section was 98%

similar to scaffold1045, and the 19,668-35,527 bp section was 97% similar to

scaffold296 (Figure 2.2). Contigs and raw reads matched within these two sections, but

not between, and therefore did not support the assembly of a continuous sequence for

contig1701 as illustrated in Figure 2.2. A section of scaffold1045 (5,968-26,642 bp)

Page 68: Identification and Characterization of Common Bacterial

53

matched to contig1701 at 1-19,626 bp, and corresponded with contig9955, contig9956,

and pbr2. The 1-5,967bp overhang region of scaffold1045 which was missing on the

contig1701 sequence corresponded to the same overhang region of contig9955. The

section of scaffold296 (113,508-130,206 bp) that matched to contig1701 at 20,885-

35,527 bp, corresponded with contig4107, contig4106, pbr4 and pbr11. The same

region of scaffold296 (113,508-130,206 bp) also had high similarity to the region

between 35,647-49,671 bp in the reverse compliment direction compared to the 20,885-

35,527 bp region. Alignment of the two regions (19,668-35,527 bp and 35,647-49,671

bp) within contig1701, which corresponded to scaffold296 (113,508-130,206 bp),

showed them to be 87% similar, with large stretches of identical sequence (>1kb)

interrupted by smaller stretches of no, or low similarity (<200 bp).

Since a single location could not be identified for the entirety of the contig1701

sequence within the OAC Rex sequence assembly, it was divided into four sections

based on homology to OAC Rex contig and scaffold matches (Figure 2.2). The first

section extended from 1-19.586 bp (contig1701-1), the second from 19,668-35,527 bp

(contig1701-2), the third from 35,647-49,671 bp (contig1701-3) and the fourth and final

section from 50,608-53,077 bp (contig1701-4).

The junctions between sections contig1701-1 and contig1701-2, and contig1701-

3 and contig1701-4 were confirmed on the basis of the similarity between the

contig1701 sequence and the consensus sequence produced when PacBio raw reads

were assembled using the contig1701 sequence as a reference. The contig1701

sequence and the assembly consensus sequence were 93% similar for the junction

between contig1701-1 and contig1701-2 sections, and were 97% similar for the junction

Page 69: Identification and Characterization of Common Bacterial

54

between contig1701-3 and contig1701-4 sections. No consensus assembly was

obtained for the junction between contig1701-2 and contig1701-3 sections. Therefore,

because the sequence from 35,527-53,077 bp was highly similar to the sequence from

1-35,527 bp, and the number of PacBio raw reads that matched the 35,527-53,077 bp

sequence was significantly lower (Figure 2.2), further analyses were performed only on

the genes predicted in the 1-35,527 bp sequence. However, characterization of CBB3

from the discarded region (39,732-42,310 bp), is reported below, and in Chapter 4,

since these analyses were performed on this gene previous to the confirmation of the

continuity of the contig1701 sequence assembly.

2.4.1.2.2 Contig1455 Assembly Confirmation

Comparison of contig1455 to the most recent assembly of the OAC Rex genome

identified high similarities to one scaffold, one contig and one raw sequence read (Table

2.7). The 1-6,365 bp region of contig1455 matched to scaffold296 (117,948-124,325)

which also corresponded to contig4106, and the 47-7,720 bp region matched to pbr11.

Alignment of contig1455 to the section within contig1701 (25,469-32,210 bp), which also

matched to contig4106 and pbr11, indicated 92% similarity, and the predicted genes

were present in similar arrangements in all three locations (Figure 2.2). This confirmed

the continuity of the contig1455 sequence assembly.

2.4.1.3 Contig1701 Gene Annotation

When FGeneSH was used to identify genes on contig1701, 16 genes were

identified within the whole sequence (1-53,077 bp) (Figure 2.2). Within the confirmed

sequence (1-35,527 bp), 10 genes were identified (Figure 2.2-2.3: Table 2.4). Promoter

elements were identified for all ten genes, but poly A tails could only be identified for

Page 70: Identification and Characterization of Common Bacterial

55

Figure 2.2 OAC Rex contig1701 sequence confirmation, with alignments between OAC Rex sequence elements and the contig1701 and contig1455 sequence assemblies. A) Frequency of OAC Rex PacBio raw read match locations within contig1701 when it was used as the reference for a sequence assembly. The number of reads mapping to the reference is indicated by the lefthand scale.B) Alignments between contig1701 and PacBio raw reads, Illumina sequencing assembled contigs and scaffolds, and contig1455. Contig1701 and contig 1455 are highlighted in blue. Sections within contig1701 where sequence integrity was in question are highlighted in green. Sequences with high homology to contig1701 (>85%) are indicated by solid lines. Sequences with low homology to contig1701 (<85%) are indicated by dashed lines. Arrows indicate genes with high homology between sequences. C) Summary of OAC Rex genome assembly elements, their percent similarity and location of similarity to contig1701 and contig1455.

C OAC Rex Sequence Element

Homologous Genes

Contig1701 Match Location

Similarity to Contig1701

Similarity to Contig1455

Contig9955 1(CBB1), 2, 3 1 – 9,980 98% - PBR2 3, 4, 5 7,300 – 16,189 87% - Contig9956 4, 5 9,756 – 19,586 98% - PBR4 6, 7 19,668 – 24,932 91% - Contig4107 7 23,030 – 25,424 99% - PBR11 7, 8(CBB2), 9, 10 24,947 – 34,161 88% 88% Contig4106 7,8(CBB2) 26,697 – 35,131 98% 98% Scaffold1045 1, 2, 3, 4, 5 1 – 19,626 98% - Scaffold296 6, 7, 8(CBB2), 9, 10 20,885 – 35,527 97% 98% Contig1455 8(CBB2), 9 25,469 – 32,210 92% -

A

B

2,100

0

1,000

Page 71: Identification and Characterization of Common Bacterial

56

genes 4-10 (Table 2.4). The predicted exon/intron splice points were confirmed for each

of the ten genes.

2.4.1.3.1 Contig1701 CGs

Three resistance gene homologous regions were identified on contig1701,

based on similarity to a reported resistance gene from O. sativa (NM_197391.1; Table

2.3). The first homologous region was 662 bp long, 57% similar to the matching contig

sequence (2E-20), and extended from the promoter region to the second coding region

of a predicted gene, which was designated candidate CBB resistance gene CBB1 (123-

2,274 bp) (Figure 2.5A; Table 2.4). CBB1 was 2,249 bp in size, with 3 exons consisting

of 804 bp of coding sequence (Figure 2.3; Table 2.4 and 6). The second homologous

region was 266 bp long, 65% similar to the matching contig sequence (3E-23), and

located between the first and second coding regions of a predicted gene, designated

CG CBB2 (27,481-31,229 bp) (Figure 2.5B; Table 2.4). CBB2 was 3,922 bp in size, with

4 exons comprising 1,452 bp of coding sequence (Figure 2.3; Table 2.4 and 6). The

third significant resistance gene homologous region was 488 bp long, 59% similar (2E-

24), and extended from the first coding region into the second coding region of a

predicted gene, designated candidate CBB resistance gene CBB3 (39,732-42,310 bp)

(Figure 2.5C; Table 2.4 and 2.6). CBB3 is 2,806 bp in size with 3 exons comprising

1,020 bp of coding sequence (Figure 2.3; Table 2.5).

2.4.1.4 Contig1455 Gene Annotation

FGeneSH identified three genes on contig1455 (Figure 2.4; Table 2.5).

Promoter elements were identified for all three genes, but poly A tails could only be

identified for genes 1-2, indicating they were likely not whole genes (Table 2.5).

Page 72: Identification and Characterization of Common Bacterial

57

Figure 2.3 Locations, coding sequences and potential functions of predicted genes on contig1701. Candidate genes CBB1, CBB2 and CBB3 are indicated in blue, yellow and red respectively. Genes with solid fills have promoter and poly A elements, and outlined genes are missing both the promoter elements and a poly A. A BLAST analysis against the G19833 genome sequence associated predicted genes with retroelement (red and candidate genes), L-ascorbate oxidase/multi-copper oxidase (pale green), and glucan-like synthase/callose synthase annotations.

Page 73: Identification and Characterization of Common Bacterial

58

Table 2.4 Locations within the contig1701 sequence assembly and genetic features of contig1701 gene annotations Gene Contig1701 Location TATA Box CAAT Core Coding Sequence Poly A

Gene 1 (CBB1) 123 – 2,274 2,294-2,299 2,367-2,371 CDS1 – 2,119-2,158 CDS2 – 1,578-1,888 CDS3 – 924-1,376

-

Gene 2 2,426 – 5,867 5,891-5,898 6,193-6,197 CDS1 – 3,965-4,145 CDS2 – 2,753-3,267

-

Gene 3 8,273 – 9,780 - - CDS1 – 9,633-9,731 CDS2 – 9,058-9,328 CDS3 – 8,700-8,871 CDS4 – 8,293-8,389

-

Gene 4 10,965 – 13,458 10,927-10,933 10,885-10,889 CDS1 – 11,550-11,601 CDS2 – 11,893-12,259 CDS3 – 13,062-13,290

13,459-13,503

Gene 5 14,960 – 20,200 20,228-20,234 - CDS1 – 18,739-18,990 CDS2 – 17,572-17,766 CDS3 – 16,676-16,780 CDS4 – 15,980-16,125 CDS5 – 15,500-15,705 CDS6 – 15,166-15,437

14,972-14,946

Gene 6 20,838 – 21,767 21,797-21,801 22,224-22,228 CDS1 – 21,334-21,504 20,839- 20817 Gene 7 22,461 – 25,632 25,458-25,465 - CDS1 – 25,013-25,027

CDS2 – 24,506-24,645 CDS3 – 23,626-23,712 CDS4 – 23,190-23,237 CDS5 – 22,617-22,833

22,524- 22,402

Gene 8 (CBB2) 27,481 – 31,229 27,447-27,452 27,308-27,312 CDS1 – 28,082-28,303 CDS2 – 28,426-28,945 CDS3 – 29,562-30,058 CDS4 – 30,891-31,103

31,164-31,196

Gene 9 31,668 – 32,062 32,090-32,095 - CDS1 – 31,679-31,966 31,668-31,678 Gene 10 32,417 – 34,101 34,128- 34,133 - CDS1 – 33,738-33,855

CDS2 – 32,506-32,741 32,413- 32,422

Page 74: Identification and Characterization of Common Bacterial

59

The predicted exon/intron splice points were confirmed for each of the three predicted

genes. CBB4 is 4,049 bp in size and includes 6 exons comprising a 1,428 bp coding

sequence (Figure 2.4-2.5D; Table 2.5-2.6).

2.4.2 OAC Rex Gene Locations within OAC Rex Genome Assembly

In attempting to confirm the sequence continuity of the contig1701 and

contig1455 (see section 2.4.1.2) sequences, no location within the most recent OAC

Rex pseudochromosomes could be found. Therefore, no conclusions about the

chromosome locations within OAC Rex could be made.

2.4.2.1 G19833 Location for Contig1701 Identified Genes

When the sequences of the ten genes identified in contig1701 were BLASTed to

the G19833 genome sequence, the genes matched to multiple locations (>100) within

and between chromosomes, with varying similarities and confidence values. The best

result for each CG resulted in genes 1-5 matched to chromosome 6 and genes 6-10

matching to chromosome 8 (Figure 2.5; Table 2.5). However, when BLAST was

performed using the entire contig1701 sequence a location within the G19833 sequence

assembly was not identified (data not shown).

Putative functions for the genes on contig1701 were assigned according to the

annotations present for the G19833 genes they were matched to. Gene 5 matched to

the same location within chromosome 6 as an annotation for a multi-copper oxidase/L-

ascorbate oxidase gene and for which expression was observed in multiple tissue

sources (Figure 2.3 and 2.6; Table 2.7). Gene 6 matched to the same location within

chromosome 8 as a hypothetical protein, for which expression was observed in the

nodules and roots (Figure 2.3 and 6; Table 2.7). Gene 7 matched to the same location

Page 75: Identification and Characterization of Common Bacterial

60

Figure 2.4 Location, coding sequence and potential function of predicted genes on contig1455 showing candidate gene CBB4 (bright green). Genes with solid fill have promoter and poly A elements, and outlined genes are missing both the promoter and poly A. A BLAST analysis against the G19833 genome sequence associated predicted genes with retroelement (red and candidate genes).

Table 2.5 Locations within the contig1455 sequence assembly and genetic features of contig 1455 gene annotations

Gene Contig 1455 Location TATA Box CAAT Core Coding Sequence Poly A Gene 1 282-706 283-290 - CDS1 – 412-549 688-706

Gene 2 (CBB4) 1,604-6,056 6,002-6,008 - CDS1 – 5,064-5402 CDS2 – 4,544-4,759 CDS3 – 4,135-4347 CDS4 – 3,908-4033 CDS5 – 2,997-3110 CDS6 – 2,491-2916

1,604-1,736

Gene 3 6,695-8,815 - 8051-8054 CDS1 – 7,619-7969 -

Page 76: Identification and Characterization of Common Bacterial

61

Figure 2.5 Candidate common bacterial blight (CBB) resistance genes. Gene organization for candidate genes CBB1 (A), CBB2 (B), CBB3 (C) and CBB4 (D). Promoters are indicated with black arrow boxes, coding regions are indicated by colored boxes, pathogen resistance gene homologies are indicated by the outline of a rectangle and the poly A tails are indicated with four black lines. The positions (bp) within contig1701 for CBB1, CBB2, and CBB3 and contig1455 for CBB4 are indicated.

Table 2.6 Summary of candidate common bacterial blight resistance gene characteristics. The sizes of the candidate genes, predicted coding sequences, protein sizes and the number of leucine rich repeats for each of the candidate genes CBB1, CBB2, CBB3 and CBB4 are indicated.

Candidate Gene

Total Gene Size (Base Pairs)

Coding Sequence Size (Base Pairs)

Protein Size (Amino Acids)

Number of LRRs

CBB1 2,249 804 267 8

CBB2 3,922 1,452 483 11

CBB3 2,806 1,020 339 10

CBB4 4,453 1,428 476 11

Page 77: Identification and Characterization of Common Bacterial

62

within chromosome 8 as an annotation for a β-glucan-like synthase/callose synthase for

which expression was not observed in the source tissues used in the RNA-seq study

performed by Schmutz et al. (2014) and reported in Phytozome (Figure 2.3 and 6; Table

2.7). The remaining genes, including CBB1-3, matched to regions within chromosomes

6 and 8 which were annotated as retrotransposons (Table 2.7). Some of the

retrotransposons had associated RNAseq data indicating expression, but no source

tissue locations were given (Table 2.7).

2.4.2.2 G19833 Location for Contig1455 Identified Genes

When the genes identified in contig1455 were BLASTed against the G19833 sequence,

they matched to multiple different locations within and between chromosomes. The

best results matched genes 1 and 2 on chromosome 8 and gene 3 on chromosome 6

(Figure 2.7; Table 2.8). Each of the three predicted genes matched to locations where

retrotransposons had been identified, but no other gene annotations were present.

Contig1455 predicted genes 2 and 3 were associated with locations where RNA seq

data indicated gene expression to be present. Two of the three predicted genes were

located in the same region as contig1701 predicted genes. CBB4 matched to the same

location as CBB2 and CBB3. Contig1455 gene 1 was matched to the same location as

contig1701 gene 10.

Page 78: Identification and Characterization of Common Bacterial

63

Figure 2.6 Sequence comparisons between OAC Rex contig1701 and syntenic regions in G19833. Contig1701 sections coloured by blue and purple blocks correspond to chromosome 6 (5.817-5.839 Mbp) and green and yellow coloured regions corresponded to chromosome 8 (49.680-49.698 Mbp) when compared (with BLAST) to the G19833 sequence on Phytozome. Predicted genes for G19833 sequences are indicated by grey arrows for genes matching to contig1701 genes and arrow outline for genes not matching contig1701 genes. Gene match locations are indicated with lines and white boxes.

Page 79: Identification and Characterization of Common Bacterial

64

Table 2.7 Predicted locations in G19833 and expression of annotated genes in OAC Rex contig1701. The positions in the G19833 assembly (Phytozome; Schmutz et al 2014) are indicated by chromosome number and location (bp) of contig1701 gene synteny, and associated gene(s) or element(s), which indicate a possible function.

Gene Number

G19833 Chromosome Location (bp) Similarity to G19833

Associated Genetic Feature G19833 Expression Details

Chromosome 6 Chromosome 8 Gene 1 (CBB1) 5,817,952 – 5,820,113 98% pvRetroSF205-2, pvRetroSF112-2,

pvRetroSF100-1

No expression

Gene 2 5,820,266 – 5,823,700 98% pvRetroSF100-1, pvRetroSF172-1 No expression

Gene 3 5,826,105 – 5,832,901 98% pvRetroSF318-1 No expression

Gene 4 5,825,866 – 5,833,506 97% pvRetroSF318-1 No expression

Gene 5 5,835,002 – 5,839,661 98% Multi-copper oxidase Expression in flowers, pods, leaves,

stem, nodules, roots

Gene 6 49,682,264 – 49,683,238 97% Unknown protein transcript Expression in nodules, roots

Gene 7 49,683,929 – 49,687,869 92% Glucan synthase No expression

Gene 8 (CBB2) 49,682,319 – 49,693,055 75% pvRetro12, pvRetroSF188-3 Expression

Gene 9 49,693,473 – 49,693,864 88% pvRetroSF188-3 Expression

Gene 10 49,694,477 – 49,699,799 69% pvRetro19 No expression

Gene 13 (CBB3) 49,688,390 – 49,692,057 95% pvRetro12, pvRetroSF188-3 Expression

Page 80: Identification and Characterization of Common Bacterial

65

Figure 2.7 Alignments of contig1455 and associated match locations within G19833. The locations where contig1455

gene annotations match within the G19833 assembly chromosomes are indicated with lines. Candidate gene CBB4 is

indicated by a green arrow, and other gene predictions within contig1455 are indicated with red arrows. The red coloured

block indicates similarity to chromosome 6 and the orange coloured block indicates similarity to chromosome 8 within the

G19833 assembly.

Page 81: Identification and Characterization of Common Bacterial

66

Table 2.8 Predicted gene locations on OAC Rex contig 1455 in bp with associated G19833 chromosome number, location (in bp), gene or element and possible function.

Gene Number

G19833 Chromosome Location

Similarity to G19833

Associated Genetic Feature

G19833 Expression

Details Chromosome 6 Chromosome 8 Gene 1 49,693,478 –

49,693,898

99% pvRetro19 No Expression

Gene 2

(CBB4)

49,688,126 –

49,691,770

98% pvRetroSF188-3

pvRetro12

Expression

Gene 3 2,980,777 –

2,981,831

95% pvRetro3 Expression

Page 82: Identification and Characterization of Common Bacterial

67

2.4.3 Contig1701 Gene and Protein Characterization

2.4.3.1 Characterization of CGs

2.4.3.1.1 CG Similarities

Alignments of the CG sequences showed that CBB3 had the highest similarity to

CBB2 and CBB4 (>90%), CBB2 was less similar to CBB4 (72%), and as a group CBB2,

CBB3, CBB4 (CBB2-4) had low similarity with CBB1 (<15%) (Table 2.9). When the

predicted coding sequences were aligned for CBB1-4 (Figure 2.8), the sequences were

less similar than the full gene sequences (Table 2.9). CBB1 was the least similar to

CBB2-4 (<11%), CBB3 had the highest similarity to CBB2 and CBB4 (>89%), and CBB2

was less similar to CBB4 (65%) (Figure 2.8; Table 2.9). The predicted protein

sequences for the CGs were 267, 483, 339 and 476 amino acids long for CBB1, CBB2,

CBB3 and CBB4, respectively (Table 2.6). Alignments between amino acid sequences

showed there to be low similarity between CBB1 and CBB2-4 (<17%), the highest

similarity was seen between CBB3 and CBB4 (84%) (Figure 2.9;Table 2.9). Large

stretches of conserved sequence could be seen in all three comparisons between

CBB2-CBB4, especially in the carboxyl end, with stretches of non-conserved, variable

amino acids at the amino-terminal end as well as roughly in the middle of the protein

sequence (Figure 2.8-9).

2.4.3.1.2 CG Domain Analysis

Domain analysis of CBB1-4 determined that there were no NBS or related motifs

present in any of the CG predicted proteins. In addition, no TIR or CC motifs could be

identified in their N terminal ends. Further analysis of the N terminal ends indicated that

no signal peptides for extracellular transport exist in the putative proteins, indicating a

Page 83: Identification and Characterization of Common Bacterial

68

low likelihood that the proteins function outside the cell. Also, no transmembrane

domains were present in the putative protein sequences. These results indicate that the

CG proteins likely function within the cell.

Although NBS motifs were absent in all four putative CG proteins, they all

contained high numbers of leucine and other allopathic amino acids, as well as an LDL

motif, with a consensus sequence of LDLLLD. The predicted amino acid sequences

contained 8, 11, 10, and 11 LRRs for CBB1 CBB2, CBB3 and CBB4, respectively

(Figure 2.9-10; Table6). These LRRs conformed to the consensus motif sequence

consisting of LxxxLxLxLxxL, (Figure 2.9) where L represents a leucine or allopathic

amino acid and x represents any amino acid. Some variability in the lengths of LRRs

and the spacings between Ls was observed.

2.4.3.1.3 Contig Sequence Retrotransposon Analysis

Different analyses for gene and protein features provided low confidence

evidence that CGs were retrotransposons. BLAST searches using contig1701 and

contig1455 predicted gene sequences against the G19833 genome sequence indicated

that many of the predicted genes on contig 1701 and all contig1455 genes were

retrotransposons (Figure 2.3-4; Table 2.4-5). However, when CGs were BLASTed

against the NCBI database, only the sequence for CBB1 (gene 1) matched to a

retrotransposon (NCBI accession FJ402925.1). NCBI BLAST performed using CG

protein sequences resulted in multiple matches to unknown/hypothetical proteins.

BLAST searches against the UniPro database indicated that all four CGs had low (30-

47%) similarity to retrotransposon proteins including Gag-pol poly protein, POL3-like

Page 84: Identification and Characterization of Common Bacterial

69

Table 2.9 Similarity between candidate common bacterial blight resistance genes. Percent similarity between candidate resistance genes was determined by aligning full gene sequences, coding sequences and protein sequences in ClustalW

Genes Compared Compared Gene Similarity

Full Gene Coding Sequence Protein Sequence CBB1 CBB2 14% 11% 17%

CBB3 15% 9% 11%

CBB4 14% 11% 13%

CBB2 CBB3 90% 89% 72%

CBB4 72% 65% 52%

CBB3 CBB4 93% 91% 84%

Page 85: Identification and Characterization of Common Bacterial

70

CBB1 ------------------------------------------------------------

CBB2 ------------------------------------------------------------

CBB3 ------------------------------------------------------------

CBB4 ATGGTAGGCAATGGTGCCTTCATACAATGCCAAGGCATATGTCCCTTGGTGGACATTTCT

CBB1 ------------------------------------------------------------

CBB2 ------------------------------------------------------------

CBB3 ------------------------------------------------------------

CBB4 CTTCAAACTTCTACATTCACCATTTCTTTTTACCTATTACCTATTGAAGGAGCAAATGTA

CBB1 ------------------------------------------------------------

CBB2 ------------------------------------------------------------

CBB3 ------------------------------------------------------------

CBB4 GTTTTGGGCATTGAATGTTTGCGCACTCTTGGGCCTACTCAGGCATATTTCTCTATACCT

CBB1 ------------------------------------------------------------

CBB2 ------------------------------------------------------------

CBB3 ------------------------------------------------------------

CBB4 AGCATTGCTTTCACCCACCAGAACCAACAAATCACCCTACAAGCTGCCTTTCCATCTACC

CBB1 ------------------------------------------------------------

CBB2 ------------------------------------------------------------

CBB3 ------------------------------------------------------------

CBB4 GCAACTTCAACCACCTACCACCAATTATGTCAATACCTATCCACCAATTCCATTGCATCC

CBB1 ------------------------------------------------------------

CBB2 -------------------------------------ATGGTAGTTGGCGCTTTTTGTGT

CBB3 -------------------------------------ATGGTACTTGGCACTTTTTGTGT

CBB4 ATTCACCTGCTTTCAATTGACAATAACCCTATTCCACCAGGTACTTGGCACTTTT-GTGT

CBB1 ------------------------------------------------------------

CBB2 TGACTACAGAGCCTTAAATGCAGCCACCATCCGTGATCGCTTCCCCATCCCTACCATCGA

CBB3 TGATTACAGAGCATTAAATGCAGCCACCATTCGTGATCACTTCCCCATTCCTACCATCGA

CBB4 TGATTACAGAGCATTAAATGCAGCCACCATTCGTGATCACTTCCCCATTCCTACCATCGA

CBB1 ------------------------------------------------------------

CBB2 TGAATTGCTCGATGAACTAGGCTCAACTACAGTTTTCACTAAGATTGATTTATACTCAGG

CBB3 TGAATTGCTCGATGAACTAGGCTCAACTACAATTTTCACTAAGATAGATTTACACTCAGG

CBB4 TGAATTGCTCGATGAACTAGGCTCAACTACAATTTTCACTAAGATAGATTTACACTCAGG

CBB1 ------------------------------------------------------------

CBB2 GTATCATCAAATTCGACTTATCCCTCAAGACACTCACAAATCAGCTTTTAGAACCATTGA

CBB3 TCCCAATTTGAAGGAACATATTGTGCACTTACAGATTATCTTAGAAGTTTTACACACTAA

CBB4 GTATGATTATATTCGACTCATCCCTCAGGAAACTCACAAATCAGCATTTAGAC-CATTGA

CBB1 ------------------------------------------------------------

CBB2 TGGACATTACGAATTTTTAGAACATATTGTGCACTTAAAGATTATCTTAGAAGTTTTACA

CBB3 AAAACATTTTGC-TAAGT------------------------------------------

CBB4 TGGACATTATGAATTTTT------------------------------------------

CBB1 -------------ATGGAAACCGCTTTTAAGACCAAATTTG--GTTATATGAGTGACTTC

CBB2 CACTAAAAAATTTTTTGCTAAGTTATCTAAGTGCAGTTTTGCCACTTCTCAAGTAAATTA

CBB3 -----------------------TATCTAAGTGCAGTTTTGCCACTTCTCAAGTTAGTTA

CBB4 -----------------------TATCTAAGTGCAGTTTTGCCACTTCTCAAGTTAGTTA

* **** ** **** * *** * **

Page 86: Identification and Characterization of Common Bacterial

71

CBB1 GTAGGGTT-TGTGGTATGTTCTAAAGGGGTGCATGTAGACAAAGAAAAGGTTGTAACAAT

CBB2 TTTGGGCCATATCATTTCGGCTAAAGGAGTAGCCCCTGATCCAGAAAAAGTTATTGCCAT

CBB3 TTTGGGTAATATCATTTTGGCTAAAGTAGTAGTCCTTGATCCAAAAAAAGTTATTGCCAT

CBB4 TTTGGGTAATATCATTTTGGCTAAAGTAGTAGTCCTTGATCCAAAAAAAGTTATTGCCAT

* *** * * * * ****** ** ** * **** *** * * **

CBB1 TCAACATTGGCCAACACCCACCAATGTTAATGAGGT-ACGAAGTTTTCATGGCCTTGCTA

CBB2 ACACAATTGGCCG-CAGCCACGTTCACTCACGGAATTACGTGGTTTTCTCAGCCTCACAG

CBB3 ACACAATTGGCCA-CATCCACGTTCACTCACGAAATTGCGTGGTTTCCTCAACCTCACAG

CBB4 ACACAATTGGCCA-CATCCACGTTCACTCACGAAATTGCGTGGTTTCCTCAACCTCACAG

** ******* ** **** * * * * ** **** * *** *

CBB1 GT--TTTTATAGGAGGTTTGTGCAAAACTTTAGTACTATTG-------------------

CBB2 GA--TTTTATAGAAAGTTTTTTTATCACTACGCCACTCTCGCCGCTCCTCTCACCGATTT

CBB3 GGATTTTTATAGAAAGTTTGTTTATCACTACACCACTCAAAA------------------

CBB4 GA--TTTTATAGAAAGTTTGTTTATCACTACACCACTCTCG-------------------

* ******** * **** * * *** ***

CBB1 ------------------------------------------------------------

CBB2 GTTGCATAATCAAAAATTTACATGGAGTGTCTCGACTCAGGAGGCCTTTACAGAATTAAA

CBB3 ----------------TTTACATGGAGTGTCTC---------------------------

CBB4 -------------------------------TC---------------------------

CBB1 -------------------------CCACACCTATAAA----------------TGCTAT

CBB2 ATTACAAATATCTCAAGTTCCTACACTACACCTACAGAATTTCTCCTTGCCTTTTGTTGT

CBB3 ------------------------GCTACACCTACAGAATTTCTCCCTACCATTTGTCGT

CBB4 ------------------------ACTCCTCTCACTGATTTTCTCCCTACCATTTGTCGT

* * * * * ** *

CBB1 AGTAA-AGAAGGATGTGGTATTCAAGTGGGGACAAGACCAAATTAAAG------CTTTTG

CBB2 GGAAACAGACGCGTCTGCAGTAGTCGTGGGGTCGTGCTCAGCCAAAAGGGGCACCCGCTA

CBB3 GGAGA-AGACGCATCTGCAGTAGCCGTGGGGTCGTGCTCAGCTAAAAGGGGCACCCGCTA

CBB4 GGAGA-AGACGCATCTGCAGTAGCCGTGGGGTCGTGCTCAGCTAAAAGGGGCACCCGCTA

* * *** * * ** * ****** * * ** **** * *

CBB1 AAACTTTAAAGGAAAAATTCCCATTTTTAGCTCTTCCTAACTTTAAAACATTTG------

CBB2 TCCTTCTTCAACAAGAAAATGTGTCCAAGGCTCCAAGCTTCATCAGTGTATGTCCGAGAG

CBB3 TCCTTCTTCAATAAGAAGATGTGTCCAAGGCTCCAAGCATCATCATT-------------

CBB4 TCCTTCTTCAATAAGAAGATGTGTCCAAGGCTCCAAGCATCATCATTCGAGATCCCCCAT

* * * ** ** * **** * * *

CBB1 --------AAATAGAATGTGATGCTTCCAACCCAAAGCACACAAAGGAACTTTATAATGA

CBB2 ATACGGACAAACAGAAGTCCTTAATCGAAGTTTGGAAGCCTACCTACAATGCTTTGTTGG

CBB3 ------------------------------------------------------------

CBB4 TTGTCAGTACATTCTGGA----AACACCTATTCAAGGCACAAGGCACAACACTAAAATTC

CBB1 TGACCTTGATTTTTCTC-------------TTATCTATCAAGAGTTTTCCAAAGGAGCAC

CBB2 TGATCACCCTCACACTTGGTTCCAATATCTCCATCTAGCGGAACTTTGGCACAATACGAC

CBB3 -------------------------------------GTATAACTTTGGCACAATACGAC

CBB4 ATTTCCACTTTACCATCTACAACAGA----CGGACAAACATAACTTTGGCACAATACGAC

* *** ** * * **

CBB1 A---CAAAGAATTTTTCATACATAATGGTTTTCT-TTTCAAAGGAAAAAGACTTTGTGTT

CBB2 ATATCATTCCGCCATTGGCACCTCACCGTTCCATGCTCTGTACGGAAGACAACCTCCAAC

CBB3 ATATCATTCTGCCATTGGCACC-CACCGTTCCATGCTCTGTACGGAAGACAACCTCCAAC

CBB4 ATATCATTCTGCCATTGGCACC-CACCGTTCCATGCTCTGTACGGAAGACAACCTCCAAC

* ** ** ** * *** * * * * ** * * *

Page 87: Identification and Characterization of Common Bacterial

72

CBB1 CCCCAAGGATCTTTAAGACAATCTCTTGTTAGGGAAGCACATGAGAGTCGGCTCGTGGGT

CBB2 CACCTTAGATCTGTTACACAGTCCACACTCAAGCACCACCATTGCAGAT--TTGTTGCAC

CBB3 CACCTTAGATTTGTTAGACAATCCACACTCAAGCACCACCATTGCAGAT--TTGTTGCAC

CBB4 CACCTTAGATTTGTTAGACAATCCACACTCAAGCACCACCATTGCAGAT--TTGTTGCAC

* ** *** * * * *** ** * * * * *** ** * **

CBB1 CACTTTAGGATAGCTAAAACTTTAGATACATTGCATGAGTATTTCTTTTGGCCACACATG

CBB2 CAACACACGCTGGTTCTCCACGCCCTCAAACAAACCCTTCGCCGAACACGTCAACGAATG

CBB3 CAACACACGCTGGTTCTCCACGCCCTTAAACTAAGCCTTCACCGGACACGTCAACGAATG

CBB4 CAACACACGCTGGTTCTCCACGCCCTTAAACTAAGCCTTCACCGGACACGTCAACGAATG

** * * * * * * * * * ** ***

CBB1 CGCAAATCTGTACATAGCTTCCGTGATAAATG-TATAGCATGTAGGAAGGCTAAATCTAA

CBB2 TGCGACCAGGCCAACCGCCACCGATCCGAACGCTCTTTTAACCCGGATGATTGGGTATGG

CBB3 TGCGACCAGGCCAATCGCCACCAATCCGAACGCTCTTTTAACTCGGATGATTGGGTGTGG

CBB4 TGCGACCAGGCCAATCGCCACCAATCCGAACGCTCTTTTAACTCGGATGATTGGGTGTGG

** * * * ** ** ** * * * * *** * * * *

CBB1 AGTACAACCCCATGGTTTTATATACCCTTCTTCCTATTCCAACAATGC-----ATTGGGT

CBB2 GTACGTCTTCAACCTTACCGCCAACAGTTGGTGGAACGTCGTCCATGTTCGAAGCTCGAC

CBB3 GGACGTCTTCAATCTTACTACCGACAGTTGATGGAGCGTCGTTTATGTTCGAAGCTCGAC

CBB4 GGACGTCTTCAATCTTACTACCGACAGTTGATGGAGCGTCGTTTATGTTCGAAGCTCGAC

* * * ** ** * * *** * *

CBB1 CGATATTTCTATGGATTTTATCTTAGGTCTTCCCA---AGACATCGAAGGGTGTGGATTC

CBB2 CCTCGTTACTCGGGTCCTTACCAAGTGCTCCGCCGCATCGGCGTTGTGGCCTACAAACTC

CBB3 CCTTGTTACTCGGGTCCTTACCAAGTGCTCCGCCGCATCGACGTTGTAGCCTACAAACTT

CBB4 CCTTGTTACTCGGGTCCTTACCAAGTGCTCCGCCGCATCGACGTTGTAGCCTACAAACTT

* ** ** ** *** * * ** * * * * * * * *

CBB1 TATATTTGTTGTAGTGGACCGATTTTCTAA------------------------------

CBB2 CGTTTACCAAGCACCTCACAAGTTCTCATGGCCTCTAGTGTGCGGTGCTTCGATGTTCTT

CBB3 CGTTTATCAAGCACCTCACAAATACACCCAGTATTCCATGTATCCCTCCTTCGTGCTTAC

CBB4 CGTTTATCAAGCACCTCACAAATACACCCAGTATTCCATGTATCCCTCCTT---------

* * * * ** * *

CBB1 ------------------------------------------------------------

CBB2 TCCCCTATTAGTCGAATGAGTGGCGGGCCACCAGGGGTGACGTGCAAAGTCACATCTGAC

CBB3 TAA---------------------------------------------------------

CBB4 ------------------------------------------------------------

CBB1 ------------------------------------------------------------

CBB2 GATTTCAAGTTAGCTTTTAAGTACAAGTTGCAGGAGTCGAGCCAGAGAAAAATTACCTAT

CBB3 ------------------------------------------------------------

CBB4 ------------------------------------------------------------

CBB1 ------------------------------------------------------

CBB2 CTGTACGCTGCGATGCCTGGTCTCAGCTATCATGGACATGCTCGTGCGGTCTAA

CBB3 ------------------------------------------------------

CBB4 ------------------------------------------------------

Figure 2.8 Alignment of candidate common bacterial blight resistance gene coding sequences. Candidate genes CBB1 (blue), CBB2 (yellow), CBB3 (red) and CBB4 (green) were aligned in ClustalW. Coding sequences are indicated with alternating coloured rectangles.

