characterization of dickeya dianthicola and pectobacterium

167
The University of Maine The University of Maine DigitalCommons@UMaine DigitalCommons@UMaine Electronic Theses and Dissertations Fogler Library Spring 5-7-2021 Characterization of Dickeya dianthicola and Pectobacterium Characterization of Dickeya dianthicola and Pectobacterium parmentieri Causing Blackleg and Soft Rot on Potato parmentieri Causing Blackleg and Soft Rot on Potato Tongling Ge University of Maine, [email protected] Follow this and additional works at: https://digitalcommons.library.umaine.edu/etd Part of the Agricultural Science Commons, Plant Biology Commons, and the Plant Pathology Commons Recommended Citation Recommended Citation Ge, Tongling, "Characterization of Dickeya dianthicola and Pectobacterium parmentieri Causing Blackleg and Soft Rot on Potato" (2021). Electronic Theses and Dissertations. 3486. https://digitalcommons.library.umaine.edu/etd/3486 This Open-Access Thesis is brought to you for free and open access by DigitalCommons@UMaine. It has been accepted for inclusion in Electronic Theses and Dissertations by an authorized administrator of DigitalCommons@UMaine. For more information, please contact [email protected].

Upload: others

Post on 03-Apr-2022

6 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Characterization of Dickeya dianthicola and Pectobacterium

The University of Maine The University of Maine

DigitalCommons@UMaine DigitalCommons@UMaine

Electronic Theses and Dissertations Fogler Library

Spring 5-7-2021

Characterization of Dickeya dianthicola and Pectobacterium Characterization of Dickeya dianthicola and Pectobacterium

parmentieri Causing Blackleg and Soft Rot on Potato parmentieri Causing Blackleg and Soft Rot on Potato

Tongling Ge University of Maine, [email protected]

Follow this and additional works at: https://digitalcommons.library.umaine.edu/etd

Part of the Agricultural Science Commons, Plant Biology Commons, and the Plant Pathology

Commons

Recommended Citation Recommended Citation Ge, Tongling, "Characterization of Dickeya dianthicola and Pectobacterium parmentieri Causing Blackleg and Soft Rot on Potato" (2021). Electronic Theses and Dissertations. 3486. https://digitalcommons.library.umaine.edu/etd/3486

This Open-Access Thesis is brought to you for free and open access by DigitalCommons@UMaine. It has been accepted for inclusion in Electronic Theses and Dissertations by an authorized administrator of DigitalCommons@UMaine. For more information, please contact [email protected].

Page 2: Characterization of Dickeya dianthicola and Pectobacterium

CHARACTERIZATION OF DICKEYA DIANTHICOLA AND PECTOBACTERIUM

PARMENTIERI CAUSING BLACKLEG AND SOFT ROT ON POTATO

By

Tongling Ge

B.S. Shandong Agricultural University, China 2014

M.S. China Agricultural University, China 2016

A DISSERTATION

Submitted in Partial Fulfillment of the

Requirements for the Degree of

Doctor of Philosophy

(in Ecology and Environmental Sciences)

The Graduate School

The University of Maine

May 2021

Advisory Committee:

Jianjun Hao, Associate Professor of Plant Pathology, School of Food and Agriculture, Advisor

Robert P. Larkin, Research Plant Pathologist (USDA-ARS)

Gregory Porter, Professor of Crop Ecology and Management, School of Food and Agriculture

Seanna L. Annis, Associate Professor of Mycology, School of Biology and Ecology

Ek Han Tan, Assistant Professor of Plant Genetics, School of Biology and Ecology

Page 3: Characterization of Dickeya dianthicola and Pectobacterium

Copyright 2021 Tongling Ge

Page 4: Characterization of Dickeya dianthicola and Pectobacterium

CHARACTERIZATION OF DICKEYA DIANTHICOLA AND PECTOBACTERIUM

PARMENTIERI CAUSING BLACKLEG AND SOFT ROT ON POTATO

By Tongling Ge

Dissertation Advisor: Dr. Jianjun Hao

An Abstract of the Dissertation Presented

in Partial Fulfillment of the Requirements for the

Degree of Doctor of Philosophy

(in Ecology and Environmental Sciences)

May 2021

Potato blackleg and soft rot (PBSR), which can be caused by Dickeya spp. and

Pectobacterium spp., is a serious problem worldwide. The recent outbreak of PBSR in the

Northeastern USA, caused primarily by D. dianthicola, has resulted in significant economic

losses since 2015. This seedborne disease is highly associated with and therefore spread by seed

tuber distribution. To understand how the outbreak occurred and where the pathogen originated,

a total of 1204 potato samples were collected from 11 northeastern states from 2015 to 2020. All

the samples were processed for bacterial isolation and DNA extraction. Dickeya dianthicola and

P. parmentieri were detected using conventional polymerase chain reaction (PCR). Dickeya

dianthicola and P. parmentieri were found in 38.1% and 53.3% of the samples, respectively, and

20.6% of samples contained both D. dianthicola and P. parmentieri. Seventeen isolates of D.

dianthicola were obtained from the samples and classified into three genotypes (Type I, II, III).

Results based on 258 samples showed that Maine mainly had Type I but no Type III, while Type

II appeared to be distributed throughout the Northeastern USA. By pan-genomic analysis, D.

dianthicola strains collected worldwide were classified into eight distinguished clades. Type I

strains had an extraordinarily high homogeneity and distinct discrimination from other countries,

Page 5: Characterization of Dickeya dianthicola and Pectobacterium

indicating a single-strain population. Virulence-related systems, such as plant cell-wall degrading

enzymes, flagellar and chemotaxis related features, two-component regulatory system, and type

I/II/III secretion systems were highly conserved, but type IV/VI secretion systems varied, in

which type I strain had an additional set T4SS cluster, implying more aggressiveness and

virulence. Thus, the PBSR outbreak was proposed to be associated with a new strain derived by

mutation.

Bacterial communities were analyzed on the samples using Illumina sequencing targeting

on the V3-V4 region of 16S rRNA gene. Genera Dickeya or Pectobacterium prevailed in the

microbial community when they each existed alone, while Dickeya surpassed Pectobacterium

when they coexisted. Among the pathogen complex, D. dianthicola was the only species in the

Dickeya genus, while species varied in the Pectobacterium genus, with P. parmentieri, P.

polaris, P. carotovora subsp. carotovora, P. carotovora subsp. odoriferum being the most

prevalent presumptive species. Isolates of the four presumptive species and P. c. subsp.

brasiliensis were identified by sequencing the gapA gene and were confirmed to be pathogenic

on potatoes. Thus, PBSR was caused by intergeneric or intrageneric species of Dickeya and

Pectobacterium that contribute collectively to the disease complex. To further investigate the

relationship between the two bacterial species and their interaction, field trials were established.

Three varieties of potato seed pieces were inoculated with bacterial suspensions of D.

dianthicola and planted in the field. Two-year results showed that there was a significant linear

correlation (P < 0.01) between relative yield loss and percentage of inoculated seed pieces.

Furthermore, D. dianthicola was more virulent than P. parmentieri in the field, but the co-

inoculation of the two species resulted in increased disease severity compared to single-species

inoculation with either pathogen.

Page 6: Characterization of Dickeya dianthicola and Pectobacterium

iii

DEDICATION

Back in the summer of 2016, I opened a new chapter at the University of Maine

(UMaine), with the beginning of a one-year work exchange in Dr. Hao’s lab. However, I stayed

here for five years as I gradually got into the spirit of life and research at UMaine. It was unlikely

to happen without the support and encouragement from my parents Ensheng Ge, Xiu’e Wang,

and my husband Shenghua Zhou. Their sacrifices allowed me to concentrate on my study and

research. The best possession that my parents gave to me was the instruction of being steadfast in

my work, which contributed so much toward conducting my research. I will cherish this

personality trait though my whole life. I also appreciate the patient waiting and ‘virtual

accompany’ of my husband. If I had any success, I would give half of the credit to him. We were

pursuing our dreams in both ends of the earth and never give up our relationship to each other. I

am looking forward to the new life when we reunite.

I love you all, as always!

Page 7: Characterization of Dickeya dianthicola and Pectobacterium

iv

ACKNOWLEGEMENTS

The first to be acknowledged is my former advisor Dr. Laixin Luo who supported me to

go to Maine. I would like to express the deepest appreciation to my advisor Dr. Jianjun Hao for

having me in his program. My life has changed since and I am constantly improving myself both

in academia and personal growth because of his patience, support, and guidance. A special

“thank you” goes to Dr. Steven B. Johnson for his all-time and all-way support, constant

encouragement and expert guidance. The project would not have been possible without the

precious samples he provided and managed. Being a thesis committee member, Dr. Robert

Larkin has contributed to this project in multiple ways including experimental design,

manuscript revision and additional efforts in technical support. I sincerely thank my other

committee members Dr. Gregory Porter, Dr. Seanna Annis, Dr. Ek Han Tan and Dr. John Singer,

for their encouragement, advice and patience throughout my graduate study. Although it’s

unfortunate that Dr. Singer could not witness my defense, I feel his inspiration is still in me. I

have received a great deal of support and assistance from members of Hao lab, Dr. Helen Jiang,

Nayara Marangoni, Alice Chesley, Fatemeh Ekbataniamiri, Andrea Hain, Kedi Li and Sylvia

Zhang. I appreciate their wonderful collaboration and friendship. Special thanks also go to

Elbridge Giggie, the most experienced and respected field technician in supporting my project. I

would like to thank Bradly Libby who helped in my greenhouse work. I am thankful that Patty

Singer has provided me a great amount of support in DNA sequencing. With a special mention to

Lihua Yang, I am very appreciative for ‘Mom’-like care, which made my life much easier and

happier without being homesick. I would never forget my special friend Val Kinerson who has

unselfishly offered me English lessons for years. A very special gratitude goes out to Maine

Potato Board, USDA-ARS, and USDA-NIFA-SCRI for providing the funding for my study.

Page 8: Characterization of Dickeya dianthicola and Pectobacterium

v

TABLE OF CONTENTS

DEDICATION.............................................................................................................................. iii

ACKNOWLEGEMENTS ........................................................................................................... iv

LIST OF TABLES ....................................................................................................................... ix

LIST OF FIGURES ..................................................................................................................... xi

CHAPTER 1 .................................................................................................................................. 1

LITERATURE REVIEW ............................................................................................................ 1

1.1 Potato ..................................................................................................................................... 1

1.1.1 Potato production ............................................................................................................ 1

1.1.2 Biology and physiology of potato growth ...................................................................... 2

1.2 Potato diseases....................................................................................................................... 3

1.3 Blackleg and soft rot of potato .............................................................................................. 6

1.3.1 Disease symptoms .......................................................................................................... 6

1.3.2 Disease cycle .................................................................................................................. 7

1.3.3 The occurrence of PBSR worldwide and outbreaks in the US ....................................... 9

1.3.4 The pathogens of PBSR ................................................................................................ 10

1.3.4.1 Taxonomy of PBSR pathogens .............................................................................. 10

1.3.4.2 Host range .............................................................................................................. 13

1.3.4.3 Pathogenesis of PBSR pathogens .......................................................................... 16

1.3.5 Management of PBSR .................................................................................................. 18

1.3.5.1 Exclusion................................................................................................................ 18

1.3.5.2 Resistant varieties .................................................................................................. 19

1.3.5.3 Protection ............................................................................................................... 20

1.4 Microbiomes........................................................................................................................ 21

1.4.1 Plant microbiome .......................................................................................................... 22

1.4.2 Plant pathobiome .......................................................................................................... 23

1.4.3 Microbial interactions ................................................................................................... 24

1.4.3.1 Competition and coexistence ................................................................................. 25

Page 9: Characterization of Dickeya dianthicola and Pectobacterium

vi

1.4.3.2 Cooperation and synergism.................................................................................... 26

1.4.4 Sequencing-based methods for studying microbiomes ................................................ 27

1.5 Objectives ............................................................................................................................ 29

CHAPTER 2 ................................................................................................................................ 31

GENOTYPING DICKEYA DIANTHICOLA CAUSING POTATO BLACKLEG AND

SOFT ROT ASSOCIATED WITH INOCULUM GEOGRAPHY IN THE UNITED

STATES ....................................................................................................................................... 31

ABSTRACT .............................................................................................................................. 31

2.1 INTRODUCTION ............................................................................................................... 32

2.2 MATERIALS AND METHODS ........................................................................................ 34

2.2.1 Sample collection ......................................................................................................... 34

2.2.2 Pathogen isolation......................................................................................................... 35

2.2.3 Pathogen detection and identification........................................................................... 36

2.2.4 Fingerprinting analysis ................................................................................................. 36

2.2.5 Primer design for genotyping ....................................................................................... 37

2.2.6 Phylogenetic analysis ................................................................................................... 37

2.3 RESULTS............................................................................................................................ 38

2.3.1 Disease detection and pathogen isolation ..................................................................... 38

2.3.2 Genotyping by fingerprinting and DNA sequencing of D. dianthicola ....................... 40

2.3.3 Identification of genotypes of D. dianthicola ............................................................... 46

2.4 DISCUSSION ..................................................................................................................... 51

CHAPTER 3 ................................................................................................................................ 55

PANGENOMIC ANALYSIS OF DICKEYA DIANTHICOLA STRAINS REVEALS THE

OUTBREAK OF BLACKLEG AND SOFT ROT ON POTATO IN USA ........................... 55

ABSTRACT .............................................................................................................................. 55