Page 88: Identification and Characterization of Common Bacterial

73

CBB1 ------------------------------------------------------------

CBB2 ------------------------------------------------------------

CBB3 ------------------------------------------------------------

CBB4 MVGNGAFIQCQGICPLVDISLQTSTFTISFYLLPIEGANVVLGIECLRTLGPTQAYFSIP

CBB1 ------------------------------------------------------------

CBB2 ----------------------------------------------------MVVGAFCV

CBB3 ----------------------------------------------------MVLGTFCV

CBB4 SIAFTHQNQQITLQAAFPSTATSTTYHQLCQYLSTNSIASIHLLSIDNNPIPPGTWHFCV

CBB1 --------------------------METAFKTKFGYMS---------------------

CBB2 DYRALNAATIRDRFPIPTIDELLDELGSTTVFTKIDLYSGYHQIRLIPQDTHKSAFRTID

CBB3 DYRALNAATIRDHFPIPTIDELLDELGSTTIFTKIDLHSGPNLKEHIVHLQIILEVL---

CBB4 DYRALNAATIRDHFPIPTIDELLDELGSTTIFTKIDLHSGYDYIRLIPQETHKSAFR---

.*:. **:. *

CBB1 --------------------------------------DFVGFVVCSKGVHVDKEKVVTI

CBB2 GHYEFLEHIVHLKIILEVLHTKKFFAKLSKCSFATSQVNYLGHIISAKGVAPDPEKVIAI

CBB3 -------------------HTKKHFAKLSKCSFATSQVSYLGNIILAKVVVLDPKKVIAI

CBB4 -------------------PLMDIMNFLSKCSFATSQVSYLGNIILAKVVVLDPKKVIAI

.::* :: :* * * :**::*

CBB1 QHWPTPTNVNEVRSFHGLASFYRRFVQNFSTIATPINAIVK-------------------

CBB2 HNWPQPRSLTELRGFLSLTGFYRKFFYHYATLAAPLTDLLHNQKFTWSVSTQEAFTELKL

CBB3 HNWPHPRSLTKLRGFLNLTGIFIESLFITTPLKIYMECLAT-------------------

CBB4 HNWPHPRSLTKLRGFLNLTGFYRKFVYHYT------TLVTP-------------------

::** * .:.::*.* .*:.:: . . : :

CBB1 ------------KDVVFKWGQDQIKAFETLKEKFPFLALPNFKTFEIECDASNP------

CBB2 QISQVPTLHLQNFSLPFVVETDASAVVVGSCSAKRGTRYPSSTRKCVQGSKLHQCMSERY

CBB3 ---------PTEFLPTICRGEDASAVAVGSCSAKRGTRYPSSIRRCVQGSKHHH------

CBB4 ---------LTDFLPTICRGEDASAVAVGSCSAKRGTRYPSSIRRCVQGSKHHHSRSPIC

: * . . *. :: . :

CBB1 ---------------------KHTKELYNDDLDFSLIYQEFSKGAHKEFFIHNGFLFKGK

CBB2 GQTEVLNRSLEAYLQCFVGDHPHTWFQYLHLAELWHNTTYHSAIGTSPFHALYGRQPPTT

CBB3 -------------------------------CITLAQYDISFCHWHPPFHALYGRQPPTT

CBB4 ---QYILETPIQGTRHNTKIHFHFTIYNRRTNITLAQYDISFCHWHPPFHALYGRQPPTT

*. * .

CBB1 RLCVPQGSLRQSLVREAHESRLVGHFRIAKTLDTLHEYFFWPHMRKSVHSFRDKCIACRK

CBB2 LDLLHSPHSSTTIADLLHQHTLVLHALKQTLRRTRQRMCDQANRHRSERSFNPDDWVWVR

CBB3 LDLLDNPHSSTTIADLLHQHTLVLHALKLSLHRTRQRMCDQANRHQSERSFNSDDWVWGR

CBB4 LDLLDNPHSSTTIADLLHQHTLVLHALKLSLHRTRQRMCDQANRHQSERSFNSDDWVWGR

: . ::. *: ** * . * :. .: ::* :**. . . :

CBB1 AKSKVQPHGFIYPSSYSN----------NALGRYFYGFYLRSSQDIEGCGFYICCSGPIF

CBB2 LQPYRQQLVERRPCSKLDPRYSGPYQVLRRIGVVAYKLRLPSTSQVLMASSVRCFDVLSP

CBB3 LQSYYRQLMERRLCSKLDPCYSGPYQVLRRIDVVAYKLRLSSTSQIHPVFHVSLLRAY--

CBB4 LQSYYRQLMERRLCSKLDPCYSGPYQVLRRIDVVAYKLRLSSTSQIHPVFHVSLL-----

:. : .* : . :. * : * *:.::

CBB1 -------------------------------------------------------

CBB2 ISRMSGGPPGVTCKVTSDDFKLAFKYKLQESSQRKITYLYAAMPGLSYHGHARAV

CBB3 -------------------------------------------------------

CBB4 -------------------------------------------------------

Figure 2.9 Alignment of candidate gene protein sequences. LRR regions highlighted for candidate genes CBB1 (blue) CBB2 (yellow), CBB3 (red) and CBB4 (green), LDL motif is indicated by a box.

Page 89: Identification and Characterization of Common Bacterial

74

A) CBB1

B) CBB2

C) CBB3

D) CBB4

[4]…

FK T K F G YM- SDF …[11]

VD K E - K V -V T -I …[8]

VN E - V R - SFHGL …[13]

I A T P I N - A IVK - …[17]

LK E KF P F - LA - L …[21]

YN DDL D FS L I - Y …[43] LV G HF R I A - KTL …[48]

LG R - Y F Y G FY- L …[20]

[15]…

I R D R F P I P TIDEL …[14]

LY S GY H - Q I R LI …[25]

ID G H Y E F - LEHI …[17]

FA T S - Q VNYLGHI …[38]

LR G - F - L S LTGF …[32]

F T E - L K L Q ISQV …[29]

LN R S LEAY - LQCF …[12]

LHQHTL - V - LHAL …[14]

YS G PY Q V - LR RI …[23]

VR C - F D V - LS PI …[28]

VTSDDF K L A F K-Y …[10]

[17]…

I R D H F P IP TI D E L …[20]

L K E H I - V HL - Q I …[10]

FATS QV S Y - L G N I …[27]

LR G F L N LTGI F - I …[4]

I T T P L K I - YMECL …[50]

L A Q - Y D I S F C H - …[32]

L H Q H- T LV L H A L …[29]

VW GRLQSY - Y R Q L

…[12]

YS G P Y Q V - L R R I …[14]

I H P - V - F HV S LL …[3]

[20]…

LQTSTF T I S F YL L …[59] LSTNS IAS IHL LS - I …[23] I RD H F P IPT IDE L …[14] LH S GY D Y - I R- L …[24] FATSQV S Y - L GNI …[27] LR G F L N LTGF - -Y …[9] LV T P L TDF - L P T I …[74] L A Q- Y D I S FCH - …[32] LHQHTL V LHAL K- L …[29] VWGR LQ SY - YRQL …[12] YS G P YQ V - LR R I …[24]

LxxxLxLxLxxL

Figure 2.10 LRR sequences of candidate genes A) CBB1, B) CBB2, C) CBB3 and D) CBB4 with LRR consensus motif indicated

Page 90: Identification and Characterization of Common Bacterial

75

reverse transcriptase, Gypsy/Ty-3 retro-element polyprotein, as well as many

uncharacterized proteins.

Domain analysis with InterPro was unable to identify protein domains within any

of the CG protein sequences. However, the program indicated the presence of protein

signatures for all four CGs other than CBB2, providing a suggested protein

type/function. For CBB1, reverse transcriptase, GAG/POL/ENV polyprotein and

DNA/RNA polymerase signatures were found. For CBB2, no protein signatures were

found. For CBB3, reverse transcriptase and DNA/RNA polymerase signatures were

found. Finally, for CBB4, reverse transcriptase and DNA/polymerase signatures were

identified.

LTR_finder was unable to identify long terminal repeats within either of the contig

sequences. However, LTRphyler identified multiple retrotransposon poly protein

domains associated with CGs. Within CBB1, both an integrase core domain (660-1028

bp) and a Gypsy reverse transcriptase domain (1,908-2,374 bp) were identified which

corresponded to the same region as CDS1 and CDS3, although they also associated

with intron sequence. Gypsy reverse transcriptase domains were identified in CBB2,

CBB3, and CBB4. For CBB2, the reverse transcriptase domain was located between

CDS1 and CDS2 (28,086-28,295 bp). For CBB3, the reverse transcriptase domain was

associated with the location of CDS2 and the adjacent intron (41,622-41,831 bp). For

CBB4, the reverse transcriptase domain corresponded to the location of CDS2 (4,572-

4,814).

Page 91: Identification and Characterization of Common Bacterial

76

Figure 2.11 Secondary structure models of predicted proteins for candidate CBB

resistance genes CBB1 (A), CBB2 (B), CBB3 (C) and CBB4 (D), using Phyre2 software

(Kelley and Sternberg 2009)

Page 92: Identification and Characterization of Common Bacterial

77

Table 2.10 Phyre2 secondary structure modelling (Kelley and Sternberg 2009) for candidate gene proteins CBB1, CBB2, CBB3 and CBB4

Candidate

Gene

Amino Acids

Modelled

Template Template Characteristic Similarity to

Template

Confidence

CBB1 14-127 D1mu2a2 Reverse transcriptase

DNA/RNA polymerase

12% 99.9%

146-262 C3I2uA_ Prototype foamy virus (pfv)2

integrase

22% 99.7%

CBB2 5-271 D1mu2a2 Reverse transcriptase

DNA/RNA polymerase

14% 100%

213-430 C3I2uA_ Prototype foamy virus (pfv)2

integrase

19% 99.9%

CBB3 5-173 D1mu2a2 Reverse transcriptase DNA/RNA

polymerase

18% 100%

193-338 C312uA_ Prototype foamy virus (pfv)2

integrase

15% 99.7%

CBB4 9-68 C3nr6A_ xenotropic murine leukemia virus-

related virus2 (xmrv) protease

21% 97.9%

75-279 D1mu2a2 Reverse transcriptase DNA/RNA

polymerase

15% 100%

218-476 C3I2uA_ Prototype foamy virus (pfv)2

integrase

12% 99.7%

Page 93: Identification and Characterization of Common Bacterial

78

2.4.3.1.4 CG Secondary Structure Prediction

Modeling of protein secondary structure using Phyre2 (Kelley and Sternberg

2009) resulted in models for each of the CG proteins (Figure 2.11). CGs CBB1-4 were

modelled with the same reverse transcriptase (D1mu2a2) and integrase (C312uA_)

proteins, and CBB4 had an additional section modelled with a protease protein

(C3nr6A_) (Table 2.11). The model for CBB2 was produced using two viral integrase

proteins (Table 2.11). The presence of the manually modelled LRR motifs were

unconfirmed by the secondary structure models produced for the CGs, except for the

small section of β-sheet structure in CBB4 (Figure 2.11).

2.4.3.2 Characterization of Genes with Putative Functions

2.4.3.2.1 Ascorbate Oxidase Gene Homologs

Gene 5 matched to a Multicopper oxidase/L-ascorbate oxidase gene at

5,835,002 – 5,839,661 bp on chromosome 6 (Figure 2.6; Table 2.7). When the full

gene sequence for gene 5 was BLASTed at NCBI, putative conserved Cupredoxin

superfamily domains, and L-ascorbate oxidase and multicopper oxidase multi-domains

were predicted (data not shown). When the amino acid sequence was BLASTed against

the PvGEA Atlas (O’Rourke et al. 2014), matches were identified in several

chromosomes to several L-ascorbate oxidase annotations, including the L-ascorbate

oxidase annotation gene 5 matched to within G19833 chromosome 6

(Phvul.006G011700). When the sequence was BLASTed to the G. max genome

(Schmutz et al. 2010) in Phytozome it identified homology to an ascorbate oxidase

annotation at 11,769,051 – 11,776,969 bp on chromosome 20 (Glyma.20G051900).

Page 94: Identification and Characterization of Common Bacterial

79

Table 2.11 Similarity between contig1701 genes of interest and identified homologues. Percent similarity between genes identified in P. vulgaris varieties OAC Rex contig1701 (gene 5), G19833 annotation Phvul.006G011700, G. max annotation Glyma.20G051900 was determined by aligning full gene sequences, coding sequences and protein sequences in ClustalW. Zucchini ascorbate oxidase protein sequence (1ASP_A; Messerschmidt et al. 1993) was also aligned to the protein sequences to indicate conservation of protein domains.

Genes Compared Similarity (%)

Full Gene Coding Sequence Protein Sequence

Gene 5 Phvul.006G011700 91 87 82

Glyma.20G051900 19 80 75

1ASP_A - - 58

Phvul.006G011700 Glyma.20G051900 17 89 90

1ASP_A - - 68

Glyma.20G051900 1ASP_A - - 69

Gene 7 G2 100 52 81

Glyma.08G361500 2 14 23

G2 Glyma.08G361500 2 41 46

Gene 6 Phvul.008G191100 97 8 5

Page 95: Identification and Characterization of Common Bacterial

80

Alignment of the three gene sequences (gene 5, Phvul.006G011700 and

Glyma.20G051900) showed high similarity between gene 5 and Phvul.006G011700

(91%), but the similarity was low between Glyma.20G051900 and the other two genes

(Table 2.11). When the coding sequences for the three genes were aligned, higher

similarities were observed, with similarity between the coding sequences >80% (Figure

2.12; Table 2.11). Alignments of the protein sequences for gene5, Phvul.006G011700,

Glyma.20G051900, and X-ray modelled Zucchini ascorbate oxidase (IASP_A), obtained

from NCBI indicated a number of regions with high similarities (Figure 2.13; Table 2.11).

When InterPro was used to identify domains within gene 5, Phvul.006G011700,

Glyma.20G051900 and 1ASP_A, multicopper oxidase type 1, 2 and 3 domains were

identified (Figure 2.13). The amino acid sequences of these domains were highly

conserved between the aligned genes. However, the domains indicated by InterPro in

gene 5 were less conserved (Figure 2.13). Phyre2 secondary structure modelling of

gene 5, Phvul.006G011700 and Glyma.20G051900 used the X-ray structure of the

Zucchini ascorbate osxidase 1ASP_A (Messerschmidt et al.1993) (Figure 2.14).

Secondary structure modelling of the putative ascorbate oxidase proteins resulted in

structures that consisted of three similar motifs (Figure 2.14).

2.4.3.2.2 β-Glucan Synthase-Like Homologs

Gene 7 matched to 49,683,929 – 49,687,869 bp in chromosome 8 of G19833,

which is associated with an EST, and multiple annotations for similarity to β-glucan

synthase-like or callose synthase genes from various other species. When the gene 7

sequence was BLASTed against NCBI, multiple matches to P. vulgaris hypothetical

proteins were

Page 96: Identification and Characterization of Common Bacterial

81

Gene 5 ------------------------------------------------------------

Phvul.006G011700 ATGGGGTTGAAAGCAGTTTTGGTTTGGTGC---ATATGGTTGGGGCTGATACAATATTCA

Glyma.20G051900 ATGGGTTTGAAAGCACTTTTTGTTTGGTGCATTATATGGTTAGGTTTAGCACATTTGTCC

Gene 5 ------------------------------------------------------------

Phvul.006G011700 GTTGGAGGAAGGGTGAGGCACTACAGGTTTGATGTGGAGTACATGATGAGAAAGCCAGAT

Glyma.20G051900 CTTGGAGGAAGAGTGAGGCACTACAAGTTTGATGTGGAGTACATGATCAGAAAGCCAGAT

Gene 5 ------------------------------------------------------------

Phvul.006G011700 TGTTTGGAACACGTTGTGATGGGAATCAATGGGCAGTTTCCAGGGCCAACTATTAGGGCT

Glyma.20G051900 TGCTTGGAACACGTTGTGATGGGAATCAACGGCCAGTTTCCAGGCCCAACTATTAGGGCT

Gene 5 ------------------------------------------------------------

Phvul.006G011700 GAAGTTGGTGACACTCTTCACATTGCTCTCACCAACAAGCTTTTCACTGAGGGAACTGTT

Glyma.20G051900 GAAGTTGGTGACATTCTTGACATTGCTCTCACCAACAAGCTTTTCACTGAGGGAACTGTT

Gene 5 ------------------------------------------------------------

Phvul.006G011700 ATTCACTGGCATGGAATCAGACAGGTTGGAACTCCTTGGGCAGATGGAACTGCTGCTATC

Glyma.20G051900 ATTCACTGGCATGGAATCAGACAGGTTGGAACTCCTTGGGCTGATGGAACTGCTGCCATC

Gene 5 ------------ATGTGTGTTGTAATGA---TTAATTGTAATTATATGGTGCA---GCCT

Phvul.006G011700 TCACAGTGTGCTATAAATCCAGGAGAAACTTTTCACTACAGGTTTATAGTTGACAGGCCT

Glyma.20G051900 TCACAGTGTGCTATAAACCCAGGAGAGGCTTTTCATTACAGGTTTACAGTTGACAGGCCT

** * * ** * * * * ** ** * ****

Gene 5 GGAACGTATTTCTACCATGGACACTATGGTATGCAAAGATCAGCAGGGTTGTATGGTTCA

Phvul.006G011700 GGAACATATTTCTACCATGGACACTATGGTATGCAAAGATCAGCAGGGTTGTATGGTTCA

Glyma.20G051900 GGTACATATTTCTATCATGGACACCATGGTATGCAAAGATCAGCTGGGTTGTATGGTTCA

** ** ******** ********* ******************* ***************

Gene 5 CTGATAGTAGATTTGCCAAAGGGACAAAACGAACCGTTTCACTATGATGATGAGTTCAAC

Phvul.006G011700 CTGATAGTAGATTTGCCAAAGGGACAAAACGAACCGTTTCACTATGATGATGAGTTCAAC

Glyma.20G051900 TTGATAGTAGATTTGCCAAAGGGACAAAACGAACCGTTTCATTACGATGGTGAATTCAAC

**************************************** ** **** *** ******

Gene 5 CTTCTTCTAAGTGATTTGTGGCACACCAGCTCACATGAACAAGAAGTTGGTCTCTCTTCC

Phvul.006G011700 CTTCTTCTAAGTGATTTGTGGCACACCAGCTCACATGAACAAGAAGTTGGTCTCTCTTCC

Glyma.20G051900 CTTCTTCTTAGTGATTTGTGGCACACCAGTTCACATGAACAGGAAGTTGGCCTCTCTTCC

******** ******************** *********** ******** *********

Gene 5 CTACCATTCAAATGGATTGGTGAACCTCAGAGTCTGTTGATCAATGGAAGGGGACAATTC

Phvul.006G011700 CTACCATTCAAATGGATTGGTGAACCTCAGAGTCTTTTGATCAATGGAAGGGGACAATTC

Glyma.20G051900 AAACCATTCAAATGGATTGGTGAACCTCAGACTCTACTCATCAATGGAAAAGGACAATTC

***************************** *** * ******* ** *********

Gene 5 AATTGTTCTCTGGCAGCTAAATTCATTAACACCACTTTACCCGAGTGCCAATTCAGAGGT

Phvul.006G011700 AATTGTTCTCTGGCAGCTAAATTCATTAACACCACTTTACCCGAGTGCCAATTCAGAGGT

Glyma.20G051900 AATTGTTCCCTTGCATCTAAATTCATCAACACAACCCTACCCCAATGCCAACTTAAAGGT

******** ** *** ********** ***** ** ***** * ****** * * ****

Gene 5 GGTGAAGAATGTGCACCTCAAATTCTTCATGTAGAGCCAAACAAGACCTACAGGATCAGA

Phvul.006G011700 GGTGAAGAATGTGCACCTCAAATTCTTCATGTAGAGCCAAACAAGACCTACAGGATCAGA

Glyma.20G051900 GGTGAAGAATGTGCCCCTCAGATTCTTCATGTGGAGCCAAACAAGACCTATAGAATCAGG

************** ***** *********** ***************** ** *****

Page 97: Identification and Characterization of Common Bacterial

82

Gene 5 ATTTCCAGCACCACTTCCTTGGCTTCACTCAACTTAGCCATTTCG---CTGAGATCTACC

Phvul.006G011700 ATTTCCAGCACCACTTCCTTGGCTTCACTCAACTTAGCCATTTCGAATCACAAACTTATT

Glyma.20G051900 ATTGCCAGCACCACTGCCCTAGCTTCACTCAACTTAGCCATTTCGAATCACAAACTTGTA

*** *********** ** * ************************ * * * *

Gene 5 TCTGT---AATTAATGGAGTTCA-ATACAATCTTTCTTTCTT-----------CTTCTTC

Phvul.006G011700 GTGGTGGAAGCTGATGGAAACTACGTTACACCATTTGTGGTTGATGATATGGATATTTAT

Glyma.20G051900 GTGGTGGAAGCTGATGGAAATTACGTTTCACCTTTTGCGGTTGATGATATTGACATCTAT

** * * ***** * * * * ** ** * *

Gene 5 TCTG---AATGTTACACAGT----------AGAGGC---------------AATTATTTG

Phvul.006G011700 TCTGGTGAAAGCTACTCAGTTCTCCTTCGCACAGATCAAGATCCAAAGAAAAACTATTGG

Glyma.20G051900 TCTGGGGAAAGTTATTCAGTGCTCCTTCGCACAGACCAAGATCCAAACAAAAATTATTGG

**** ** * ** **** * ** ** **** *

Gene 5 TTCTT--------------------------------------------CAAATAATGGA

Phvul.006G011700 CTTTCAATTGGAGTTAGAGGAAGAAA---ACCTAACACACCACAAGGTCTAACCATTCTA

Glyma.20G051900 CTATCAATCGGGGTGAGAGGAAGAAGGGCACCCAACACCCCACAAGGCCTAACCATTCTA

* * ** * * *

Gene 5 AATCACAAACTTATTGTGGTGGAAGCTGATGGAA-------ACTACGTTACAC---CATT

Phvul.006G011700 AACTACAAGACAATTTCTGCCTCAGTTTTTCCAACTTCTCCACCACCCCTCACACCCCTT

Glyma.20G051900 AACTATAAGCCCATTTCTGCCTCAATTTTCCCAATTTCTCCACCACCCATCACTCCCATT

** * ** *** * * * ** ** ** *** * **

Gene 5 TG---TGGTT----------------GATGATATGGATATTTATTCTGGTGAAAGCTACT

Phvul.006G011700 TGGAATGATTTTGAACGTAGCAAAGCATTCACCAAGAAAATCATTGCCAAGATGGGAACC

Glyma.20G051900 TGGAATGATTTTGAGCGCAGCAAAGCGTTCACCAAGAAAATCATTGCGAAAATGGGGACC

** ** ** * * ** * * *** * * **

Gene 5 C----------------------------AGTTCTCCTTCGCACAGATCAAGATC-----

Phvul.006G011700 CCACAACCTCCAAAACGTTCTGATCGTACAATTTTTCTCCTCAACACCCAAAACCGTGTT

Glyma.20G051900 CCACAACCTCCAAAACGCTCTGATCGTACAATTTTCCTCCTCAACACCCAAAATTTGCTT

* * ** * ** * ** *** *

Gene 5 ----------CAAAGAAAACTATTGGC---------TTTCAATTG--GACAACCCCTTAC

Phvul.006G011700 GCTGGATTCACAAAATGGGCTATTAACAATGTGTCCCTAACATTGCCAGCAACCCCTTAC

Glyma.20G051900 GATGGATTCACCAAATGGGCCATTAACAATGTGTCCCTAACATTGCCGCCAACCCCTTAC

* ** * *** * * **** ***********

Gene 5 TTGGGCTCCATCAAGTTCAAACTAA-CAATGCTTTTGATCAAACACCTCCACCTGTTACC

Phvul.006G011700 TTGGGCTCCATCAAGTTCAAACTAAACAATGCTTTTGATCAAACACCTCCACCTGTTACC

Glyma.20G051900 TTGGGTTCCATCAAATTCAAAATAAATAATGCCTTTGACAAAACTCCCCCACCAGTTACC

***** ******** ****** *** ***** ***** **** ** ***** ******

Gene 5 TTCCCTCAAGACTATGACATTTTCAACCCTCCTGTGAACCCTAATTCAACCATTGGCAAT

Phvul.006G011700 TTCCCTCAAGACTATGACATTTTCAACCCTCCTGTGAACCCTAATTCAACCATTGGCAAT

Glyma.20G051900 TTCCCACAAGATTATGACATCTTCAACCCTCCTGTGAACCCTAATGCATCTATTGGCAAT

***** ***** ******** ************************ ** * *********

Gene 5 GGGGTGTACAAGTTCAACCTTAACGAGGTTGTTGATGTGATATTGCAAAATGCAAATCAA

Phvul.006G011700 GGGGTGTACAAGTTCAACCTTAACGAGGTTGTTGATGTGATATTGCAAAATGCAAATCAA

Glyma.20G051900 GGGGTGTACATGTTCAACCTTAATGAAGTTGTTGATGTGATCTTGCAAAATGCAAACCAA

********** ************ ** ************** ************** ***

Page 98: Identification and Characterization of Common Bacterial

83

Gene 5 TTGAGTGGAAATGGTAGTGAGATTCACCCTTGGCACTTGCATGGACATGACTTTTGGGTT

Phvul.006G011700 TTGAGTGGAAATGGTAGTGAGATTCACCCTTGGCACTTGCATGGACATGACTTTTGGGTT

Glyma.20G051900 TTATCTGGGAGTGGTAGTGAGATTCACCCTTGGCACTTGCATGGACATGACTTTTGGGTT

** *** * *************************************************

Gene 5 TTGGGGTATGGAGAAGGCAAATTCAAACAAGGTGATGAGAAGAAATTCAACTTGACACAT

Phvul.006G011700 TTGGGGTATGGAGAAGGCAAATTCAAACAAGGTGATGAGAAGAAATTCAACTTGACACAT

Glyma.20G051900 TTGGGGTATGGAGAAGGGAAATTCAAACCAAGTGATGAGAAGAAATTCAATTTGACACAT

***************** ********** * ******************* *********

Gene 5 GCACCATTGAGGAACACAGCAGTGATATTTCCATATGGTTGGACTGCATTAAGGTTTAAG

Phvul.006G011700 GCACCATTGAGGAACACAGCAGTGATATTTCCATATGGTTGGACTGCATTAAGGTTTAAG

Glyma.20G051900 GCACCGTTGAGGAACACTGCTGTGATATTTCCATATGGTTGGACTGCTTTGAGGTTTAAG

***** *********** ** ************************** ** *********

Gene 5 GCAGATAATCCAGGAGTTTGGGCCTTCCATTGCCACATTGAACCACATTTGCACATGGGA

Phvul.006G011700 GCAGATAATCCAGGAGTTTGGGCCTTCCATTGCCACATTGAACCACATTTGCACATGGGA

Glyma.20G051900 GCTGATAACCCAGGAGTTTGGGCCTTCCATTGTCACATTGAACCCCATTTGCACATGGGA

** ***** *********************** *********** ***************

Gene 5 ATGGGTGTGATTTTTGCTGAGGGTGTCCACAAAGTTGGAAAAATTCCAAGGGAAGCACTA

Phvul.006G011700 ATGGGTGTGATTTTTGCTGAGGGTGTCCACAAAGTTGGAAAAATTCCAAGGGAAGCACTA

Glyma.20G051900 ATGGGTGTGATTTTTGCTGAGGGTGTTCACAAAGTTGGAAAAATCCCAAGAGATGCACTA

************************** ***************** ***** ** ******

Gene 5 AATTGTGGGCTTGTTGGGAAGAAGCTAGTAGAAAATGGACACTACTGA

Phvul.006G011700 AATTGTGGGCTTGTTGGGAAGAAACTAGTAGAAAATGGACACTACTGA

Glyma.20G051900 ACTTGTGGGCTTACTGGGAATAAGCTAGTAGGAAATAGACACTATTGA

* ********** ****** ** ******* **** ******* ***

Figure 2.12 Alignment and annotation of ascorbate oxidase coding sequences. Exons are indicated in alternating dark and light coloured boxes. Gene 5 is indicated in green, Gene 16 is indicated in purple, Phvul.006G011700 is indicated in red, and Glyma.20G051900 is indicated in blue.

Page 99: Identification and Characterization of Common Bacterial

84

Gene5 ------------------------------------------------------------

Phvul.006G011700 -MGLKAVLVWCIWLGLIQYSVGGRVRHYRFDVEYMMRKPDCLEHVVMGINGQFPGPTIRA

Glyma.20G051900 MGLKALFVWCIIWLGLAHLSLGGRVRHYKFDVEYMIRKPDCLEHVVMGINGQFPGPTIRA

1ASP_A ----------------------SQIRHYKWEVEYMFWAPNCNENIVMGINGQFPGPTIRA

Gene5 ----------------------------------------------MCVVMINCNYMVQP

Phvul.006G011700 EVGDTLHIALTNKLFTEGTVIHWHGIRQVGTPWADGTAAISQCAINPGETFHYRFIVDRP

Glyma.20G051900 EVGDILDIALTNKLFTEGTVIHWHGIRQVGTPWADGTAAISQCAINPGEAFHYRFTVDRP

1ASP_A NAGDSVVVELTNKLHTEGVVIHWHGILQRGTPWADGTASISQCAINPGETFFYNFTVDNP

.: : .*

Gene5 GTYFYHGHYGMQRSAGLYGSLIVDLPKGQNEPFHYDDEFNLLLSDLWHTSSHEQEVGLSS

Phvul.006G011700 GTYFYHGHYGMQRSAGLYGSLIVDLPKGQNEPFHYDDEFNLLLSDLWHTSSHEQEVGLSS

Glyma.20G051900 GTYFYHGHHGMQRSAGLYGSLIVDLPKGQNEPFHYDGEFNLLLSDLWHTSSHEQEVGLSS

1ASP_A GTFFYHGHLGMQRSAGLYGSLIVDPPQGKKEPFHYDGEINLLLSDWWHQSIHKQEVGLSS

**:***** *************** *:*::******.*:****** ** * *:*******

Gene5 LPFKWIGEPQSLLINGRGQFNCSLAAKFINTTLPECQFRGGEECAPQILHVEPNKTYRIR

Phvul.006G011700 LPFKWIGEPQSLLINGRGQFNCSLAAKFINTTLPECQFRGGEECAPQILHVEPNKTYRIR

Glyma.20G051900 KPFKWIGEPQTLLINGKGQFNCSLASKFINTTLPQCQLKGGEECAPQILHVEPNKTYRIR

1ASP_A KPIRWIGEPQTILLNGRGQFDCSIAAKYDSN-LEPCKLKGSESCAPYIFHVSPKKTYRIR

*::******::*:* .: : .. * *:::*.*.*** *:**.*:******

Gene5 ISSTTSLASLNLAISLRSTSVINGVQYNLSFFFFSECYTVEAIICSSNNGNHKLIVVEAD

Phvul.006G011700 ISSTTSLASLNLAISN-----------------------------------HKLIVVEAD

Glyma.20G051900 IASTTALASLNLAISN-----------------------------------HKLIVVEAD

1ASP_A IASTTALAALNFAIGN-----------------------------------HKLIVVEAD

*:***:**:**:**. *:*:*****

Gene5 GNYVTPFVVDDMDIYSGESYSVLLRTDQDPKKTIGFQLDN--------------------

Phvul.006G011700 GNYVTPFVVDDMDIYSGESYSVLLRTDQDPKKNYWLSIGVRGRKPN-TPQGLTILNYKTI

Glyma.20G051900 GNYVSPFAVDDIDIYSGESYSVLLRTDQDPNKNYWLSIGVRGRRAPNTPQGLTILNYKPI

1ASP_A GNYVQPFYTSDIDIYSGESYSVLITTDQNPSENYWVSVGTRARHPN-TPPGLTLLNYLPN

**** ** ..*:***********: ***:*.:. ..:.

Type III

Type I

Page 100: Identification and Characterization of Common Bacterial

85

Gene5 ------------------------------------------------------------

Phvul.006G011700 SASVFPTSPPPLTPLWNDFERSKAFTKKIIAKMGTPQPPKRSDRTIFLLNTQNRVAGFTK

Glyma.20G051900 SASIFPISPPPITPIWNDFERSKAFTKKIIAKMGTPQPPKRSDRTIFLLNTQNLLDGFTK

1ASP_A SVSKLPTSPPPQTPAWDDFDRSKNFTYRITAAMGSPKPPVKFNRRIFLLNTQNVINGYVK

Gene5 -------------PLLGLHQVQTNNAFDQTPPPVTFPQDYDIFNPPVNPNSTIGNGVYKF

Phvul.006G011700 WAINNVSLTLPATPYLGSIKFKLNNAFDQTPPPVTFPQDYDIFNPPVNPNSTIGNGVYKF

Glyma.20G051900 WAINNVSLTLPPTPYLGSIKFKINNAFDKTPPPVTFPQDYDIFNPPVNPNASIGNGVYMF

1ASP_A WAINDVSLALPPTPYLGAMKYNLLHAFDQNPPPEVFPEDYDIDTPPTNEKTRIGNGVYQF

* ** : : :***:.*** .**:**** .**.* :: ****** *

Gene5 NLNEVVDVILQNANQLSGNG----------------------------------------

Phvul.006G011700 NLNEVVDVILQNANQLSGNGSEIHPWHLHGHDFWVLGYGEGKFKQGDEKKFNLTHAPLRN

Glyma.20G051900 NLNEVVDVILQNANQLSGSGSEIHPWHLHGHDFWVLGYGEGKFKPSDEKKFNLTHAPLRN

1ASP_A KIGEVVDVILQNANMMKENLSETHPWHLHGHDFWVLGYGDGKFSAEEESSLNLKNPPLRN

::.*********** :. .

Gene5 -TVIFPYGWTALRFKADNPGVWAFHCHIEPHLHMGMGVIFAEGVHKVGKIPREALNCGLV

Phvul.006G011700 TAVIFPYGWTALRFKADNPGVWAFHCHIEPHLHMGMGVIFAEGVHKVGKIPREALNCGLV

Glyma.20G051900 TAVIFPYGWTALRFKADNPGVWAFHCHIEPHLHMGMGVIFAEGVHKVGKIPRDALTCGLT

1ASP_A TVVIFPYGWTAIRFVADNPGVWAFHCHIEPHLHMGMGVVFAEGVEKVGRIPTKALACGGT

.*********:** ***********************:*****.***:** .** ** .

Gene5 GKKLVENGHY-

Phvul.006G011700 GKKLVENGHY-

Glyma.20G051900 GNKLVGNRHY-

1ASP_A AKSLINNPKNP

.:.*: * :

Figure 2.13 Alignment and annotation of L-ascorbate oxidase amino acid sequences. Amino acid sequences were aligned for predicted ascobate oxidase like genes from P. vulgaris varieties OAC Rex (gene 5) and G19833 (Phvul.006G011700; Schmutz et al. 2014), a predicted G. max ascorbate oxidase gene (Glyma.20G051900; Schmutz et al. 2010) and an ascorbate oxidase protein from Cucurbita pepo var. melopepo, for which the secondary structured was resolved by X-ray (1ASP_A; Messerschmidt et al. 1992). Type 1, 2 and 3 multicopper oxidase domains are outlined and labelled. A copper binding site is highlighted in grey.

Type II

Page 101: Identification and Characterization of Common Bacterial

86

Figure 2.14 Secondary structure models of OAC Rex and G19833 L-ascorbate oxidase protreins. Secondary structure models for Gene 5 (A) from contig1701, G19833 gene annotation Phvul.006G011700 (B) and G. max gene annotation Glyma.20G051900 (C) were produced using Phyre2 (Kelley and Sternberg 2009). Ascorbate oxidase type I, II and III domains are indicated by circles. All models were created using the X-ray structure of a zucchini ascorbate oxidase (D) as a template (Messerschmidt et al. 1993)

Page 102: Identification and Characterization of Common Bacterial

87

Table 2.12 Phyre2 secondary structure modelling (Kelley and Sternberg 2009) for gene 5, Phvul.006G011700, Glyma.20G051900 and 1ASP_A

Gene Name Amino Acids

Modelled

Template Template

Characteristic

Similarity to

Template

Confidence

Gene 5 11-408 C1asqB_ Ascorbate oxidase

(Messerschmidt et al. 1993)

57% 100%

Phvul.006G011700 22-570 C1asqB_ Ascorbate oxidase

(Messerschmidt et al. 1993)

69% 100%

Glyma.20G051900 23-572 C1asqB_ Ascorbate oxidase

(Messerschmidt et al. 1993)

70% 100%

1ASP_A 1-552 C1asqB_ Ascorbate oxidase

(Messerschmidt et al. 1993)

100% 100%

Page 103: Identification and Characterization of Common Bacterial

88

identified, as well as to callose synthase 3-like predictions from other species, including

Medicago truncatula (data not shown). BLAST against the pvGEA EST database

identified several β-glucan synthase-like gene matches on several chromosomes,

including chromosome 8 (data not shown). However, none of these matches

corresponded to 49,683,929 – 49,687,869 bp location in chromosome 8.

Since there was no transcript or gene annotation within G19833 where gene 7

matched, a ~12 kb sequence (49,679,901 – 49,691,730 bp on chromosome 8) was

selected surrounding the match location and FGeneSH analysis identified 4 genes.

One gene (G2) corresponded to the gene 7 match location within the G19833 sequence

(49683929 – 49687948 bp; Figure 2.6). Gene 7 and G2 gene sequences were then

BLASTed against the G. max genome sequence in Phytozome (Schmutz et al. 2010),

which identified homology to gene annotation Glyma.08GG361500 on chromosome 8

(47,300,299 – 47,324,252 bp).

Alignment of gene 7, G2 and Glyma08G361500 DNA sequences showed high

similarity between gene 7 and G2, but low similarity to Glyma08G361500 (Table 2.12).

Alignment of the coding sequences indicated low similarity between all three genes

(Table 2.11). When the protein sequences were aligned, the highest similarity to

Glyma08G361500 was seen for both genes, however, they were still <50% similar

(Appendix 1; Table 2.11).

When the protein sequences were BLASTed at InterPro, callose synthase family

related signatures were identified in G2 and Glyma08G361500. G2 also contained

three transmembrane domains which corresponded to the location of similar domains in

Glyma08G361500 (Appendix 1). Glyma08G361500 was classified as a callose

Page 104: Identification and Characterization of Common Bacterial

89

Figure 2.15 Secondary structure model for predicted β glucan synthase-like genes. Phyre 2 (Kelley and Sernberg 2009) secondary structure models were created for P. vulgaris gene 7 (A) from OAC Rex contig1701, and G2 (B) from G19833, and G. max gene annotation Glyma.08G361500 (C).

Table 2.13 Phyre2 secondary structure modelling (Kelley and Sternberg 2009) for gene 7, G2, and Glyma.08G361500

Gene Name Amino Acids

Modelled

Template Template

Characteristic

Similarity to

Template

Confidence

Gene 7 36-98 D1h2vc3 Alpha-alpha superhelix 17% 52.3%

G2 105-225 C4gmjE_ RNA binding protein 17% 54.4%

Glyma.08G361500 66-449 C4hg6A_ Cellulose synthase

subunit a

14% 96.8%

Page 105: Identification and Characterization of Common Bacterial

90

synthase based on the whole peptide sequence, and as a glycosyl transferase family 48

(glucan synthase) member based on amino acids 1,059-1,815. Domains identified in

Glyma08G361500 included a vacuolar protein sorting-associated protein Vta1/callose

synthase N-terminal domain (amino acids 45-175), a 1,3-β-glucan synthase subunit

FKS1-like domain (amino acids 321-436), two large and one small cytoplasmic domains

and two transmembrane domains (Appendix 1). No family or domain signatures were

predicted for gene 7 (Appendix 1).

When Phyre2 was used to predict protein secondary structure for gene 7, G2 and

Glyma08G361500, there was very low similarity to previously resolved proteins (Figure

2.15; Table 2.13). Secondary structure modelling for gene 7 and G2 produced low

confidence (<55%) models for small sections of the protein sequences (Figure 2.15A-B;

Table 2.12). Glyma.08G361500 was modelled using a cellulose synthase subunit a,

and was a 96.8%confident model, with 14% similarity to the model template (c4hg6A_)

(Figure 2.15C; Table 2.13).

2.4.3.2.3 Hypothetical Gene Annotations

Gene 6 matched to a hypothetical protein annotation at 49,682,264 – 49,683,238 bp on

chromosome 8 (Figure 2.5-6; Table 2.6). The matched region in G19833 was

associated with RNA seq data obtained from the nodules and roots, EST and transcript

information related to the hypothetical gene annotation Phvul.008G191100.1. The

coding sequence predicted in FGeneSH for gene 6 corresponded to a region beside this

annotation, but still corresponded to low level RNA seq expression data (49,682,789 –

49,682,959 bp). Neither the coding sequence associated with Phvul.008G191100.1,

Page 106: Identification and Characterization of Common Bacterial

91

nor the coding sequence of gene 6 had high similarity to known protein domains when

BLASTed against the NCBI database.

When the gene 6 and Phvul.008G191100 DNA sequences were BLASTed against the

G. max sequence in Phytozome (Schmutz et al. 2010), no homologs were found.

Alignment of the DNA sequences for gene 6 and Phvul.008G191100.1 indicated that

gene 6 was 97% similar to Phvul.008G191100 (Table 2.11). However, alignment of

coding and protein sequences indicated low similarity (8% and 5% respectively)

between gene 6 and Phvul.008G191100 (Appendix 2; Table 2.11) .

Analysis with InterPro identified no protein family or domain signatures within the

protein sequences. Phyre2 modelling (Appendix 3) produced a low confidence (15.2%)

model that had low similarity (33%) to an iron regulated transcription activator (c4ImgD)

for gene 5 (Appendix 4). The model produced for Phvul.008G191100 also had low

confidence (32%) (Appendix 4).

2.4.4 Isolation and Amplification of CGs

Amplification of bands from genomic OAC Rex DNA using the designed primers

for CGs CBB2 and CBB3 (Table 2.14) produced multiple bands ranging from 850 –

5,000 bp (Figure 2.16 A-D). To reduce amplification of non-specific bands, BiBAC2

library clone 21 DNA was used as the template, which produced a single band when the

same primers were used (Figure 2.16 E-H). The PCR fragments were cloned into E. coli

(Figure 2.17) and 1 kb of the forward (5’) and reverse (3’) ends of the clones were

sequenced. BLAST of the 5’ and 3’ end sequences for all sequenced clones matched

to contig1455 at 2,236-3,401 bp and 4,526-5,566 bp respectively (Figure 2.18). When

the contig1455 sequence surrounding the 5’ and 3’ 1 kb sequences was analyzed using

Page 107: Identification and Characterization of Common Bacterial

92

FGeneSH and GeneMark, three genes were predicted (discussed in section 2.4.1.3.2;

Figure 2.4; Table 2.5). Additionally, when contig1455 was compared to the region

surrounding CGs CBB2 and CBB3, it was 99% similar to each region, with small (<100

bp) sequence differences, so the gene was included as a CG (CBB4).

Page 108: Identification and Characterization of Common Bacterial

93

Table 2.14 Primers used to amplify candidate genes CBB2 and CBB3 from OAC Rex DNA Primer Forward Sequence Reverse Sequence Band Size

(bp)

CBBS2-14 CATTTCTCACCACCTCCACCTACCC GAAAGTCTTTGAGGGTTAGGTGAGTGG 2,553

CBBS2-15 CACCTATAAATACCTAATGCCCAAACACC GTGAGTGAAGTAGTGCAGAGTTAGGTGG 2,808

CBBS3-8 CACCTATAAATACCTAATGCCCAAACACC GTTGGAGAATTGTTAGGTGTGAGGTTTTG 2,391

CBBS3-9 CACATTTCTCACCACCACCTCCACCTAC GTTTTGGGGTTGGGAATCTTTGGTG 2,204

Figure 2.16 Candidate gene PCR amplification from OAC Rex genomic DNA and BiBAC2 library DNA sources. Amplification of candidate genes from OAC Rex genomic DNA exhibited non-specific binding A) CBBS3-8, B) CBBS3-9, C) CBBS2-14, and D) CBBS2-15. Amplification of candidate genes from OAC Rex BiBAC2 library clone 021 E) CBB3-8, F) CBB3-9, G) CBB2-14, and H) CBB2-15 produced single bands. No bands were produced when water was used as a negative control (NC)

Page 109: Identification and Characterization of Common Bacterial

94

Figure 2.17 Common bacterial blight resistance candidate gene clones. Amplification of ~3.5 kb band using candidate gene primers for CBBS3-8 (A), CBBS3-9 (B), CBBS2-14 (C), CBBS2-15 from Escherichia coli clones. E. coli clones contained DNA amplified from OAC Rex BiBAC2 clone 021 using primers CBBS3-8 and CBBS3-9 designed for candidate gene CBB3, and CBBS2-14 and CBBS2-15 designed for candidate gene CBB2.

Figure 2.18 Location of clone sequence match within contig1455. Sequence of 1kb flanking ends of CBB candidate gene Escherichia coli clones matched to CBB4 5’ and 3’ ends when BLAST search was performed. The 1 kb sequences matched to 2,236-3,401 bp and 4,526-5,566 bp within the contig1455 sequence, and are indicated with a blue box.

Page 110: Identification and Characterization of Common Bacterial

95

2.5 Discussion

2.5.1 CG Characteristics

2.5.1.1 R Gene Characteristics of CGs

Several characteristics that are exhibited by CBB resistance CGs CBB1, CBB2,

CBB3 and CBB4 have previously been reported in literature. The organization of

multiple CGs and genes with potential disease resistance functions on contig1701

(Figure 2.2B and 2.3; Table 2.4, Table 2.7) are typical of resistance gene family clusters

(Michelmore and Meyers 1998; Graham et al. 2002; Meyers et al. 2003; Ameline-

Torregrosa et al. 2008). This clustering was accompanied by high similarity between

two of the CGs on contig1701 (CBB2 and CBB3) and the single CG identified on

contig1455 (CBB4) (Figure 2.2B). High similarity between resistance genes has been

frequently reported, especially in gene families where genes are present in close

proximity (Anderson et al. 1997; Meyers et al. 2003; Song et al. 2003). Both

characteristics are thought to be the result of frequent recombination events, which

result in high variability in the pathogen-interaction region (LRR) and may change

pathogen recognition capabilities (Anderson et al. 1997; Meyers et al. 1998; Song et al.

2003; Leister, 2004; Bakker et al. 2006).

Another important feature of the CGs is the presence of LRR domains, which are

found in many resistance related protein classes such as NBS-LRR, eLRR, RLKs and

RLPs (Meyers et al. 2003; Wang et. al. 2008b). Although it was not reflected in the

secondary structure model (Figure 2.11), LRR motifs were identified in the protein

sequence for each of the CG proteins (Figure 2.10). In addition, CBB2-4 contained an

Page 111: Identification and Characterization of Common Bacterial

96

LDL motif, but didn’t contain end motifs similar to those described for CC-NBS-LRR and

TIR-NBS-LRR proteins in Arabidopsis (Meyers et al. 2003).

Although there are similarities between the CBB2, CBB3 and CBB4 gene

sequences (Table 2.9), they were sufficiently variable to encode proteins which were

even more variable (Figure 2.9). Both the overall protein sizes and the numbers of LRR

motifs identified within the sequences varied (Figure 2.10; Table 2.6), resulting in

similarities between proteins as low as 11%, in the case of CBB1 and CBB2 (Table 2.9).

However, an overall consensus sequence (LxxxLxLxLxxL) was still identified, that was

applicable to all four predicted CG proteins (Figure 2.10). The finding that the motif is

not an exact match to LRR motifs reported previously (LxxLxLxxL), is consistent with

reports of high variability in LRR motif organization in NBS-LRR genes, within and

between species (Kajava 1998; Monosi et al. 2004; Zhou et al. 2004b)

2.5.1.2 Unique/Unusual Characteristics of CGs

Although the characteristics discussed above provide evidence to support the

hypothesis that they play a role in disease resistance, there are some differences which

indicate otherwise. When the protein sequences were analyzed both manually and via

predictive programs, NBS, CC, TIR, kinase, transmembrane and eLRR motifs were not

identified, and no signal peptide was indicated for any of the CGs. The sizes of the

loops between LRR motifs were more variable than expected when compared to the

LRR domains of NBS-LRRs and PGIPs (Di Matteo et al. 2003; Meyer et al. 2003).

Additionally, the secondary structure predicted for the CGs failed to support the manual

prediction of LRR domains (Figure 2.11). It was expected that the secondary structure

of the CGs would resemble the resolved secondary structures for eLRR protein PGIP,

Page 112: Identification and Characterization of Common Bacterial

97

and receptor like kinase RPK2 (Di Matteo et al. 2003; Song et al. 2014). These proteins

exhibit LRRs that align to form a stacked backbone, with the inter-LRR sequence

forming loops. Since there are few plant pathogen-resistance LRR proteins modelled, it

might be assumed that the failure to model CBB1-4 protein sequences into the

expected LRR secondary structure model is due to the lack of a good reference

structure for the proteins analysed with the modelling program. However, when a PGIP

was used as the model template, the expected secondary structure model still could not

be obtained. This evidence indicates that the CBB1-4 may not be LRR proteins, and

originate from another classification of protein.

2.5.1.3 Retrotransposon Characteristics of CGs

Retrotransposons are characterized by the ability to make copies of themselves

using RNA as the transposition medium, reverse transcribe and re-insert into the

genome (Wicker et al. 2007). BLAST search and secondary modelling of the CG

proteins identified low similarities to viral integrases, reverse transcriptase and

DNA/RNA polymerase proteins and GAG/POL/ENV polyproteins (see section 2.4.3.1.3;

Table 2.11). These are all proteins found in various types of retrotransposons, although

the similarity to the templates was low (Table 2.11), with percent similarity reaching no

higher than 22% for an 116 amino acid sequence for CBB4. This is likely to be an

indication of the lack of information about the secondary structure of plant

retrotransposon polyproteins.

When the CGs and other predicted genes from the area surrounding the CGs

were BLASTed against the G19833 sequence, many of the genes matched to areas

within the G19833 genome that were marked as retrotransposons (Figure 2.3; Table

Page 113: Identification and Characterization of Common Bacterial

98

2.7). In several cases, including for CBB2 and CBB3, there was also RNASeq data

associated with the retrotransposons, indicating that they were active (Table 2.7).

Retrotransposons have been shown to integrate preferentially into genic regions (Miyao

et al. 2003; Le et al. 2007). It is therefore not surprising that multiple gene predictions

associated with retrotransposons were present in a contig1701. Since the contig was

selected based on its presumed association with disease resistance marker PV-ctt001,

the presence of multiple possible resistance-related genes and retrotransposons

indicate the likelihood that part of a multigene family is represented in the sequence.

Retrotransposons have previously been shown to be involved in disease

resistance by interacting with resistance genes. When transposable elements

transpose into the location of a resistance gene, they may disrupt gene function which

might lead to reduced or abolished resistance (Song et al. 1997; Hernández-Pinzón et

al. 2009). Gene 7 may be an example of this, since it matched to the site of a putative

callose synthase gene on Chromosome 8 in the G19833 sequence assembly, but was

shown to be similar to only a portion of the large gene expressed in G. max (Figure 2.6,

2.15 and 2.16; Table 2.7 and 2.12). Studies have also shown that transposable

elements transpose in front of or between two genes and cause a chimeric protein to be

formed (Kashkush et al. 2003; Frost et al. 2004). Transposition can also change

promoter function and epigenetic controls (Hayashi et al. 2009; McDowell and Meyers

2013; Tsuchiya and Eulgem 2013). Considering that transposable elements were

reported to comprise 45% of the G19833 sequence (Schmutz et al. 2014), the likelihood

of an interaction between transposable elements and resistance genes is high, and is

an area to expand research into pathogen resistance.

Page 114: Identification and Characterization of Common Bacterial

99

2.5.2 Additional CGs on Contig1701

Since the predicted CGs CBB1-4 may be retrotransposons, the characterization

of three other genes within contig1701, which have potential disease resistance roles,

were explored. The genes were classified as ascorbate oxidase (gene 5), callose

synthase (gene 7) and hypothetical gene (gene 6) (see section 2.4.3.2; Appendices 1-4;

Figure 2.3 and 2.12-2.15; Table 2.7 and 2.11). Only gene 5, which matched an

ascorbate oxidase gene in G19833 chromosome 6, was likely to be a functional gene

based on the gene and protein analyses performed (see section 2.4.3.2).

The other two potential genes (callose synthase, hypothetical) either were not

very similar in sequence or secondary structure to previously identified genes of their

classification (Appendix 1-4; Figure 2.15; Table 2.11 and 2.13), or they were similar to

other uncharacterized genes and so were uninformative (Appendices 2-4). Gene 6 and

gene 7 may be fragments of genes that were produced by recombination events that

are often associated with the evolution of resistance genes (Michelmore and Meyers

1998; Meyers et al. 2005; Liu et al. 2007; Joshi and Nayak 2013). The association

between multiple predicted genes within the contig1701 with retro-element annotations

within the G19833 (Figure 2.3 and 2.5; Table 2.7) sequence suggests that transposable

element activity caused gene fragmentation in this stretch of the bean genome.

2.5.2.1 Ascorbate Oxidases and Disease Resistance

Ascorbate oxidases (AOs) are cell wall bound proteins which, are members of

the multi-copper oxidase family and are involved in binding and oxidizing ascorbic acid

(AA) in the apoplast (Messerschmidt et al. 1989). AA is involved primarily in providing

protection against reactive oxygen species (ROS) which are involved in the modulation

Page 115: Identification and Characterization of Common Bacterial

100

of stress (Smirnoff 2000) and defense responses (Pastori et al. 2003; Pavet et al. 2005)

have been shown to have a role in stress and pathogen resistance in various plant

species (Pavet et al. 2005; Pignocchi et al. 2006; Sanmartin et al. 2007; Fotopoulos et

al. 2008).

The presence of a predicted ascorbate oxidase gene within contig1701 is

suggestive of a potential role in resistance to CBB. The involvement of AA in protection

against ROS indicates that gene 5, as an AO, would regulate PCD associated with cell

death (Pavet et al. 2005), which is not a characteristic of the CBB resistance reaction.

However, the truncated domain predictions for gene 5 (Figure 2.13) may cause

incomplete binding and oxidation to occur, which has the potential to affect the

accumulation of ROS and the resultant hypersensitive response. Since this gene was

characterized at the bioinformatics level only, further studies would show more

definitively whether gene 5 is a functional ascorbate oxidase.

2.5.3 CG Location within Bean Genome

When the genes identified in the contig1701 sequence assembly were BLASTed

against the G19833 sequence, there were high numbers of significant matches (> e -63)

to many chromosomes for most of the predicted genes. The best matches (Figure 2.6;

Table 2.7) were to chromosomes 6 (5,817,952 – 5,839,661 bp) and 8 (49,682,264 –

49,698,171 bp). Since the contig1701 sequence was produced from BiBAC clones that

were selected based on the presence of disease resistance marker PV-ctt001, it was

expected that the contig and the genes predicted on it would match best to the area

surrounding its location (517,577-517,741 bp) on chromosome 4 (Tar’an et al. 2001;

Perry 2010; Perry et al. 2013). This is also the location of a large cluster of NBS-LRR

Page 116: Identification and Characterization of Common Bacterial

101

genes at the end of chromosome 4, which is associated with a variety of resistance-

related genes (Geffroy et al. 2009; Schmutz et al. 2014)

An explanation for this discrepancy is that the contig1701 sequence was

assembled using sequence data which originates from both Chromosome 6 and 8, and

assembled into one continuous sequence by human or program error. However, the

junction between the section that matched to chromosome 6 (contig1701-1) and the

section that matched to chromosome 8 (contig1701-2) was confirmed by PacBio reads.

Additionally, the creation of a new reference assembly which had read coverage of 8-

2,037 per 1 kb, indicating that it is likely that the sequence is continuous for the 1-35 kkb

length of contig1701 (see section 2.4.1.2.1; Figure 2.2A). The presence of the PV-

ctt001 marker within the BiBAC2 library clone from which CG CBB4 was isolated

(Rex021F10; Figure 2.1, 2.16 and 2.17) indicates that the contig1701 sequence is from

chromosome 4. This suggests the possibility that sequence rearrangement has

relocated sections of chromosomes 6 and 8 to chromosome 4 in the process of

integrating P. acutifolius DNA into the common bean genome. Considering the

apparent presence of retrotransposons in both the contig1455 and contig1701

sequence assemblies discussed above, this is likely to have been facilitated by

transposable element action, however, further studies are required for confirmation.

The difficulty in identifying a location for the full contig1701 sequence within the

OAC Rex genome assembly is that it is located within a region that has yet to be

completely assembled. Perry et al. (2013) compared the location of the PV-ctt001

marker and its surrounding area within OAC Rex with the G19833 sequence (Schmutz

et al. 2014). Although there was high conservation in the genomic region between

Page 117: Identification and Characterization of Common Bacterial

102

varieties, there was a 12 bp sequence difference between PV-ctt001 in OAC Rex, and

the corresponding location in G19833. There was also a large section (~200 kb) of the

G19833 sequence which did not match any of the assembled contigs within the OAC

Rex assembly. Since CBB resistance in OAC Rex resulted from an interspecific cross

between P. vulgaris and P. acutifolius (Parker 1985; Scott and Michaels 1992; Michaels

et al. 2006), it is possible that the area in question has higher amounts of the P.

acutifolius DNA, which is not present in the G19833 genome. However, the genomic

comparisons performed by Perry et al. (2013) indicated that differences between OAC

Rex and G19833 sequences were limited to single genes or up to 3 genes, as opposed

to a large stretch of highly different sequence.

The matching patterns observed by Perry et al. (2013) are similar to what was observed

in the current study for the comparison between section contig1701-1 and the the

G19833 sequence. The genes found in contig1701-1 (genes 1-5) matched to a segment

of G19833 chromosome 6 that contained all 5 genes in the same orientations.

However, an additional gene was predicted in the G19833 sequence between

contig1701 genes 3 and 4 (Figure 2.6). In contrast, there were multiple differences in

the number and size of genes identified in section contig1701-2 compared to those

identified in the associated region within the G19833 sequence assembly (genes 6-10).

This may be a reflection of the presence of multiple retrotransposon like elements in this

sequence, whith gene predictions identifying multiple sections of transposable element

polyproteins.

Page 118: Identification and Characterization of Common Bacterial

103

2.5.3.1 CBB Resistance Markers and CG G19833 Genome Location

Two CBB resistance QTLs, SU91 and UBC420, have previously been reported

on chromosomes 6 and 8 (Pedraza et al. 1997; Yu et al. 2000b; Vandemark et al. 2008;

Shi et. al 2011). The QTLs associated with these markers have also been shown to

have epistatic effects which lead to increased disease resistance when both markers

are present (Shi et al. 2011; Durham et al. 2013). SU91 was located initially to

chromosome 8 by linkage mapping (Pedraza et al. 1997). Subsequently, genome

sequence analysis by Perry et al. (2013) physically located the marker sequence to a

contig within the current OAC Rex sequence. Although the marker itself was absent

from the G19833 sequence assembly, the sequence surrounding SU91 in the OAC Rex

contig matched to the segment between 58,994,870-59,444,870 bp in chromosome 8

of G19833. UBC420 was located to chromosome 6 by linkage mapping in various

populations (Yu et al. 2000b; Vandemark et al. 2008; Durham et al. 2013), as well as by

association mapping performed by Shi et al. (2011). BLAST of the UBC420 marker

sequence (EF553635.1) against G19833 results in a match from 8,124,578-8125483

bp, and alignment between the two indicates 77% similarity.

The genes predicted in section contig1701-1 matched to ~5.8 Mbp which is

approximately 2.3 Mbp away from the UBC420 location on chromosome 6. In the case

of contig1701-2 genes, they matched to ~49.7 Mbp which is approximately 10 Mbp

away from SU91 on chromosome 8 (Perry et al. 2013). Schmutz et al. (2014) reported

the recombination rate of euchromatic arms to be 220 kb/cM, which indicates that the

regions to which the contig1701-1 and contig1701-2 genes have similarity to are 10.45

cM and 45.45 cM from the location of UBC420 on chromosome 6 and SU91 on

Page 119: Identification and Characterization of Common Bacterial

104

chromosome 8. Thus, the likelihood that CGs CBB1, CBB2, CBB3 and CBB4 are the

resistance genes associated with these CBB resistance markers is very small.

However, it is interesting that similar sequences are found in chromosomes 4, 6 and 8;

all of which have been associated with CBB resistance in various studies.

Page 120: Identification and Characterization of Common Bacterial

105

2.6 Conclusion

Candidate common bacterial blight (CBB) resistance genes CBB1-4 were

identified in the sequence of OAC Rex BiBAC2 library clones associated with CBB

resistance QTL marker PV-ctt001. Characterization of the candidate genes resulted in

the prediction of amino acid sequences which had leucine rich repeat (LRR)

characteristics including a consensus LRR motif sequence (LxxxLxLxLxxL). BLAST

comparisons of CG sequences to CBB susceptible common bean variety G19833

genome sequence identified similarities to retrotransposons in chromosome 6 and 8.

Protein sequence BLAST and secondary structure modelling indicated low similarity

between CGs and retrotransposon proteins. Characterization of gene predictions in the

sequence surrounding CGs CBB1-4 identified gene 5 as an ascorbate oxidase, which

has potential pathogen resistance implications.

Page 121: Identification and Characterization of Common Bacterial

106

Chapter 3: Detached leaf Xap susceptibility assay in Arabidopsis thaliana

3.1 Abstract

Efforts are currently underway to identify CGs for CBB resistance in common

bean and test them. As an alternative to transformation of common bean to observe

CG efficacy, a testing method using a model system would be beneficial. Here we

describe methods for inoculating detached Arabidopsis rosette leaves with different Xap

isolates, maintenance of the leaves in a tissue culture system over 192 hours, and

assaying disease progression in the inoculated tissue. Symptom progression and

assays of colony forming units (CFUs)/ unit of leaf tissue from inoculated leaf tissue

samples indicated that one isolate (Xap18) is more aggressive on Arabidopsis leaves

than other isolates. However, significant variation in establishing a disease state with

this system suggests that optimization of the protocol is needed to obtain consistent and

reproducible results.