3.1 INTRODUCTION ............................................................................................................... 56

3.2 MATERIALS AND METHODS ........................................................................................ 59

Page 10: Characterization of Dickeya dianthicola and Pectobacterium

vii

3.2.1 Bacterial strains ............................................................................................................ 59

3.2.2 Whole genome sequencing (WGS) and assembly ........................................................ 61

3.2.3 Genome annotation ....................................................................................................... 61

3.2.4 Pangenome analysis ...................................................................................................... 61

3.2.5 Comparison among USA strains of D. dianthicola genomes ....................................... 62

3.3 RESULTS............................................................................................................................ 63

3.3.1 Genomic assemblies ..................................................................................................... 63

3.3.2 Pangenome.................................................................................................................... 65

3.3.3 SNPs-based phylogeny ................................................................................................. 67

3.3.4 Analyses of USA strains ............................................................................................... 69

3.4 DISCUSSION ..................................................................................................................... 75

CHAPTER 4 ................................................................................................................................ 80

PATHOGEN COMPLEX AND MICROBIAL ASSOCIATION OF POTATO

BLACKLEG AND SOFT ROT IN NORTHEASTERN AMERICA .................................... 80

ABSTRACT .............................................................................................................................. 80

4.1 INTRODUCTION ............................................................................................................... 81

4.2 MATERIALS AND METHODS ........................................................................................ 83

4.2.1. Disease samples ........................................................................................................... 83

4.2.2 Metagenomic sequencing ............................................................................................. 84

4.2.3 Bacterial isolation and identification ............................................................................ 84

4.2.4 Bacterial isolation and identification ............................................................................ 85

4.2.5 Pathogenicity assay....................................................................................................... 86

4.2.6 Bacterial growth on CVP medium ................................................................................ 87

4.3 RESULTS............................................................................................................................ 88

4.3.1 Bacterial communities associated with PBSR .............................................................. 88

4.3.2 Distribution of Dickeya and Pectobacterium spp. ........................................................ 91

4.3.3 Presumptive bacterial species in genera Dickeya and Pectobacterium ........................ 94

4.3.4 Pathogenicity and growth of Pectobacterium spp. ....................................................... 97

4.4 DISCUSSION ................................................................................................................... 102

Page 11: Characterization of Dickeya dianthicola and Pectobacterium

viii

CHAPTER 5 .............................................................................................................................. 105

INTERACTION BETWEEN DICKEYA DIANTHICOLA AND PECTOBACTERIUM

PARMENTIERI IN POTATO INFECTION OF POTATO UNDER FIELD CONDITIONS

..................................................................................................................................................... 105

ABSTRACT ............................................................................................................................ 105

5.1 INTRODUCTION ............................................................................................................. 106

5.2 MATERIALS AND METHODS ...................................................................................... 108

5.2.1 Frequency and distribution of Dickeya dianthicola and Pectobacterium parmentieri in

naturally infected potato ...................................................................................................... 108

5.2.2 Tuber inoculation using vacuum infiltration .............................................................. 109

5.2.3 Effects of bacterial inoculation on potato growth ...................................................... 109

5.2.4 Interaction between D. dianthicola and P. parmentieri in potato infection ............... 110

5.2.5 Statistical analysis....................................................................................................... 111

5.3 RESULTS.......................................................................................................................... 111

5.3.1 Frequency and distribution of Dickeya dianthicola and Pectobacterium parmentieri in

naturally infected potato ...................................................................................................... 111

5.3.2 Effects of bacterial inoculation on potato growth ...................................................... 113

5.3.3 Interaction between D. dianthicola and P. parmentieri in potato infection ............... 113

5.4 DISCUSSION ................................................................................................................... 118

DISSERTATION SUMMARY ................................................................................................ 122

REFERENCES .......................................................................................................................... 124

BIOGRAPHY OF THE AUTHOR ......................................................................................... 152

Page 12: Characterization of Dickeya dianthicola and Pectobacterium

ix

LIST OF TABLES

Table 1-1. Development of etiology and taxonomy of blackleg and soft rot bacteria. ................. 12

Table 1-2. Confirmed hosts of Dickeya spp. and Pectobacterium spp. causing blackleg and soft

rot of potato. .................................................................................................................................. 15

Table 2-1. Detection of Dickeya dianthicola in potato samples collected from Northeastern US,

and a smaller amount from Midwestern and Southern regions. ................................................... 39

Table 2-2. Dickeya dianthicola isolates obtained in this study. .................................................. 40

Table 2-3. Dickeya dianthicola strains retrieved from NCBI database. ....................................... 42

Table 2-4. Primers (F: forward; R: reversed) used for genotypic analysis of Dickeya dianthicola.

....................................................................................................................................................... 46

Table 2-5. Genotype detection of Dickeya dianthicola strains using polymerase chain reaction

with primers Trac_4, Anti_1, and DIA-A. .................................................................................... 48

Table 3-1. Dickeya dianthicola strains used in this study. ........................................................... 60

Table 3-2. Genome statistics of Dickeya dianthicola strains. ....................................................... 64

Table 4-1. Bacterial detection in potato samples using polymerase chain reaction (PCR) with

primers pelADE targeting on Dickeya spp. and PEC targeting on both Dickeya and

Pectobacterium spp. ...................................................................................................................... 83

Table 4-2. Dickeya and Pectobacterium spp. strains used for pathogen-bacteria interactions. .... 88

Table 4-3. Relative abundance (%) of shared and exclusive genera of bacteria in four groups (D,

DP, P, and N) of potato samples showing symptoms of blackleg and soft rot. ............................ 92

Table 4-4. Relative abundance (%) of shared and exclusive genera of bacteria in four groups (D,

DP, P, and N) of potato samples showing symptoms of blackleg and soft rot. ............................ 95

Page 13: Characterization of Dickeya dianthicola and Pectobacterium

x

Table 4-5. Cavity diameter (cm) of bacterial species belonging to genera Dickeya and

Pectobacterium estimated by measuring the colony diameter on crystal violet pectate agar plates

incubated at 28 C in darkness for 24 and 48 hours. .................................................................. 101

Table 5-1. Detection of Dickeya and Pectobacterium spp. in symptomatic potato stems or tubers

using polymerase chain reaction (PCR). ..................................................................................... 112

Table 5-2. Chi-square test on tissue distribution of Dickeya dianthicola (DDi) and

Pectobacterium parmentieri (PPa) on potato tissues showing blackleg and soft rot symptoms. 112

Table 5-3. Effects of infection level of seed pieces with Dickeya dianthicola on yield and

emergence of potato in 2020. ...................................................................................................... 114

Table 5-4. Effects of seed tuber inoculation with Dickeya dianthicola (DDi), Pectobacterium

parmentieri (PPa) and non-inoculated (NI) on emergence and yield of potato varieties in years

2018 and 2019. ............................................................................................................................ 118

Page 14: Characterization of Dickeya dianthicola and Pectobacterium

xi

LIST OF FIGURES

Figure 1-1. Potato production in the United States (USDA 2019). ................................................ 2

Figure 1-2. Symptoms of blackleg and soft rot of potato.. ............................................................. 7

Figure 1-3. Flowchart of the chapters. .......................................................................................... 30

Figure 2-1. Gel electrophoresis of BOX PCR on Dickeya dianthicola strains ............................. 43

Figure 2-2. Phylogenetic tree using maximum-likelihood analysis on Dickeya dianthicola strains

collected in this study.................................................................................................................... 44

Figure 2-3. Phylogenetic tree using maximum-likelihood analysis on Dickeya dianthicola strains

collected in this study and NCBI .................................................................................................. 45

Figure 2-4. Multiple genome analysis of Dickeya dianthicola strains using MAUVE for

identifying a region for primer design. ......................................................................................... 47

Figure 2-5. Genotypic distribution of Dickeya dianthicola in the Northeastern United States from

2015 to 2019. ................................................................................................................................ 50

Figure 3-1. Pangenome statistics of 32 Dickeya dianthicola strains. ........................................... 66

Figure 3-2. Geographic and phylogenetic relationship of Dickeya dianthicola strains used in this

study. ............................................................................................................................................. 68

Figure 3-3. Whole genome alignment and comparison of orthologous genes from clades I, II and

III of Dickeya dianthicola from USA. .......................................................................................... 70

Figure 3-4. BLAST Ring Image Generator (BRIG) analysis of genomes of Dickeya dianthicola

strains clustered in clades I, II and III against reference strain ME23. ......................................... 71

Figure 3-5. Schematic map of type IV secretion system in clades I, II and III of Dickeya

dianthicola from USA................................................................................................................... 73

Page 15: Characterization of Dickeya dianthicola and Pectobacterium

xii

Figure 3-6. Schematic map of type VI secretion system in clades I, II and III of Dickeya

dianthicola .................................................................................................................................... 74

Figure 4-1. Number of operational taxonomic unit (OTU) and Shannon evenness index of

bacterial community in potato tissues. .......................................................................................... 90

Figure 4-3. Relative abundances of Dickeya and Pectobacterium spp. by geographic locations or

states (A and C) and sample groups (B and D) of PBSR. ............................................................ 93

Figure 4-4. Phylogenetic tree constructed based on partial sequences the gapA gene (850 bp)

using maximum-likelihood algorithm with strains of Pectobacterium spp., including P.

carotovora subsp. odoriferum (PCO), P. polaris (PPO), P. parmentieri (PPA), P. carotovora

subsp. carotovora (PCC), and P. carotovora subsp. brasiliensis (PCB) ...................................... 98

Figure 4-5. Pathogenicity test of Pectobacterium spp. Species include P. carotovora subsp.

brasiliensis (PCB), P. carotovora subsp. odoriferum (PCO), P. parmentieri (PPA), and P.

polaris (PPO) on potato ‘Lamoka’ stems ..................................................................................... 99

Figure 4-6. Pathogenicity test of Pectobacterium spp. Species include P. carotovora subsp.

brasiliensis (PCB), P. carotovora subsp. odoriferum (PCO), P. parmentieri (PPA), and P.

polaris (PPO) on potato ‘Lamoka’ tubers. .................................................................................. 100

Figure 5-1. Regression analysis between relative emergence loss/relative yield loss and infection

percentage of potato. ................................................................................................................... 115

Figure 5-2. Disease progress of blackleg in potato ‘Lamoka’, ‘Atlantic’ and ‘Shepody’ grown

from seed pieces inoculated with either Dickeya dianthicola (DDi), Pectobacterium parmentieri

(PPa), or DDi + PPa under field conditions in 2018 (upper panels) and 2019 (bottom panels).

Non-inoculated (NI) plants were used for control. ..................................................................... 116

Page 16: Characterization of Dickeya dianthicola and Pectobacterium

1

CHAPTER 1

LITERATURE REVIEW

1.1 Potato

1.1.1 Potato production

Potato (Solanum tuberosum L.) is the most important vegetable crop in the world,

ranking number four as a main food crop, following rice (Oryza sativa L.), wheat (Triticum

aestivum L.) and corn (Zea mays L.), feeding a big part of the global population (Johnson 2008).

Potato originated from the Andes Mountains of South America (Salaman and Burton 1985), and

was introduced to Europe with Spain being the first entry port as early as 1576 (Brown 1993). In

1719, potato was brought to the United States of America (USA) from Ireland and was first

grown in New Hampshire (Brown 1993). Due to its adaptation to temperate growing conditions,

potatoes have been grown in over 100 countries throughout the world (Hijmans 2001). The

popularity and importance of potato as a food crop is because it produces greater nutrition on

smaller amounts of land than most other crops, and has high food value, therefore can be used as

a staple food (F. A. O. 2008).

USA is the fifth largest potato producer in the world, following China, India, Russia, and

Ukraine (Hijmans 2001). As in many other countries, potato is the top vegetable crop in the US,

and its importance has been documented by the US Department of Agriculture since 1866. There

are three growing seasons for potato nationwide in USA, depending on the location, including

spring, summer and fall, with the usual time of harvest being January to June, July to mid-

September, and August to November respectively (USDA 2019). According to the latest potato

report, 92% of the total production or 90% of the harvested area in 2018 came from the fall

season, and the total potato production totaled 450 million cwt, with harvested area of 1.01

Page 17: Characterization of Dickeya dianthicola and Pectobacterium

2

million acres (USDA 2019). Production from the spring and summer seasons each make up 4%

of the total (USDA 2019).

Idaho and Washington are the top two US states in potato production and harvested area,

and together produce half of the total US potato crop (Figure 1.1). Potato production in the state

of Maine ranked the 10th in 2018, while the harvested area ranked 6th (Figure 1.1). Maine has

48,500 acres of harvested potato with total production of 15 million cwt, 67% of which is used

for processing, 22% for seed and 11% for table stock (NASS 2018).

Figure 1-1. Potato production in the United States (USDA 2019).

1.1.2 Biology and physiology of potato growth

A potato crop goes through several stages, including sprout development, vegetative

growth, tuber initiation, tuber bulking, and tuber maturation (Dean 1994; Thornton 2020).