Page 122: Identification and Characterization of Common Bacterial

107

3.2 Introduction

3.2.1 Common Bean Improvement

Breeding for the improvement crop species is a widely used tactic, especially in

the case of disease resistance. A major focus of common bean improvement has been

resistance to the bacterial disease common bacterial blight (CBB). CBB affects

worldwide production of P. vulgaris and is caused by the bacterial pathogen Xap and

fuscans variant Xff (Vauterin et al. 1995). Since little natural CBB resistance is

exhibited in common bean, introgression of resistance has occurred primarily by

interspecific crosses with resistant lines of P. acutifolius (Parker, 1985) and P.

coccineus (Park and Dhanvantari 1987; Miklas et al. 1994). Introgression of loci

associated with resistance into common bean lines has been achieved through

backcrossing, disease screening and marker assisted selection (MAS) and has

resulted in the development of CBB-resistant common bean varieties, including the

navy bean OAC Rex (Michaels et al. 2006). However, the use of markers for selecting

for resistance is limited by the possibility of the gene for resistance segregating from the

marker, creating the potential for false positives. The identification of the gene(s) that

confers resistance to CBB in common bean and subsequent development of a gene-

based marker would reduce the testing time and the frequency of false positives

involved in developing new disease resistant varieties.

3.2.2 CBB in A. thaliana

To test the involvement of a particular candidate gene in CBB resistance, a

testing method for the gene(s) in planta is desired. However, common bean has proven

to be a difficult species to transform (Gepts et al. 2008). This makes the use of gene

Page 123: Identification and Characterization of Common Bacterial

108

transformation to test CGs in susceptible varieties a difficult, if not impossible strategy,

and necessitates the use of other means of testing CGs for efficacy. Although common

bean is the primary host of CBB, a study by Perry (2010) indicated that the model plant

A. thaliana can be manually inoculated with Xap to produce CBB-like symptoms.

Chlorotic lesions appeared on Arabidopsis leaves inoculated with a mixture of Xap

isolates by 96 hours post inoculation (HPI). The lesions expanded and caused leaf

curling and eventually considerable tissue damage by 192 HPI. Symptoms also began

to appear around the margins of non-inoculated leaves in proximity to inoculated leaves,

which developed into symptoms similar to those of inoculated leaves. The

demonstration of Arabidopsis as a host makes it possible to test candidate CBB

resistance genes in the model system. By first transforming Arabidopsis with genes of

interest, the transformed plants can then be inoculated with Xap, and tested for

symptoms.

Although inoculation of leaves while still on the plant was shown to produce CBB-

like symptoms on Arabidopsis, for the purpose of testing transformed Arabidopsis

plants, a more precise, sterile and contained environment is preferred. The goal of this

study was to develop a tissue culture system to maintain detached rosette leaves post-

inoculation for the purpose of observing disease progression. This approach has the

potential to reduce the damage to the leaves and plant, as well as the likelihood of

introducing contaminating diseases during inoculation. This will help to improve the

ability to observe CBB symptom progression in Arabidopsis plants transformed with

candidate CBB resistance genes.

Page 124: Identification and Characterization of Common Bacterial

109

3.3 Materials and Methods

3.3.1 Bacterial and Plant Growth

A. thaliana ecotype Columbia seeds were suspended in distilled water at

approximately 50 seeds/mL and dispersed into 24-cell flats containing LA4 (Green

Island Distributors, Riverhead, N.Y.) that had been pre-soaked with deionized water and

drained. Flats were covered with clear plastic domes and maintained in a growth

chamber providing 16 h light (150 µmol/m2/s), 8 h dark on a daily cycle at 22 ºC.

Seeds from Phaseolus vulgaris varieties Nautica (susceptible) and OAC Rex

(resistant) were inserted in 6” round pots filled with LA4 and covered with a 1” layer of

medium vermiculite (Green Island Distributors, Riverhead, N.Y.). Pots were soaked with

deionized water and maintained in a growth chamber providing 16 h light and 8 h dark

cycle at 22oC.

Bacterial inoculum was generated by spreading 50 mL of Xap isolates 12, 18, 98

and 118, separately, on yeast salt 1% agar media plates and maintaining them at 28 ºC

for 48 h. Xap cultures were sub-cultured as streak plates for each isolate, and were

incubated for an additional 48 h at 28oC. The resultant colonies were then washed off

the plates into a 50 mL Falcon tube with 20 mL of distilled sterile water. The cell density

of the plate rinse was determined and diluted with distilled sterile water to an OD600 of

0.6 (1x107 colony forming units (CFU)/g) as needed.

3.3.2 Plant Wounding and Inoculation

Healthy rosette leaves were detached from 3-4 week old Arabidopsis plants

using a scalpel so that approximately 2 cm of petiole remained attached to the leaf.

Healthy P. vulgaris leaflets were detached from 3-4 week old plants. Inoculation was

Page 125: Identification and Characterization of Common Bacterial

110

performed by wounding each leaf or leaflet with a 10 µl plastic pipette tip once, close to

the mid-vein, and ejecting 10 µl of inoculum onto the wound site. Wound sites were

inoculated with 10 µl of either sterile distilled water or one of the four Xap isolates

(Xap12, Xap18, Xap98, Xap118) collected in Ontario (Park and Dhanvantari 1987). The

liquid inoculum was left on the wound site of each inoculated leaf to ensure infection.

Leaves representative of the 0 HPI were carefully transferred to a light blue background,

pictures were taken for IA, and then a sample of 1 cm2 of the tissue surrounding the

inoculation site was cut from the leaf.

Additional inoculated Arabidopsis, Nautica and OAC Rex leaves were placed in

tissue culture for further observations at 0, 48, 96 and 144 HPI. The Arabidopsis leaves

were placed in petri plates containing 0.7% water agar media. Both bean varieties were

placed in PhytatrayTM II boxes (Sigma-Aldrich Col. LLC) containing 0.7% water agar.

For both Arabidopsis and P. vulgaris, the leaves were kept upright at a 45o angle by

inserting the petiole into the solid media (Chen et al. 2006) to prevent the leaf blade

from contacting the agar. Phytatrays and petri plates were wrapped with Parafilm to

maintain a high relative humidity and placed into a 28 ºC tissue culture growth cabinet

set to 16 h light (150 µmol/m2/s) and 8 h dark in a randomized complete block design

where each experiment was a block. At 48 hour intervals post inoculation, pictures and

~1cm leaf samples were taken as described above, for assessment of disease symptom

progression over time.

Page 126: Identification and Characterization of Common Bacterial

111

3.3.3 Assessment of Disease Symptoms

3.3.3.1 Image Collection and Analysis

Disease symptom assessment was accomplished by performing Image Analysis

(IA) on inoculated leaves using the APS Assess 2.0 computer program (The American

Phytopathological Society, St. Paul, MN). Image collection of the inoculated common

bean leaves and subsequent IA was performed as described by Xie et al. (2012) to

determine the percent of the leaf area showing symptoms. The same procedure was

followed for inoculated Arabidopsis leaves. However the distance of the camera from

the Arabidopsis leaves was adjusted to 20 cm, to reflect the smaller size of the

Arabidopsis leaves. Manual IA was performed to determine the area of chlorosis and/or

necrosis for inoculated leaves at 0, 48, 96,144 and 192 HPI. Although there were no

symptoms at 0 HPI, pictures were taken to provide a visual comparison for later time

points when the symptoms had developed.

3.3.3.2 CFU Analysis

Samples of approximately 1 cm2 of the tissue surrounding the inoculation site

were taken from each leaf and homogenized in 0.9 ml of sterile distilled water (Perry,

2010). Serial dilutions of 103-fold and 105 -fold were made for samples taken at 0 HPI

for all inoculum types. The dilutions were plated on XCP media (10 g/L KBr, 0.25 g/L

CaCl2, 10 g/L potato starch, 10 g/L peptone, 10 ml/L Tween 80, 3 mg/L fluorouracil and

0.16 mg/L tobramycin pH 7.0; Remeeus and Sheppard 2006) for a final dilution of 105

and 107, and plates were incubated at 28oC for 48 hours. For samples taken at 48,

96,144 and 192 HPI, 105 and 107 fold serial dilutions were made for all inoculum types.

The dilutions were plated on XCP media plates for a final dilution of 107 and 109 and

Page 127: Identification and Characterization of Common Bacterial

112

incubated at 28oC for 48 hours. Visible colonies >2 mm in diameter were marked at 48

hours, and plates were incubated for an additional 96 hours. Colonies with visible

media clearing were then counted to provide the CFU value for the corresponding

inoculum type and time post inoculation.

3.3.4 Statistical Analyses

Variance analysis, lsmeans estimates and comparisons were performed on IA

lesion percent values and CFUs from a single replication of the experriment using

PROC MIXED in SAS 9.4 (SAS institute Inc., Cary, NC). Isolate, genotype and time

were all considered fixed effects. Tests for normality prompted log transformation of the

data, however further analysis was performed using untransformed data since

transformation did not improve the normality of the residuals. The data was sliced to

separate the effects of the different isolates, plant genotypes and time points post

inoculation on the interaction of all factors. ANOVA was performed to determine the

significance of the sliced effects. Letter grouping to show the significant differences

between lsmeans was performed using the PDMIX800 macro (Saxton 2000) with data

sliced for time to make comparisons between all lsmeans at the different time points.

For all analyses, an interaction was considered significant when the p-value was less

than 0.05.

Page 128: Identification and Characterization of Common Bacterial

113

3.4 Results

3.4.1 Analysis of Variance

Analysis of variance for OAC Rex and Nautica CFU and IA results showed that

the inoculum, genotype effects were significant (Appendix 5). Interactions between

inoculum and genotype, inoculum and time and genotype and time were also significant

(Appendix 5). Interaction between inoculum, genotype and time was not significant

(Appendix 5). Analysis of variance for Arabidopsis CFU and IA results showed that

inoculum and time effects were significant, and that the interaction between inoculum

and time was significant (Appendix 6).

3.4.2 Visual Observations and IA

3.4.2.1 Common Bean Visible Leaf Symptoms

Nautica and OAC Rex leaves inoculated with water showed no visible disease

symptoms at all sampling time-points, and lsmeans of percent of leaf with disease

symptoms (hereafter known as symptoms) showed no significant differences between

time points or variety (Appendix 7-9; Figure 3.1A; Tables 3.1-3.2 and 3.8-3.9). Xap

inoculated leaves from both common bean varieties were not significantly different for

inoculation with any of the four isolates or water at 0, 48 and 96 HPI (Appendix 7-9;

Figure 3.1B-E). At 144 HPI, significant differences were seen between the symptoms

seen in Nautica and OAC Rex leaves when inoculated with Xap isolates (Figure 3.1;

Table 3.1). OAC Rex leaves showed no significant difference between water and Xap

inoculation (Figure 3.1; Table 3.1). Xap and water inoculated Nautica leaf symptoms

were significantly different when isolates Xap98 and Xap118 were used, with the most

severe symptom being observed in the Xap98 inoculated sample (36.21%) (Figure 3.1

Page 129: Identification and Characterization of Common Bacterial

114

Figure 3.1. Symptom progression in inoculated P. vulgaris leaves. Percent leaf area showing symptoms at 0, 48, 96, 144

and 192 hours post inoculation (HPI). Inoculations were performed on leaves of susceptible (Nautica – dark colour) and

resistant (OAC Rex - light) varieties with water (A - Black) or one of four X. axonopodis pv. phaseoli (Xap) isolates; Xap12

(B - Red), Xap18 (C - Purple), Xap98 (D - Blue) and Xap118 (E - Green). Error bars represent the standard error of the

values of the lsmeans (+ 6.18).

0

10

20

30

40

50

60

70

80

0 48 96 144 192

Sym

pto

ms (%

Leaf

Are

a)

Hours Post-Inoculation (HPI)

A) WATER

0

10

20

30

40

50

60

70

80

0 48 96 144 192

Sym

pto

ms (%

Leaf

Are

a)

Hours Post-Inoculation (HPI)

B) Xap12

0

10

20

30

40

50

60

70

80

0 48 96 144 192

Sm

pto

ms (%

Leaf A

rea)

Hours Post-Inculation (HPI)

C) Xap18

0

10

20

30

40

50

60

70

80

0 48 96 144 192

Sm

pto

ms (%

Leaf A

rea)

Hours Post-Inoculation (HPI)

D) Xap98

0

10

20

30

40

50

60

70

80

0 48 96 144 192

Sym

pto

ms

(% L

eaf

Are

a)

Hours Post-Inoculation (HPI)

E) Xap118

Page 130: Identification and Characterization of Common Bacterial

115

A, D, E; Table 3.1). At the last time point (192 HPI) there were significant differences

between Nautica and OAC Rex for inoculation with each Xap isolate (Figure 3.1; Table

3.2). Inoculation of OAC Rex leaves, resulted in symptoms that were significantly

different from water inoculation only when isolate Xap118 was used (31.65%) (Figure

3.1 A, E; Table 3.2). All isolates produced symptoms greater than water inoculation in

Nautica leaves, with the highest percentage of affected leaf being observed in the

samples inoculated with Xap118 (66.71%) (Figure 3.1 A-E; Table 3.2).

3.4.2.2 Arabidopsis Visible Leaf Symptoms

Arabidopsis leaves inoculated with water showed no visible disease symptoms at

all sampling time-points, and lsmeans of symptoms showed no significant differences

between time points (Appendix 7-9; Figure 3.2A; Tables 3.1-3.2 and 3.7). Xap

inoculated Arabidopsis leaves were not significantly different than the water controls or

each other with any of the four isolates at 0, 48 and 96 HPI (Appendix 7-9; Figure 3.2;

Table 3.7). At 144 HPI, all Xap isolates produced symptoms that were significantly

higher than the water control in Arabidopsis leaves, with the most severe symptoms

being produced by Xap98 (24.50%) (Figure 3.2; Table 3.1). However, at 192 HPI, the

symptoms in Arabidopsis leaves that were induced by Xap98 (2.10%) were not

significantly different than in the water control (Figure 3.2; Table 3.2). Symptoms

induced by the other isolates at 192 HPI were significantly higher than the water control

for the other isolates with the highest mean symptoms being produced by Xap18

(19.30%) (Figure 3.2; Table 3.2).

Page 131: Identification and Characterization of Common Bacterial

116

Figure 3.2. Symptom progression in inoculated A. thaliana rosette leaves. Percent leaf

area showing symptoms for Arabidopsis rosette leaves at 0, 48, 96, 144 and 192 hours

post inoculation (HPI). Inoculations were performed using water (black) as a control,

and one of four X. axonopodis pv. phaseoli (Xap) isolates Xap12 (red) Xap18 (purple)

Xap98 (blue) and Xap118 (green). Error bars represent the standard error of the values

of the lsmeans (+ 3.06).

0

5

10

15

20

25

30

0 48 96 144 192

Sym

pto

ms (%

Leaf

Are

a)

Hours Post-Inoculation (HPI)

Page 132: Identification and Characterization of Common Bacterial

117

Table 3.1. LSmeans for common bacterial blight symptoms on OAC Rex, Nautica and Arabidopsis leaves at 144 hours post inoculation (HPI). Percent of leaf showing disease symptoms for OAC Rex, Nautica and Arabidopsis inoculated with water, Xap12, Xap18, Xap98 and Xap118 at 144 HPI

Symptoms (%)

Inoculum

Genotype Water Xap12 Xap18 Xap98 Xap118 Genotype Mean

OAC Rex 1.87 XZ

c 2.72 X

c 3.14 X

c 4.72 X

c 7.90 X

c 4.07 WY

B

Nautica 1.88 X c 3.78

X c 6.38

X c 36.2

X a 14.8

X bc 12.6

W A

Arabidopsis 1.68V x 16.6

V yz 13.9

V y 24.5

V z 9.79

V xz 13.3

U

ZLsmeans followed by the same lower case letter were not significantly different at p=0.05

YLsmeans followed by the same upper case letter were not significantly different at p=0.05

XDenotes lsmeans with a standard error of +127.80

WDenotes lsmeans with a standard error of +57.1532

VDenotes lsmeans with a standard error of +3.06

UDenotes lsmeans with a standard error of +1.37

Table 3.2. LSmeans for common bacterial blight symptoms on OAC Rex, Nautica and Arabidopsis leaves at 192 hours post inoculation (HPI). Percent of leaf showing disease symptoms for OAC Rex, Nautica and Arabidopsis inoculated with water, Xap12, Xap18, Xap98 and Xap118 at 192 HPI

Symptoms (%)

Inoculum

Genotype Water Xap12 Xap18 Xap98 Xap118 Genotype Mean

OAC Rex 1.62YZ

f 10.7Y

def 2.69Y f 5.27

Y ef 31.7

Y bc 10.4

X B

Nautica 7.07Y def 20.4

Y cd 57.1

Y a 38.7

Y b 66.7

Y a 38.0

X A

Arabidopsis 1.04 V

x 9.3 V xy 19.3

V z 2.10

V x 12.6

V yz 8.87

U

ZLsmeans followed by the same lower case letter were not significantly different at p=0.05

YLsmeans followed by the same upper case letter were not significantly different at p=0.05

XDenotes lsmeans with a standard error of +127.80

WDenotes lsmeans with a standard error of +57.1532

VDenotes lsmeans with a standard error of +3.06

UDenotes lsmeans with a standard error of +1.37

Page 133: Identification and Characterization of Common Bacterial

118

3.4.3 Leaf Sample CFUs

3.4.3.1 Common Bean CFUs

CFUs produced from leaf samples taken from water-inoculated Nautica and OAC

Rex leaves were not significantly different at all time-points (Appendix 10; Figure 3.3 A;

Table 3.3-3.6). There were also no significant differences between CFU produced from

either bean genotype inoculated with water or the four Xap isolates at 0 HPI (Appendix

10; Figure 3.3 B-E; Table 3.3). At 48, 96, 144 and 192 HPI, there were significant

differences between the CFUs for OAC Rex and Nautica leaf samples inoculated Xap

isolates (Figure 3.3; Table 3.3-3.6). At 48 HPI, OAC Rex CFU means were significantly

different from water when Xap18 and Xap98 were used, with the highest mean being

produced by Xap18 (7.64 x 107 CFU) (Figure 3.3 C-D; Table 3.3). All isolates

produced CFU means that were significantly different from water in Nautica leaves, with

the highest being produced by inoculation with Xap18 (1.07 x 108 CFU) (Figure 3.3;

Table 3.3). At 96 HPI, OAC Rex CFU means were significantly different from water

inoculation when Xap18 and Xap98 were used, with the highest mean being produced

by Xap18 (7.03 x 107 CFU) (Figure 3.3 C-D; Table 3.4). Inoculation of Nautica leaves

by all Xap isolates produced CFUs that were significantly different from water

inoculation, with the highest mean being produced by Xap118 (1.10 x 108 CFU) (Figure

3.3; Table 3.4). At 144 HPI, OAC Rex leaf samples produced significantly more CFU

than the water control only when inoculated with Xap18 (4.57 x 107 CFU) (Figure 3.3 C;

Table 3.5). All isolates produced more CFUs than the water inoculation in the Nautica

leaf samples, with the highest mean being produced by Xap18 (1.18 x 108 CFU) (Figure

3.3; Table 3.5). At 192 HPI all isolates produced CFUs that were significantly higher

Page 134: Identification and Characterization of Common Bacterial

119

Figure 3.3. Pathogen growth in inoculated P. vulgaris leaf samples. Colony forming units (CFU) for leaf samples taken

from susceptible (Nautica -dark) and resistant (OAC Rex - light) varieties, respectively at 0, 48, 96, 144 and 192 hours

post inoculation (HPI). Inoculation was performed with water (A - Black) or one of four X. axonopodis pv. phaseoli (Xap)

isolates; Xap12 (B - Red) Xap18 (C - Purple) Xap98 (D - Blue) and Xap118 (E - Green). Error bars represent the

standard error of the values of the lsmeans (+ 142).

0

200

400

600

800

1000

1200

1400

0 48 96 144 192Colo

ny F

orm

ing U

nits (

CF

U x

10

5)

Hours Post-Inoculation (HPI)

A) WATER

0

200

400

600

800

1000

1200

1400

0 48 96 144 192Colo

ny F

orm

ing U

nits (

CF

U x

10

5)

Hours Post-Inoculation (HPI)

B) Xap12

0

200

400

600

800

1000

1200

1400

0 48 96 144 192Colo

ny F

orm

ing U

nit (

CF

U x

10

5)

Hours Post-Inoculation (HPI)

C) Xap18

0

200

400

600

800

1000

1200

1400

0 48 96 144 192Colo

ny F

orm

ing U

nits (

CF

U x

10

5)

Hours Post-Inoculation (HPI)

D) Xap98

0

200

400

600

800

1000

1200

1400

0 48 96 144 192Colo

ny F

orm

ing U

nits (

CF

U x

10

5)

Hours Post-Inoculation (HPI)

E) Xap118

Page 135: Identification and Characterization of Common Bacterial

120

than observed for the water inoculated leaf samples for both OAC Rex and Nautica

(Figure 3.3; Table 3.6). The highest mean CFUs produced were produced by

inoculation with Xap18, which had 4.57 x 107 CFU and 1.18 x 108 CFU for OAC Rex

and Nautica, respectively (Figure 3.3 C; Table 3.6).

3.4.3.2 Arabidopsis CFUs

Arabidopsis leaves inoculated with water showed no visible disease symptoms at

all sampling time-points, and lsmeans of symptoms showed no significant differences

between time points (Appendix 10; Figure 3.4; Table 3.3-3.6). At 0 HPI, none of the

Xap isolates produced CFU means that were significantly different from water

inoculation (Appendix 10; Figure 3.4 B-E). At 48 HPI, only Xap98 produced a CFU

mean that was significantly higher than the water sample (4.76 x 107 CFU) (Figure 3.4

D; Table 3.3). At 96 HPI, all Xap isolates leaves produced CFU means that were

significantly higher than water inoculated leaves, with the highest mean being produced

by Xap18 (5.76 x 107 CFU) (Figure 3.4; Table 3.4). At 144 HPI, all Xap isolates

produced CFU means that were significantly higher than water inoculated leaves, with

the highest mean being produced by Xap18 (6.76 x 107 CFU) (Figure 3.4; Table 3.5). At

192 HPI, all Xap isolates leaves produced CFU means that were significantly higher

from water inoculated leaves, with the highest mean being produced by Xap18 (1.04 x

108 CFU) (Figure 3.4; Table 3.6).

Page 136: Identification and Characterization of Common Bacterial

121

Figure 3.4. Pathogen growth in inoculated A. thaliana leaf samples. Colony forming

units (CFU) for Arabidopsis leaves 0, 48, 96, 144 and 192 hours post inoculation (HPI).

Inoculation was performed with water (black), Xap12 (red), Xap18 (purple), Xap98

(blue), or Xap118 (green). Error bars represent the standard error of the values of the

lsmeans (+ 96.3).

0

200

400

600

800

1000

1200

0 48 96 144 192

Colo

ny F

orm

ing U

nits (

CF

U x

10

5)

Hours Post-Inoculation (HPI)

Page 137: Identification and Characterization of Common Bacterial

122

Table 3.3. LSmeans for colony forming units for X. axonopodis pv. phaseoli (Xap) bacteria at 48 hours post inoculation (HPI). CFUs were determined by incubating 105 dilutions of ground leaf samples taken from inoculated leaves from resistant (OAC Rex) and susceptible (Nautica) P. vulgaris varieties, and Arabidopsis. Samples were taken at 48 HPI after inoculation with Water, Xap12, Xap18, Xap98 or Xap118.

CFU (x 105)

Inoculum

Genotype Water Xap12 Xap18 Xap98 Xap118 Genotype Mean

OAC Rex 108 XZ

de 348 X cde 764

X ab 704

X abc 8.67

X e 387

WY B

Nautica 107 X de 676

X abc 1070

X a 920

X a 452

X bcd 645

W A

Arabidopsis 49.7 Y y 261

Y yz 307

Y yz 476

Y z 297

Y yz 278

W

ZLsmeans followed by the same lower case letter were not significantly different at p=0.05

YLsmeans followed by the same upper case letter were not significantly different at p=0.05

XDenotes lsmeans with a standard error of +142

WDenotes lsmeans with a standard error of +63.6

VDenotes lsmeans with a standard error of +96.3

UDenotes lsmeans with a standard error of +43.1

Table 3.4. LSmeans for colony forming units for X. axonopodis pv. phaseoli (Xap) bacteria at 96 hours post inoculation (HPI). CFUs were determined by incubating 105 dilutions of ground leaf samples taken from inoculated leaves from resistant (OAC Rex) and susceptible (Nautica) P. vulgaris varieties, and Arabidopsis. Samples were taken at 96 HPI after inoculation with Water, Xap12, Xap18, Xap98 or Xap118.

CFU (x 105)

Inoculum

Genotype Water Xap12 Xap18 Xap98 Xap118 Genotype Mean

OAC Rex 48.7 XZ

de 101 X de 703

X abc 445 cd 171

X de 294

WY B

Nautica 3.00 X e 667

X bc 1070

X a 957

X ab 1100

X a 761

W A

Arabidopsis 26.7 V x 206

V yz 567

V z 420

V yz 310

V yz 306

U

ZLsmeans followed by the same lower case letter were not significantly different at p=0.05

YLsmeans followed by the same upper case letter were not significantly different at p=0.05

XDenotes lsmeans with a standard error of +142

WDenotes lsmeans with a standard error of +63.6

VDenotes lsmeans with a standard error of +96.3

UDenotes lsmeans with a standard error of +43.1

Page 138: Identification and Characterization of Common Bacterial

123

Table 3.5. LSmeans for colony forming units for X. axonopodis pv. phaseoli (Xap) bacteria at 144 hours post inoculation (HPI). CFUs were determined by incubating 105 dilutions of ground leaf samples taken from inoculated leaves from resistant (OAC Rex) and susceptible (Nautica) P. vulgaris varieties, and Arabidopsis. Samples were taken at 144 HPI after inoculation with Water, Xap12, Xap18, Xap98 or Xap118.

CFU (x 105)

Inoculum

Genotype Water Xap12 Xap18 Xap98 Xap118 Genotype Mean

OAC Rex 57.7 XZ

c 183 X bc 457

X b 207

X bc 204

X bc 222

WY B

Nautica 0.333 X

bc 425 X b 1180

X a 1130

X a 1050

X a 756

W A

Arabidopsis 0V x 321

V y 667

V z 532

V yz 295

V y 363

U

ZLsmeans followed by the same lower case letter were not significantly different at p=0.05

YLsmeans followed by the same upper case letter were not significantly different at p=0.05

XDenotes lsmeans with a standard error of +142

WDenotes lsmeans with a standard error of +63.6

VDenotes lsmeans with a standard error of +96.3

UDenotes lsmeans with a standard error of +43.1

Table 3.6. LSmeans for colony forming units for X. axonopodis pv. phaseoli (Xap) bacteria at 192 hours post inoculation (HPI). CFUs were determined by incubating 105 dilutions of ground leaf samples taken from inoculated leaves from resistant (OAC Rex) and susceptible (Nautica) P. vulgaris varieties, and Arabidopsis. Samples were taken at 192 HPI after inoculation with Water, Xap12, Xap18, Xap98 or Xap118.

CFU (x 105)

Inoculum

Genotype Water Xap12 Xap18 Xap98 Xap118 Genotype Mean

OAC Rex 34.7 XZ

c 789 X ab 863

X ab 804

X ab 592

X b 617

WY A

Nautica 0X c 1100

X a 1180

X a 1070

X a 625

X b 794

Y A

Arabidopsis 0V w 319

V x 1040

V z 316

V x 660

V y 467

U

ZLsmeans followed by the same lower case letter were not significantly different at p=0.05

YLsmeans followed by the same upper case letter were not significantly different at p=0.05

XDenotes lsmeans with a standard error of +142

WDenotes lsmeans with a standard error of +63.6

VDenotes lsmeans with a standard error of +96.3

UDenotes lsmeans with a standard error of +43.1

Page 139: Identification and Characterization of Common Bacterial

124

Table 3.7. Inoculated A. thaliana leaf symptom progression. Arabidopsis rosette leaves were inoculated with sterile distilled water control and X. axonopodis pv. phaseoli (Xap) isolates 12, 18, 98 and 118. Pictures were taken at 0, 48, 96, 144 and 192 hours post inoculation (HPI)

Inoculum

Hours Post Inoculation

0 48 96 144 192

Water

Xap12

Xap18

Xap98

Xap118

Page 140: Identification and Characterization of Common Bacterial

125

Table 3.8. Inoculated common bacterial blight susceptible P. vulgaris variety leaf symptom progression. Trifoliate Nautica leaves were inoculated with sterile distilled water control and X. axonopodis pv. phaseoli (Xap) isolates 12, 18, 98 and 118. Pictures were taken at 0, 48, 96, 144 and 192 hours post inoculation (HPI)

Inoculum

Hours Post Inoculation

0 48 96 144 192

Water

Xap12

Xap18

Xap98

Xap118

Page 141: Identification and Characterization of Common Bacterial

126

Table 3.9. Inoculated common bacterial blight resistant P. vulgaris variety leaf symptom

progression. Trifoliate OAC Rex leaves were inoculated with sterile distilled water

control and X. axonopodis pv. phaseoli (Xap) isolates 12, 18, 98 and 118. Pictures

were taken at 0, 48, 96, 144 and 192 hours post inoculation (HPI)

Inoculum

Hours Post Inoculation

0 48 96 144 192

Water

Xap12

Xap18

Xap98

Xap118

Page 142: Identification and Characterization of Common Bacterial

127

3.5 Discussion

3.5.1 Symptom Development

3.5.1.1 Common Bean Symptom Evaluation by IA

The highest disease symptoms were generally produced in Nautica, with

significant variability caused by the isolate used. Nautica was included in this study as

a Xap susceptible common bean check variety. Therefore, the increased symptoms

observed in Nautica leaves compared to OAC Rex leaves inoculated with the same Xap

isolate was consistent with the expectation that this variety would show more severe

symptoms in a shorter time after inoculation. Although a direct comparison is not

possible due to the differences between experimental parameters and the resistant

varieties used, a comparison of the symptom severity scores observed in the current

study with detached leaves to those reported for a similar growth chamber study (Xie

et al. 2012) with intact plants, suggests that the disease reaction was more severe in

the current detached leaf assay. In the previous study, the percent lesion area was

6.03% for Nautica and 0.22-0.29% for the resistant varieties (MBE7, Rexeter and Apex)

14-18 days post inoculation (336-432 HPI) (Xie et al.2012). In contrast, Nautica and

OAC Rex symptoms covered 38.0% and 10.4% of the leaves, respectively, at 192 HPI

in the current study (Table 3.2).

There are several potential explanations for the observed differences between

the results reported by Xie et al. (2012) and those obtained in the current study. The

most immediate explanation is that the environment provided by the tissue culture is

more suited for symptom development. However, it is also possible that there are

differences in the reaction to pathogens in detached versus attached leaves. This was

Page 143: Identification and Characterization of Common Bacterial

128

observed previously in Arabidopsis by Liu et al. (2007b), when attached and detached

leaves were inoculated with either Colletotrichum linicola A1 or C. higginsianum,

resulting in opposite responses. Inoculation with C. linicola A1 did not affect attached

leaves, but it did infect detached leaves. In contrast, inoculation with C. higginsianum

did affect attached leaves, but not when they were detached.

Other possible causes for the increased symptom severity in the detached leaf

assay are differences in the sizes and developmental stages of the leaves. In both the

common bean study performed by Xie et al. (2012) and the Arabidopsis study

performed by Perry (2010), inoculated leaves remained on the plant for the duration of

the experiment. This allowed the leaves to continue to grow, whereas, the detached

leaves in the current study remained approximately the same size, which would affect

the relative symptom size. Since the media used in this study lacked any additional

nutrients, the detached leaves may have had an impaired ability to produce a resistance

response due to lack of resources.

In the current study, trifoliate leaves with different sizes and developmental

stages were used, whereas, Xie et al. (2012) used only unifoliate leaves. Not only were

the trifoliate leaves smaller than the unifoliate leaves, some of the leaves used for

inoculation were quite new, which may have made them more susceptible to Xap

infection than the unifoliate leaves. This is consistent with the study performed by Jung

et al. (1997), which found that there was a significant strain by organ interaction, and

that certain QTL only contributed to resistance in specific organs and developmental

stages. The Random Amplified Polymorphic DNA (RAPD) marker D13.1000, in

Page 144: Identification and Characterization of Common Bacterial

129

particular, was only associated with resistance to Xap isolate DR-7 in the first trifoliate

leaves, and not in later trifoliate leaves, pods or seeds.

3.5.1.2 Arabidopsis Symptom Evaluation by IA

Xap induced symptoms on Arabidopsis leaves were variable compared to those

induced on common bean leaves of both varieties. The mean symptom percent for

Arabidopsis was higher than the means for both Nautica and OAC Rex at 144 HPI,

indicating that it’s response to Xap inoculation was similar to the susceptible Nautica

(Table 3.1). This contrasted to the symptom severity observed in Arabidopsis leaves

at 192 HPI, which was lower than that observed for Nautica or OAC Rex (Table 3.2).

This change from a susceptible symptom mean to a resistant one, over time, is likely

caused by the changes observed in two isolates in particular. The symptoms produced

by isolates Xap12 and Xap98 were the highest at 144 HPI, and then dropped

significantly at 192 HPI (Figure 3.1 B, D; Table 3.1-3.2). However, the symptoms

produced by Xap18 and Xap118 increased from 144 to 192 HPI (Figure 3.1 C, E; Table

3.1-3.2).

This change from high to low symptoms in Xap12 and Xap98 inoculated

Arabidopsis leaves may be a reflection of the ability of these isolates to infect

Arabidopsis. It is also possible that the conditions of those detached leaves were

particularly conducive to disease progression such as increased humidity or light

availability, or that contamination from another pathogen occurred. However, since

only one set of assays was performed, and the results are not entirely consistent with

those observed in the study by Perry (2010), more repetitions of the experiment are

needed to determine if this is a significant trend, or a random occurrence.

Page 145: Identification and Characterization of Common Bacterial

130

3.5.1.3 Symptom Evaluation of CFUs

When compared to the collected IA data, the CFU means were more variable

than the data based on symptom development. However, all genotypes showed

significant differences between Xap- inoculation means and water inoculation means as

early as 48 HPI. Additionally, Nautica generally had the highest CFU values compared

to OAC Rex, which is consistent with the trends observed by IA. This supports the

assumption that the symptoms are caused by the proliferation of the pathogen in the

tissue.

Nevertheless, some inconsistencies like the observation that the highest CFU

values were induced by Xap18 on all three leaf types. Although the highest symptoms

observed in Arabidopsis were also produced by inoculation with isolate Xap18, the

highest values induced in leaves from both common bean varieties were from

inoculation with isolate Xap118. The difference in common bean and Arabidopsis

symptom development in response to inoculation with Xap18 is further emphasized by

the lack of difference between the symptoms produced in OAC Rex leaves inoculated

by water and Xap18. It is also interesting to note that leaves from all genotypes

inoculated with Xap18 tended to be a lighter green than leaves inoculated with other

isolates (Figures 3.7-3.9), which indicates that Xap18 affects a larger area within the

leaf, instead of causing localized lesions. This suggests that Xap18 symptoms included

more than localized lesions, and that IA didn’t entirely capture Xap18 symptoms

because the IA parameters measured lesions (chlorosis and necrosis), and the lighter

green colour was not taken into account. These symptom and CFU differences

Page 146: Identification and Characterization of Common Bacterial

131

suggest that differences in the interaction between the genotype and Xap isolate are at

play.

The variability in Xanthomonas host specificity, both within and between species

and pathovars, has been well documented (Rademaker et al. 2005: Sun et al. 2006;

Bogdanove et al. 2011), as well as specifically for the Xap-P. vulgaris interaction

(Schuster et al. 1983; Mkandawire et al. 2004; Duncan et al. 2010; Fourie and

Herselman 2011). Within the four Xap isolates used in Ontario breeding efforts, isolates

Xap98 and Xap118 have been observed to be the most virulent on common bean

varieties. It is likely that human error is responsible for the discrepancy between the

Xap18 IA and CFU results observed in both common bean genotypes. Just as the IA

results may be lower than they should be (discussed above), the CFUs may be higher

than actual due to inexperience in counting. Since the IA and CFU means for Xap18

inoculated Arabidopsis leaves progressed in a similar way (Figure 3.2; Figure 3.4), it is

most likely that the suggested higher virulence of this strain on Arabidopsis is reliable.

However, replication of the study is required to determine the significance of these

results.

3.5.2 Suitability of Detached Arabidopsis Leaf CBB Assay

3.5.2.1 Variability in CFU Observations

Considering the variability of the CFU counts over the course of the study, it must

be concluded that CFU count cannot be taken as an independent measure of isolate

virulence. However, the study has limited scope due to the failure to replicate the

results using a single isolate (Xap12) (data not shown). Additional studies with different

parameters might reduce the variability in the results by providing more data, and

Page 147: Identification and Characterization of Common Bacterial

132

reducing human error. Regardless of the limitations of the CFU method to compare

virulence between isolates within the current study, it does demonstrate that Xap

accumulates in inoculated leaf samples over time. Therefore, when combined with IA

values, CFU values can be used as an indicator of proliferation within the inoculated

sample.

The failure of the experiments with detached Arabidopsis leaves and a single

Xap isolate (Xap12) to replicate the symptoms observed with attached Arabidopsis

leaves may have been caused by the isolate selection. This isolate initially produced

symptoms in Arabidopsis rosette leaves that were visibly similar to those observed by

Perry (2010) at 144 HPI (Table 7). However, these symptoms have not been observed

in subsequent experiment repetitions, and statistical analysis of the initial results

showed that Xap12 produces low IA symptom means and CFU counts in Arabidopsis,

as well as OAC Rex and Nautica, in comparison to other isolates (Figure 3.1-3.4).

A more appropriate choice of Xap isolates for further replications of this study

would be Xap18, which was shown to produce the highest symptoms in Arabidopsis at

192 HPI (19.3%). This isolate also had a more consistent increase in symptoms and

CFU means over time compared to Xap12. Symptoms produced by Xap12 increased

quickly and then dropped, whereas Xap18 inoculation produced a more steady increase

in symptoms (Figure 3.2). This was even more evident with the CFU means, where

Xap18 CFU means increased at a relatively steady rate over time, whereas Xap12 CFU

means fluctuated around an apparent plateau from 48 to 192 HPI (Figure 3.4).

Page 148: Identification and Characterization of Common Bacterial

133

3.5.2.2 Detached Leaf vs. Attached Leaf Symptom Progression

The symptom progression in detached Arabidopsis leaves did not follow the

pattern of symptom development that was observed in inoculated leaves on the plant,

where chlorotic symptoms developed around 96 HPI and increased to the point of

considerable damage by 192 HPI (Perry 2010). However, the much slower

development of symptoms in the detached leaves, which were not significantly different

from water until 144 HPI, and varied in severity at 192 HPI depending on which isolate

was used, may have been due to the amount of inoculum and delivery method used in

this study compared to the one performed by Perry (2010). This study inoculated the

detached leaves with 10 µl of inoculum, with a directed injury to the adaxial side of the

leaf, where the inoculum was left on top of the injury location. This contrasts with the

~200 µl of inoculum used in the previous study, which was forced into the intercellular

matrix by infiltration on the abaxial side of the leaf. The use of the infiltration method

would likely have caused injury not only to the site of inoculation, but the region

surrounding it, causing artificially high symptoms compared to the single small injury

location used in this study.

The increased damage potential and resultant effect on symptom development of

the infiltration method on attached leaves was one of the motivations for the

development of the detached leaf tissue culture method. It also provided a more

controllable environment, which was hypothesized to decrease the variability and

increase reproducibility of the results, as was observed by Chen et al. (2006) in their

detached leaf assay for observing Fusarium graminearum infection in Arabidopsis.

However, Chen et al. (2006) required the use of an external disease factor

Page 149: Identification and Characterization of Common Bacterial

134

deoxynivalenol (DON) to provide consistency to the symptoms caused by wounding and

inoculation with F. graminearum, and a similar option is not available for Xap

inoculation.

A possible combination of the infiltration and detached leaf methods would be to

inoculate detached Arabidopsis leaves with Xap isolates using the infiltration method. A

similar technique was reported for peach detached leaves by Randhawa and Civerolo

(1985). In their study, they used paper towels to cushion and support detached peach

leaves while they were infiltrated with Xanthomonas campestris pv. pruni. They

reported that this resulted in little damage to the leaves, suggesting that infiltration of

detached Arabidopsis leaves may be possible with the use of cushioning such as

sterilized filter paper.

Another possible explanation of the differences in symptoms seen in the previous

study (Perry 2010) and the current study is that the latter used a mixture of all four

isolates to inoculate the leaves on the plant, whereas, the current study used individual

isolates. The use of the bacterial mixture may have resulted in exposure of the leaf

tissue to several effectors, simultaneously, thus enabling infection of Arabidopsis, even

though it is not a natural host. This has been documented before in literature, such as

the infection of Arabidopsis by non-host Pseudomonas syringae pv. tomato DC3000 (Li

et al. 2005), where the tomato pathovar overcomes non-host defense by the secretion

of multiple effectors.

Page 150: Identification and Characterization of Common Bacterial

135

3.6 Conclusion

The use of a detached leaf assay indicate the susceptibility of A. thalianarosette

leaves inoculated with Xap isolates could be an important tool when testing CBB

resistance genes in the model plant system. This study demonstrated that CBB

symptoms and increased Xap CFUs accompany the inoculation of detached

Arabidopsis and common bean leaves with Xap isolates at 192 HPI. However, variability

in disease response compared to previous common bean and Arabidopsis Xap

inoculation studies prevent conclusions to be made about the efficacy of this bioassay.

More work needs to be performed to optimize this method.

Page 151: Identification and Characterization of Common Bacterial

136

Chapter 4: Candidate Common Bacterial Blight Resistance Gene Expression in a

Resistant (OAC Rex) and Susceptible (Seaforth) Variety of Phaseolus vulgaris

4.1 Abstract

Resistance to common bacterial blight, caused by Xanthomonas axonopodis pv.

phaseoli (Xap) is a quantitative trait in common bean (P. vulgaris), introduced from

tepary bean (P. acutifolius) and selected for in common bean to produce resistant

varieties such as OAC Rex (Michaels et al. 2006). Although selection for QTLs by their

associated markers has been successful in producing resistant varieties of common

bean, direct selection for a gene underlying the QTL would provide greater speed and

power of selection. This study aimed to measure the expression of candidate CBB R

genes CBB2, CBB3 and CBB4 in resistant (OAC Rex) and susceptible (Seaforth)

varieties after inoculation with water (control) or Xap, for up to 168 HPI. To measure

expression, leaf samples were collected at 0, 12, 24, 36, 48, 72, 96, 120, and 168 HPI,

RNA was isolated and cDNA synthesized via RT-PCR was examined by gel

electrophoresis. Expression of CGs was not observed consistently, suggesting that

none of the CGs are involved in resistance to CBB, or that there were some limitations

to the experimental protocol that prevented expression of the CGs to be measured.

Further studies are needed to confirm these results, and to identify and test more CGs.

Page 152: Identification and Characterization of Common Bacterial

137

4.2 Introduction

4.2.1 Significance of Common Bean

Common dry bean (P. vulgaris) is a widely grown legume crop species, which is

produced in Canada for its dry grain and fresh pods. Edible dry beans are a

recommended part of a healthy human diet (Health Canada, www.hc-sc.gc.ca)

because they contain high levels of protein (Yañez et al. 1995), dietary fibre and

complex carbohydrates, vitamins such as folate, and minerals (Tosh and Yada 2010).

They are also low in fats and high in resistant starches. In Canada, common bean is

grown in Ontario, Manitoba, and Alberta (Goodwin 2005). Most of the production is

primarily exported to the US and Europe, as only a small portion is consumed in

Canada.

Common bean production is restricted due to a variety of abiotic factors such as

frost, humidity and soil quality as well as biotic factors, in particular weeds and diseases

(Goodwin 2005). A combination of management practices and variety selection are

employed to reduce the effect of these factors on yield quality and quantity, however

there is room for improvement, especially for increased disease resistance.

4.2.2 CBB in Common Bean

One of the major diseases of common dry bean, which affects producers world-

wide is CBB, caused by the bacterial pathogen Xap and its fuscans variety Xff (Vauterin

et al. 1995). The disease affects producers yearly, with severity dependent on the

susceptibility of the variety and environmental factors, such as humidity and

temperature, with yield reduction between 10-45% being observed (Saettler 1989;

Gillard et al. 2009). Since the symptoms may develop on the leaves, stems, pods and

Page 153: Identification and Characterization of Common Bacterial

138

seeds of infected plants (Vidaver 1993), CBB not only affects yield potential, but the

quality of the marketable portion of the plant – the seeds and pods.

4.2.3 CG Testing in Common Bean

P. vulgaris has little natural resistance to Xap. However, close relatives of the

species, P. acutifolius (tepary bean) and P. coccineus (scarlet runner bean) have

resistant genotypes (Mohan 1982; Drijfhout and Blok 1987; Singh and Muñoz 1999).

These relatives have been used to introduce resistance to CBB into P. vulgaris lines,

and have resulted in the production of varieties which are considered resistant, as they

exhibit delayed and/or reduced symptoms. CBB resistance is a quantitative trait in

common bean, with multiple major and minor quantitative trait loci (QTLs) being

mapped in many populations (Miklas et al. 2000b; Tar’an et al. 2001; Blair et al. 2003;

Cordoba et al. 2010; Shi et al. 2011). However, it has been shown that P. acutifolius

has a high level of leaf and pod resistance which is controlled by one dominant gene

and displays a characteristic hypersensitive response (Drijfhout and Blok 1987).

Although mapping of CBB resistance QTLs has helped to identify regions of the

genome that are associated with the trait, specific genes and their expression patterns

in response to Xap inoculation have yet to be identified. The aim of this study was to

measure expression from candidate CBB R genes (predicted in Chapter 2) in response

to the inoculation by Xap in resistant and susceptible bean cultivars. The study tested

the hypothesis that CBB R genes would be upregulated in leaf samples inoculated with

Xap in resistant P. vulgaris compared to water inoculated resistant, and water and Xap-

inoculated susceptible P. vulgaris.

Page 154: Identification and Characterization of Common Bacterial

139

4.3 Materials and Methods

4.3.1 Plant and Inoculum Preparation

4.3.1.1 Plant Growth

A total of 50 plants each for P. vulgaris resistant variety (OAC Rex) and

susceptible variety (OAC Seaforth) were potted in 5” round pots filled with pre-

moistened LA4 mix (Green Island Distributors, Riverhead, N.Y.) at 5 seeds per pot, and

covered with medium vermiculite. The pots were soaked with deionized water with

fertilizer (20-20-20 N-P-K), and maintained in a growth room providing 16h light (150

µmol/m2/s), 8h dark cycle at 22oC until the uni-foliate stage (~10 days). At this point, all

plants from each variety were inoculated, placed in a misting chamber, and the

temperature was raised to 28oC daytime temperature and 20oC night temperature, with

80% humidity. A second replication was performed which included 90 plants to

accommodate additional time points.

4.3.1.2 Inoculum Preparation

Bacterial inoculum was generated by spreading 50 mL of Xap isolate 98 (Xap98)

on 5 yeast salt media plates with 1% agar, and maintaining them at 28oC for 48 hours.

To generate a liquid inoculum, the colonies were washed off the plates with distilled

sterile water and diluted with distilled sterile water to an OD600 of 0.3. Control inoculum

was prepared by autoclaving distilled water.

4.3.1.3 Plant Inoculation

Plant inoculations were performed using the multiple needle method (Yu et al.

2000b). Both uni-foliate leaves of each plant were pierced by needles into a sterile

Page 155: Identification and Characterization of Common Bacterial

140

sponge soaked in inoculum. The needles were removed and inoculum soaked sponges

were squeezed gently against either side of the punctured leaf.

4.3.2 Sample Collection and Analysis

4.3.2.1 Sample Collection and Preparation

Single whole-leaf samples were taken from three plants per treatment (Xap98

and water control) per time point for OAC Rex and Seaforth varieties respectively.

Samples were removed from the plant and immediately frozen in liquid nitrogen to

preserve RNA present in the leaf. Samples were taken at 0, 12, 24, 48, 72, 96, 120 HPI

for the first replication (January 2013) and 0, 12, 24, 36, 48, 72, 96, 120 and 168 HPI for

the second replication (February 2013). Frozen leaf samples were maintained in a -

80oC freezer until grinding and RNA isolation.

Whole leaf samples were ground to a fine powder using a mortar and pestle

chilled to -80oC by liquid nitrogen. Ground tissue samples were transferred to liquid

nitrogen chilled 1.5 mL centrifuge tubes in ~100 mg aliquots and stored in a -80oC

freezer until further analysis.

4.3.2.2 RNA Isolation

RNA was isolated from ~100 mg samples of ground leaf tissue using Qiagen

RNeasy Plant RNA isolation kit. Aliquots of ground leaf tissue were placed in liquid

nitrogen to maintain -80oC temperature and added to 450 µl of RLT buffer – β-

mercatoethanol mixture (1 mL RLT buffer and 10 µL β-mercaptoethanol) then vortexed

for ~5 seconds. The mixture was then added to the QIAshredder column and

centrifuged for 2 minutes at 15,000 RPM. The supernatant was then mixed with ~225

µL 100% ethanol, the mixture was added to an RNeasy Mini spin column and

Page 156: Identification and Characterization of Common Bacterial

141

centrifuged for 15 seconds at 10,000 RPM. The flow-through was discarded, 700 µL of

RW1 buffer was added to the RNeasy spin column and centrifuged for 15 seconds at

10,000 RPM. The flow-through was discarded and 500 µL of RPE buffer was added to

the spin column and centrifuged at 15 seconds at 10,000 RPM. This step was repeated

with centrifugation lasting for 2 minutes. The spin column was then transferred to a 1.5

µL tube, 50 µL of RNase-free water was added to the column membrane and the tube

was centrifuged for 1 minute at 10,000 RPM to elute the RNA.

The RNA concentrations of the eluted samples were measured by applying 1 µL

of the RNA sample to the Nanodrop ND-1000 Spectrophotometer (Thermo Fisher

Scientific Inc. © 2013, Wilmington DE) pedestal. RNA quality was determined by both

by the 260/280 (>1.90) and 260/340 (> 1.80) absorbance ratios, and by visualization of

band patterns when ~1 µg of each sample was run in 1% agarose gel electrophoresis at

100 V. The sample was considered to be good quality when the most intense band was

present at ~1100 bp, with the second most intense band present at ~700 bp.

DNAse digestion was performed using the Ambion® Turbo DNA free (Life

Technologies Inc., Burlington, Ontario) kit and protocol. A total of 10 µg of RNA for

each sample was mixed with 5 µL of DNAse buffer and 1 µL of Turbo DNAse, mixed

and incubated for 25 minutes at 37oC in a water bath. Post-incubation, 5 µL of DNAse

inactivation buffer was added to the sample and incubated for 2 minutes at room

temperature, then centrifuged for 1.5 minutes at 10,000 RPM. The supernatant was

transferred to a new tube, and the concentration and quality of the RNA was determined

by applying 1 µL of the sample to the Nanodrop pedestal. Only samples with

Page 157: Identification and Characterization of Common Bacterial

142

absorbance ratios higher than 1.90 at 260/280 and 1.80 at 260/230 were used for cDNA

synthesis.

4.3.2.3 cDNA Transcription

Isolated RNA was transcribed to cDNA using the qScriptTM cDNA SuperMix. A

standard of 1 µg of RNA template was used in each reaction and incubated as per kit

instructions in a PCR cycler. Transcription of cDNA from RNA was confirmed by using 2

µL of the cDNA product as the nucleotide template in a PCR reaction using CG specific

primers and actin primers. The products were visualized via 1% agarose gel

electrophoresis, which was run for 1.25 hours at 100 volts.

4.3.2.4 Primer Design and PCR Amplification

CG primers were designed using PRIMER3 software

(http://frodo.wi.mit.edu/primer3) to amplify sections of the predicted coding region for

each of the CGs. Primers were designed to amplify sequences either between or within

predicted exons of CGs (Figure 4.1). The primers that produced the most intense and

distinct bands from genomic DNA were used for RNA expression analysis (Table 4.1).

PCR reactions consisted of 2.5 µl 10x PCR Buffer without MgCl2, 2.5 µl MgCl2, 1 µl 10

mM dNTP, 1 µl each for reverse and forward primers, 0.5 µl Jumpstart Taq Polymerase

(Sigma-Alderich Co.), 14.5 µl sterile distilled H2O and 2 µl DNA template for a total

reaction size of 25 µl. The PCR reaction cycling parameters were initial denaturation at

94oC for 1 minute, 25 cycles of denaturation (94oC for 30 seconds), annealing

(temperature dependent on primer, 30 seconds) and extension (72oC for 1 minute), a

final extension cycle for 10 minutes at 72oC and held at 4oC until gel electrophoresis.

PCR reaction products were loaded with a 100 bp ladder as size reference into a 1%

Page 158: Identification and Characterization of Common Bacterial

143

agarose 1x TBE gel stained with ethidium bromide and run at 100 volts for 1.5 hours in

1x TBE buffer. Visualization was performed using UV light.

Page 159: Identification and Characterization of Common Bacterial

144

Figure 4.1. Locations of RNA expression primers within CGs. For CG CBB2, yellow

arrows indicate the locations of reverse and forward primers. For CG CBB3, red arrows

indicate the locations of reverse and forward primers. For CG CBB4, green arrows indicate

the locations of reverse and forward primers. The red and yellow arrow indicates a

shared primer location between CBB2 and CBB3.

Page 160: Identification and Characterization of Common Bacterial

145

Table 4.1. Candidate common bacterial blight R gene primer sequences for real time quantitative RT-PCR. Expected genomic and CDS template product sizes and annealing locations within the CG sequences are indicated.

Target Gene

Primer Name

Reverse Sequence Forward Sequence Genomic Band Size

CDS Band Size

CG Location

CBB2

CBB2R1 CCTTAAACTAAGCCTTCACC GGCTCAACTACAATTTTCAC 1,627 bp - CDS2-CDS6

CBB2CS CCGAACGCTCTTTTAACTCG CGACTCATCCCTCAGGAAAC 1,635 bp - CDS2-CDS6

CBB3

CBB3R1 GGAAACACCTATTCAAGG GTTTATCACTACACCACTC 582 bp - CDS3-CDS4

CBB3CS CCGAACGCTCTTTTAACTCG TGCTCGATGAACTAGGCTCA 1,470 bp 719 bp CDS2-CDS5

CBB4

CBB4R3 CAACCTCCAACCACCTTAG GTTTTGCCACTTCTCAAG 1,503 bp - CDS3-CDS6

CBB4CS TATGTTCGAAGCTCGACCCT GCACTCTTGGGCCTACTCAG 2,662 bp 1,320 bp CDS1-CDS6

ACTIN

Act-R/F (Chen et al.

2008)

TTTCCTTGCTCATTCTGTCCG GAAGTTCTCTTCCAACCATCC - 180 bp -

Page 161: Identification and Characterization of Common Bacterial

146

4.4 Results

4.4.1 CG Expression

4.4.1.1 CBB Susceptible Genotype (Seaforth)

Initial RNA isolation and cDNA synthesis from Seaforth ground leaf samples from

the first experiment replication showed amplification of actin, CBB2, CBB3 and CBB4

for samples taken at 12HPI and 24HPI with primer sets CBB2R1, CBB3R1, CBB4R2

(Figure 4.2). Amplification of a 200bp band by actin primers was observed for both

water and Xap-inoculated cDNA as well as for the OAC Rex control, and there was no

amplification of actin from the water control (Figure 4.2). Amplification of a 550bp band

by the CBB2 primers was observed at 12 and 24HPI for Xap inoculated samples, but

not for water inoculated samples (Figure 4.2). Amplification of CBB2 was observed in

the OAC Rex genomic DNA control but not the water control (Figure 4.2). CBB3

primers amplified a 300 bp band from samples taken at both 12 and 24 HPI, when

inoculated by water and Xap98, as well as the OAC Rex genomic DNA (Figure 4.2).

There was no amplification from the water control (Figure 4.2). Amplification of a 200

bp band by CBB4 primers was observed in samples taken at 12 and 24 HPI when

inoculated by Xap98 and water (Figure 4.2). In the case of the water-inoculated 24 HPI

sample, the band was indistinct, and covered a larger range of sizes (100-200 bp). A

200 bp band was amplified from OAC Rex genomic DNA, but not the water control

(Figure 4.2).

When isolation of RNA and synthesis of cDNA was performed using leaf samples from

the first experiment replication for time points 0HPI, 48HPI, 72HPI, 96HPI, and 120 HPI,

no amplification of CGs CBB2, CBB3 and CBB4 was observed when

Page 162: Identification and Characterization of Common Bacterial

147

Figure 4.2. Initial CG expression in inoculated P. vulgaris leaves. Expression of CGs CBB2, CBB3 and CBB4, were compared to expression of actin check in resistant (OAC Rex) and susceptible (Seaforth) common bean varieties. The inoculum used was X. axonopodis pv. phaseoli isolate 98 (X) and sterile distilled water (C). RNA was isolated from leaf samples taken from water and Xanthomonas-inoculated OAC Rex and Seaforth plants at 12 (12HPI) and 24 (24HPI) hours post-inoculation. OAC Rex genomic DNA and sterile distilled water were used as positive (+) and negative (-) PCR controls, respectively. CG bands were produced using primers CBB2R1, CBB3R1 and CBB4R2 for CBB2, CBB3 and CBB4, respectively. Band size is indicated by a white arrow for actin, yellow arrow for CBB2, red arrow for CBB3 and green arrow for CBB4.

Page 163: Identification and Characterization of Common Bacterial

148

Figure 4.3. Subsequent CG expression in second experiment replication of P. vulgaris leaf inoculation. Expression of CGs CBB2, CBB3 and CBB4 were compared to expression of actin check in resistant (OAC Rex) and susceptible (Seaforth) common bean varieties. The inoculum used was X. axonopodis pv. phaseoli isolate 98 (X) and sterile distilled water (C). RNA was isolated from leaf samples taken from water and Xanthomonas-inoculated OAC Rex and Seaforth plants at 0, 12, 24, 36, and 48 hours post-inoculation (HPI). OAC Rex genomic DNA and sterile distilled water were used as positive (+) and negative (-) PCR controls, respectively. CG bands were produced using primers CBB2CS, CBB3CS and CBB4CS for CBB2, CBB3 and CBB4, respectively. Band size is indicated by a white arrow for actin, yellow arrow for CBB2, red arrow for CBB3 and green arrow for CBB4.

Page 164: Identification and Characterization of Common Bacterial

149

primers CBB2R1, CBB3R1 and CBB4R2 were used (data not shown). This was

consistent for both water and Xap- inoculated samples. Multiple repetitions of the RNA

isolation and cDNA synthesis procedures for 12 and 24 HPI samples from both the first

and second experiment replications produced no PCR no amplification of CGs CBB2,

CBB3 and CBB4 when primers CBB2R1, CBB3R1 and CBB4R2 were used (data not

shown). CG primers amplified bands of the previously observed size for CBB2 (550bp),

CBB3 (300bp) and CBB4 (200bp) from Seaforth genomic DNA (data not shown).

The bands produced by actin primers with cDNA and genomic DNA templates

were either of inconsistent intensity, or not present (data not shown). Therefore,

multiple repetitions of cDNA synthesis were performed using samples from all time

points from both experiments until consistent quality was obtained. The most consistent

cDNA quality was achieved when absorbance ratios above 1.90 (260/280) and 1.80

(260/340) were observed for isolated RNA post DNAse treatment. Consistency of

cDNA quality was identified by the presence of actin bands for all samples, although

variability in the intensity was still observed (data not shown).

To determine if primers from different locations within the CG sequence were

capable of amplifying bands from cDNA, new primers were designed (Table 4.1; Figure

4.1). Newly designed primers (CBB2CS, CBB3CS and CBB4CS) produced no

amplification of CGs from synthesized cDNA from samples taken at 0, 12, 24, 48, 72,

96, and 120 HPI for the first experiment replication (Appendix 11) and 0, 12, 24, 36, 48,

72, 96, 120, and 168 HPI for the second experiment replication (Figure 4.3, samples

from 0-48 HPI). These primers produced bands of the expected size for CBB2 (1,635

bp), CBB3 (1,470 bp) and CBB4 (2,662 bp) from Seaforth genomic DNA (Figure 4.3).

Page 165: Identification and Characterization of Common Bacterial

150

Amplification of actin was observed for all time points and inoculum types, as well as

Seaforth genomic DNA control, but no amplification was observed with the water control

(Figure 4.2-4.3).

4.4.1.2 CBB Resistant Genotype (OAC Rex)

Initial RNA isolation and cDNA synthesis from OAC Rex ground leaf samples

from the first replication showed amplification of actin, CBB3 and CBB4 for samples

taken at 12HPI and 24HPI with primer sets CBB2R1, CBB3R1, CBB4R2 (Figure 4.2).

Amplification of a 200bp band by actin primers was observed for both water and Xap-

inoculated cDNA as well as for the OAC Rex control, and there was no amplification of

actin from the water control (Figure 4.2). There was no amplification of a 550bp band

by the CBB2 primers at either the 12 or 24HPI for water and Xap-inoculated OAC Rex

leaf samples (Figure 4.2). Amplification of CBB2 was observed in the OAC Rex

genomic DNA control but not the water control (Figure 4.2). CBB3 primers amplified a

300 bp band from samples taken at both 12 and 24 HPI, when inoculated by Xap98, as

well as from the OAC Rex genomic DNA (Figure 4.2). There was faint amplification of a

300 bp band from the water inoculated sample at 24HPI, and no amplification at 12HPI

or from the water control (Figure 4.2). Amplification of a 200 bp band by CBB4 primers

was observed in samples taken at 12 and 24 HPI when inoculated by Xap98 as well as

the OAC Rex genomic DNA control (Figure 4.2). Water-inoculated samples and the

water control produced no observed amplification of the 200 bp band (Figure 4.2).

When isolation of RNA and synthesis of cDNA was performed using leaf samples

from the first experiment replication for time points 0HPI, 48HPI, 72HPI, 96HPI, and 120

HPI for OAC Rex leaf samples, no amplification of CGs CBB2, CBB3 and CBB4 was

Page 166: Identification and Characterization of Common Bacterial

151

observed when primer sets CBB2R1, CBB3R1, CBB4R2 were used (data not shown).

This was consistent for both water and Xap- inoculated samples. Repetition of RNA

isolation and cDNA synthesis for 12 and 24 HPI samples from both the first and second

replicates produced no PCR amplification of CGs CBB2, CBB3 and CBB4 with the

same primers (data not shown). CG primers amplified bands of the previously observed

size for CBB2 (550bp), CBB3 (300bp) and CBB4 (200bp) from OAC Rex genomic DNA

(data not shown).

The bands produced by actin primers with cDNA and genomic DNA templates

were either of inconsistent intensity, or not present (data not shown). Therefore,

multiple repetitions of cDNA synthesis were performed using samples from all time

points from both experiments until consistent quality was obtained. The most consistent

cDNA quality was achieved when absorbance ratios above 1.90 (260/280) and 1.80

(260/340) were observed for isolated RNA post DNAse treatment. Consistency of

cDNA quality was identified by the presence of actin bands for all samples, although

variability in the intensity was still observed (data not shown).

Since the CGs primers (CBB2R1, CBB3R1, and CBB4R3) were still unable to

amplify bands from the cDNA, new primers were designed (CBB2CS, CBB3CS and

CBB4CS) (Figure 4.1; Table 4.1). When new primers were used (CBB2CS, CBB3CS

and CBB4CS), no amplification was observed for all time points for both water and Xap-

inoculated leaves over both replicates (Appendix 11; Figure 4.3; Table 4.1). These

primers produced bands of the expected size for CBB2 (1,635 bp), CBB3 (1,470 bp)

and CBB4 (2,662 bp) from OAC Rex genomic DNA (Figure 4.3). Amplification of actin

was observed from OAC Rex cDNA samples at all time points and inoculation types, as

Page 167: Identification and Characterization of Common Bacterial

152

well as from OAC Rex genomic DNA, but no amplification was observed from water

control (Figure 4.3).

Page 168: Identification and Characterization of Common Bacterial

153

4.5 Discussion

4.5.1 Observed CG Expression

Despite initial results at 12 and 24 HPI suggesting that there were differences in

expression of candidate CBB CGs CBB2, CBB3 and CBB4 between resistant (OAC

Rex) and susceptible (Seaforth) varieties, these results could not be repeated both with

the initial (CBB2R1, CBB3R1, CBB4R2) and subsequent primers (CBB2CS, CBB3CS,

CBB4CS) (Figures 4.2-4.3). The inability of these primers to show CG expression may

be caused by the location of CGs. Since the design and use of the primers used in this

experiment (Table 4.1), predictions for exon locations within CBB2, CBB3 and CBB4

have changed, based on the same program used previously to identify exons

(FgeneSH). This re-analysis also caused the incorporation of a previously disregarded

CG (CBB1) into bioinformatics analyses due to improved gene feature prediction (see

Chapter 2). It was also shown that the contig1701 sequence was not continuous, which

resulted in the ~20 kb sequence where CBB3 was predicted to be discarded (see

Chapter 2).

In the case of primers CBB2R1, CBB2CS, CBB3R1 and CBB4R3, one of the pair

of primers is located outside of predicted coding regions, which is the likely cause of the

inability to amplify from cDNA (Figure 4.1). This corresponds to the observed

differences between expected and actual band sizes produced (Figure 4.2). The initial

bands produced by CBB2R1 (~550 bp), CBB3R1 (~300 bp) and CBB4R3 (~170 bp)

were the same regardless of whether cDNA or genomic DNA were used. These bands

were also significantly smaller than the than the expected genomic DNA size (Figure

Page 169: Identification and Characterization of Common Bacterial

154

4.2; Table 4.1). This indicates the likelihood that the initial results were produced by

DNA contamination, and that the primers were amplifying an unexpected target.

It is also important to note that the both primer sets designed for CBB2 have at

least one primer located outside the exon (Figure 4.2). Consequently, these primers

would not successfully indicate the presence of CBB2 coding sequence within the

synthesized cDNA (Table 4.1). This raises the possibility that although current results

indicate that CBB3 and CBB4 are not expressed as a result of Xap inoculation, CBB2

may have been expressed, but the designed primers lacked the ability to show this.