Potato tubers are dormant due to hormonal control, after first being harvested, and it takes

several weeks or months to break the dormancy. Under favorable environmental conditions (e.g.,

0

50

100

150

200

250

300

350

Idah

o

Was

hin

gto

n

Nort

h D

akota

Wis

con

sin

Colo

rado

Mai

ne

Mic

hig

an

Ore

gon

Min

nes

ota

Cal

iforn

ia

Flo

rida

Neb

rask

a

Tex

as

New

York

Nort

h C

aroli

na

Monta

na

Illi

no

is

Mis

souri

Vir

gin

ia

Kan

sas

New

Jer

sey

Mar

yla

nd

Ala

ska

Oth

er s

tate

s

Area harvested (1000 acre) Total production (1,000,000 cwt)

Page 18: Characterization of Dickeya dianthicola and Pectobacterium

3

soil temperature above 64 F), potato tubers are ready to sprout and grow once they break

dormancy. Lenticels on the surface of potato tubers provide pathways for gas exchange with the

outside, and the activities of lenticels are affected by soil water content (Adams 1975b;

Wigginton 1973). Lenticels enlarge in soil with high water content, which favors entry of some

opportunistic pathogens (Adams 1975a; Adams and Lapwood 1978). Mother (seed) tubers

provide nutrition for initial growth before the young plants have a well-established root system

(Davies 1984). Any seed-borne pathogens can be transported to plants during this period.

Elongation of stems and growth of leaves, along with a well-established root system constitute a

sophisticated system of nutrition exchange and maintain healthy plant establishment. Under

appropriate growing conditions, daughter tubers start to initiate from the tips of stolon and

rapidly bulk. Interruption of the leaf canopy (damage caused by foliar diseases), which produce

photosynthate, can result in reduced tuber growth rates (Tekalign and Hammes 2005). Excessive

irrigation and inappropriate nutrient management can also affect tuber growth, resulting in

greater susceptibility to tuber diseases (Ojala et al. 1990). Tubers are close to mature as potato

vines die. During tuber maturation, skin or periderm thickens and hardens, which can block entry

of pathogens to the tuber (Singh et al. 2020). Properly matured tubers also have greater

resistance to pathogens in storage. However, tubers will have higher respiration rates when they

are over mature, and then are more susceptible to decay in storage (Marquez-Villavicencio et al.

2011).

1.2 Potato diseases

The growth of potatoes can be affected by various environmental factors, such as

temperature, soil type, water, cultivar, as well as all kinds of disease. Under certain conditions,

potato tubers serve as a nutritive source for many pathogens. There are at least 80 pests and

Page 19: Characterization of Dickeya dianthicola and Pectobacterium

4

diseases caused by insects, nematodes, viruses, bacteria, fungi and oomycetes, which can affect

all parts of a potato plant, especially harvested tubers that are the most economically important

parts (Fiers et al. 2012).

Tuber diseases can directly cause economic losses by reducing yield and quality of

tubers. Tuber diseases include viral diseases such as net necrosis caused by potato leafroll virus

(PLRV), dry rot caused by the fungal pathogen Fusarium spp., soft rot caused by

Pectobacterium spp. and Dickeya spp. and ring rot caused by Clavibacter michiganensis subsp.

sepedonicus, as well as Pythium leak and pink rot caused by Pythium spp. and Phytophthora

erythroseptica respectively (Gudmestad et al. 2007; van der Wolf and De Boer 2007). These

pathogens are responsible for significant losses in the field as well as storage problems. Although

common scab caused by Streptomyces scabies and powdery scab caused by Spongospora

subterranea have little or no influence on total yield, they can severely blemish the tuber

periderm, detract for tuber appearance, and drastically reduce the value in the fresh market (Dees

and Wanner 2012; Harrison et al. 1997). More importantly, S. subterranea is a vector for potato

mop top virus (PMTV) (Jones and Harrison 1969).

Foliar diseases affect potato production by disturbing photosynthesis and reducing plant

growth (Waggoner and Berger 1987). Foliar diseases mainly show up from the tuber initiation to

harvest stages. One example of a foliar disease is late blight (Phytophthora infestans), which is

best known for causing the Irish Potato Famine, resulting in the starvation deaths of one million

people and at least another one million people had to leave their homeland (Donnelly 2002;

Zwankhuizen et al. 1998). Late blight starts with small pale to dark green spots and spreads to

entire leaflets, eventually killing the entire plant when not controlled (Zwankhuizen et al. 1998).

In addition, other foliar diseases can also be constraints to production. For example, early blight

Page 20: Characterization of Dickeya dianthicola and Pectobacterium

5

(Alternaria solani), white mold (Sclerotinia sclerotiorum) and grey mold (Botrytis cinerea)

cause discoloration and concentric or irregular rings on the leaf surface, which affect potato

production (Stevenson et al. 2001).

Stem and vascular diseases include stem canker/black scurf (Rhizoctonia solani), white

mold (Sclerotinia sclerotiorum), early dying/Verticillium wilt (Verticillium spp.) and blackleg

(Pectobacterium spp. and Dickeya spp.) (Stevenson et al. 2001). Stem canker is the most

common fungal disease of potatoes. Besides canker symptoms on stems, R. solani can produce

black sclerotia on the tuber surface (black scurf) reducing the value in the fresh market (Fiers et

al. 2012). Stem lesions of white mold begin when stems are in contact with soil containing

surviving pathogen spores. White mold can girdle potato stems and cause vine wilt or death.

Pectobacterium spp. and Dickeya spp. can colonize host xylem interrupting the transportation of

water and dissolved minerals (De Boer and Rubio 2004). Once the community population of

Pectobacterium spp. and Dickeya spp. reaches a certain level, these bacteria will secrete pectic

enzymes to degrade cell wall structure and cause the whole plant to collapse (Põllumaa et al.

2012).

Disease establishment requires a susceptible host, an available pathogen, and favorable

environmental conditions (Scholthof 2007). Pathogens causing diseases on potatoes may

originate from seed pieces, plant debris in soil, weed hosts, or even contaminated machinery like

seed cutters or storage conditions (Stevenson et al. 2001). Many fungal or oomycetes pathogens

produce survival structures to remain in the soil throughout the non-growing season. These

include oospores of Phytophthora spp., sclerotia of Rhizoctonia and Verticillium spp., spore balls

of Spongospora subterranea, and chlamydospores of Fusarium spp. Pathogens surviving on/in

seed tubers can infect plants systemically and transmit to daughter tubers, which are called seed-

Page 21: Characterization of Dickeya dianthicola and Pectobacterium

6

borne pathogens. Bacterial diseases, such as ring rot, blackleg and soft rot are common seed-

borne potato diseases. Survival structures will serve as a primary inoculum and start the disease

cycle when reaching a susceptible plant under favorable conditions. In the growing season,

disease can transmit from infected plants to healthy ones by wind, rain, insects, and irrigation

water (Binyam 2014; Radcliffe and Ragsdale 2002; Scholthof 2007).

1.3 Blackleg and soft rot of potato

1.3.1 Disease symptoms

Blackleg and soft rot of potato (PBSR) are named based on symptoms on different tissues

of potato. Blackleg is a consequence of decayed stems, while soft rot is a symptom shown on

tubers both prior- and post-harvest (Perombelon and Kelman 1980). PBSR is primarily a seed-

borne disease. Seed tubers carrying the pathogens may be decayed at an earlier stage in the field

and result in missed emergence, which is often termed as pre-emergence blackleg or blanking

(Figure 1-2). Successfully emerged seedlings from infected seed pieces could be characterized

by blackened decay on the basal stem with length ranging from less than one inch up to the entire

stem, followed by rapid wilting and yellowing of leaves (Perombelon and Kelman 1980).

Blackleg symptoms can develop during any stage of plant development. The stolon attachment

site of tubers produced from blackleg-infected plants may be first blackened and decayed, and

the entire tuber may decay as the disease progresses under moist conditions or may remain intact

until conditions favorable for disease development return. Internal soft rotted tissues are wet and

mushy, cream to tan-colored with a black border between the diseased and healthy tissues

(Figure 1-2). Symptoms of tuber soft rot can also become sunken, dry and hard under dry

conditions. In seed potato production, the disease can cause important losses mainly due to

downgrading and rejection of seed lots in certification schemes. The extent of losses varies

Page 22: Characterization of Dickeya dianthicola and Pectobacterium

7

widely from country to country and is influenced by the climate as well as the conditions of

growth and storage.

Figure 1-2. Symptoms of blackleg and soft rot of potato. Symptoms include low emergence in

the field (left panel), blackleg on stems (middle), and soft rot on tuber (right).

1.3.2 Disease cycle

It is now generally accepted that PBSR can be caused by seed-borne bacteria, such as D.

dianthicola and P. atroseptica, also soil-borne bacteria, such as D. solani and P. carotovora

(Parkinson et al. 2015; Perombelon and Kelman 1980). Therefore, PBSR can be initiated from

multiple sources. Soil-borne bacteria can survive in soil for a long time and penetrate seed tubers

through natural openings or wounds. Seed tubers harboring the soil-borne and seed-borne

bacteria play a critical role in the disease cycle (Ansermet et al. 2016; De Boer and Rubio 2004).

When a contaminated or infected seed potato is planted, symptoms will develop or not

depending on weather conditions. If pathogens are confined to the lenticels, decay of the seed

tuber occurs first and may cause non-emergence in the field. Pathogens can also move through

the vascular tissue directly into the growing plant, which will form blackleg and collapse the

stem if bacterial populations become great enough. However, when conditions are favorable for

the growth of the potato plant, no disease may occur even when blackleg-causing bacteria are

Page 23: Characterization of Dickeya dianthicola and Pectobacterium

8

present. Whether diseases show up during growth or not, daughter tubers generated from the

contaminated mother tubers may carry blackleg bacteria. In storage, the contaminated tubers may

decay or remain symptomless. Then the disease cycle starts again if the symptomless but

contaminated tubers are used for planting in the following season (De Boer and Rubio 2004).

PBSR pathogens are believed to be dormant when present in potato tubers under stressed

environmental conditions, such as low storage temperature. One dormant status, called the viable

but non-culturable state (VBNC), has been described for more than 80 bacterium species (Pinto

et al. 2015). Bacteria can convert to a VBNC state when nutrient starved or in a stressed

environment, caused by chemicals, nutrient stress, temperature, etc. Cells converted to VBNC

state underwent a non-recoverable stage of existence, but still remain viable (Xu et al. 1982).

Although the VBNC state cannot be detected by conventional culture techniques because of the

non-culturability, they do take up nutrients, produce new biomass and maintain active

metabolism (Mederma et al. 1992). Furthermore, cells in VBNC state can still be pathogenic

after release from stress or when triggered by some compounds (Pinto et al. 2015).

The disease cycle of these pathogens is facilitated by their epiphytic and saprophytic

survival for extended periods of time in the phyllosphere, rhizosphere, soil, and plant debris. The

soft rot bacteria cannot survive over winter in soil, with the length of survival restricted to 1

week to 6 months, depending on environmental conditions such as soil temperature, moisture

and pH (Czajkowski et al. 2011; Perombelon and Kelman 1980). However, soft rot bacteria can

survive longer on plants, such as weeds and volunteers (Czajkowski et al. 2011). Although some

of the PBSR pathogens, such as Dickeya zeae, D. dadantii and D. aquatica have been frequently

isolated from surface water in the US and Europe (Cother et al. 1992; Duprey et al. 2019;

Ekbataniamiri 2020; Parkinson et al. 2014), indicating a long-time survival, their low titer may

Page 24: Characterization of Dickeya dianthicola and Pectobacterium

9

not allow a direct infection of potato plants in fields. In other words, it is not proven that

blackleg bacteria can be spread through soil or water between multiple cropping seasons.

1.3.3 The occurrence of PBSR worldwide and outbreaks in the US

PBSR, especially caused by Dickeya dianthicola and D. solani, have emerged in

European countries in the past decades and severely affected local potato industries (Toth et al.

2011; van der Wolf et al. 2014). Dickeya dianthicola was first detected causing stunting and

slow wilting of Dianthus in the early 1950s in Denmark, then in the Netherlands and the UK

(Hellmers 1958; Parkinson et al. 2015). The presence of D. dianthicola strains has been reported

on potato or ornamental plants in more than 10 countries, including USA, UK, Belgium, France,

Morocco, Netherlands, Pakistan, Switzerland, Israel, Serbia etc. (Jiang et al. 2016; Marković et

al. 2020; Nasaruddin et al. 2019; Oulghazi et al. 2017; Rosenzweig et al. 2016; Sarfraz et al.

2018; Toth et al. 2011).

In 2013 and 2014, problems of low potato emergence were observed in Maine fields by

Dr. Steven B. Johnson and his colleagues. In the following years, potato plants with blackened

and decayed stems were recorded in the field, along with severe tuber decay in storage (Johnson

2016). Similar situations also happened in other eastern states in the U.S., including New York,

New Jersey, Massachusetts, Pennsylvania, Virginia (Charkowski 2018; Ma et al. 2018; Patel et

al. 2019). Twenty-four states in the U.S. have now been identified with PBSR symptoms.

Dickeya dianthicola was identified as the predominant bacterial pathogen in the outbreak

(Jiang et al. 2016). Pectobacterium parmentieri was later detected in the diseased potato stems

and tubers (Ge et al. 2018b). As a seed-borne disease, related pathogens are believed to be

distributed and spread through seed transportation. Spread of contaminated seed potatoes from

Maine, as the biggest potato seed producer for the northeastern area, was considered to be

Page 25: Characterization of Dickeya dianthicola and Pectobacterium

10

responsible for the outbreak. Tracing and analysis of the geographically distributed pathogens

could help explain the inoculum of the outbreak. Furthermore, it was also believed that D.

dianthicola was present in seed potatoes and farms in the affected states for at least a few years.