4.5.2 Expected CG Expression Patterns

Disease resistance gene expression patterns have been previously characterized

in multiple plants and in different plant-pathogen models, with examples including

common bean, coffee, orange, Arabidopsis and tobacco (Anderson et al. 2004; Ganesh

et al. 2005; Ahn et al. 2011; Borges et al. 2012; Rodríguez et al. 2014). Expression

patterns can take a variety of forms and may be different depending on the plant, tissue,

pathogen and genes involved as was evidenced by the study performed by Borges et

al. (2012). Different expression patterns were observed for multiple genes in the 96 HPI

with Colletotrichum lindemuthianum (anthracnose). These responses to pathogen

recognition include constant expression (PR16a), up-regulation from low expression

(PR2), rapid induction of unexpressed genes (PR1b), and down-regulation (PGIP).

Another example of resistance gene expression patterns includes the measurement of

PAL in P. vulgaris inoculated with the same pathogen, C. lindemuthianum (Fraire-

Velázquez and Lozoya-Gloria 2003). In the resistant reaction, PAL expression was

strong for the first 2 HPI, decreased until 16 HPI and then increased more significantly.

Page 170: Identification and Characterization of Common Bacterial

155

Large scale mRNA expression profiling performed by Tao et al. (2003),

demonstrated that in the Arabidopsis-Pseudomonas syringae patho-system, increased

gene expression of ~ 2000 genes occurred at 6-9 HPI in the resistant interaction as

opposed to 30 HPI in the susceptible one. Although the CBB resistant source for OAC

Rex, P. acutifolius accession (P1440795), exhibits a hypersensitive response to CBB

(Parker 1985; Drijfhout and Blok 1987), the response to CBB in OAC Rex is

characterized by an increased time to development of CBB symptoms, and decreased

severity of symptoms. Considering that the resistant reaction in OAC Rex causes

retardation in disease progression and not complete immunity, it is hypothesized that

this is due to slower pathogen recognition and/or response. Based on this hypothesis, it

is therefore expected that the R gene(s) involved would be expressed less strongly or at

a later time post-inoculation when compared to the same response in P. acutifolius.

Page 171: Identification and Characterization of Common Bacterial

156

4.6 Conclusion

This study aimed to observe three candidate CBB resistance genes for changes

in their expression over time in common bean when resistant (OAC Rex) and

susceptible (OAC Seaforth) varieties were inoculated with either water or Xap.

Expression of the CGs was not observed within the parameters of the experiment.

Further studies of these CGs and further identification and testing of CGs in the area of

the PV-ctt001 QTL region will help to progress current knowledge about the

mechanisms for CBB resistance and the genes involved.

Page 172: Identification and Characterization of Common Bacterial

157

Chapter 5: Future Directions and Recommendations

5.1 OAC Rex Genome Sequence Assembly

Since a location within the OAC Rex genome has not been identified for

contig1701, it is important to continue to search for this. This is highly dependent on the

completion of the full OAC Rex genome sequence. The scaffolds and contigs identified

in BLAST searches discussed above have yet to be included in the pseudochromosome

assembly, making it difficult to identify the location and surrounding region of the

candidate genes reported in this study. Once a location for the candidate genes is

determined, it will be possible to explore further their position relative to the PV-ctt001

marker as well as other markers and disease resistance genes in the surrounding area.

Despite the possibility that this marker was associated to CBB resistance as the result

of a false positive (see Chapter 1), the marker provides a targeted region to look at

genome details such as the repetitive nature surrounding identified disease resistance

genes.

5.2 Candidate CBB R Gene Analysis

5.2.1 Identification and Characterization of New CGs

The evidence indicating the possible retrotransposon characteristics of candidate

genes CBB1-4 (see Chapter 2), and the lack of candidate gene expression (see

Chapter 3) suggest that identification of new candidate genes would further the aims of

identifying CBB resistance genes more effectively. Several potential locations for the

identification of candidate resistance genes exist, including the sequence in the regions

surrounding CBB resistance markers that was analyzed by Perry et al. (2013).

Page 173: Identification and Characterization of Common Bacterial

158

The sequence characterized by Perry et al. (2013) surrounding CBB resistance

QTLs PV-ctt001 (Chromosome 4), SU91 (Chromosome 8) and SAP6 (Chromosome 10)

provide a source for identification of additional candidate resistance genes based on

predicted function. A total of 114 genes reported within 5 OAC Rex contigs that

matched to the PV-ctt001 region of chromosome 4. Of the identified genes, several

categories have potential as the source of disease resistance, including

defense/signalling/stress (21.9%), unknown (7.9%) and hypothetical (27.5%) genes.

The defense/signalling/stress gene predictions especially would be an excellent target

group, considering the large percent of the predicted genes it comprises, as well as the

fact that they are easily comparable to previously characterized genes. Testing of

multiple candidate genes identified in the region surrounding PV-ctt001 to observe

expression changes could be performed by real time PCR. Alternatively, genome wide

gene expression observations could be performed using RNAseq.

In a previous study, Shi et al. (2012) identified 16 candidate genes associated

with CBB resistance marker SU91, which is present in OAC Rex Pv08, to test for

increased marker precision. Although two candidate gene markers were reported to

have improved association to CBB resistance, this was only related to disease ratings,

and not gene expression. The use of these candidate gene markers in RNA expression

studies would have more power in confirming the involvement of the associated

candidate resistance genes in resistance to CBB. An interesting line of inquiry could

explore candidate resistance gene expression identified in both the PV-ctt001 and SU91

QTL regions.

Page 174: Identification and Characterization of Common Bacterial

159

5.2.2 CG Expression Analysis

The inclusion of additional collection time points and tissue sources would

increase the ability to make observations about candidate gene expression patterns.

Multiple studies have shown the first 24 hours post-inoculation to be a critical period for

the induction of disease resistance-related genes, with peaks of induction occurring at a

variety of different time points within that critical period (Fraire-Velázquez and Lozoya-

Gloria 2003; Tao et al. 2003; Ahn et al. 2011; Borges et al. 2012), However, the

expected increased time to resistance gene induction, for the reasons discussed above,

suggests that sample collection after a longer time period (perhaps up to 36 HPI), with

higher frequency (perhaps every 2-6 hours) would provide the best coverage and

increase the likelihood of detecting changes in expression.

5.3 Model Plant CBB Inoculation Procedure

Disease progression was shown in Xap-inoculated Arabidopsis leaves (see

Chapter 3), showing that Arabidopsis is a host when manually inoculated. However,

repetitions of the initial experiment have been unsuccessful. It is therefore imperative

that the study be repeated, to decrease variability and improve the power of any

conclusions made. Since the subsequent repetitions used only isolate Xap12, which

was shown to be a less virulent isolate, future replications should use Xap18, which

produced the highest symptoms and CFU count in Arabidopsis. Future inoculations of

detached leaves should also be performed using a mixture of the inoculation method

described (see Chapter 3), and the infiltration method used by Randhawa and Civerolo

(1985) in peach. The use of a mixture of isolates as an additional inoculum to compare

to inoculation with a single isolate (i.e. Xap18) would also indicate whether a mixture of

Page 175: Identification and Characterization of Common Bacterial

160

effectors contributed from the four isolates is what is required for non-host resistance to

be overcome.

Page 176: Identification and Characterization of Common Bacterial

161

Chapter 6: Summary

Disease susceptibility is one of the primary limitations to yield for all plant species

produced across the globe. One of the major diseases affecting common bean (P.

vulgaris L.) yield quantity and quality is the disease CBB, caused by bacteria pathogen

Xap and its fuscans variant Xff. The resistant variety OAC Rex provides reduced

disease symptoms and increased yield under disease pressure (Michaels et al. 2006;

Gillard et al. 2009). The source of its resistance is a wild relative P. acutifolius (tepary

bean) accession P1440795 (Parker 1985). Although resistance QTLs have been

identified and associated with markers in OAC Rex, such as PV-ctt001 (Tar’an et al.

2001; Perry, 2010; Perry et al. 2013) and SU91 (Perry et al. 2013), no resistance genes

have been identified to date. It is important to determine the genes involved in

resistance to CBB to increase the current understanding of the resistance reaction, and

also to improve selection capabilities for breeding new resistant varieties.

The current study aimed to identify candidate CBB resistance genes within

contigs for BAC clone Rex021F10 of OAC Rex selected with the PV-ctt001marker, and

test them in a variety of ways. The characterization of genes identified through

homology to previously reported candidate genes revealed that retrotransposon

elements comprised a high proportion of the sequence of the selected contig sequences

(contig1701 and contig1455). Retrotransposons integrate most often in genic regions

(Miyao et al. 2003; Le et al. 2007), and have been shown to interact with resistance

genes by disrupting gene function (Song et al. 1997; Hernández-Pinzón et al. 2009),

causing chimeric proteins (Kashkush et al. 2003; Frost et al. 2004) changing promoter

Page 177: Identification and Characterization of Common Bacterial

162

function and modifying the epigenetome (Hayashi et al. 2009; McDowell and Meyers,

2013; Tsuchiya and Eulgem, 2013). Characterization of the identified candidate genes

revealed retrotransposon features including similarity to GAG/POL poly proteins,

reverse transcriptases, and Gypsy/Ty-3 retro-element proteins. Although the similarities

to these proteins were not high (30-47%), when combined with the presence of multiple

other retrotransposons identified within the same contig indicates the likelihood the CGs

are retrotransposons. The presence of several genes in the associated contig

sequence with potential disease resistance functions (i.e ascorbate oxidase, callose

synthase, hypothetical protein), provides additional areas for testing additional

candidate CBB resistance genes.

Candidate CBB resistance gene expression was tested in susceptible and

resistant common bean genotypes. However, no gene expression was observed in

cDNA at multiple time points for candidate genes CBB2-CBB4. This could be due to

human error during primer design. However, the subsequent characterization of the

candidate genes as retrotransposons indicates that expression is not likely to be

measured, even with experimental parameter alterations. A more promising target is

the putative ascorbate oxidase gene identified in contig1701, since this gene family has

previously been shown to be involved in regulating reactive oxygen species levels,

which are associated with stress and disease resistance responses (Pavet et al. 2005;

Pignocchi et al. 2006; Sanmartin et al. 2007; Fotopoulos et al. 2008).

Since genetic manipulation of common bean is difficult (i.e. genetic

transformation), testing candidate CBB resistance gene function within resistant

common bean genotypes is an unattractive option. The development of a detached leaf

Page 178: Identification and Characterization of Common Bacterial

163

A. thaliana Xap-inoculation method was aimed at providing an alternative method of

testing candidate gene function in a susceptible model. Although optimization is still

required, symptom progression was shown in the model plant, and provides a more

controlled environment to test candidate genes in a more targeted way.

Page 179: Identification and Characterization of Common Bacterial

164

8.0 References

Abedon, B. G. and Tracy, W. F. 1996. Corngrass 1 of maize (Zea mays L. delays development of adult plant resistace to common rust (Puccinia sorghi Schw.) and European corn borer (Ostrinia nuvilalis Hubner). J. of Heredity. 87: 219-223

Aggour, A.R., Coyne, D.P. and Vidaver, A.K. 1989. Comparison of leaf and pod disease reactions of beans (Phaseolus vulgaris L.) inoculated by different methods with strains of Xanthomonas campestris pv. phaseoli (Smith) dye. Euph. 43: 143-152

Agriculture and Agri-Food Canada. 2011. Canada: Pulse and special crops outlook. Obtained Aug 23 2011 from World Wide Web: http://www.agr.gc.ca/pol/mad-dam/index_e.php?s1=pubs&s2=spec&s3=php&page=spec_2011-07-05

Health Canada. 2014. Canada’s Food Guide. Obtained December 9 2014 from World Wide Web: http://www.hc-sc.gc.ca/fn-an/food-guide-aliment/choose-choix/meat-viande/index-eng.php

Agrios, G.N. 2005. Chapter six. How plants defend themselves against pathogens. In 5th Ed. Plant Pathology. Elsevier Academic Press. Burlington, MA, USA

Ahn, E.-R., Sarowar, S., Jeon, Y., Pai, H.-S. 2011. Molecular characterization of NLP function in Nicotiana benthamiana. J. Plant Biol. 54: 199-208

Alghisi, P. and Favaron, F. 1995. Pectic degrading enzymes and plant parasite interactions. Eur. J. Plant Pathol. 101: 365-375

Ameline-Torregrosa, C., Wang, B.-B., O’Bleness, M.S., Deshpande, S., Zhu, H., Roe, B., Young, N.D., and Cannon, S.B. 2008. Identification and characterization of nucleotide-binding site-leucine-rich repeat genes in the model plant Medicago truncatula. Plant Physiol. 146(1): 5-21

Anderson, P.A., Lawrence, G.J., Morrish, B.C., Ayliffe, M.A., Finnegan, E.J. and Ellis, J.G. 1997. Inactivation of the flax rust resistance gene M associated with loss of a repeated unit within the leucine-rich repeat coding region. Plant Cell. 9(4): 641-651

Page 180: Identification and Characterization of Common Bacterial

165

Anderson, J.P., Badruzsaufari, E., Schenk, P.M., Manners, J.M., Desmond, O.J., Ehlert, C., Maclean, D.J., Ebert, P.R. and Kazan, K. 2004. Antagonistic interaction between abscisic acid and jasmonate-ethylene signaling pathways modulates defense gene expression and disease resistance in Arabidopsis. Plant Cell. 16: 3460-3479

Arnold K, Bordoli L, Kopp J, and Schwede T. 2006. The SWISS-MODEL Workspace: A web-based environment for protein structure homology modelling. Bioinformatics. 22:195-201

Asai, T., Tena, G., Plotnikova, J., Willmann, M.R., Chiu, W.-L., Gomez-Gomez, L., Boller, T., Ausubel, F.M. and Sheen, J. 2002. MAP kinase signalling cascade in Arabidopsis innate immunity. Nature. 415:977-983

Bakker, E.G., Toomajian, C., Kreitman, M. and Bergelson. 2006. A genome-wide survey of R gen polymorphisms in Arabidopsis. Plant Cell. 18(8):1803-1818

Bernoux, M., Ve, T., Williams, S., Warren, C., Hatters, D., Valkow, E., Zhang, X., Ellis, J. G., Kobe, B. and Dodds, P. N. 2011. Structural and functional analysis of a plant resistance protein TIR domain reveals interfaces for self-association, signaling and autoregulation. Cell Host Microbe. 9(3): 200-211

Bewley, J.D., Black, M. and Halmer, P. 2006. The encyclopedia of seeds: Science, technology and uses. CAB International 875 Massachusetts Ave., Cambridge MA, USA. pg 35-36

Bitocchi, E., Nanni, L., Bellucci, E., Rossi, M., Giardini, A., Spagnoletti Zeuli, P., Logozzo, G., Stougaard, J., McClean, P., Attene, G., and Papa, R. 2012. Mesoamerican origin of the common bean (Phaseolus vulgaris L.) is revealed by sequence data. PNAS. www.pnas.org/cgi/doi/10.1073/pnas.1108973109

Blair, M. W., Pedraza, F., Buendia, H. F., Gaitán-Solís, E., Beebe, S. E. and Tohme, J. 2003. Development of a genome-wide anchored microsatellite map for common bean (Phaseolus vulgaris L.). Theor. Appl. Genet. 107:1362-1374

Block, A., Guangyong, L., Fu, Z. Q. and Alfano, J. R. 2008. Phytopathogen type III effector weaponry and their plant targets. Curr. Opin. Plant Biol. 11: 396-403

Page 181: Identification and Characterization of Common Bacterial

166

Bogdanove, A.J., Koebnik, R., Lu, H., Furutani, A., Angiuoli, S.V., Patil, P.B., Van Sluys, M.-A., Ryan, R.P, Meyer, D.F., Han, S.-W., Aparna, G., Rajaram, M., Delcher, A.L., Phillippy, A.M., Puiu, D., Schatz, M.C., Shumway, M., Sommer, D.D., Trapnell, C., Benahmed, F., Dimitrov, G., Madupu, R., Radune, D., Sullivan, S., Jha, G., Ishihara, H., Lee, S.-W., Pandey, A., Sharma, V., Sriariyanun, M., Szurek, Vera-Cruz, C.M., Dorman, K.S., Ronald, P.C., Verdier, V., Dow, J.M., Sonti, R.V., Tsuge, S., Brednel, V.P., Rabinowicz, P.D., Leach, J.E., White, F.F. and Salzberg, S.L. 2011. Two new complete genome sequences offer insight into host and tissue specificity of plant pathogenic Xanthomonas spp. J. Bacteriol. 193(19):5450

Bollini, R. and Chrispeels, M. J. 1978. Characterization and subcellular localization of

vicilin and phytohemagglutinin, the two major reserve proteins of Phaseolus vulgaris L. - Planta 142: 291–298.

Bollini, R., Van der Wilden, W., and Chrispeels, M.J. 1982. A precursor of the reserve-protein, phaseolin, is transiently associated with the endoplasmic reticulum of developing Phaseolus vulgaris cotyledons. Physiologia Plantarum. 55: 82-92

Bonas, Ul, Schulte, R., Fenselau, S, Minsavage, G. V., Staskawcz,j B. J. and Stall, R. E. 1990.

Borges, A., Melotto, M., Tsai, S. M., and Gregorio Gomes Caldas, D. 2012. Changes in spatial and temporal gene expression during incompatible interaction between common bean and anthracnose pathogen. J. Plant Physiol. 169:1216-1220

Botella, M. A., Parker, J. E., Frost, L. N., Bittner-Eddy, P. D., Beynon, J. L., Daniels, M. J., Holub, E. B. and Jones, J. D. G. 1998. Three genes of the Arabidopsis RPP1 complex resistance locus recognize distinct Peronospora parasitica avirulence determinants. Plant Cell 10: 1847-1860

Bressani, R. 1983. Research needs to upgrade the nutritional quality of common beans (Phaseolus vulgaris). Qualitas Plantarum Plant foods Human Nutrition. 32: 101-110

Campa, A., Giraldez, R. and Ferreira, J. J. 2011. Genetic analysis of the resistance to eight anthracnose races in the common bean differential cultivar Kaboon. Pytopath. 101(6):757-764

Page 182: Identification and Characterization of Common Bacterial

167

Century, K. S., Lagman, R. A., Adkisson, M., Morlan, J., Tobias, R., Schwartz, K., Smith, A., Love, J., Ronald, P. C., and Whalen, M. C. 1999. Developmental control of Xa21-mediated disease resistance in rice. Plant Journal. 20(2): 231-236

Chan SL, Mukasa T, Santelli E, Low LY and Pascual J. 2010. The crystal structure of a TIR domain from Arabidopsis thaliana reveals a conserved helical region unique to plants. Protein Sci. 19:155–161

Chang, J.H., Tai, Y.S., Bernal, A.J., Lavelle, D.T., Staskawicz, B.J., and Michelmore, R.W. 2002. Functional analyses of the pto resistance gene family in tomato and the identification of a minor resistance determinant in a susceptible haplotype. Mol. Plant-Microbe Interact. 15:281-291

Chen, X., Steed, A., Harden, C. and Nicholson, P. 2006. Characterization of Arabidopsis thaliana-Fusarium graminearum interactions and identification of variation in resistance among ecotypes. Mol Plant Pathol. 7(5):391-403

Chen, Q., Han, Z., Jiang, H., Tian, D. & Yang, S. 2010b. Strong positive selection drives rapid diversification of R-Genes in Arabidopsis relatives. J Mol Evol. 70: 137-148

Chinchilla, D., Bauer, Z., Regenass, N., Boller, T. and Felix, G. 2006. The Arabidopsis receptor kinase FLS2 binds flg22 and determines the specificity of flagellin perception. Plant Cell. 18: 465-476

Chinchilla, D., Zipfel, C., Robatzek, S., Kemmerling, B., Nüremberger, T., Jones, J.D.G., Felix, G. and Boller, T. 2007. A flagellin-induced complex of the receptor FLS2 and BAK1 initiates plant defence. Nature. 448: 497-500

Collmer, A.. and Keen, N. T. 1986. The role of pectic enzymes in plant pathogenesis. Annu Rev. Phytopathol. 24: 383-409

Cook, B. J., Clay, R. P., Bergmann, C. W., Albersheim, P. and Darvill, A. G. 1999. Fungal polygalacturonases exhibit different substrate degradation patterns and differ in their susceptibilities to polygalacturonase-inhibiting proteins

Córdoba, J. M., Chavarro, C., Schlueter, J. A., Jackson, S. A. and Blair, M. W. 2010. Integration of physical and genetic maps of common bean through BAC-derived microsatellite markers. BMC Gen. 11: 436

Page 183: Identification and Characterization of Common Bacterial

168

Dangl, J. L. and Jones, J. D. G. 2001. Plant pathogens and integrated defence responses to infection. Nature. 411: 826-833

Darrasse, A., Bureau, C., Samson, R., Morris, C. E. and Jacques, M.-A. 2007. Contamination of bean seeds by Xanthomonas axonopodis pv. phaseoli associated with low bacterial densities in the phyllosphere under field and greenhouse conditions. Eur. J. Plant Pathol. 119: 203-215

David, P., Chen, N.W.G., Pedrosa-Harrand, A., Thareau, V., Sévignac, M., Cannon, S. B., Debouck, D., Langin, T., and Geffroy, V. 2009. A nomadic subtelomeric disease resistance gene cluster in common bean. Plant Physiol. 151(3): 1048-1065

Debouck, D. G., Rubiano, H. and del Carmen Menéndez, M. S. 1988. Determinate climbers among Argentinean materials of Phaseolus vulgaris L. Ann. Rep. Bean Improv. Coop. 31:189-190

Debouck, D., G., Toro, O., Paredes, O. M., Johnson, W. C. and Gepts, P. 1993. Genetic diversity and ecological distribution of Phaseolus vulgaris in northwestern South America. Econ Bot. 47:408-423

De Lorenzo, G., D’Ovidio, R. and Cervone, F. 2001. The role of polygalacturonase-inhibiting proteins (PGIPs) in defense against pathogenic fungi. Annu. Rev. Phytopath. 39:313-335

De Lorenzo, G. and Ferrari, S. 2002. Polygalacturonase-inhibiting proteins in defense against phytopathogenic fungi. Curr. Opin. Plant Biol. 5: 295-299

Desaki, Y., Miya, A., Venkatesh, B., Tsuyumu, S., Yamane, H., Kaku, H., Minami, E., and Shibuya, N. 2006. Bacterial lipopolysaccharides induce defense responses associated with programmed cell death in rice cells. Plant Cell Physiol. 47(11): 1530-1540

Desiderio, A., Aracri, B., Leckie, F., Mattei, B., Salvi, G., Tigelaar, H., Van Roekel, J.S., Baulcombe, D.C., Melchers, L.S., De Lorenzo, G. and Cervone, F. 1997 Polygalacturonase-inhibiting proteins (PGIPs) with different specificities are expressed in Phaseolus vulgaris. Mol Plant Microbe Interact. 10: 852–860

Page 184: Identification and Characterization of Common Bacterial

169

Develey-Rivière, M.-P. and Galiana, E. 2007. Resistance to pathogens and host developmental stage: A multifaceted relationship within the plant kingdom. New Phytol. 175(3):405-416

de Wit, P. J., Brandwagt, B. F., van den Burg, H.A., Cai, X., van der Hoorn, R. A., de Jong, C. F., van Klooster, J., de Kock, M. J., Krujit, M, Lindhout, W. H., Luderer, R., Takken, F. L., Westerink, N., Vervoort, J. J., Joosten, M. H. 2002.The molecular basis of co-evolution between Cladosporium fulvum and tomato. Antonie Van Leeuwenhoek. 81: 409-412

Di Matteo, A., Federici, L., Mattei B., Salvi, G., Johnson, K. A., Savino, C., De Lorenzo, G., Tsernoglou, D. and Cervone, F. 2003. The crystal structure of polygalacturonase-inhibiting protein (PGIP), a leucine-rich repeat protein involved in plant defense. PNAS. 100 (17): 10124-10128

Dodds,P.N., Lawrence,G.J. and Ellis,J.G. 2001. Six amino acid changes confined to the leucine-rich repeat b-strand/b-turn motif determine the difference between the P and P2 rust resistance specificities in flax. Plant Cell. 13: 163-178

Dorey, S., Baillieul, F., Pierrel, M.-A., Saindrenan, P., Fritig, B. and Kauffmann, S. 1997. Spatial and temporal induction of cell death, defense genes and accumulation of salicylic acid in tobacco leaves reacting hypersensitively to a fungal glycoprotein elicitor. MPMI. 10(5): 646-655

Dow, J. M., Clarke, B. R., Milligan, D. E., Tang, J.-L. and Daniels, M. J. 1990. Extracellular proteases from Xanthomonas campestris pv. campestris, the black rot pathogen. App. Env. Micro. 56 (10): 2994-2998

Drijfhout, E. and Blok, W.J. 1987. Inheritance of resistance to Xanthomonas campestris pv. phaseoli in tepary bean (Phaseolus acutifolius). Euphytica 36:803-808

Duncan, R.W., Singh, S.P. and Gilbertson, R.L. 2010. Interaction of common bacterial blight bacteria with disease resistance quantitative trait loci in common bean. Phytopathol.101(4): doi:10.1094/PHYTO-03-10-0095

Durham, K. M., Xie, W., Kangfu, Y., Pauls, K. P., Lee, E. and Navabi, A. 2013. Interaction of common bacterial blight quantitative trait loci in a resistant inter-cross population of common bean. Plant Breeding. Doi10.1111/pbr.12103

Page 185: Identification and Characterization of Common Bacterial

170

Ellendorff, U., Zhang, Z. and Thomma, B.P.H.J. 2008. Gene silencing to investigate the roles of receptor-like proteins in Arabidopsis. Plant Sign. Behav. 3(10):893-896

Ellis, J. G., Lawrence, G. J., Luck, J. E. and Dodds, P. N. 1999. Identification of regions of alleles of the flax rust resistance gene L that determine differences in gene-for-gene specificity. Pl Cell. 11: 495-506

Ellis, J., Dodds, P., Pryor, T. 2000. Structure, function and evolution of plant disease resistance genes. Curr. Opin. Plant Biol. 3: 278-284

Fabro, G., Steinbrenner, J., Coates, M., Ishaque, N., Baxter, L., Studholme, D. J., Körner, Allen, R. L., Piquerez, S. J. M., Cardoso-Rougon, A., Greenshields, D., Lei, R., Badel, J. L., Caillaud, M.-C., Sohn, K.-H., van den Acherveken, G., Parker, J. E., Beynon, J., and Jones, J. D. G. 2011. Multiple candidate effectors from the oomycete pathogen Hyaloperonospora arabidopsidis supress host plant immunity. PLOS Pathogens. 7(11): e1002348

Federici, L. Di Matteo, A., Fernandez-Recio, J., Tsernoglou, D. and Cervone, F. 2006. Polygalacturonase inhibiting proteins: players in plant innate immunity?. Trends in Plant Sci. 11:

Felix, G., Duran, J. D., Volko, S. and Boller. 1999. Plants have a sensitive perception system for the most conserved domain of bacterial flagellin. Plant J. 18(3): 265-276

Ferreira, J. J., Campa, A., and Kelly, J. D. 2013. Chapter 9: Organization of conferring resistance to anthracnose in common bean. Ed. Varshney, R. K. and Tuberosa, R. in Translational Genomics for Crop Breeding

Ferrier-Cana, E., Geffroy, V., Macadré, C., Creusot, F., Imbert-Bolloré, P., Sévignac, M. and Langin, R. 2003. Characterization of expressed NBS-LRR resistance gene candidates from common bean. Theor. Appl. Genet. 106:251-261

Finnegan, D. J. 1989. Eukaryotic transposable elements and genome evolution. Trends Genet. 5 (4): 103-107

Fotopoulos, V., De Tullio, M.C., Barnes, J. and Kanellis, A.K. 2008. Altered stomatal dynamics in ascorbate oxidase over-expressing tobacco plants suggest a role for dehydroascorbate signalling. J. Exp. Bot. 59(4): 729-737

Page 186: Identification and Characterization of Common Bacterial

171

Fourie, D. and Herselman, L. Pathogenic and genetic variation in Xanthomonas axonopodis pv. phaseoli and its fuscans variant in Southern Africa. African Crop Sci. 19(4): 393-407

Fraire-Velázquez, S. and Lozoya-Gloria, E. 2003. Differential early gene expression in Phaseolus vulgaris to Mexican isolates of Colletotrichum lindemuthianum in incompatible and compatible interactions. Phys Mol Plant Pathol. 63: 79-89

Frost, D., Way, H., Howles, P., Luck, J., Manners, J., Hardham, A., Finnegan, J. and Ellis, J. 2004. Tobacco transgenic for the flax rust resistance gene L expresses allel-specific activation of defense responses. Mol Plant Microbe Interact. 17: 224-232

Fu, D., Uauy, C., Distelfeld, A., Blechl, A., Epstein, L., Chen, X., Sela, H., Fahima, T., Dubcovsky, J. 2009. A kinase-START gene confers temperature-dependent resistance to wheat stripe rust. Science 323: 1357-1360

Gabriel, D.W., Kingsley, M.T., Hunter, J.E., Gottwyld, T. 1989. Reinstatement of Xanthomonas citri and X. phaseoli to species and reclassification of all X. campestris pv. citri strains. Int. J. Syst. Bacteriol. 39:14-22

Ganesh, D., Petitot, A.-S., Silva, M. C., Alary, R., Lecouls, A.-C. and Fernandez, D. 2006. Monitoring of the early molecular resistance responses of coffee (Coffea arabica L.) to the rust fungus (Hemileia vastatrix) using real-time quantitative RT-PCR. Plant Sci. 170:1045-1051

Geffroy, V., Sicard, D., De Oliviera, J.C.F., Sevignac, M., Cohen, S., Gepts, P., Neema, C., Langin, T. and Dron, M. 1999. Identification of an ancestral resistance gene cluster involved in the coevolution process between Phaseolus vulgaris and its fungal pathogen Colletotrichum lindemuthianum. Mol. Plant Microbe Interact. 12:774-784

Geffroy, V., Sévignac, M., De Oliviera, J. C. F., Skroch, P., Toquet, P., Gepts, P., Langin, T and Dron, M. 2000. Inheritance of partial resistance against Colletotrichum lindemuthianum in Phaseolus vulgaris and co-localization of QTLs with genes involved in specific resistance. Mol Plant-Microbe Interact. 13:287-296

Page 187: Identification and Characterization of Common Bacterial

172

Gepts, P., Aragão,F. J. L., de Barros, E., Blair, M. W., Brondani, R., Broughton, W., Galasso, I., Hernández, G., Kami, J., Lariguet, P., McClean, P., Melotto, M., Miklas, P., Pauls, P., Pedrosa-Harand, A., Porch, T., Sánchez, F., Sparvoli, F. and Yu, K. 2008. 5.1.1 “Economic, agronomic, and societal importance of phaseolus beans” in Genomics of Tropical Crop Plants edited by Moore, P. H. and Ming, R. Springer Science + Business Media, LLC, 233 Spring Street, New York, NY 10013, USA

Gepts, P. and Debouck, D. 1991. Origin, domestication and evolution of the common bean (Phaseolus vulgaris L.). In: van Schoonhoven, A. and Voysest, O. (Editors) Common beans: research for crop improvement. Wallingford/Colombia: CAB International/CIAT, pp 7-53.

Gepts, P. 1988. Phaseolin as an evolutionary marker. In ‘Genetic resources of Phaseolus beans’. (Ed. P Gepts) pp. 215-241. Kluwer Academic Publishers: Dordrecht, the Netherlands

Gepts, P. and Bliss F. A. 1986. Phaseolin variability among wild and cultivated common beans (Phaseolus vulgaris) from Columbia. Economic Botany. 40: 469-478

Gepts, P. and Bliss, F. A. 1984. Enhanced available methionine concentration associated with higher phaseolin levels in common bean seeds. Theor Appl. Genet. 69:47-53

Gillard, C.L., Conner, R.L., Howard R.J., Pauls, K.P., Shaw L. and Tar’an, B. 2009. The performance of dry bean cultivars with and without common bacterial blight resistance in field studies across Canada. Can J. Plant Sci. 89: 405-416

Glazebrook. J. 2005. Contrasting mechanisms of defense against biotrophic and necrotrohic pathogens. Annu Rev Phytopath. 43:205-227

Glémin, S. and Bataillon, T. 2009. A comparative view of the evolution of grasses under domestication. New Pytol. 183(2): 273-290

Gómez-Gómez, L., Felix, G. and Boller, T. 1999. A single locus determines sensitivity to bacterial flagellin in Arabidopsis thaliana. Plant J. 18(3): 277-284

Gómez-Gómez, L. and Boller, T. 2000. FLS2: An LRR receptor–like kinase involved in the perception of the bacterial elicitor flagellin in Arabidopsis. Cell Press.