Abundant rainfall in 2013 and 2014 spread the pathogen, but cool temperatures caused dormancy

or latency. Warmer temperatures in 2015 resulted in the significant disease outbreak on

commercial farms (Johnson 2016).

Pectobacterium species are typical soil inhabitants and are well adapted to living as

saprophytes in the soil until a host plant is nearby. Therefore, Pectobacterium spp. always exist

in rot samples as secondary invaders associated with other major pathogens (Charkowski 2018).

Increasingly diversified Dickeya and Pectobacterium likely result from the evolution of gene

repertoires, differences in pathogenicity, and the presence or absence of different virulence

factors (Duprey et al. 2019).

1.3.4 The pathogens of PBSR

1.3.4.1 Taxonomy of PBSR pathogens

PBSR can be caused by multiple species belonging to the closely related Dickeya and

Pectobacterium species (Charkowski 2018; Czajkowski et al. 2011; De Boer and Rubio 2004).

The bacteria are Gram-negative, non-sporing, and facultative anaerobes. They are characterized

by producing large quantities of extracellular pectic enzymes which can macerate plant tissues

along with a wide range of other plant cell wall-degrading enzymes (PCWDEs) (Collmer and

Keen 1986).

PBSR was initially recognized to be caused by several species in the genus Erwinia,

including E. carotovora subsp. carotovora, E. carotovora subsp. atroseptica and E.

chrysanthemi (Perombelon and Kelman 1980). These species were called soft rot Erwinia

Page 26: Characterization of Dickeya dianthicola and Pectobacterium

11

because they share the ability of producing large quantities of pectic enzymes, which

distinguished these species from other Erwinia species. Since 1981, E. carotovora was further

differentiated into three subspecies, including E. carotovora subsp. betavasculorum isolated

from vascular necrosis of sugar beet in 1981, E. carotovora subsp. wasabiae isolated from

wasabi samples in 1987, and E. carotovora subsp. odoriferum from slimy rot witloof chicory in

1992 (Gallois et al. 1992; Goto and Matsumoto 1987; Thomson et al. 1981). In 1998, by

analyzing the 16S rRNA gene, Hauben et al. proposed to place the above six into one cluster, and

reassigned them into genus Pectobacterium (Hauben et al. 1998). Later on, the three subspecies

of P. carotovora, P. carotovora subsp. betavasculorum, P. carotovora subsp. wasabiae, P.

carotovora subsp. atroseptica were then elevated to species level, which are P. betavasculorum,

P. wasabiae, P. atroseptica now (Gardan et al. 2003). Up to then, soft rot Erwinia included P.

carotovora subsp. carotovora, P. carotovora subsp. odoriferum, P. atroseptica, P. chrysanthemi,

P. betavasculorum, P. wasabiae. In 2004, a new species, P. carotovora subsp. brasiliensis was

isolated from potato in Brazil (Duarte et al. 2004). Pectobacterium wasabiae isolated from

potato plants was renamed as P. parmentieri based on whole genome analysis in 2016 (Khayi et

al. 2016). New species P. polaris was isolated from potato in 2017 (Dees et al. 2017). Currently,

soft rot bacteria include P. carotovora subsp. carotovora, P. carotovora subsp. odoriferum, P.

carotovora subsp. brasiliensis, P. atroseptica, P. chrysanthemi, P. betavasculorum, P.

parmentieri and P. polaris (Table 1-1). However, the Pectobacterium complex also contains far

more than the discussed species.

In 1998, P. chrysanthemi was divided into six pathovars (Lelliott and Dickey 1984;

Young et al. 1978), which were then transferred into a new genus Dickeya, including D.

chrysanthemi, D. paradisiaca, D. dadantii, D. dianthicola, D. dieffenbachiae and D. zeae based

Page 27: Characterization of Dickeya dianthicola and Pectobacterium

12

Table 1-1. Development of etiology and taxonomy of blackleg and soft rot bacteria.

Before

1981

1981 1987 1992 1998 2003 2004 2005 2012 2014 2016 2017 Conclusion

Erwinia chrysanthemi

Brenneria paradisiaca Dickeya paradisiaca D. paradisiaca

Dick

eya sp

p.

Pectobacterium chrysanthemi

D. chrysanthemi D. chrysanthemi

D. zeae D. zeae

D. dadantii D. dadantii subsp. dadantii D. dadantii

D.

diffenbachiae

D. dadantii subsp. diffenbachiae

D. dianthicola D. dianthicola

D. solani

D. solani

D. aquatica D. aquatica

D. fangzhongdai D. fangzhongdai

E. carotovora subsp. atroseptica

P. carotovora

subsp.

atroseptica

P. atroseptica

P. atroseptica

Pecto

ba

cterium

spp

.

E. carotovora subsp. carotovora P. carotovora subsp. carotovora P. carotovora

subsp.

carotovora

E. carotovora subsp.

betavasculorum

P. carotovora

subsp.

betavasculoru

m

P. betavasculorum P.

betavasculorum

E. carotovora subsp.

wasabiae

P. carotovora

subsp.

wasabiae

P. wasabiae P. parmentieri

P. wasabiae

P. parmentieri

E. carotovora

subsp.

odoriferum

P. carotovora subsp. odoriferum P. carotovora

subsp. odoriferum

P. carotovora subsp. brasiliensis P. carotovora

subsp.

brasiliensis

P.

polaris

P. polaris

Page 28: Characterization of Dickeya dianthicola and Pectobacterium

13

on 16S rRNA gene-based phylogenetic analyses in 2005 (Samson et al. 2005). The above six

Dickeya spp. were then reduced to five by dividing D. dadantii into two subspecies and placing

all strains in D. dieffenbachiae as a subspecies within D. dadantii (D. dadantii subsp. dadantii

and D. dadantii subsp. dieffenbachiae) (Brady et al. 2012). More recently, three additional

Dickeya species have been described: D. solani, isolated from potatoes in Europe and hyacinth

(Sławiak et al. 2009; van der Wolf et al. 2014), D. aquatica from freshwater rivers

(Ekbataniamiri 2020; Parkinson et al. 2014), and D. fangzhongdai from pear trees displaying

symptoms of bleeding canker in China (Tian et al. 2016). So far, the species in genus Dickeya

include D. chrysanthemi, D. dadantii, D. dianthicola, D. solani, D. zeae, D. aquatica, D.

paradisiaca and D. fangzhongdai (Table 1-1).

1.3.4.2 Host range

Host range and host specificity of pathogens are conventionally determined by their

serological and temperature characteristics, which reflects their geographic distribution (Ma et al.

2007). Dickeya spp. and Pectobacterium spp. are considered as broad-host-range pathogens, with

the exceptions of P. atroseptica and D. solani that are restricted mostly to potatoes, and D.

paradisiaca which is restricted only to plantain (Table 1-2). Most Dickeya spp. infect a wide

range of crops in tropical and subtropical regions, as well as greenhouse crops grown in

temperate regions. Since potato is widely cultivated in various regions, Dickeya spp. have

become important in potato production. Dickeya dianthicola is a serious pathogen that is known

to occur in many countries such as Australia, European countries, Israel, and the U.S. and others,

and has resulted in significant yield losses to potato crops (Charkowski 2018).

Page 29: Characterization of Dickeya dianthicola and Pectobacterium

14

In addition to potatoes, D. dianthicola can also infect other crops and ornamentals, such

as carnation, lily, chrysanthemum, dahlia, hyacinth, dianthus, chicory and globe artichoke.

Dickeya zeae causes maize stalk rot, rice foot rot and banana soft rot diseases in many countries

and regions (Nassar et al. 1994; Pu et al. 2012; Samson et al. 2005; Sinha and Prasad 1977;

Zhang et al. 2014). Dickeya zeae has not been found and isolated from potatoes in the U.S.,

although isolation from irrigation water and inoculation tests have shown that D. zeae can be

pathogenic to potatoes (Ekbataniamiri 2020). Dickeya dadantii causes soft rot on many crops and

ornamentals as well as potato (Ngadze et al. 2010). Pathogenicity of D. aquatica has been

confirmed on cucumber, tomato and potato under high concentration, indicating that it is a

potential pathogen of potato. However, D. aquatica is much more pathogen to aquatic

charophyte hosts due to the lack of xylanases and xylose (Duprey et al. 2019). Since D. aquatica

was only isolated from surface water near irrigation ponds with very low concentrations, it may

not result in potato infection in the field. As the newest Dickeya species, D. fangzhongdai was

first reported on pear trees and later was isolated from decayed onions, phalaenopsis and

poinsettia (Alič et al. 2018).

In contrast to Dickeya spp., Pectobacterium spp. have a wide distribution in both the

temperate and tropical zones and are pathogenic to a much wider range of crops. For example, P.

carotovora subsp. carotovora has a wide host range including potato, cabbage, carrot, celery,

cucumber, maize, and hyacinth (Ma et al. 2007). Pectobacterium carotovora subsp. brasiliensis

was first found in Brazil, and later in New Zealand, Canada, and South Africa to be pathogenic

to multiple hosts, including cucumber, sugar beet, cabbage, Chinese cabbage, tomato, pepper,

potato, eggplant, nepenthes and zucchini (Duarte et al. 2004). Pectobacterium odoriferum was

first isolated from chicory with odorous smell, and then also found in multiple crops, including

Page 30: Characterization of Dickeya dianthicola and Pectobacterium

15

leek, celery, parsley, carrot, onion, and potato (Gallois et al. 1992; Ma et al. 2007).

Pectobacterium parmentieri can infect potato, horseradish, carrot, cabbage, maize, sugar beet,

and calla lily (Khayi et al. 2016). Pectobacterium betavasculorum was viewed as a pathogen

restricted to sugar beet, but later was isolated from sunflower and artichoke (Gardan et al. 2003).

Table 1-2. Confirmed hosts of Dickeya spp. and Pectobacterium spp. causing blackleg and soft

rot of potato.

Species Host

D. solani Potato

D. paradisiaca Plantain

D. zeae Maize, rice, banana

D. chrysanthemi Chrysanthemum, dendranthema, grandiflorum, dianthus caryophyllus, potato,

etc.

D. dianthicola Carnation, lily, chrysanthemum, dahlia, hyacinth, dianthus chicory, globe

artichoke, potato, etc.

D. dadantii Pepper, eggplant, tomato, tobacco, sweet potato, broccoli, radishes, celery,

carrot, sugar cane, sorghum, rice, pineapple, urn plant, onion, orchids, tulip,

chicory, chrysanthemums, carnations, hyacinths, etc.

P. atroseptica Potato

P. polaris Potato

D. aquatica No host, but can be pathogenic to many crops

D. fangzhongdai Onion, pear trees, phalaenopsis, poinsettia

P. betavasculorum Sugar beet, sunflower, artichoke

P. carotovora

subsp. odoriferum

Chicory, leek, celery, parsley, carrot, onion, potato

P. parmentieri Carrot, cabbage, maize, sugar beet, calla lily, horseradish, potato etc.

P. carotovora

subsp. brasiliensis

Cucumber, sugar beet, cabbage, tomato, pepper, eggplant, nepenthes, zucchini,

potato

P. carotovora

subsp.

carotovora

Arum, cabbage, carrot, celery, cotton, cucumber, cyclamen, delphinium,

hyacinth, maize, sugar cane, tobacco, potato

Page 31: Characterization of Dickeya dianthicola and Pectobacterium

16

1.3.4.3 Pathogenesis of PBSR pathogens

Disease establishment of PBSR starts with the entry of the bacteria into the host tissue

through a natural opening or wound. Under favorable conditions, bacteria multiply to a high titer.

In order to infect host plants, pathogenic bacteria translocate various chemicals on/in host

through different secretion systems. Secreted chemicals included adhesins, effectors, biofilm-

related chemicals, pectinolytic enzymes, and quorum-sensing signals. There are six types of

secretion systems that are more or less related to the pathogenicity of soft rot bacteria.

Bacteria firstly attach to the plant surface with the formation of a biofilm by secreting

adhesins such as extracellular polysaccharides (Yousef and Espinosa-Urgel 2007). The adhesion

is governed by some genes in association with the type I secretion system (T1SS) gene cluster.

For example, the hecA gene in D. dadantii encodes adhesin of the hemagglutinin (Rojas et al.

2002). In P. atroseptica, genes encoding the components of the T1SS were identified along with

the gene encoding a multirepeat adhesin protein (MRP) (Pérez‐Mendoza et al. 2011). Through

T1SS, cytoplasmic proteins of various sizes, such as the adhesins, are secreted directly from the

cytoplasm across the inner and outer membranes in a periplasm-independent manner (Tseng et

al. 2009).

After attaching to the plant surface, the plant cell wall is the first structural barrier that

soft rot bacteria must overcome during infection. To overcome this barrier, soft rot bacteria

produce multiple PCWDEs, including pectin methylesterases, pectate lyases, pectin lyases,

pectin acetylesterases, and poly-galacturonases, most of which are translocated from bacterial

cells through the type II secretion system (T2SS) (Charkowski et al. 2012; Salmond 1994; Toth

et al. 2006).

Page 32: Characterization of Dickeya dianthicola and Pectobacterium

17

Effector proteins translocated from bacterial cells to host plants play an important role in

their virulence. Effectors that define the outcome of host-pathogen interaction are secreted by the

type III secretion system (T3SS) directly into host cells. Once inside host cells, effectors

manipulate host defenses and promote bacterial growth. The utilization of T3SS has been found

in D. dadantii, P. carotovora, P. atroseptica, and D. zeae (Bell et al. 2002; Ham et al. 1998; Li

et al. 2012; Rantakari et al. 2001; Yang et al. 2002). However, not all Pectobacterium spp.

require the same secretion system. For example, T3SS is absent in P. parmentieri. In addition to

the well-known type I, II, and III secretion systems, the function of type IV, V and VI secretion

systems are also believed to support the infection process by soft rot bacteria (Bell et al. 2004;

Glasner et al. 2008; Li et al. 2012; Toth et al. 2003).

Genomic studies of P. atroseptica, P. carotovora subsp. brasiliensis, and D. zeae have

revealed the existence of a type IV secretion system (T4SS) (Glasner et al. 2008; Li et al. 2012;

Mattinen et al. 2008), but the full function in virulence has yet to be identified. The putative

T4SS gene virB4 only has been found to affect the lesion development by P. atroseptica on

potato stems (Bell et al. 2004). Knowledge of the biological function of the type V secretion

system (T5SS) and type VI secretion system (T6SS) in soft rot bacteria is still limited.

Quorum sensing (QS) is an intercellular signaling mechanism allowing bacterial cells to

sense the population density and then modulate cellular behaviors when the population reaches a

certain threshold. Activities associated with QS include production of virulence factors,

formation of biofilms, control of motility, and biosynthesis of secondary metabolites (Decho et

al. 2010; von Bodman et al. 2003). Novel biocontrol strategies based on the disruption of

bacterial communication through QS are being developed. In the soft rot bacteria, QS is largely

dependent on N-acylhomoserine lactone- (AHL-) type signaling molecules, followed by

Page 33: Characterization of Dickeya dianthicola and Pectobacterium

18

autoinducer-2 (AI-2) (Põllumaa et al. 2012). Pectobacterium and Dickeya spp. are quite different

in that AHL-mediated QS plays an important role in virulence of P. carotovora and P.

atroseptica but is less obvious in the virulence of D. dadantii (Chatterjee et al. 2010; Ham et al.

2004).

1.3.5 Management of PBSR

An integrative strategy, combining the use of certified seed tubers, hygienic practices

during harvesting and grading, and avoidance of PBSR introduction during planting and in

storage, has been partially effective, but has not resulted in a broad eradication of the PBSR in

the potato production chain (Perombelon and Kelman 1980; Toth et al. 2011). Management

approaches, including exclusion, use of resistant varieties, and protection through chemical and

biological control, have been developed and implemented with varying degrees of success.

1.3.5.1 Exclusion

Exclusion strategies are often regulatory and involve the use of seed certification,

quarantine, and plant inspection to prevent disease spread. Seed certification schemes are widely

used to manage PBSR and have been partially successful. This method involves sampling

representative tubers from seed lots, testing tuber tissues for pathogens and estimating blackleg

risk assessment (Perombelon 2000). However, latent infection of progeny tubers from

symptomless plants can easily escape diagnosis, which may increase the risk of exclusion failure.

Minitubers grown in a controlled pathogen-free environment are used to obtain pathogen-free

seed (Farran and Mingo-Castel 2006; Rolot and Seutin 1999). Avoidance of wounding at any

steps during harvesting and grading is important to reduce the risks of introducing bacteria

through wounds.

Page 34: Characterization of Dickeya dianthicola and Pectobacterium

19

Potato plants in a seed production system are usually subjected to field inspections twice

during the growing season. During field inspections, removal of diseased plants, including

daughter tubers contribute to reducing PBSR inoculum. Moreover, removal of decayed tubers

during harvesting and grading can also reduce the spreading and smearing of the bacteria in the

seed lots (Elphinstone and Pérombelon 1986). Spreading of the causal agents across plants or

fields can be controlled by sanitation procedures in storage facilities, as well as on cutters or

other equipment (Pérombelon et al. 1979).

1.3.5.2 Resistant varieties

It is challenging to breed a highly resistant potato variety due to limitations of available

germplasm. Breeding for resistant varieties has so far had relatively little success. No hybrid

lines that are immune to PBSR have been commercialized to date (Czajkowski et al. 2011).

Screening for resistance is not straightforward, as cultivar resistance or susceptibility to PBSR is

not always measurable. Moreover, the extremely complicated pathogen community capable of

causing PBSR makes the target of breeding much more difficult. Some white-skined potato

varieties such as ‘Shepody’ and ‘Atlantic’ show a good level of PBSR tolerance (Ge et al.

2020a). It is also encouraging that the newly released variety ‘Caribou Russet’ bred by the

University of Maine breeding program has shown good resistance against PBSR (Ekbataniamiri

2020).

Physiological features, such as mature tubers with a well-developed periderm will reduce

risk of wounding. Using true seeds, derived from sexual crosses which are believed to be free

from blackleg and soft rot bacteria, can be an alternative way to reduce the incidence of

infections (Chujoy and Cabello 2007). A high calcium content in crops is often positively related

Page 35: Characterization of Dickeya dianthicola and Pectobacterium

20

to increased resistance against bacterial diseases, including potato blackleg (McGuire and

Kelman 1984; Pagel and Heitefuss 1989).

1.3.5.3 Protection

Physical, chemical, and biological controls can be considered in protecting potatoes from

PBSR. Physical factors involve hot water, steam, dry hot air and ultraviolet and solar radiation

(Bartz and Kelman 1985; Bdliya and Haruna 2007; Ranganna et al. 1997; Robinson and Foster

1987; Shirsat et al. 1991). These methods have been widely used for several decades. However,

their limitations are the inability to kill plant pathogenic bacteria located deep inside the tubers or

plant xylem without negative effects on plant growth. Also, bulky seed tubers prepared for each

growing season result in substantial operational costs and the difficulty of expanding for large-

scale use (Czajkowski et al. 2011).

Chemical control strategies are usually used based on the eradication of the pathogen

and/or the creation of unfavorable environmental conditions (e.g. low or high pH, etc.) to inhibit

disease development (Czajkowski et al. 2013a). Chemical treatments applied on seed tubers are

the most efficient way to control PBSR. However, extensive use of antibiotics is no longer

allowed for large-scale field applications due to the high risk of development of antibiotic

resistance in bacterial pathogens. Some chemicals have indirect effects by enhancing plant

growth or altering plant physiology. For example, salicylic acid can induce the systemic acquired

resistance defense mechanism, which has been effective at reducing infection symptoms caused

by D. solani in tissue-culture-grown potato plants (Czajkowski et al. 2015a). Reduction of PBSR

causal agents by physical and chemical treatments is only partially effective and is still in

experimental development (Czajkowski et al. 2011; van der Wolf and De Boer 2007).

Page 36: Characterization of Dickeya dianthicola and Pectobacterium

21

Biological agents have been studied for PBSR control. Pseudomonas spp., including P.

putida and P. fluorescens are frequently tested as biological agents (Des Essarts et al. 2016;

Kastelein et al. 1999). Some other bacteria have been also studied (Czajkowski et al. 2012;

Garge and Nerurkar 2017). Bacteriophages have been tested for the control of bacterial plant

diseases for more than 30 years, such as the control of Agrobacterium tumefaciens, Xanthomonas

campestris pv. pruni, X. oryzae etc. (Jones et al. 2007). The advantages of bacteriophages over

other biocontrol agents include specificity to their bacterial hosts and non-pathogenicity for

humans and animals (Jones et al. 2007). Bacteriophages were isolated and used to infect Erwinia

and Dickeya spp., which can infect bacterial host effectively (Czajkowski 2016; Czajkowski et

al. 2015b; Czajkowski et al. 2014; Smolarska et al. 2018; Soleimani-Delfan et al. 2015).

Unfortunately, no commercial biocontrol products are currently available for the control of

PBSR.

1.4 Microbiomes

The microbiome is defined as a characteristic microbial community occupying a

reasonable well-defined habitat which has distinct physio-chemical properties (Berg et al. 2020).

A microbial community is a multi-species assemblage, in which microorganisms share a

common living space in natural environments, such as soil, water, and animal or plant host, and

interact with each other. In agroecosystems, soil microbial communities can influence the

physiology and development of associated plants (Mendes et al. 2013). Plant microbiota

comprise all microorganisms in/on plants, which is also called the plant microbiome and is

comprised of all microbial genomes. Plant-associated microbiota includes diverse microbes such

as archaea, bacteria, fungi, nematodes, and protists, and the micro biota is a key determinant of

plant health and productivity (Berendsen et al. 2012; Compant et al. 2019; Turner et al. 2013).

Page 37: Characterization of Dickeya dianthicola and Pectobacterium

22

1.4.1 Plant microbiome

The rhizosphere (roots and underground parts), phyllosphere (plant aerial surfaces) and

endosphere (internal tissues) are the most investigated microbial reservoirs on plants. The nature

of plant-associated microbiota is plant genotype dependent. The rhizosphere is a region of rich

microbial diversity, which is influenced by deposition of plant exudates, mainly dominated by

the bacterial phyla Acidobacteria, Verrucomicrobia, Bacteroidetes, Proteobacteria,

Planctomycetes and Actinobacteria (Fierer 2017; Kent and Triplett 2002). Dominant microbes

may be diverse as the structure of plant-associated microbiota in rhizosphere is largely plant

genotype dependent (Chaparro et al. 2014; Hartmann et al. 2009; Ladygina and Hedlund 2010;

Reinhold-Hurek et al. 2015). Microbes in the rhizosphere can obtain carbon sources from root

exudates, sloughed root cells or the release of mucilage. The root exudates, in particular, contain

a variety of compounds and are key determinants of rhizosphere microbiome structure (Bertin et

al. 2003). In addition to root exudates, microbes in the rhizosphere can degrade sloughed root

cells and mucilage and obtain a large amount of material, such as plant cell wall polymers

(cellulose or pectin), as food source, (Dennis et al. 2010). In contrast to the rhizosphere, the

phyllosphere niche is relatively nutrient poor and easily subjected to extremes of temperature,

radiation and moisture.

Microbial inhabitants of the rhizosphere and phyllosphere are considered as epiphytes.

Microbes residing within plant tissues (the endosphere) whether in leaves, roots or stems, are

referred to as endophytes. Endophytes are thought to be a sub-population of the rhizosphere

microbiome, partially because not all microorganisms from the rhizosphere can enter plants

(Compant et al. 2010). Endophytes are generally non-pathogenic, but they also include latent

Page 38: Characterization of Dickeya dianthicola and Pectobacterium

23

pathogens that can cause disease under certain environmental circumstances and/or with certain

host genotypes (James and Olivares 1998; Monteiro et al. 2012).

Microbial communities and their interactions with plant hosts are of high significance to

plant health. Plant hosts and the associated microorganisms mutually influence their lifestyles.

Plant hosts provide nutrient sources and niche space for microbial growth, while microbes in

these niches can establish beneficial, neutral or detrimental (pathogenic) associations with their

host plants. Microorganisms benefit plant health with a wide range of effects, including disease

suppression, priming of the plant immune system through microbe-associated molecular patterns

(MAMPs) or pathogen-associated molecular patterns (PAMPs). They can also increase nutrient

acquisition, tolerance to abiotic stresses, adaptation to environmental variation, or the

establishment of mycorrhizal associations (Garbaye 1994; Haney et al. 2015; Mendes et al.

2011; Ritpitakphong et al. 2016; Rolli et al. 2015; Van der Ent et al. 2009; Van Der Heijden et

al. 2016; Zamioudis et al. 2015).

1.4.2 Plant pathobiome

Within the context of microbial community interactions, pathobiome is defined by

Vayssier-Taussat et al. (2014) to represent the pathogenic agent integrated within its biotic

environment. In this case, “pathobiome” has been proposed to replace “pathogens”, indicating a

leveraged understanding of plant diseases from “one microbe-one disease” to “multi-microbes-

one disease” (Vayssier-Taussat et al. 2014). Pathobiome plays a direct role of triggering or

influencing disease progression by establishing multiple interactions with other microorganisms

within plant microbiota. Such microbial consortia represent the pathogenic agent integrated

within its biotic environment and make disease establishment/development much more complex

(Mina et al. 2020). Pathobiome can be viewed as an approach to study plant diseases by looking

Page 39: Characterization of Dickeya dianthicola and Pectobacterium

24

at more factors though it may not replace one-pathogen for one disease, because non-pathogenic

microbes can be considered as one kind of environmental factor.

Currently in plant pathology, the role and interactions of pathobiome have been focused

on disease establishment or development, rather than being focused on a single species of

disease-causing agent. The latter overlooks the crucial biotic complexity encountered in plant

hosts. This microflora is usually a ubiquitous constituent of the plant host. This microbiota can

cooperate with pathogens via specific pathways, such as quorum sensing signal sharing,

metabolic sharing/complementarity, resulting in a mutualistic interaction (Egland et al. 2004;

Kim et al. 2008). The mutualistic interactions between pathogens and other microorganisms can

result in a more aggressive disease when co-inoculations are made compared with single

inoculations (da Silva et al. 2014).

1.4.3 Microbial interactions

Microbiome is an important component of plant health. Microorganisms have various

interactions, which are essential to maintain the stability and functioning of microbial

communities, including plant pathogens (Singer 2010). The prevalence of multiple infections in

the field indicates that the extent of complex interactions, known as co-infection, can be

significant in some disease systems (Tollenaere et al. 2016).