Page 188: Identification and Characterization of Common Bacterial

173

Gómez-Gómez, L. and Boller, T. 2002. Flagellin perception: A paradigm for innate immunity. Trends in Pl Sci. 7(6): 251-256

Goodwin, M. 2005. Crop profile for dry bean in Canada. Agriculture and Agri-Food Canada

Goritschnig, S., Krasileva, K. V., Dahlbeck, D. and Staskawicz, B. J. 2012. Computational prediction and molecular characterization of an oomycete effector and the cognate Arabidopsis resistance gene. PLOS Genet. 8(2): e1002502

Gos, G., Slotte, T. and Wright, S. I. 2012.Signatures of balancing selection are maintained at disease resitance loci following mating system evolution and a population bottleneck in the genus Capsella. BMC Evol. Biol. 12:152

Graham, M., A., Marek, L.F., and Shoemaker, R.C. 2002. Organization, expression and evolution of a disease resistance gene cluster in soybean. Genet. 162: 1961-1977

Gunn, N., Byrne, J. and Holub E. B. 2002. Outcrossing of two homothallic isolates of Peronospora parasitica and segregation of avirulence matching six resistance loci in Arabidopsis thaliana. In Advances in Downy Mildew Research. Ed. Spencer-Phillips, P. T. N., Gisi, U. and Lebeda A. 185-188. Springer, Netherlands

Gupta, Y.P. 1987. Anti-nutritional and toxic factors in food legumes: a review. Plant Foods Hum Nutr. 37:201-228

Gutierrez, J.R., Balmuth, A.L., Ntoukakis, V., Mucyn, T.S., Gimenez-Ibanez, S., Jones, A. M.E. and Rathjen, J.P. 2009. Prf immune complexes of tomato are oligomeric and contain multiple Pto-like kinases that diversify effector recognition. Plant J. 61(3): 507-518

Hayashi, K. and Yoshida, H. 2009. Refunctionalization of the ancient rice blast disease resistance gene Pit by the recruitment of a retrotransposon. Plant J. 57:413-425

Hayward, A.C.1993. 1 The hosts of Xanthomonas, in Xanthomonas edited by Swing, J.G. and Civerolo, E.L., Chapman & Hall, 2-6 Boundary Row, London SE1 8HN, UK

Page 189: Identification and Characterization of Common Bacterial

174

Hernández-Pinzón, I., de Jesús, E., Santiago, N. and Casacuberta, J.M. 2009. The frequent transcriptional readthrough of the tobacco Tnt1 retrotransposon and its possible implications for the control of resistance genes. J. Mol. Evol. 68(3): 269-278

Huang, Q and Allen, C. 2000. Polygalacturonases are required for rapid colonization and full virulence of Ralstonia solanacereaum on tomato plants. Physiol. Mol. Plant Pathol. 57:77-83

Huard-Chauveau, C., Perchepied, L., Debieu, M., Rivas, S., Kroj, T., Kars, I., Bergelson, J., Roux, F. and Roby, D. 2013. An atypical kinase under balancing selection confers broad-spectrum disease resistance in Arabidopsis. PLoS Genet 9(9): e1003766. doi:10.1371/journal.pgen.1003766

Hulbert, S. H., Webb, C. A., Smith, S. M. and Sun Q. 2001. Resistance gene complexes: Evolution and utilization. Annu. Rev. Phytopathol. 39: 285-312

Hwang, C. F., Bhakta, A. V., Truesdell, G. M., Pudo, W. M., and Williamson, V. M. 2000. Evidence for a role of the N terminus and leucine-rich repeat region of the Mi gene product in regulation of localized cell death. Plant Cell. 12: 1319-1329

Hwang, B. H., Bae, H., Lim, H.-S., Kim, K. B., Kim, S. J., Im, M.-H., Park, B.-S., Kim, D. S. and Kim, J. 2010. Overexpression of polygalacturonase-inhibiting protein 2 (PGIP2) of Chinese cabbage (Brassica rapa ssp. pekinensis) increased resistance to the bacterial pathogen Pectobacterium carotovorum ssp. Carotovorum. Plant Cell Tiss Organ Cult. 103:293-305

Jaffe, D.B., Butler J., Gnerre, S., Mauceli, E., Lindblad-Toh, K., Mesirov, J. P., Zody, M.C. and Lander, E.S. 2003. Whole-genome sequence assembly for mammalian genomes: Arachne 2. Genome Research. 13:91-96

Jones, T. M., Anderson, A. J. and Albersheim, P. 1972. Host-pathogen interactions IV. Studies on the polysaccharide-degrading enzymes secreted by Fusarium oxysporum f. sp. lycopersici. Phys Plant Path. 2: 153-166

Jones, D.A., Thomas, C.M, Hammond-Kosack, K.E., Balint-Kurti, P.J., Jones, J.D.G. 1994. Isolation of the tomato Cf-9 gene for resistance to Cladosporium fulvum by transposon tagging. Science. 266:789-793

Page 190: Identification and Characterization of Common Bacterial

175

Jones D. A. and Jones, J. D. G. 1997. The role of leucine-rich repeat proteins in plant defenses. Adv. Bot. Res. 24:90-167

Jones, J. D. G. and Dangl, J. L. 2006. The plant immune system. Nature. 444(16): 323-329

Jorgensen. T.H. 2012. The effect of environmental heterogeneity on RPW8-mediated resistance to powdery mildews in Arabidopsis thaliana. Ann. Bot. doi:10.1093/aob/mcr320, available online at www.aob.oxfordjournals.org

Joshi, R.K. and Nayak, S. 2013. Perspectives of genomic diversification and molecular recombination towards R-gene evolution in plants. Physiol Mol Biol Plants. 19(1):1-9

Jung, G., Skroch, P.W., Coyne, D. P., Nienhuis, J., Arnaud-Santana, E., Ariyarathne, H.M., Kaepller, S. M., Bassett. 1997. Molecular-marker-based genetic analysis of tepary bean-derived common bacterial blight resistance in different developmental stages of common bean. J. Amer. Soc. Hort. Sci. 122(3): 329-337

Kajava, A.V. 1998. Structural diversity of leucine-rich repeat proteins. J. Mol Biol. 277 (3): 519-527

Kajava, A. V. and Kobe, B. 2002. Assessment of the ability to model proteins with leucine-rich repeats in light of the latest structural information. Prot. Science: 11: 1082-1090

Kamoun, S. and Kado, C.I. 1990. A plant-inducible gene of Xanthomonas campestris pv. campestris encodes an exocellular component required for growth in the host and hypersensitivity on nonhosts. J. Bacteriol. 172(9): 5165-5172

Kashkush, K., Feldman, M. and Levy, A.A. 2003. Transcriptional activation of retrotransposons alters the expression of adjacent genes in wheat. Nat Gen. 33(1):102-106

Kelley L. A. and Sternberg M. J. E. 2009. Proteins tructure prediction on the web: A case study using the Phyre server. Nature Protocols 4: 363-371

Kelly, J. D., and Vallejo, V. 2004. A comprehensive review of the major genes conditioning resistance to anthracnose in common bean. HortScience 39, 1196—1207.

Page 191: Identification and Characterization of Common Bacterial

176

Kim, K. D., Shin, J., H., Van, K., Kim, D. H. and Lee, S.-H. 2009. Dynamic rearrangements determine genome organization and useful traits in Soybean. Genome Anal. 151:1066-1076

Kobe B and Deisenhofer J. 1993. Crystal structure of porcine ribonuclease inhibitor, a protein with leucine-rich repeats. Nature. 366:751-756.

Kobe, B. and Kajava, A. W. 2001. The leucine-rich repeat as a protein recognition motif. Curr Op in Struct Biol. 11:725-732

Koch, M. F., and Mew, T. W. 1991. Effects of plant age and leaf maturity on the quantitative resistance of rice cultivars to Xanthomonas campestris pv. oryzae. Plant Disease. 75: 901-904

Krasileva, K. V., Dahlbeck, D., and Staskawicz, B. J. 2010. Activation of an Arabidopsis resistance protein is specified by the in planta association of its leucine-rich repeat domain with the cognate oomycete effector. Plant Cell. 22: 2444-2458

Krasileva, K. V., Zheng, C., Leonelli, L., Goritschnig, S., Dahlbeck, D., and Staskawicz, B. J. 2011. Global analysis of Arabidopsis/downy mildew interactions reveal prevalence of incomplete resistance and rapid evolution of pathogen recognition. PLoS ONE. 6(12): e28765

Kruijt, M., de Kock, M.J.D. and de Wit, P.J.G.M. 2005. Receptor-like proteins involved in plant disease resistance. Mol. Plant Pathol. 6(1):85-97

Kuykendall, D., Shao, J. and Trimmer. K. 2009. A nest of LTR retrotransposons adjacent the disease resistance-priming gene NPR1 in Beta vulgaris L. U.S. hybrid H20. Int J. Pl Gen. doi:10.1155/2009/576742

Larkan, N.J., Lydiate, D.J., Parkin, I.A.P., Nelson, M.N., Epp, D.J., Cowling, W.A., Rimmer, S.R. and Borhan, M.H. 2013. The Brassica napus blackleg resistance gene LepR3 encodes a receptor-like protein triggered by the Leptospaeria maculans effector AVRLM1. New Phytol. 197:595-605

Page 192: Identification and Characterization of Common Bacterial

177

Larsen, R.J.A. 2005. Discovery and utilization of molecular markers for genetic studies of common bacterial blight resistance and seed coat colour in Phaseolus vulgaris L. Thesis, University of Guelph

Lawrence, M.C., Suzuki, E., Varghese, J.N., Davis, P.C., Van Donkelaar, A., Tulloch and Colman, P.M. 1990. The three-dimensional structure of the seed storage protein phaseolin at 3 A resolution. EMBO J. 9(1): 9-15

Lawrence, M.C., Izard, T., Beuchat, M., Blagrove, R.J., and Colman, P.M. 1994. Structure of phaseolin at 2-2 A resolution: Implications for a common vicilinlegumin structure and the genetic engineering of seed storage proteins. J. Mol Biol. 238(5): 748-776

Le, Q., Melayah, D., Bonnivard, E., Petit, M., Grandbastien, M. 2007. Distribution dynamics of the Tnt1 retrotransposon in tobacco. Mol. Genet. Genomics. 278(6): 639-651

Leckie, F., Mattei, B., Capodicasa, C., Hemmings, A., Nuss, L., Aracri, B., De Lorenzo, G. and Cervone, F. 1999. The specificity of polygalacturonase-inhibiting protein (PGIP): a single amino acid substitution in the solvent-exposed β-strand/β-turn region of the leucine-rich repeats (LRRs) confers a new recognition capability. EMBO. 18(9): 2352-2363

Lee, S.-K, Song, M.-Y., Seo, Y.-S., Kim, H.-K., Ko, S., Cao, P.-J., Suh, J.-P., Yi, G., Roh, J.-H., Lee, S., An, G., Hahn, T.-R., Wang, G.-L., Ronald, P., and Jeon, J.-S. 2009. Rice Pi5-mediated resistance to Magnaporthe oryzae requires the presence of two coiled-coil-nucleotide-binding-leucine-rich repeat genes. Genetics. 181:1627-1638

Leister, D. 2004. Tandem and segmental gene duplication and recombination in the evolution of plant disease resistance genes. Trends in Gen. 20(3): 116-122

Lescot, M., Rombauts, S., Zhange, J., Aubourg, S., Mathé, C., Jansson, S., Rouzé, P. and Boerjan, W. 2004. Annotation of a 95-kb Populus deltoides, genomic sequence reveals a disease resistance gene cluster and novel class I and class II transposable elements. Theor Appl Genet. 109:10-22

Li, B., and Xu, X. 2002. Infection and development of apple scab (Venturia inaequalis) on old leaves. J. of Phytopath. 150: 687-691

Page 193: Identification and Characterization of Common Bacterial

178

Li, J. Wen, J., Lease, K.A., Doke, J.T., Tax, F. E. and Walker, J. C. 2002. BAK1, an Arabidopsis LRR receptor like protein kinase, interacts with BRI1 and modulates brassinosteroid signaling. Cell. 110(2): 213-222

Li, X., Lin, H., Zhang, W., Zou, Y., Zhang, J., Tang, X. and Zhou, J.-M. 2005. Flagellin induces innate immunity in nonhost interactions that is suppressed by Psuedomonas syringae effectors. PNAS. 102(36): 129990-122995

Liu, G., Kennedy, R., Greenshields, D.L., Peng, G., Forseille, L., Selvaraj, G. and Wei Y. 2007a. Detached and attached Arabidopsis leaf assays reveal distinctive defense responses against hemibiotrophic Colletotrichum spp. Mol Plant Microbe Interact. 20(10):1308-1319

Liu, J., Liu, X., Dai, L., and Wang, G. 2007b. Recent progress in elucidating the structure, function and evolution of disease resistance genes in plants. J. Genet and Genom. 34(9): 765-776

Liu, S.Y., Yu, K., Haffner, M., Park, S.J., Banik, M., Pauls, K.P. and Crosby, W. 2010. Construction of a BAC library and a physical map of a major QTL for CBB resistance of common bean (Phaseolus vulgaris L.). Genetica 138: 709-716

Lorković , Z.J., Wieczorek Kirk, D.A., Lambermon, M.H.L., and Filipowicz, W. 2000. Pre-mRNA splicing in higher plants. Trends Plant Sci. 5, 160–167.

Lu, D., Mu, S., Gao, Zhang, Y., Shan, L. and He, P. 2010. A receptor-like cytoplasmic kinase, BIK1, associates with a flagellin receptor complex to initiate plant innate immunity. PNAS. 107(1): 496-501

Luehrsen, K.R., S. Taha, and V. Walbot. 1994. Nuclear pre-mRNA splicing in higher plants. Prog. Nucleic Acids Res. Mol. Biol. 47: 149-193.

Ma, Y. and Bliss, F. A. 1978. Seed proteins of common bean. Crop Sci. 18: 431-437

Mackay, T. F. C. 2001. The genetic architecture of quantitative traits. Annu Rev. Genet. 35: 303-339

Page 194: Identification and Characterization of Common Bacterial

179

Maekawa, T., Cheng, W., Spiridon, L. N., Töller, A., Lukasik, E., Saijo, Y., Liu, P., Shen, Q.-H., Micluta, M. A., Somssich, I. E., Takken, F. L. W., Petrescu, A.-J., Chai, J. and Shulze-Lefert, P. 2011. Coiled-coil domain-dependent homodimerization of intracellular barley immune receptors defines a minimal functional module for triggering cell death. Cell Host & Microbe. 9:187-199

Mamidi, S. Rossi, M., Anam, D., Moghaddam, S., Lee, R., Papa, R. and McClean, P. 2011. Investigation of the domestication of common bean (Phaseolus vulgaris) using multilocus sequence data. Funct. Plant. Biol. 38: 953-967

Manfredini, C., Sicilia, F., Ferrari, S., Pontiggia, D., Salvi, G., Caprari, C., Lorito, M. and De Lorenzo, G. 2005. Polygalacturonase-inhibiting protein 2 of Phaseolus vulgaris inhibits BcPG1, a polygalacturonase of Botrytis cinerea important for pathogenicity, and protects transgenic plants from infection. Phys. Mol Plant Path. 67(2): 108-115

Marcel, T.C., Varshney, R.K., Barbieri, M., Jafary, H., de Kock, M.J.D., Graner, A., and Niks, R.E. 2007. A high-density consensus map of barley to compare the distribution of QTLs for partial resistance to Puccinia hordei. And of defence gene homologues. Theor. Appl. Genet. 114:487-500

Mazzola, M., Leach, J. E., Nelson, R. and White, F. F. 1994. Analysis of the interaction between Xanthomonas oryzae pv. oryzae and the rice cultivars IR24 and IRBB21. Phytopath. 84:392-397

McClean, P.E., Mamidi, S., McConnell, M., Chikara, S. and Lee, R. 2010. Synteny mapping between common bean and soybean reveals extensive blocks of shared loci. BMC Genomics 11:184

McDowell, J.M. and Meyers, B.C. 2013. A transposable element is domesticated for service in the plant immune system. PNAS. 110(37): 14821-14822

McHale, L., Tan, X., Koehl, P. and Michelmore, R. W. 2006. Plant NBS-LRR proteins: Adaptable guards. Gen. Biol. 7:212

Melotto, M., Underwood, W., Koczan, J., Nomura, K., He, W. Y. 2006. Plant stomata function in innate immunity against bacterial invasion. Cell. 126: 969-980

Page 195: Identification and Characterization of Common Bacterial

180

Messerschmidt, A., Rossi, A., Ladnstein, R., Huber, R., Bolognesi, M., Gatti, G., Marchesini, A., Petruzzelli, R. and Finazzi-Agró. 1989. X-ray crystal structure of the blue oxidase ascorbate oxidase from Zucchini: Analysis of the polypeptide fold and a model of the copper sites and ligands. J. Mol. Biol. 206:513-529

Messerschmidt, A., Luecke, H., and Huber, R. 1993. X-ray structures and mechanistic implications of three functional derivatives of ascorbate oxidase from zucchini: Reduced, peroxide and azide forms. J. Mol. Biol. 230:997-1014

Mestre, P. and Baulcombe, D. C. 2006. Elicitor-mediated oligomerization of the tobacco N disease resistance protein. Plant Cell. 18: 491-501

Meyers, B. C., Chin, D. B., Shen, K.A., Svaramakrishnan, S., Lavelle, D.O., Zhang, Z. and Michelmore, R. W. 1998. The major resistance gene cluster in lettuce is highly duplicated and spans several megabases. Plant Cell.15:1817-1832

Meyers, B. C., Morgante, M., and Michelmore, R.W. 2002. TIR-X and TIR-NBS proteins: Two new families related to disease resistance TIR-NBS-LRR proteins encoded in Arabidopsis and other plant genomes. Plant J. 32(1):77-92

Meyers, B. C., Kozik, A., Griego, A., Kuang and Michelmore, R. W. 2003. Genome-wide analysis of NBS-LRR-encoding genes in Arabidopsis. Plant Cell. 15: 809-834

Meyers, B.C., Kaushik, S. and Sekhar Nandety, R. 2005. Evolving disease resistance genes. Curr Op in Plant Biol. 8:129-134

Michelmore, R. W. and Meyers B. C. 1998. Clusters of resistance genes in plants evolve by divergent selection and a birth-and-death process. Genome Research. 8: 1113-1130

Michaels, T. E., Smith, T. H., Larsen, J., Beattie, A. D. and Pauls, K. P. 2006. OAC Rex common bean. Can. J. Plant Sci. 86: 733-736

Miklas, P. N, Beaver, J. S., Grafton, K. F. and Freytag, G. F. 1994. Registration of TARS VCI-4B multiple disease resistant dry bean germplasm. Crop Sci. 34: 1415

Page 196: Identification and Characterization of Common Bacterial

181

Miklas, P.N., Smth, J.R., Riley, R., Grafton, K.F., Singh, S.P., Jung, G. and Coyne, D.P. 2000a. Marker-assisted breeding for pyramided resistance to common bacterial blight in common bean. Annu Rep. Bean. Improv. Coop. 43: 39-40

Miklas, P. N., Stone, V., Daly, M. J., Stavely, J. R., Steadman, J. R., Bassett, M. J., Delorme, R., and Beaver, J. S. 2000b. Bacterial, fungal and viral disease resistance loci mapped in recombinant inbred common bean population ‘Dorado/Xan176). J. Am. Soc. Hortic. Sci. 125:476-481

Miklas. P. N., Coyne, D. P., Grafton, K. F., Mutlu, N., Reiser, J., Lindgren, T. and Singh, S. P. 2003. A major QTL for common bacterial blight resistance derives from the common bean great northern landrace cultivar Montana No. 5. Euphytica 131:137–146

Miklas P.N., Smith, J.R. and Singh, S.P. 2006. Registration of common bacterial blight resistant dark red kidney bean germplasm line USDK-CBB-15. Crop Sci. 46:1005-1006

Mitchell, A., Chang, H.-Y, Daugherty, L., Fraser, M., Hunter S., Lopez, R., McAnulla, C., McMenamin, C., Nuka, G., Pesseat, S., Sangrador-Vegas, A., Scheremetjew, M., Rato, C., Yong, S.-Y., Bateman, A., Punta, M., Attwood, T. K., Sigrist, C.J.A., Redaschi, N., Rivoire, C., Xenarios, I., Kahn, D., Guyot, D., Bork, P., Letunic, I., Gough, J., Oates, M., Haft, D., Huang, H., Natale, D.A., Wu, C.H., Orengo, C., Sillitoe, I., Mi, H., Thomas, P.D. and Finn, R.D. 2014. The InterPro protein families database: the classification resource after 15 years. Nuc. Acids Res. doi: 10.1093/nar/gku1243

Miyao, A., Tanaka, K., Murata, K., Sawaki, H., Takeda, S., Abe, K., Shinozuka, Y., Onosato, K., and Hirochika, H. 2003. Target site specificity of the Tos17 retrotransposon shows preference for insertion within genes and against insertion in retrotransposon-rich regions of the genome. Plant Cell. 15:1771-1780

Mkandawire, A.B.C., Mabagala, R.B., Guzmán, P., Gepts, P. and Gilbertson, R.L. 2004. Genetic diversity and pathogenic variation of common blight bacteria (Xanthomonas campestris pv. phaseoli and X. campestris pv. phaseoli var. fuscans) suggests pathogen coevolution with the common bean. Phytopathol. 94(6): 593-603

Moffett. P., Farnham, G., Peart, J. and Baulcombe, D. C. 2002. Interaction between domains of a plant NBS-LRR protein in disease resistance-related cell death. EMBO. 21(17): 4511-4519

Page 197: Identification and Characterization of Common Bacterial

182

Mohan, S.T. 1982. Evaluation of Phaseolus coccineus Lam. germplasm for resistance to common bacterial blight of bean. Turrialba 32:489-490

Monosi, B., Wisser, R.J., Pennill, L. and Hulbert, S.H. 2004. Full-genome analysis of resistance gene homologues in rice. Theor Appl Genet. 109: 1434-1447

Mucyn, T.S., Clemente, A., Andriotis, V.M., Balmuth, A.L., Oldroyd, G.E., Staskawicz, B.J. and Rathjen, J.P. 2006. The tomato NBARC-LRR protein Prf interacts with Pto kinase in vivo to regulate specific plant immunity. Plant Cell. 18:2792-2806

Murthy, K., Deb, A., Goonesekera, S., Szabó and Salzman, A. 2003. Identification of conserved domains in Salmonella muenchen flagellin that are essential for its ability to activate TLR5 and to induce an inflammatory response in vitro. J. Biol Chem. 279: 5667-5675

Muthamilarasan M., and Prasad, M. 2013. Plant innate immunity: An updated insight into defense mechanism. J. Biosci. 38:433-449

Mutlu, N., Vidaver, A. K., Coyne, D. P., Steadman, J. R., Lambrecht, P.A. and Reiser, J. 2008. Differential pathogenicity of Xanthomonas campestris pv. phaseoli and X. fuscans subsp. fuscans strains on bean genotypes with common blight resistance. APS. 92(4): 546-554

Noël, L. Moores, T.L., van Der Bizen, E.A., Parniske, M., Daniels, M.J., Parker, J.E. and Jones, J.D. 1999. Pronounced intraspecific haplotype divergence at the RPP5 complex disease resistance locus of Arabidopsis. Plant Cell. 11(11): 2099-2112

Nüremberger, T., Brunner, F., Kemmerling, B. and Piater, L. 2004. Innate immunity in plants and animals: striking similarities and obvious differences. Immunol. Rev. 198:249-266

Opio, A. F., Allen, D. J. and Teri, J. M. 1996. Pathogenic variation in Xanthomonas campestris pv. phaseoli, the causal agent of common bacterial blight in Phaseolus beans. Plant Path. 45 (6): 1126-1133

Page 198: Identification and Characterization of Common Bacterial

183

Ori N, Eshed Y, Paran I, Presting G, Aviv D, Tanksley S, Zamir D and Fluhr R .1997. The I2C1 family from the wilt disease resistance locus I2 belong to the nucleotide binding, leucine-rich repeat superfamily of plant resistance genes. Plant Cell 9:521–532

O'Rourke, J.A., Iniguez, L. P., Fu, F., Bucciarelli, B., Miller, S.S., Jackson, S.A., McClean, P.E., Li, J., Dai, X., Zhao, P.X., Hernandez, G. and Vance, C.P. 2014. An RNA-Seq based gene expression atlas of the common

bean. BMC Genomics. 15:866.

Padmanabhan, M., Cournoyer, P. and Dinesh-Kumar, S. P. 2009. The leucine-rich repeat domain in plant innate immunity: a wealth of possibilities. Cell. Micro. 29(2): 191-198

Park, S. J. and Dhanvantari, B. N. 1987. Transfer of common blight (Xanthomonas campestris pv. phaseoli) resistance from Phaseolus coccineus Lam. to P. vulgaris L. through interspecific hybridization. Can. J. Plant Sci. 67:685-695

Parker, J. P. K. 1985. Interspecific transfer of common bacterial blight resistance from Phaseolus acutifolius A. Gray to Phaseolus vulgaris L. M.Sc. thesis, University of Guelph, Guelph, ON

Parker, J. E., Coleman, M. J., Szabò, V., Frost, L. N., Schmidt, R., van der Biezen, E., Moores, T., Dean, C. Daniels, M. J., and Jones, J. D. G. 1997. The Arabidopsis downy mildew resistance gene RPP5 shares similarity to the Toll and interleukin-1 receptors with N and L6. Plant Cell. 9:879-894

Parniske, M., Hammond-Kosack, K. E., Golstein, C., Thomas, C. M., Jones, D. A., Harrison, K., Wulff, B. B. H. and Jones, J. D. G. 1997. Novel disease resistance specificities result from sequence exchange between tandemly repeated genes at the Cf-4/9 locus of tomato. Cell. 91: 821-832

Pastori, G.M., Kiddle, G., Antoniw, J., Bernard, S., Veljovic-Jovanovic, S., Verrier, P.J., Noctor, G., Foyer, C.H. 2003. Leaf vitamin C contents modulate plant defense transcripts and regulate genes that control development through hormone signaling. Plant Cell.15: 939-951

Pavet, V., Olmos, E., Kiddle, G., Mowla, S, Kumar, S., Antoniw, J., Alvarez, M.E. and Foyer, C.H. 2005. Ascorbic acid deficiency activates cell death and disease resistance responses in Arabidopsis thaliana. Plant Physiol. 139:1291-1303

Page 199: Identification and Characterization of Common Bacterial

184

Pedraza, F., Gallego, G., Beebe, S., and Tohme, J. 1997. Marcadores SCAR y RAPD para la resitencia a la bacteriosis comun (CBB). In: Singh S.P., Voysest O., (eds) Taller de mejoramiento de frijol para el Siglo XXI: bases para una estrategia para America Latina. CAT, Cali, pp 130-134

Perry, G. 2010. Common bacterial blight resistance in Phaseolus vulgaris: An examination of the major resistance QTLs of OAC-Rex and HR67, Ph.D. thesis, University of Guelph, Guelph ON

Perry, G., DiNatale, C., Xie, W., Navabi, A., Reinprecht, Y., Crosby, W., Yu, K., Shi, C. and Pauls, K. P. 2013. A comparison of the molecular organization of genomic regions associated with resistance to common bacterial blight in two Phaseolus vulgaris genotypes. Frontiers in Plant Sci. doi: 10.3389/fpls.2013.00318

Petersen, T. N., Brunak, S., von Heijne, G. and Nielsen, H. 2011. Signal 4.0: Discriminating signal peptides from transmembrane regions. Nature Methods. 8: 785-786

Pignocchi, C., Kiddle, G., Hernández, I., Foster, S.J., Asensi, A., Taybi, T., Barnes, J. and Foyer, C.H. 2006. Ascorbate oxidase-dependent changes in the redox state of the apoplast modulate gene transcript accumulation leading to modified hormone signaling and orchestration of defense processes in tobacco. Plant Phys. 141(2): 423-435

Protsenko, M. A., Buza, N. L., Krinitsyna, A. A., Bulanseva, E. A. and Korableva, N. P. 2008. Polygalacturonase-inhibiting protein is a structural component of plant cell wall. Biochem. 73 (10): 1317-1329

Pulse Canada. 2007. Canada and the global pulse industry. Retrieved October 7 2014 from the World Wide Web. http://www.pulsecanada.com/uploads/wv/nr/wvnr5qGWIi8Rqbl22UyY8g/Canadas-Pulse-Industry-in-the-Global-Market-Sep-07.pdf

Pulse Canada. 2015. Canada’s growing regions. Retrieved June 24, 2015 from the World Wide Web. http://www.pulsecanada.com/canadas-growing-regions

Rademaker, J.L.W., Louws, F.J., Schultz, M.H., Rossbach., U., Vauterin, L., Swings, J. and de Bruijn, F.J. 2005. A comprehensive species to strain taxonomic framework for Xanthomonas. Phytopathol. 95:1098-1111

Page 200: Identification and Characterization of Common Bacterial

185

Rafiqi, M., Bernoux, M., Ellis, J. G. and Dodds, P. N. 2009. In the trenches of plant pathogen recognition: Role of NB-LRR proteins. Sem Cell & Dev Biol. 20: 1017-1024

Rairdan, G. J., Collier, S. M., Sacco, M. A., Baldwin, T. T., Boettrich, and Moffett, P. 2008. The coiled-coil and nucleotide binding domains of the potato Rx disease resistance protein function in pathogen recognition and signaling. Plant Cell. 20:739-751

Randhawa, P.S. and Civerolo, E. L. 1985. A detached-leaf bioassay for Xanthomonas campestris pv. pruni. Phytopath. 75:1060-1063

Rehmany, A. P. Gordon, A., Rose, L. E., Allen, R. L., Armstrong, M. R., Whisson, S. C., Kamoun, S., Tyler, B. M., Birch, P. R. J. and Beynon, J. L. 2005. Differential recognition of highly divergent downy mildew avirulence gene alleles by RPP1 resistance genes from two Arabidopsis lines. Plant Cell 17: 1839-1850

Remeeus, P.M. and Sheppard, J. W. 2006. Proposal for a new method for detecting Xanthomonas axonopodis pv. phaseoli on bean seeds. ISTA Method Validation Reports. 3:1-11

Rodríguez, A., Shimada, T., Cervera M., Alquézar, B., Gadea, J., Góme-Cadenas, A., De Ollas, C.J., Rodrigo, M.J., Zacarías, L. and Peña, L. 2014. Terpene down-regulation triggers defense responses in transgenic orange leading to resistance against fungal pathogens. Plant Physiol. 164(1):321-339

Rudolph, K. Infection of the plant by Xanthomonas. 1993. In: Swings, J.G., Civerolo, E.L., editors. Xanthomonas. London: Chapman and Hall. P. 193-264

Ryals, J. A., Neuenschwander, U. H., Willits, M. G., Molina, A., Steiner, H.-Y. and Hunt, M. D. 1996. Systemic acquired resistance. Plant Cell. 8: 1809-1819

Saettler, A. W. 1989. Chapter 11: Common bacterial blight. In Bean production problems in the tropics. Eds. Howard F. Schwartz, Marcial A. Pastor Corrales. CIAT. Pp 261-283

Sakamoto, T., Bonnin, E., and Thibault, J.-F. 2003. A new approach for studying interaction of the polygalacturonase-inhibiting proteins with pectins. Biochim. Biophys. Acta. 1621:280-284

Page 201: Identification and Characterization of Common Bacterial

186

Sanmartin, M., Pateraki, I., Chatzopoulou, F., and Kanellis, A.K. 2007. Differential expression of the ascorbate oxidase multigene family during fruit development and in response to stress. Planta. 225: 873-885

Saxton, A.M. 2000. PDMIX800. University of Tennessee, Knoxville TN.

Schaad, N. W., Postnikova, E., Lacy, G. H., Sechler, A., Agarkova, I., Stromberg, P.E., Stromberg, V.K. and Vidaver, A.K. 2006. Emended classification of xanthomonad pathogens on citrus. Syst. Appl. Microbiol. 29:690-695

Schmutz, J., Cannon, S. B., Schlueter, J., Ma, J., Mitros, T., Nelson, W., Hyten, D. L., Song, Q., Thelen, J.J., Cheng, J., Xu, D., Hellsten, U., May, G. D., Yu, Y., Sakurai, T., Umezawa, T., Bhattacharyya, M.K., Sandhu, D., Valiyodan, B., Lindquist, E., Peto, M., Grant, D., Shu, S., Goodstein, D., Barry, K., Futrell-Griggs, M., Abernathy, B., Du, J., Tian, Z., Zhu, L., Gill, N., Joshi, T., Libault, M., Sethuraman, A., Zhang, X.-C., Shinozaki, K., Nhuyen, H.T., Wing, R.A., Cregan, P., Specht, J., Grimwood, J., Rokshar, D., Stacey, G., Shoemaker, R.C. and Jackson, S.A. 2010. Genome sequence of the paleopolyploid soybean. Nature. 463:doi:10.1038/nature08670

Schmutz J, McClean PE, Mamidi S, Wu GA, Cannon SB, Grimwood J, Jenkins J, Shu S, Song Q, Chavarro C, Torres-Torres M, Geffroy V, Moghaddam SM, Gao D, Abernathy B, Barry K, Blair M, Brick MA, Chovatia M, Gepts P, Goodstein DM, Gonzales M, Hellsten U, Hyten DL, Jia G,et. al., 2014 A reference genome for common bean

and genome-wide analysis of dual domestications, Nature Genetics, published online 2014

June 08

Schuleter JA, Goicoechea JL, Collura K, Gill N, Lin JY, Yu Y, Kundra D, Zuccolo A, Vallejos CE, Munoz-Torres M, Blair M, Tohme J, Tomkins J, McClean P, Wing RA and Jackson SA (2008). BAC-end-sequence analysis and a draft physical map of the common bean (Phaseolus vulgaris L.). Tropical Plant Biology 1:40-48.