Co-infection has become of particular interest since it tends to alter the course of disease

and the overall virulence of multiple pathogens (Tollenaere et al. 2016). Before co-infection was

studied, host-pathogen interaction was the only well-studied key interaction when investigating

disease development or other topics (Brown 2015), while pathogen-pathogen interaction, and

host-multiple-pathogen interactions have been less studied. Here, pathogen-pathogen interactions

will be reviewed in detail, and can be classified into three types: 1) competition, in which

Page 40: Characterization of Dickeya dianthicola and Pectobacterium

25

competing pathogens exclude competitors from resource-dense niches by physical or chemical

approaches; 2) cooperation, whereby pathogens benefit from each other by sharing or

exchanging substances essential for pathogenesis or survival; 3) coexistence, whereby pathogens

can stably coexist through niche specialization (Abdullah et al. 2017).

1.4.3.1 Competition and coexistence

Theoretically, in a defined space with sufficient nutrients, many microbes including

pathogens can stably coexist (Monod 1949). Because host resources are limited, pathogens often

enter into fierce competition, fundamentally for growth- and fitness-limiting resources (West et

al. 2006). The limited resources, such as nutrients, can be utilized by one pathogen and then

other pathogens sharing the similar nutritional requirements will be restricted. However, some

species may dominate within a space of suboptimal nutrition. Competition under such conditions

may result in the selection for more virulent species, or conversely, all associated pathogens may

suffer from insufficient resources (Rankin et al. 2007). In addition to competing for resources,

more aggressive forms of competition between pathogens include direct chemical exclusion,

such as secreting toxins to prevent competitors’ growth (Mousa et al. 2015), or mechanical

aggression, like forming multilayer physical barriers or special structure to limit the growth or

infection of competitors (Mousa et al. 2016). Competition can also happen indirectly by

mediating the plant defense system against other pathogens (Na and Gijzen 2016).

To avoid or reduce the severity of the competition, negotiation among pathogens can

occur. Niche specializations in space and time enable pathogen species to coexist stably when

localized in different tissues of the host, or by arriving at different time intervals (Fitt et al.

2006).

Page 41: Characterization of Dickeya dianthicola and Pectobacterium

26

1.4.3.2 Cooperation and synergism

Besides competition and peaceful coexistence, plant pathogens may also cooperate to

enhance their pathogenicity. This relationship has been determined in many pathosystems, such

as biophysical transport, in which bacteria are transported by Didymella bryoniae resulting in co-

infection of Styrian oil pumpkin (Grube et al. 2011). Scanning electronic microscopy (SEM) or

confocal microscopy combined with fluorescence labeled microbes can visually demonstrate the

spatial distribution or interaction, such as clear niche specialization or adherence (Murray et al.

2014).

Microbes not only interact through direct contact. cooperation crossing kingdoms can

also occur through biochemical means. An example is Rhizopus microsporus causing rice blight

which owes its pathogenicity to phytotoxins secreted by its endosymbiont Burkholderia sp.

(Partida-Martinez and Hertweck 2005). QS is also one example of bacterial cooperation in which

a group of bacteria share the extracellular signals to regulate certain processes together (Cook et

al. 2013). Therefore, QS signals (AHLs, AI-2) are important chemical factors for synergy during

infection (Armbruster et al. 2010; Duan et al. 2003). Between closely related species,

complementation of growth substrates is a common strategy and prominent when resources are

deficient (Hoek et al. 2016). Certain microbes are only host destructive when they cooperate with

other independent pathogens, such as the endophytes which are considered as nonpathogenic or

only causing nonvisible symptoms (Partida-Martinez and Heil 2011). Complex dynamics

involving other microbes may alter the properties of endophytes.

Plant diseases can be the result of multispecies with synergistic disease effects. In many

cases, plants may not show severe disease symptoms when infected by a single microbe

(pathogen), but are subjected to severe symptoms when co-infected by multiple microbial

Page 42: Characterization of Dickeya dianthicola and Pectobacterium

27

species due to synergistic interaction (Lamichhane and Venturi 2015). Such synergistic

interactions in plants may be of crucial importance for the understanding of microbial

pathogenesis and evolution. This increases the complexity of the disease and has to be taken into

consideration in the development of more effective control strategies (Lamichhane and Venturi

2015). Tomato pith necrosis is caused by at least eight bacterial species Pseudomonas cichorii,

P. corrugata, P. viridiflava, P. mediterranea, P. fluorescens, Pectobacterium atroseptica,

Pectobacterium carotovora, and Dickeya chrysanthemi (Alivizatos 1985; Dhanvantari and Dirks

1987; Goumans and Chatzaki 1998; Malathrakis and Goumas 1987; Paula Wilkie and Dye

1974; Saygili et al. 2008; Scarlett et al. 1978). Disease severity is greatly enhanced when co-

infection of multiple bacterial species occurs (Kůdela et al. 2011; Moura et al. 2004).

1.4.4 Sequencing-based methods for studying microbiomes

Historically, microbiomes have been studied using culture-based techniques, which allow

researchers to isolate and characterize microbes for biological studies of microorganisms. But it

has limitations because the majority of microorganisms are not culturable (Steen et al. 2019).

Due to the complexity of polymicrobial disease, their study has been somewhat overlooked when

only using the culture-based techniques. In recent decades, culture-independent techniques have

been extensively applied to explore microbiomes and disease complex. Sequencing has

facilitated the study of microbiome and pathobiome as well as the underlying mechanisms of

plant diseases (Lamichhane and Venturi 2015).

Metagenomics allow the identification of microbes in situ. At the microbiome level,

huge datasets (e.g., metagenomics and metabolomics) are often generated using homogenized,

whole-tissue samples, by sequencing the variable regions of the ubiquitous genes. For example,

16S ribosomal RNA (rRNA) and internal transcribed space (ITS) have been universally used for

Page 43: Characterization of Dickeya dianthicola and Pectobacterium

28

a high-resolution (species- and strain- level) taxonomic identification of all members in a

specific microbiome (Chakravorty et al. 2007; Schoch et al. 2012).

Sequencing technologies have evolved rapidly over the last 40 years. Since Sanger

sequencing technology was invented as the first generation, we have moved to the third

generation sequencing (Sanger and Coulson 1975; Sanger et al. 1977). Sanger sequencing has

the read-length ability of up to 1000 base pair (bp) with accuracy as high as 99.999% (Shendure

and Ji 2008). To enhance the capacity and speed, next generation sequencing (NGS) was

developed, which is symbolized by Roche’s 454 technology, Illumina’s Solexa, HiSeq

technology, and ABI’s Solid technology (Mardis 2008). NGS produces short reads ranging from

75 to 300 bp, but it allows us to genotype hundreds to thousands of samples in a single

experiment (Mardis 2017; Shendure and Ji 2008). Moreover, the tremendously low cost and

shortened sequencing NGS makes possible the analyses of tens of thousands of microbial

identities in a microbiome (Guttman et al. 2014). Metagenomic information from hypervariable

regions in the 16S/18S ribosomal RNA (rRNA) amplicons for bacteria or ribosomal internal

transcribed spacer (ITS) for fungi are so far preferred in microbiome research (Hiergeist et al.

2015). Many pipelines have emerged to compute the large data sets of NGS, with the commonly

used pipelines as quantitative insights into microbial ecology (QIIME) and Mothur (López-

García et al. 2018). In NGS, operating taxonomic units (OTUs) and amplicon sequence variant

(ASV) approaches are standard to determine the taxonomic origin of microbial sequences

(Callahan et al. 2017).

The third-generation sequencing, also known as long-read sequencing is the latest DNA

sequencing method that produces substantially longer reads than the NGS technique (Bleidorn

2016). The relatively long reads allow sequencing near-complete viral or small bacterial

Page 44: Characterization of Dickeya dianthicola and Pectobacterium

29

genomes with an accuracy of as high as 97 to 99% (Greninger et al. 2015). The sequencing

platform includes the PacBio SMRT (Single Molecule Real Time) and the Oxford Nanopore

(Weirather et al. 2017). The main advantage for the third-generation sequencing technologies in

metagenomics is the speed of sequencing and length of reads in comparison to NGS, but high

error rates need to be overcome in future improvement (Petersen et al. 2019). Combining NGS

and third generation sequencing have been frequently used for accurate and complete whole

genome sequencing. For example, long read sequencing by Oxford nanopore can be polished

with short reads by Illumina sequencing (Bouchez et al. 2018; Kranz et al. 2017).

1.5 Objectives

Because potato is the most important vegetable in USA, as well as in the world, the

impact of diseases, such as PBSR is substantial. The outbreak has caused huge yield and

economic losses to the growers in multiple states. Maine as a major seed producer in the

northeastern region has been heavily affected by the outbreak of the seedborne PBSR since 2015.

We have found that D. dianthicola is the primary pathogen in the PBSR outbreak, but that many

Pectobacterium spp. have also been involved. Thus, this research project focused on these

bacterial species. To understand PBSR outbreak, I was interested in how the disease initiated,

what factors affected the disease epidemic, and eventually how to prevent and control PBSR in

the future. As such, I had the following questions and hypotheses.

1). What is the origin of bacterial pathogen D. dianthicola causing the outbreak? The

hypothesis was that D. dianthicola was clonal and that the outbreak was caused by

dispersal of infected seed tubers.

Page 45: Characterization of Dickeya dianthicola and Pectobacterium

30

2). How do D. dianthicola strains causing this outbreak differ to known strains from those

other countries? The outbreak was caused by endemic D. dianthicola strains, which could

well adapt to the northeastern environmental conditions.

3). Was D. dianthicola the only causal agent of the outbreak or were multiple bacteria

involved? Potato plants exposed to multiple species of bacteria that are potentially causal

agents of PBSR.

4). How do these pathogen complexes perform during disease development? Multiple species

of causal agents have advantages in pathogen infection and disease development.

With the above questions raised, my dissertation was arranged as shown in Figure 1-3.

Figure 1-3. Flowchart of the chapters.

Page 46: Characterization of Dickeya dianthicola and Pectobacterium

31

CHAPTER 2

GENOTYPING DICKEYA DIANTHICOLA CAUSING POTATO BLACKLEG AND

SOFT ROT ASSOCIATED WITH INOCULUM GEOGRAPHY IN THE UNITED

STATES

This chapter has been previously published in Plant Disease:

Ge, T.L., Jiang, H.H., Johnson, S.B., Larkin, R.P., Charkowski, A.O., Secor, G., and Hao, J.J. (2020)

Genotyping Dickeya dianthicola causing potato blackleg and soft rot associated with inoculum geography

in the United States. Plant Disease. DOI: 10.1094/PDIS-10-20-2138-RE.

ABSTRACT

An outbreak of blackleg and soft rot of potato, caused primarily by the bacterial pathogen

Dickeya dianthicola, has resulted in significant economic losses in the Northeastern United

States since 2015. The spread of this seedborne disease is highly associated with seed

distribution, therefore the pathogen likely spread with seed tubers. To describe the blackleg

epidemic and track inoculum origins, a total of 1183 potato samples were collected from 11

northeastern states from 2015 to 2019. Of these samples, 39.8% tested positive for D.

dianthicola. Seventeen isolates of Dickeya dianthicola were recovered from these samples and

the genetic diversity of these isolates was examined. Fingerprinting with BOX-PCR and

phylogenetic analysis based on 16S rDNA and gapA sequences indicated that D. dianthicola

isolates were divided into three genotypes, denoted Type I, II, and III. Ninety five percent of

samples from Maine were Type I. Type II was found in Maine only in 2015 and 2018. Type II

was present throughout the five years in some states at a lower percentage than Type I. Type III

was found in Pennsylvania, New Jersey and Massachusetts, but not in Maine. Therefore, Type I

appears to be associated with Maine, but Type II appeared to be distributed throughout the

Page 47: Characterization of Dickeya dianthicola and Pectobacterium

32

Northeastern United States. The Type II and rarer Type III strain were closer to the D.

dianthicola type strain isolated from the United Kingdom. This work provides evidence that the

outbreak of blackleg of potato in the Northeastern United States was caused by multiple strains

of D. dianthicola. The geographic origins of these strains remain unknown.

2.1 INTRODUCTION

The bacterial disease blackleg and soft rot of potato (Solanum tuberosum) is caused by

multiple species of Dickeya and Pectobacterium (Toth et al. 2011). Both genera are broad host

range pathogens and Dickeya infects plants in at least 35% of plant orders (Perombelon and

Kelman 1980). At least seven species of Dickeya, including D. dianthicola, D. zeae, D. solani,

D. chrysanthemi, D. dadantii, D. aquatica, and D. paradisiaca, cause similar decay symptoms in

plants, although with different levels of aggressiveness (Brady et al. 2012; Samson et al. 2005;

van der Wolf et al. 2014). Dickeya dianthicola (syn. Erwinia chrysanthemi) was described for

the first time as a soft-rotting-causative agent of potato tubers in 1980 (Cother et al. 1992), and

later recognized as a blackleg-causing agent in Europe (Toth et al. 2011).

An outbreak of blackleg in potato started in Maine and many other states in the

Northeastern region in 2015 (Hao et al. 2016; Jiang et al. 2016; Johnson 2016; Ma et al. 2018;

Patel et al. 2019; Rosenzweig et al. 2016). Symptoms included non-emergence after planting,

blackened and rotted stems and wilting canopy during the growing season, and decayed and

rotten potato tubers in storage (Johnson 2017). Severe losses caused by the outbreak were

mainly associated with the bacterial pathogen D. dianthicola (Hao et al. 2016; Johnson 2016;

Ma et al. 2019). Pectobacterium species were frequently present as part of a pathogen complex.

It is now generally believed that potato blackleg caused by Dickeya and Pectobacterium are

predominately tuber-borne diseases, therefore infected seed tubers play an important role in the

Page 48: Characterization of Dickeya dianthicola and Pectobacterium

33

disease cycle and elimination of infected seed tuber lots can aid in disease management

(Pérombelon 1992). Planting contaminated seed can cause stand loss or stem infection. However,

symptoms may not occur even when conditions are favorable, which allows the pathogen to

spread asymptomatically as seed tubers are produced, and results in larger outbreaks in

subsequent years (Pérombelon 1992). In storage, the contaminated tubers can develop soft rot

symptoms or remain symptomless. The disease cycle continues if the symptomless but

contaminated tubers are used for planting (Pérombelon 1992).

Application of genotypic approaches have made separation of bacteria species more

precise. Fingerprinting methods based on the widespread distribution of repetitive DNA elements

in genomes have been widely used as a tool to increase the taxonomic resolution at the species

level (Lanoot et al. 2004; Versalovic et al. 1991). Other methods, such as phylogenetic analysis

of 16S ribosomal RNA (rRNA), genes containing hypervariable regions, or other central

metabolism genes, can also be used to determine taxonomic relationships among different

bacteria species and strains (Van de Peer et al. 1996). Single nucleotide polymorphism (SNP)

analysis reveals a relatively high variation in gene gapA (coding for the glyceraldehyde-3-

phosphate dehydrogenase A) among Dickeya and Pectobacterium spp.(Cigna et al. 2017). A

high level of strain discrimination has been detected using variable number tandem repeat

analysis among D. dianthicola strains isolated from multiple hosts, including Dianthus

caryophyllus, and Dahlia sp. and potato (Parkinson et al. 2015).

Although potatoes are grown in most U.S. states, only a few states produce large amounts

of seed potatoes. These seed potatoes are subject to certification protocols that include multiple

inspections throughout the growing season and in storage and a postharvest test (Frost et al.

2013; Powelson 2007). This system results in high quality seed potatoes, but if new pathogens

Page 49: Characterization of Dickeya dianthicola and Pectobacterium

34

enter the system, it also can allow for swift dissemination of new diseases across the U. S. Maine

is a major source of potato seed to the Eastern states. At the beginning of the Dickeya outbreak in

2014, seed potato certification programs in the Northeastern U. S. (New York and Maine) did not

test seed potatoes for blackleg pathogens, but evidence indicates tight connection between seed

contamination and field disease development as blackleg is a seedborne disease. However, no

solid data were available in determining where the inoculum originated from except observing

the geographic incidence of disease (Johnson, 2016). Since then, testing has become routine.

The purpose of the study was to understand the origin of the primary inoculum causing

the blackleg outbreak in the Northeastern U. S. and to determine if the pathogen was clonal and

of a single origin, likely in one of the seed potatoes producing states, or if multiple D.

dianthicola taxa were present and if this pathogen was likely endemic to the Northeastern U. S.

For the present study, analyses included the BOX-PCR technique, and phylogenetic analysis of

16S rRNA and central metabolism genes to conduct a genotypic classification of D. dianthicola

strains isolated from various geographical locations to address this question.

2.2 MATERIALS AND METHODS

2.2.1 Sample collection

Symptomatic potato stems and tubers were collected from 2015 to 2019 during the

growing season as well as from storage, resulting in 1,183 samples collected. Samples were

primarily collected from commercial potato production fields by growers, consultants, or

vegetable pathologists, then directly shipped to our lab for diagnostic analyses. Each sample

represented an individual potato plant in one field. These samples were predominantly from

Maine (epidemic center), with the rest being from the eastern states of the US, including

Page 50: Characterization of Dickeya dianthicola and Pectobacterium

35

Massachusetts, Pennsylvania, New York, Rhode Island, New Jersey, Virginia, North Carolina,

Maryland, Florida, Delaware and Michigan (Table 2-1).

2.2.2 Pathogen isolation

Bacterial isolates were obtained from blackened or rotted potato stems, decayed tubers, or

healthy-appearing tubers. Stem samples of potato were surface disinfected with 10% bleach

(6.25% NaOCl) for three minutes and rinsed in sterile water for three minutes. Approximately

two-centimeter segments were cut from potato stems near the transition between healthy and

diseased tissues. Five milliliters of sterile water were added to a sterile container, and then the

tissue segments were submerged, and held in the water for 5 min to allow the bacteria to release

from the tissue. To isolate bacteria from dormant tubers, a peel around the circumference of the

stem end of the tuber was submerged in sterile water in Ziploc bags. The bags were sealed after

air removal and incubated at 28 C for 48 hours.

Tissue saps derived from the above preparations were streaked on a crystal violate

pectate (CVP) plate (Hélias et al. 2012). The plates were incubated at 28 C over 24 hours.

Preliminary tests showed that D. dianthicola colonies formed later and did not grow as large as

other enterobacteria on CVP. Therefore, only small colonies that formed in cavities in CVP were

transferred to new CVP plates for isolation. Purified cultures were then transferred to tryptic soy

broth (TSB) and incubated at 28C overnight.

Pathogenicity assay was conducted with all the isolates by tuber inoculation. Four tubers

of potato ‘Lamoka’ were used per bacterial isolate as four replicates. An aliquot of 10 L of

bacterial suspension in tryptic soy broth (106 CFU/ml) was added to a pre-poked hole on the

tuber and sealed with Parafilm. The liquid medium without bacteria was used for control. The

Page 51: Characterization of Dickeya dianthicola and Pectobacterium

36

treated tubers were incubated at 23 1 C for 3 days before disease was evaluated. Disease was

evaluated by measuring the length and width of a lesion from a dissected tuber.

2.2.3 Pathogen detection and identification

Genomic DNA was extracted from the stem- or tuber-derived sap using the FastDNA

SPIN Kit (MP Biomedical, Santa Ana, CA, USA). Genomic DNA was extracted from the pure

cultures for pathogen identification using the QiAamp DNA Mini Kit (QIAGEN, Germany).

DNA concentration was measured with NanoDrop 2000 (Thermo Fisher Scientific, Wilminton,

DE, USA) and adjusted to 10 ng/L. Polymerase chain reaction (PCR) was performed using

Dickeya-specific (genus level) primers pelADE (Nassar et al. 1996) and D. dianthicola-specific

primers DIA-A (Pritchard et al. 2013). PCR was performed in a 25 L total volume reaction

containing 17.9 L of sterile water, 5 L of 5x PCR buffer, with MgCl2 at a final concentration

of 0.25 mM dNTPs, 0.1 M of each pair of primers, 0.5 U GoTaq DNA Polymerase and 1 L of

DNA. The PCR reagents were purchased from Promega (Madison, MI). Primers were

synthesized in Integrated DNA Technologies (www.idtdna.com/pages).

2.2.4 Fingerprinting analysis

BOX-A1R-based repetitive extragenic palindromic-PCR (BOX-PCR) was used to

analyze genotypes of the bacterial isolates. The BOX element (BOX1A) was amplified using the

BOXA1R primer 5′-CTACGGCAAGGCGACGCTGACG-3′ (Koeuth et al. 1995). PCR

amplifications were performed as described above, except using 0.2 M of the primer. The

thermal cycler was programmed for an initial denaturation of 7 min at 95°C, followed by 30

cycles of 1 min at 94°C, 1 min at 53°C and 8 min at 65°C, with a final elongation of 15 min at

65°C. PCR amplicons were examined by electrophoresis of 6 L aliquots mixed with 2 L of

10 gel green (Biotium, Fremont, CA) through 1.5% agarose gels in TAE buffer.

Page 52: Characterization of Dickeya dianthicola and Pectobacterium

37

2.2.5 Primer design for genotyping

Whole genome sequences of Dickeya dianthicola strain RNS04.9 (accession number:

CP017638.1) and strain ME23 (CP031560.1) were downloaded from the NCBI GenBank

(https://www.nlm.nih.gov/). Synteny analysis was performed using the MAUVE software

(Darling et al. 2004). Based on the genome comparison, strain-specific genes were selected to

design primers. Chromosome regions (ME23: TraC_4 1576824-1577768, RNS04.9: DDI_1427

1444469-1446052) was used for primer design using integrated DNA technologies. Primers

Anti_1 is specific to RNS04.9 which generates a 434-base-pair (bp) product, while primers

TraC_4 is specific to ME23 which generates a 418 bp product (Table 2-4). The two new

designed primers were tested for optimal annealing temperature by using gradient PCR, and

55C was determined for both pairs of primers. The specification of the two new pairs of primers

were validated by running PCR with strains in Type I and Type III. All the blackleg samples that

were DIA-A positive were re-analyzed using the new primer pairs Anti_1 and TraC_4. Number

of each genotype in each state was shown with the reanalyzed PCR results using the software

Tableau 10.2 (Tableau, Seattle, WA, USA).

2.2.6 Phylogenetic analysis

Segments of the 16S rRNA gene and segments of partial three central metabolism genes

(dnaX, recA and gapA) were amplified using PCR with primer pairs 27F/1492R, dnaXf/dnaXr,

recAf/recAr and gapA-7F/gapA-938R, respectively (Table 2-2). Amplicons generated by PCR

were purified using UltraClean 15 DNA purification kit (MO BIO, Carlsbad, CA). Amplicons

generated from dnaX, recA and gapA genes were directly sequenced using the corresponding

primer pairs. Amplicons generated from 16S rDNA were sequenced with four primers 519R,

907R, 1492R and 907F to guarantee the sequencing quality. Sequence data of gapA from 16

Page 53: Characterization of Dickeya dianthicola and Pectobacterium

38

additional D. dianthicola were obtained from NCBI GenBank (Table 2-3). Phylogenetic analysis

of 16S rDNA and gapA was performed among the strains isolated in this study using MEGA7

(Kumar et al. 2016). While only gapA was used for phylogenetic analysis for comparison among

worldwide strains. A maximum likelihood phylogenetic tree was constructed with 500 bootstrap

replicates.

2.3 RESULTS

2.3.1 Disease detection and pathogen isolation

A total of 1183 samples of diseased potato plants with typical blackleg symptoms were

collected from 11 northeastern states in US. Most of the samples (680) came from Maine, where

disease incidence was higher than other states. Among these samples, 42%, 52%, 19% and 27%

diseased samples collected from 2015, 2016, 2017 and 2018 were D. dianthicola positive, but no

D. dianthicola was found in 2019 samples (Table 2-1). Representative states where a large

number of samples (>100) were collected included Pennsylvania and Massachusetts, where D.

dianthicola was detected in 39% and 46% of the total samples from each state, respectively.

Compared to Maine, the number of samples collected from Delaware, New Jersey, New York,

Rhode Island and Virginia were lower, which ranged from 12 to 69, while percent of D.

dianthicola positive samples ranged from 29% to 53%. The total number of samples received

each year declined from 512 to 61 in the years 2015 to 2019. This is likely due to greatly

increased testing by seed potato growers and buyers that resulted in elimination of infected seed

lots from the potato industry during the course of this study.

D. dianthicola grows slower than species in the closely related genus Pectobacterium and

Pectobacterium is often present in samples with Dickeya. As a result, isolation of D. dianthicola

was challenging and D. dianthicola isolates were recovered from only a small percentage of

Page 54: Characterization of Dickeya dianthicola and Pectobacterium

39

samples that tested positive for Dickeya. In total, seventeen isolates of D. dianthicola were

obtained from plant samples collected between 2015 and 2016 (Table 2-2). Results of

pathogenicity assay showed that no differences were found among all isolates of D. dianthicola.

Table 2-1. Detection of Dickeya dianthicola in potato samples collected from Northeastern US,

and a smaller amount from Midwestern and Southern regions.

State 2019 2018 2017 2016 2015 Total

Delaware - - 1 (1) 7 (5) 4 (3) 12 (9)

Maine 7 x (0 y) 34 (9) 43 (8) 144 (77) 452 (170) 680 (264)

Maryland - - - 2 (2) - 2 (2)

Massachusetts 12 (4) 12 (0) 18 (12) 65 (32) 3 (3) 110 (51)

Michigan - - 9 (1) - - 9 (1)

New Jersey - - 25 (3) 33 (10) 11 (7) 69 (20)

New York 3 (0) 1 (0) 1 (0) 19 (10) 6 (5) 30 (15)

North Carolina - - - 6 (1) - 6 (1)

Pennsylvania 34 (8) 34 (5) 54 (27) 34 (10) 36 (24) 192 (74)

Rhode Island - - 18 (3) 25 (15) - 43 (18)

Virginia 5 (1) 1 (0) 13 (7) 11 (8) - 30 (16)

Total 61 (13) 82 (14) 182 (62) 346 (170) 512 (212) 1183

(471)

“-“: no sample available.

x total number of samples collected.

y the number of samples showing D. dianthicola positive.

Page 55: Characterization of Dickeya dianthicola and Pectobacterium

40

Table 2-2. Dickeya dianthicola isolates obtained in this study.

Genotype Bacterial

Strain

State Plant tissue Target

gene

GenBank

Accession No.

Target

gene

GenBank

Accession No.