Scott,M.E. and T.E. Michaels. 1992. Xanthomonas resistance of Phaseolus interspecific cross selections confirmed by field performance. HortScience 27:348-350

Shi, C., Navabi, A. and Yu, K. 2011. Association mapping of common bacterial blight resistance QTL in Ontario bean breeding populations. BMC Plant Biol. 11: 52

Page 202: Identification and Characterization of Common Bacterial

187

Shi C., Yu, K., Xie, W., Perry, G., Navabi, A., Pauls, K.P., Miklas, P. N. and Fourie, Deidre. 2012. Development of candidate gene markers associated to common bacterial blight resistance in common bean. Theor Appl Genet. 125(7):1525-1537

Schuster, M.L., Coyne, D.P., Behre, T., and Leyna, H. 1983. Sources of Phaseolus species resistance and leaf and pod differential reactions to common blight. HortSci. 18:901-903

Sicilia, F., Fernandez-Recio, J., Caprari, C., De Lorenzo, G., Tsernoglou, D., Cervone, F. and Federici, L. 2005. The polygalacturonase-inhibiting protein PGIP2 of Phaseolus vulgaris has evolved a mixed mode of inhibition of endopolygalacturonase PG1 of Botrytis cinerea. Plant Phys. 139:1380-1388

Singh, S. P. 1982. Key for identification of different growth habits of Phaseolus vulgaris L. Centro Internacional de Agricultura Tropical, Apartado Aéreo 6713, Cali Colombia

Singh, S. P., Gepts, P. and Debouck, D. G. 1991. Races of common bean (Phaseolus vulgaris, Fabaceae). Economic Botany. 45(3): 379-396

Singh, S. P. and Muñoz, C. G. 1999. Resistance to common bacterial blight among Phaseolus species and common bean improvement. Crop Sci 39: 80-89

Smirnoff, N. 2000. Ascorbic acid: Metabolism and functions of a multi-facetted molecule. Curr. Opin. Plant Biol. 3:229-235

Song, W.-Y., Pi, L.-Y., Wang, G.-L., Gardner, J., Holsten, T., and Ronald, P.C. 1997. Evolution of the rice Xa21 disease resistance fene family. Plant Cell. 9:1279-1287

Song, J., Bradeen, J.M., Naess, S.K., Raasch, J.A., Wielgus, S.M., Haberlach, G.T., Liu, J., Kuang, H., Austin-Phillips, S., Buell, C.R., Helgeson, J.P. and Jiang, J. 2003. Gene RB cloned from Solanum bulbocastanum confers broad spectrum resistance to potato late blight. PNAS. 100: 9128-9133

Song, D., Li, G., Song, F., and Zheng, Z. 2008. Molecular characterization and expression analysis of Os BISERK1, a gene encoding a leucine-rich repeat receptor-like kinase, during disease resistance responses in rice.

Page 203: Identification and Characterization of Common Bacterial

188

Song W, Han ZF, Sun YD, and Chai JJ. 2014. Crystal structure of a plant leucine rich repeat protein with two island domains. Sci China Life Sci, 57: 137–144

St. Clair, D. A. 2010. Quantitative disease resistance and quantitative resistance loci in breeding. Annu. Re. Pytopathol. 48: 247-268

Sun, X., Cao, Y., Yang, Z., Xu, C., Li, X., Wang, S. and Zhang, Q. 2004. Xa26, a gene conferring resistance to Xanthomonas oryzae pv. oryzae in rice, encodes an LRR receptor kinase-like protein. The Plant Journal. 37: 517-527

Sun, W., Dunning, F.M., Pfund, C., Weingarten, R., and Bent. 2006. Within-species flagellin polymorphism in Xanthomonas campestris pv. campestris and its impact on elicitation of Arabidopsis flagellin sensing2-dependent defenses. Plant Cell. 18:764-779

Taguchi, F., Shimizu, R., Nakajima, R., Toyoda, K., Shiraishi, T. and Ichinose Y. 2003. Differential effects of flagellins from Pseudomonas syringae pv. tabaci, tomato and glycinea on plant defense response. Plant Phys. Bioch. 41(2): 165-174

Tameling, W.I., Elzinga, S.D., Darmin, P.S., Vossen, J.H., Takken, F.L., Haring, M.A., and Cornelissen, B.J. 2002. The tomato R gene products I-2 and Mi-1 are functional ATP binding proteins with ATPase activity. Plant Cell 14: 2929–2939.

Tang, P., Zhang, Y., Sun, X., Tian, D., Yang, S. and Ding, J. 2010. Disease resistance signature of the leucine-rich repeat receptor-like kinase genes in four plant species. Plant Science. 179: 399-406

Tao, Y., Xi, Z., Chen, W., Glazebrook, J., Chang, H.-S., Han, B., Zhu, T., Zou, G., and Katagiri, F. 2003. Quantitative nature of Arabidopsis responses during compatible and incompatible interactions with the bacterial pathogen Pseudomonas syringae. Plant Cell. 15: 317-330

Tar’an, B., Michaels, T. E. and Pauls, K. P. 2001. Mapping genetic factors affecting the reaction to Xanthomonas axonopodis pv. phaseoli in Phaseolus vulgaris L. under field conditions. Genome. 44: 1046-1056

Tornero, P., Mayda, E., Gómez, M. D., Cañas, L., Conejero, V. & Vera, P. 1996. Characterization of LRP, a leucine-rich repeat (LRR) protein from tomato plants that is processed during pathogenesis. Plant J. 10(2): 315-330

Page 204: Identification and Characterization of Common Bacterial

189

Torres, M. A., Jones, J. D. and Dangl, J. L. 2005. Pathogen-induced, NADPH oxidase derived reactive oxygen intermediates suppress spread of cell death in Arabidopsis thaliana. Nature Genet. 37: 1130–1134

Tosh, S. M. and Yada, S. 2010. Dietary fibres in pulse seeds and fractions: Characterization, functional attributes, and applications. Food Res Int. 43: 450-460

Tsuchiya, T. and Eulgem, T. 2013. Mutations in EDM2 selectively affect silencing states of transposons and induce plant developmental plasticity. Nature Scientific Reports. 3: 1701 DOI: 10.1038/srep01701

Urrea, C.A., Miklas, P.N. and Beaver, J.S. 1999. Inheritance of resistance to common bacterial blight in for tepary bean lines. J. Amer. Soc. Hort. Sci. 124(1): 24-27

Vallejos, C. E., Astua-Monge, G., Jnes, V., Plyler, T.R., Sakiyama, N. S., and Mackenzie, S. A. 2006. Genetic and molecular characterization of the I locus of Phaseolus vulgaris. Gen. 172:1229-12242

Van der Biezen, E. A., Freddie, C. T., Kahn, K., Parker, J. E. and Jones, J. D. G. 2002.Arabidopsis RPP4 is a member of the RPP5 multigene family of TIR-NBSLRR genes and confers downy mildew resistance through multiple signalling components. Plant J. 29(4): 439-451

Van Gijsefem, F., Gough, C., Zischek, C., Niqueux, E., Ariat, M., Genin, S., Barberis, P., German, S., Castello, P. and Boucher, C.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

Vandemark, G. J., Fourie. D. and Miklas, P. N. 2008. Genotyping with real-time PCR reveals recessive epistasis between independent QTL conferring resistance to common bacterial blight in dry bean. Theor Appl Genet. 117:513-522

Vandemark, G. J., Fourie, D., Larsen, R. C. and Miklas, P. N. 2009. Interactions between QTL SAP6 and SU91 on resistance to common bacterial blight in red kidney bean and pinto bean populations. Euph. 170:371-381

Page 205: Identification and Characterization of Common Bacterial

190

Vauterin, L., Hoste, B., Kersters, K. and Swings, J. 1995. Reclassification of Xanthomonas. IJSEM. 45(3):472-489

Veronese, P., Nakagami, H., Bluhm, B., AbuQamar, S., Chen, X., Salmeron, J., Dietrich, R.A., Hirt, H. and Mengiste, T. 2006. The membrane-anchored Botrytis-Induced Kinase1 plays distinct roles in Arabidopsis resistance to nectrotrophic and biotrophic pathogens. Plant Cell. 18(1):257-273

Vidaver, A.K. 1993. 1.5 Xanthomonas campestris pv. phaseoli: Cause of common bacterial blight of bean, in Xanthomonas edited by Swing, J.G. and Civerolo, E.L., Chapman & Hall, 2-6 Boundary Row, London SE1 8HN, UK

Wang, Z.-X., Yano, M., Yamanouchi, U., Iwamoto, M., Monna, L., Hayasaka, H., Katayose, Y. and Sasaki, T. 1999. The Pib gene for rice blast resistance belongs to the nucleotide binding and leucine-rich repeat class of plant disease resistance genes. Plant J. 19(1):55-64

Wang, L., Rong, W. and He, C. 2008a.Two Xanthomonas extracellular polygalacturonases, PghAxc and PghBxc, are regulated by type III secretion regulators HrpX and HrpG and are required for virulence. MPMI 21(5): 555-563

Wang, G., Ellendorff, U., Kemp, B., Mansfield, J.W., Forsyth, A., Mitchell, K., Bastas, K., Liu, C.-M., Woods-Tör, A., Zipfel, C., de Wit, P.J.G.M., Jones, J.D.G., Tör, M. and Thomma, B.P.H.J. 2008b. A genome-wide functional investigation into the roles of receptor-like proteins in Arabidopsis. Plant Physiol. 147:503-517

Wang, Y.-H. and Warren, J. T. 2010. Mutations in retrotransposon AtCOPIA4 compromises resistance to Hyaloperonospora parasitica in Arabidopsis thaliana. Gen Mol Biol. 33(1): 135-140

Weller, D. M. and Saettler, A. W. 1976. Chemical control of common and fuscous bacterial blights (Xanthomonas phaseoli, Xanthomonas phaseoli fuscans) in Michigan navy (pea) beans. Plant Dis. Rept. 60: 793-797

Wicker, T., Sabot, F., Hua-Van, A., Bennetzen, J. L., Capy, P., Chalhoub, B., Flavell, A., Leroy, P., Morgante, M., Panaud, O., Paux, E., SanMiguel, P. and Schulman A. H. 2007. A unified classification system for eukaryotic transposable elements. Nat Rev Gen. 8:973-982

Page 206: Identification and Characterization of Common Bacterial

191

Xiao, S., Brown, S., Patrick, E., Brearley, C. and Turner, J. G. 2003. Enhanced transcription of the Arabidopsis disease resistance Genes RPW8.1 and RPW8.2 via a salicylic acid-depedent amplification circuit is required for hypersensitive cell death. Plant Cell.15: 33-45

Xie, W., Yu, K., Pauls, K.P. and Navabi, A. 2012. Application of image analysis in studies of quantitative disease resistance, exemplified using common bacterial blight-common bean pathosystem. Phytopath. 102(4): 434-442

Xu, Z. and Wang, H. 2007. LTR_FINDER: An efficient tool for the prediction of full-length LTR retrotransposons. Nucleic Acid Res. 35: doi:10.1093/nar/gkm286

Yañez, E., Zacarias, I., Aguayo, M., Vasquez, M. and Guzman, E. 1995. Nutritive value evaluated on rats of new cultivars of common beans (Phaseolus vulgaris) released in Chile. Plant Foods for Human Nutrition. 47: 301-307

Yang, X., Deng, F. and Ramonell, K.M. 2012. Receptor-like kinases and receptor-like proteins: Keys to pathogen recognition and defense signaling in plant innate immunity. Front Biol. 7(2): 155-166

Yi, H. and Richards, E.J. A cluster of disease resistance genes in Arabidopsis is coordinately regulated by transcriptional activation and RNA silencing. Plant Cell. 19(9): 2929-2939

Yu, K., Park, S. J., Zhang, B., Haffner, M and Poysa, V. 2004. An SSr marker in the nitrate reductase gene of common bean is tightly linked to a major gene conferring resistance to common bacterial blight. Euphytica. 138: 89-95

Yu, K., Park S.J., Poysa, V. and Gepts, P. 2000a. Integration of simple sequence repeat markers into a molecular linkage map of common bean (Phaseolus vulgaris L.). Am. Genet. Assoc. 91:429-434

Yu, K., Park, S.J. and Poysa, V. 2000b. Marker-assisted selection of common beans for resistance to common bacterial blight: efficiency and economics. Plant Breed. 119: 411-416

Zaumeyer, W.J. and Thomas, H.R. 1957. A monographic study of bean diseases and methods for their control. United States Department of Agriculture Technical Bulletin 868, 255pp.

Page 207: Identification and Characterization of Common Bacterial

192

Zhang, Z. and Thomma, B.P.H.J. 2013. Structure-function aspects of extracellular leucine-rich repeat-containing cell surface receptors in plants. J. Integr Pant Biol. 55(12):1212-1223

Zhou, F., Menke, F. L., Yoshioka, K., Moder, W., Shirano, Y. and Klessig, D. F. 2004a. High humidity suppresses ssi4-mediated cell death and disease resistance upstream of MAP kinase activation, H2O2 production and defense gene expression. Plant J. 39: 920-932

Zhou, T., Wang, Y., Chen, J.-Q., Araki, H., Jing, Z., Jiang, K., Shen, J., Tian, D. 2004b. Genome-wide identification of NBS genes in japonica rice reveals significant expansion of divergent non-TIR NBS-LRR genes. Mol Gen Genomics. 271:402-415

Zipfel, C. and Felix, G. 2005. Plants and animals: A different taste for microbes? Curr. Opin Plant Biol. 8:353-360

Zipfel, C., Kunze, G., Chinchilla, D., Caniard, A., Jones, J. D., Boller, T., and Felix, G. 2006. Perception of the bacterial PAMP EF-Tu by the receptor EFR restricts Agrobacterium-mediated Transformation. Cell. 125: 749-760

Zipfel, C. 2008. Pattern-recognition receptors in plant innate immunity. Curr. Op. in Immunology. 20:10-15

Page 208: Identification and Characterization of Common Bacterial

193

7.0 Appendices

Gene7 ------------------------------------------------------------

G2 ------------------------------------------------------------

Glyma.08G361500 MSSSRGGAGPSSEAPPPRRIMRTQTAGNLGESVIDSEVVPSSLVEIAPILRVANEVEKTH

Gene7 ------------------------------------------------------------

G2 ------------------------------------------------------------

Glyma.08G361500 PRVAYLCRFYAFEKAHRLDPNSSGRGVRQFKTALLQRLERENDPTLKGRVKKSDAREMQS

Gene7 ---------------------------------------MAAWQATFTKPSRNPYATFTN

G2 ---------------------------------MPPWLPDRCDCPPSGPEYIPDVLELSV

Glyma.08G361500 FYQHYYKKYIQALQNAADKADRAQLTKAYNTANVLFEVLKAVNMTQSMEVDREILETQDK

.

Gene7 PSPDLRSSDLHEILTLTPLSAIAASVQG--------------------------------

G2 EFLLGCSSDLHEILTLTPLSAIAASVQG--------------------------------

Glyma.08G361500 VAEKTEILVPYNILPLDPDSANQAIMRFPEIQAAVYALRNTRGLPWPKDFKKKKDEDILD

::**.* * ** * ::

Gene7 ------------------------------------------------------------

G2 ------------------------------------------------------------

Glyma.08G361500 WLGSMFGFQKHNVANQREHLILLLANVHIRQFPKPDQQPKLDERALTEVMKKLFKNYKKW

Gene7 ------------------------------------------------------------

G2 ------------------------------------------------------------

Glyma.08G361500 CKYLGRKSSLWLPTIQQEVQQRKLLYMGLYLLIWGEAANLRFMPECLCYIYHHMAFELYG

Gene7 ------------------------------------------------------------

G2 ------------------------------------------------------------

Glyma.08G361500 MLAGNVSPMTGENVKPAYGGEDEAFLRKVVTPIYNVIAKEAARSKKGRSKHSQWRNYDDL

Gene7 --CFR-------------------------------------------------------

G2 --CFRSVDCFRLGWPMRADADFFCWPVEQLNFDKSN------------------------

Glyma.08G361500 NEYFWSADCFRLGWPMRADADFFCLPAEKLVFDKSNDDKPPSRDRWVGKVNFVEIRSFWH

*

Gene7 ----------------------------DPSAIIDVNVFKKVLNVFIT-AILKLGQ----

G2 -----------------AMIIVAWNGTGDPSAIIDVNVFKKVLNVFIT-AILKLGQAILD

Glyma.08G361500 MFRSFDRMWSFFILCLQAMIVVAWNGSGDPSAIFNGDVFKKVLSVFITAAILKFGQAVLD

*****:: :******.**** ****:**

Gene7 ------------------------------------------------------------

G2 VILSCKSQWSMSMHVKLGYILKVVSAAAWVIVLSVSYAYTWENPPRFAQTIQSWFGSNSN

Glyma.08G361500 VILSWKAQWSMSLYVKLRYILKVVSAAAWVIVLSVTYAYTWDNPPGFAQTIKSWFGSGGS

N-Terminal

1,3-β Glucan Synthase Subunit FKS1-like

Page 209: Identification and Characterization of Common Bacterial

194

Gene7 -------------------------------------------------GSLDSTRGMHE

G2 SPSFFIMVVVVYLSPNMLAAMLFLFPLIHR-----------------FLESLDSTRGMHE

Glyma.08G361500 SAPSLFILAVVVYLSPNMLAAIFFLIPFIRRHLERSNYRIVMLMMWWSQPRLYVGRGMHE

* *****

Gene7 STFSLFKK----------------------------------------------------

G2 SIFSLFKK----------------------------------------------------

Glyma.08G361500 SAFSLFKYTMFWVLLIITKLAFSYYIEIKPLVGPTKAIMSVKITTFQWHEFFPHARNNIG

* *****

Gene7 ------------------------------------------------------------

G2 ------------------------------------------------------------

Glyma.08G361500 VVIALWAPIILVYFMDTQIWYAIFSTLFGGIYGAFRRLGEIRTLGMLRSRFQSLPGAFNA

Gene7 ------------------------------------------------------------

G2 ------------------------------------------------------------

Glyma.08G361500 SLIPEETNEPKKKGLKATLSRRFPEISSNKGKEAARFAQLWNQIITSFRDEDLINDREMN

Gene7 ------------------------------------------------------------

G2 ------------------------------------------------------------

Glyma.08G361500 LLLVPYWADTQLDLIQWPPFLLASKIPIALDMAKDSNGKDRELKKRIAADNYMSCAVREC

Gene7 -SSLCSPIKYGP---------------------------------------QCLSLCTKK

G2 -SSLCSPIKYGP---------------------------------------QCLSLCTKK

Glyma.08G361500 YASFKSIIKHLVQGEREIPVIEYMFDEVDKNIETDKLISEFRMSALPSLYAQFVELTQYL

:*: * **: * :.*

Gene7 MKLKVRDAHNLLIGIQ----------VVVNYDDNFHEPDVASG----------APSPNQN

G2 MKLKVRDAHNLLIGIH----------VVVNYDDNFHEPDVASG----------APSPNQN

Glyma.08G361500 LNNDPKDRDNVVILFQDMLEVVTRDIMMEDQDQIFSLVDSSHGGTGHEGMLHLEPEPHHQ

:: . :* .*::* :: :: : *: * * : * *.*:::

Gene7 IRSSGGSKTN--------------------------------------------------

G2 IRSTGGSKTN--------------------------------------------------

Glyma.08G361500 LFASEGAIKFPIEPLTAAWTEKIKRLHLLLTTKESAMDVPSNLEARRRISFFSNSLFMDM

: :: *: .

Gene7 ------------------------------------------------------------

G2 ------------------------------------------------------------

Glyma.08G361500 PMAPKVRNMLSFSVLTPYYTEEVLFSLHDLDSQNEDGVSILFYLQKIYPDEWNNFLERVK

Gene7 ------------------------------------------------------------

G2 ------------------------------------------------------------

Glyma.08G361500 STEEDIKGSEFDELVEERRLWASYRGQTLTRTVRGMMYYRKALELQAFLDMAKDEDLMEG

Gene7 ------------------------------------------------------------

G2 ------------------------------------------------------------

Glyma.08G361500 YKAMENSDDNSRGERSLWTQCQAVADMKFTYVVSCQQYGIDKRSGSLRAQDILRLMTRYP

Gene7 ------------------------------------------------------------

G2 ------------------------------------------------------------

Glyma.08G361500 SLRVAYIDEVEEPVQDSKKKINKVYYSCLVKAMPKSNSPSEPEQNLDQIIYKIKLPGPAI

Gene7 ------------------------------------------------------------

G2 ------------------------------------------------------------

Glyma.08G361500 LGEGKPENQNHAIIFTRGEGLQTIDMNQDNYMEEALKMRNLLQEFLKKHDGVRFPSILGL

Page 210: Identification and Characterization of Common Bacterial

195

Gene7 ------------------------------------------------------------

G2 ------------------------------------------------------------

Glyma.08G361500 REHIFTGSVSSLAWFMSNQETSFVTIGQRLLANPLKVRFHYGHPDVFDRLFHLTRGGVSK

Gene7 ------------------------------------------------------------

G2 ------------------------------------------------------------

Glyma.08G361500 ASKVINLSEDIFAGFNSTLREGNVTHHEYIQVGKGRDVGLNQISMFEAKIANGNGEQTLS

Gene7 ------------------------------------------------------------

G2 ------------------------------------------------------------

Glyma.08G361500 RDVYRLGHRFDFFRMLSCYFTTVGFYFSTLITVLTVYVFLYGRLYLVLSGLEEGLSTQKA

Gene7 ------------------------------------------------------------

G2 ------------------------------------------------------------

Glyma.08G361500 IRDNKPLQVALASQSFVQIGVLMALPMLMEIGLERGFRTALSEFILMQLQLAPVFFTFSL

Gene7 ------------------------------------------------------------

G2 ------------------------------------------------------------

Glyma.08G361500 GTKTHYFGRTLLHGGAKYRPTGRGFVVFHAKFADNYRLYSRSHFVKGIELMILLVVYEIF

Gene7 ------------------------------------------------------------

G2 ------------------------------------------------------------

Glyma.08G361500 GHSYRSTVAYILITASMWFMVGTWLFAPFLFNPSGFEWQKIVDDWTDWNKWISNRGGIGV

Gene7 ------------------------------------------------------------

G2 ------------------------------------------------------------

Glyma.08G361500 LPEKSWESWWEEEQEHLQYSGMRGIIVEILLSLRFFIYQYGLVYHLNITKKGTKSFLVYG

Gene7 ------------------------------------------------------------

G2 ------------------------------------------------------------

Glyma.08G361500 ISWLVIFVILFVMKTVSVGRRKFSANFQLVFRLIKGMIFLTFVSILVILIALPHMTVQDI

Gene7 ------------------------------------------------------------

G2 ------------------------------------------------------------

Glyma.08G361500 VVCILAFMPTGWGMLQIAQALKPVVRRAGFWGSVKTLARGYEIVMGLLLFTPVAFLAWFP

Gene7 --------------------------------------

G2 --------------------------------------

Glyma.08G361500 FVSEFQTRMLFNQAFSRGLQISRILGGQRKERSSRNKE

Appendix 1. Predicted β-glucan synthase amino acid sequence alignment. ClustalW peptide sequence alignment for contig1701 gene 7, predicted gene in G19833 (G2), and G. max gene annotation (Glyma08G361500). Callose synthase family domains are indicated by boxes. Transmembrane domains are highlighted in grey.

Page 211: Identification and Characterization of Common Bacterial

196

gene6 ----MSKTTAIQSRIIRLT---LPDRKYQNNLHNHLKLSKPQSSHSYRSYRRNLRKPHHN

Phvul.008G191100 MHFRMFKPFGLNHAGTKCWGVFGNKVVFENGRVDLVEVFTVAQRKTWSWVTVKEKSAAFS

* *. .:: : . ::*. : ::: . . ::: : :.. ..

gene6 YQN--------------

Phvul.008G191100 YSDWCLDPICCMRYLKV

*.:

Appendix 2. Alignment of P. vulgaris hypothetical proteins. ClustalW alignment of P. vulgaris hypothetical gene predictions were identified from OAC Rex and G19833 genome sequence assemblies.

Appendix 3. Protein secondary structure models for hypothetical proteins. Secondary structure models were created in Phyre2 for P. vulgaris hyothetical gene 6 (A) identified in OAC Rex contig1701, and Phvul.008G191100 (B) identified in G19833 chromosome 8.

Appendix 4. Phyre2 secondary structure modelling (Kelley and Sternberg 2009) for gene 6 and Phvul.008G191100

Gene Name Amino Acids

Modelled Template

Template

Characteristic

Similarity to

Template Confidence

Gene 6 23-49 C4ImgD_ Transcriptional activator 33% 15.2%

Phvul.008G191100 11-33 D2f9ca1 Single-stranded left-handed

beta-helix fold 29% 32.0%

29-65 D1m0da_ Restriction endonuclease-like 27% 12.5%

Page 212: Identification and Characterization of Common Bacterial

197

Appendix 5. Analysis of variance for inoculated detached common bean (P. vulgaris) leaves maintained in a water agar tissue culture system. Significance (p < 0.05) of fixed treatment, treatment interaction and error effects are shown. Genotypes include resistant (OAC Rex) and susceptible (Nautica) common bean varieties. Leaves were inoculated with one of four Xanthomonas axonopodis pv. phaseoli isolates (Xap12, Xap18, Xap98 or Xap118) or water control. Leaves were maintained in a water agar tissue culture system at 28oC for up to 192 hours post inoculation (HPI) and samples were taken at different time points post inoculation (0, 48, 96, 144 and 192 HPI). Source df F-value P > F

Inoculum 4 36.55 <0.0001

Genotype 1 47.42 <0.0001

Time 4 14.73 <0.0001

Inoculum x Genotype 4 3.89 0.0059

Inoculum x Time 16 2.57 0.0025

Genotype x Time 4 4.72 0.0017

Inoculum x Genotype x Time 16 1.41 0.1530

Error 41

Appendix 6. Analysis of variance for inoculated detached Arabidopsis thaliana rosette leaves maintained in a water agar tissue culture system. Significance (p < 0.05) of fixed treatment, treatment interaction and error effects are shown. Leaves were inoculated with one of four Xanthomonas axonopodis pv. phaseoli isolates (Xap12, Xap18, Xap98 or Xap118) or water control. Leaves were maintained in a water agar tissue culture system at 28oC for up to 192 hours post inoculation (HPI) and samples were taken at different time points post inoculation (0, 48, 96, 144 and 192 HPI).

Source df F-value P > F

Inoculum 4 21.57 <0.0001

Time 4 11.21 <0.0001

Inoculum x Time 16 2.58 0.0055

Error 26

Page 213: Identification and Characterization of Common Bacterial

198

Appendix 7. LSmeans for common bacterial blight symptoms on OAC Rex, Nautica and Arabidopsis leaves at 0 hours post inoculation (HPI). Percent of leaf showing disease symptoms for OAC Rex, Nautica and Arabidopsis leaves inoculated with Water, Xap12, Xap18, Xap98 and Xap118 at 0 HPI

Genotype

Symptoms (%)

Inoculum

Water Xap12 Xap18 Xap98 Xap118 Genotype

Mean

OAC Rex 4.26 x10-2 YZ

a 7.26 x10-2 Y a 7.37 x10

-2 Y a 8.53 x10

-2 Y a 4.33 x10

-2 Y a 6.35 x10

-2 WX

A

Nautica 0.120Y a 0.154

Y a 7.29 x10

-2 Y a 6.23 x10

-2 Y a 5.91 x10

-2 Y a 9.37 x10

-2 W

A

Arabidopsis 0.297 V z 0.236

V z 0.318

V z 0.186

V z 0.367

V z 0.281

U

ZLsmeans followed by the same lower case letter were not significantly different at p=0.05

YLsmeans followed by the same upper case letter were not significantly different at p=0.05

XDenotes lsmeans with a standard error of +127.80

WDenotes lsmeans with a standard error of +57.1532

VDenotes lsmeans with a standard error of +3.06

UDenotes lsmeans with a standard error of +1.37

Appendix 8. LSmeans for common bacterial blight symptoms on OAC Rex, Nautica and Arabidopsis leaves at 48 hours post inoculation (HPI). Percent of leaf showing disease symptoms for OAC Rex, Nautica and Arabidopsis leaves inoculated with Water, Xap12, Xap18, Xap98 and Xap118 at 48 HPI

Genotype

Symptoms (%)

Inoculum

Water Xap12 Xap18 Xap98 Xap118 Genotype

Mean

OAC Rex 0.138YZ

a 0.315Y a 9.64 x10

-2 y a 0.276

Y a 0.405

Y a 0.246

WX A

Nautica 0.156Y a 0.335

Y a 0.349

Y a 0.177

Y a 0.657

Y a 0.335

W A

Arabidopsis 0.255 V z 1.26

V z 0.433

V z 0.367

V z 0.973

V z 0.656

U

ZLsmeans followed by the same lower case letter were not significantly different at p=0.05

YLsmeans followed by the same upper case letter were not significantly different at p=0.05

XDenotes lsmeans with a standard error of +127.80

WDenotes lsmeans with a standard error of +57.1532

VDenotes lsmeans with a standard error of +3.06

UDenotes lsmeans with a standard error of +1.37

Page 214: Identification and Characterization of Common Bacterial

199

Appendix 9. LSmeans for common bacterial blight symptoms on OAC Rex, Nautica and Arabidopsis leaves at 96 hours post inoculation (HPI). Percent of leaf showing disease symptoms for OAC Rex, Nautica and Arabidopsis leaves inoculated with Water, Xap12, Xap18, Xap98 and Xap118 at 96 HPI

Genotype

Symptoms (%)

Inoculum

Water Xap12 Xap18 Xap98 Xap118 Genotype

Mean

OAC Rex 0.627YZ

a 1.42Y a 3.56

Y a 3.25

Y a 3.18

Y a 2.41

WX A

Nautica 2.10Y a 3.20

Y a 4.74

Y a 2.40

Y a 2.34

Y a 2.96

W A

Arabidopsis 0.460V z 4.44

V z 5.46

V z 4.74

V z 1.02

V z 3.22

U

ZLsmeans followed by the same lower case letter were not significantly different at p=0.05

YLsmeans followed by the same upper case letter were not significantly different at p=0.05

XDenotes lsmeans with a standard error of +127.80

WDenotes lsmeans with a standard error of +57.1532

VDenotes lsmeans with a standard error of +3.06

UDenotes lsmeans with a standard error of +1.37

Appendix 10. LSmeans for colony forming units for X. axonopodis pv. phaseoli (Xap) bacteria at 0 hours post inoculation (HPI). CFU was determined by incubating 105 dilutions of ground leaf samples taken from inoculated leaves from resistant (OAC Rex) and susceptible (Nautica) P. vulgaris varieties, and Arabidopsis. Samples were taken at 0 HPI after inoculation with Water, Xap12, Xap18, Xap98 or Xap118.

Genotype

CFU (x 105)

Inoculum

Water Xap12 Xap18 Xap98 Xap118 Genotype

Mean

OAC Rex 0XZ

a 173 X a 0

X a 452

X a 147

X a 154

WY A

Nautica 0X a 280

X a 430

X a 142

X a 0

X a 170

Y A

Arabidopsis 0V z 0

V z 217

V z 122

V z 31.7

V z 74.1

U

ZLsmeans followed by the same lower case letter were not significantly different at p=0.05

YLsmeans followed by the same upper case letter were not significantly different at p=0.05

XDenotes lsmeans with a standard error of +174

WDenotes lsmeans with a standard error of +77.9

VDenotes lsmeans with a standard error of +96.3

UDenotes lsmeans with a standard error of +43.1

Page 215: Identification and Characterization of Common Bacterial

200

Appendix 11. Candidate gene expression in first experiment replication of Phaseolus vulgaris leaf inoculation. Expression of candidate genes CBB2, CBB3 and CBB4 were compared to expression of actin check in resistant (OAC Rex) and susceptible (Seaforth) common bean varieties. The inoculum used was X. axonopodis pv. phaseoli isolate 98 (X) and sterile distilled water (C). RNA was isolated from leaf samples taken from water and Xanthomonas-inoculated OAC Rex and Seaforth plants at 0, 12, 24, and 48 hours post-inoculation (HPI). OAC Rex genomic DNA and sterile distilled water were used as positive (+) and negative (-) PCR controls, respectively. Candidate gene bands were produced using primers CBB2CS, CBB3CS and CBB4CS for CBB2, CBB3 and CBB4, respectively. Band size is indicated by a white arrow for actin, yellow arrow for CBB2, red arrow for CBB3 and green arrow for CBB4.

Page 216: Identification and Characterization of Common Bacterial

201