Type I ME30 Maine Potato stem 16S rRNA MN732907 gapA MN735191

Type I 16ME21T Maine Potato tuber 16S rRNA MN732910 gapA MN735196

Type I 16ME22T Maine Potato tuber 16S rRNA MN732909 gapA MN735195

Type I 16SBJ16 Maine Potato stem 16S rRNA MN732908 gapA MN735190

Type I 16LI01 New York Potato stem 16S rRNA MN732913 gapA MN735202

Type I 16LI02 New York Potato stem 16S rRNA MN732914 gapA MN735201

Type I 16LI04 New York Potato stem 16S rRNA MN732905 gapA MN735200

Type I 16JP03 Virginia Potato stem 16S rRNA MN732915 gapA MN735189

Type I 16JP05 Virginia Potato stem 16S rRNA MN732916 gapA MN735187

Type I 16MA08CT Massachusetts Potato tuber 16S rRNA MN732912 gapA MN735199

Type I 16MA15T Massachusetts Potato tuber 16S rRNA MN732906 gapA MN735198

Type I PA24 Pennsylvania Potato stem 16S rRNA MN732917 gapA MN735188

Type I 16MB01 New Jersey Potato stem 16S rRNA MN732920 gapA MN735192

Type II 16PA07 Pennsylvania Potato stem 16S rRNA MN732911 gapA MN735197

Type II 59W Maine Water 16S rRNA MN732921 gapA MN735203

Type II 16NJ11 New Jersey Potato stem 16S rRNA MN732919 gapA MN735194

Type III 16NJ12 New Jersey Potato stem 16S rRNA MN732918 gapA MN735193

2.3.2 Genotyping by fingerprinting and DNA sequencing of D. dianthicola

Three unique genomic fingerprint patterns were discriminated with BOX-PCR (Figure 2-

1, Table 2-2). The fingerprint patterns separated the 17 examined D. dianthicola strains into

three genotypes, which were designated as Type I, II and III.

Following BOX-PCR fingerprinting, D. dianthicola strains were further analyzed with

Sanger sequencing of 27F/1492R PCR amplicons, which covered nearly the whole region of 16S

rDNA. PCR amplification yielded products of approximately 1,500 bp when visualized on

agarose gels. According to the published whole genomes of D. dianthicola, there are seven

copies of 16S rRNA gene. Sanger sequencing reactions produced chromatograms with

overlapping peaks at some nucleotide sites as expected due to 16S rRNA polymorphisms. Both

in strain ME23 (CP031560.1) and RNS04.9 (CP017638.1), 20 out of 1535 bp nucleotide sites

were shown same polymorphisms within the seven copies, while another 11 sites of

polymorphisms were only present in strain RNS04,9 and another 3 sites only present in strain

ME23 (data was not shown here). Seven copies of 16S rRNA genes in strains collected in this

Page 56: Characterization of Dickeya dianthicola and Pectobacterium

41

study were separated by sequence alignment with seven copies of 16s rRNA genes from

annotated genomes (CP031560.1 and CP017638.1). The allele with the highest variability was

selected for phylogenetic analysis. Dickeya dianthicola strains ME23 and RNS04.9 were used as

references, and D. dadantii 3937, D. solani IPO2222 and Erwinia amylovora CFBP1430 were

used as outgroups.

There were three single nucleotide polymorphisms (SNPs) in the region of gapA having

850 bp in length. Both dnaX and recA sequences were 100% identical among all isolates.

Therefore, sequences of 16S rRNA and gapA were used to genotype the strains. Strains in Type I

were grouped together by phylogenetic analysis of sequences of both 16S rRNA and gapA. Type

II and III were not separated by phylogenetic analysis, which may be a result of the higher

similarity between these two genotypes (Figure 2-2).

Because 16S rDNA sequences were not available for some D. dianthicola strains in

NCBI, gapA was again reanalyzed by adding more strains from countries out of USA, including

UK, France, Belgium, Morocco, Pakistan, Switzerland (Table 2-3). To confine the results to

country-to-country comparison, specific geographic information of American states was not

included here. The type I Dickeya dianthicola isolates from USA were clustered together in a

group that did not include strains from Europe or Africa. The four Type II and Type III isolates

were clustered with D. dianthicola strains isolated from other countries (Figure 2-3).

Page 57: Characterization of Dickeya dianthicola and Pectobacterium

42

Table 2-3. Dickeya dianthicola strains retrieved from NCBI database.

NCBI reference

strain Country Isolated from

Collection

Year

GenBank

Accession No.

ME23 USA Potato 2015 CP031560.1

DE440 USA Potato 2016 PJJB01000011.1

WV516 USA Potato 2016 PJJC01000019.1

Ddi1 USA Potato No record Unknown

CFBP2015 France Potato 1975 VZQG01000010.1

SS70 Pakistan Potato 2017 QESZ01000027.1

CFBP1888 France Potato 1978 VZQE01000002.1

RNS 11-47-1-1A France Potato 2011 VYSC01000033.1

CFBP2982 France Potato 1988 VZQF01000020.1

S4.16.03.P2.4 Morocco Potato 2016 QZDN01000072.1

MIE34 Switzerland Potato 2013 VZQH01000012.1

S4.16.03.LID Morocco Potato 2016 QZDO01000064.1

GBBC2039 Belgium Potato No record CM001838.1

IPO980 Netherlands Potato No record CM002023.1

NCPPB3534 Netherlands Potato 1987 CM001840.1

NCPPB453 UK Dianthus caryophyllus 1957 CM001841.1

RNS04.9 France Potato 2004 CP017638.1

Page 58: Characterization of Dickeya dianthicola and Pectobacterium

43

Figure 2-1. Gel electrophoresis of BOX PCR on Dickeya dianthicola strains. Most strains were isolated from potato tissues in 2016,

except ME30 and PA24 were isolated in 2015, and 59W was isolated from irrigation water in 2017. Strains were separated into three

types by DNA band patterns of BOX PCR.

Page 59: Characterization of Dickeya dianthicola and Pectobacterium

44

Figure 2-2. Phylogenetic tree using maximum-likelihood analysis on Dickeya dianthicola strains

collected in this study, reference strains (IPO2222, 3939, ME23 and RNS04.9) and Erwinia

amylovora CFBP1430 (outgroup) obtained from the GenBank. The tree was constructed based

on integrated sequences of partial sequences the gapA (850 bp) and 16S rDNA (1440 bp).

Numbers next to the branching points indicate the relative support from 500 bootstrap replicates,

with only scores above 60 are shown. The number (0.010) of the scale represents the percentage

(1%) of genetic variation for the length of the scale.

Page 60: Characterization of Dickeya dianthicola and Pectobacterium

45

Figure 2-3. Phylogenetic tree using maximum-likelihood analysis on Dickeya dianthicola strains

collected in this study and NCBI, representing seven countries, including USA, UK, France,

Belgium, Morocco, Pakistan, Switzerland. Labels in the phylogenetic tree were made up by

domestic location and strain ID. Dickeya dadantii 3937, D. solani IPO222 and Erwinia

amylovora CFBP1430 were referred as outgroup strains. The tree was constructed based on

integrated sequences of partial sequences the gapA (850 bp). Numbers next to the branching

points indicate the relative support from 500 bootstrap replicates, with only scores above60 are

shown. The number (0.020) of the scale represents the percentage (2%) of genetic variation for

the length of the scale.

Page 61: Characterization of Dickeya dianthicola and Pectobacterium

46

2.3.3 Identification of genotypes of D. dianthicola

Reference strain ME23 was assigned as Type I, while reference strain RNS04.9 was

assigned as Type III by the phylogenetic analysis of 16S rRNA and gapA genes. By aligning the

whole-genome sequences of reference strains ME23 and RNS04.9, gene traC_4 producing DNA

primase TraC was selected as a strain-specific region for ME23, while DDI_1427 producing

antirestriction protein was selected as a unique region of RNS04.9 (Figure 2-4). Two pairs of

primers specific to Type I (Trac_4) and III (Anti_1) were designed accordingly (Table 2-4).

Table 2-4. Primers (F: forward; R: reversed) used for genotypic analysis of Dickeya dianthicola.

Primer Sequence (5’-3’) Target Annealing

Tm (C)

References

pelADE F: GATCAGAAAGCCCGCAGCCAGAT

R: CTGTGGCCGATCAGGATGGTTTTGTCGTGC

Dickeya spp. 58 (Nassar et al. 1996)

TraC_4 F: ACGGCACGACAGTGATTT

R: AACTCGGCGATCAACTCTTC

Type I D.

dianthicola

55 This study

Anti_1 F: CGCATCGTAGAGAAAGGTCAA

R: GACTGAGGTTCAGGCAGTTTAG

Type III D.

dianthicola

55 This study

DIA-A F: GGCCGCCTGAATACTACATT

R: TGGTATCTCTACGCCCATCA

Type I, II and III

D. dianthicola

55 (Pritchard et al. 2013)

gapA 7-F: ATCAAAGTAGGTATCAACGG

938-R: TCRTACCARGAAACCAGTT

gapA 54 (Cigna et al. 2017)

16S

rRNA

27F: AGAGTTTGATCCTGGCTCAG

519R: GWATTACCGCGGCKGCTG

907R: CCGTCAATTCMTTTRAGTTT

1492R: GGTTACCTTGTTACGACTT

16S rDNA 59 (Frank et al. 2008;

Mao et al. 2012)

Primer pairs Trac_4 (Type I specific) and Anti_1 (Type III specific) were confirmed by

PCR amplification of all 17 strains. The results were consistent with the fingerprinting results

(Table 2- 5, Figure 2-1). All Type I strains showed positive results with Trac_4, while Type III

strains showed positive result with Anti_1. Sufficient genomic data were lacked to construct

primers that could identify Type II strains. Therefore, an exclusive method was used to

determine if samples were likely infected with Type II strains which might be a heterogeneous

Page 62: Characterization of Dickeya dianthicola and Pectobacterium

47

Type II group. To be specific, strains or samples that were DIA-A positive, Trac_4 negative and

Anti_1 negative were classified into Type II.

Figure 2-4. Multiple genome analysis of Dickeya dianthicola strains using MAUVE for

identifying a region for primer design. A: Pairwise alignment of genomes of D. dianthicola

strains ME23 (on the top) and RNS04.9 (on the bottom). The red boxes represent a homologous

region between the genomes that are internally free from genomic rearrangement. B: Zoomed-in

image of the regions in the red boxes in panel A, which was used for primer design.

Page 63: Characterization of Dickeya dianthicola and Pectobacterium

48

Table 2-5. Genotype detection of Dickeya dianthicola strains using polymerase chain reaction

with primers Trac_4, Anti_1, and DIA-A.

Genotype Strain Trac_4 Anti_1 DIA-A

Type I ME23 (reference) Not available Not available Not available

Type I ME30 + - +

Type I 16ME21T + - +

Type I 16ME22T + - +

Type I 16SBJ16 + - +

Type I 16LI01 + - +

Type I 16LI02 + - +

Type I 16LI04 + - +

Type I 16JP03 + - +

Type I 16JP05 + - +

Type I 16MA08CT + - +

Type I 16MA15T + - +

Type I PA24 + - +

Type I 16MB01 + - +

Type II 16PA07 - - +

Type II 59W - - +

Type II 16NJ11 - - +

Type III 16NJ12 - + +

Type III RNS04.9 (reference) Not available Not available Not available

After validation of the primers Trac_4 and Anti_1, samples collected from 2015 to 2019

were re-analyzed using DIA-A, Trac_4 and Anti_1. Out of the 471 samples that were positive

for D. dianthicola confirmed by PCR with the DIA-A primers, since DNAs of many samples

were progressively degraded, 256 DNA samples showing a high quality were selected and

further analyzed using Trac_4 and Anti_1. Of 71 samples collected in Maine in 2015 and re-

analyzed, Type I D. dianthicola was detected in 67 (94%) of the samples. The remaining 4

samples appeared to be infected with Type II or a yet to be described type. No samples from

Maine in 2015 were Type III (Figure 2-5). Re-analyzed samples collected after 2015, including

106 in 2016, 53 in 2017, 13 in 2018 and 13 in 2019, represented all the listed northeastern states,

although the majority was from 2016 and 2017. Samples from New York, Rhode Island and

Page 64: Characterization of Dickeya dianthicola and Pectobacterium

49

Delaware in 2016 and 2017 were all associated only with Type I D. dianthicola. Samples from

2018 and 2019 in those states were not analyzed because of the quality of DNA obtained from

the sample was insufficient. Although Type I was predominant in Pennsylvania, New Jersey and

Virginia samples, there were 10% to 33% Type II (or a yet-to-be described type) D. dianthicola

detected in samples from these states from 2016 to 2019. Type III was first observed in samples

in 2016, comprising 33% and 40% of the positive samples in Pennsylvania and New Jersey,

respectively. All samples collected from Massachusetts in 2016 and 2017 contained Type I D.

dianthicola, while all four positive samples collected in 2019 were Type III. Overall, Type I D.

dianthicola was the predominant genotype through the five years, comprising 95% of positive

samples in Maine, and 83% of the samples from all other states. The type II genotype or another

uncharacterized genotype was continuously present in at least one state every year through the

five years, but at relatively lower percentages than Type I. Type III was only detected in 2016 in

Pennsylvania and New Jersey, and 2019 in Massachusetts (Figure 2-5).

Page 65: Characterization of Dickeya dianthicola and Pectobacterium

50

Figure 2-5. Genotypic distribution of Dickeya dianthicola in the Northeastern United States

from 2015 to 2019. Total of 256 isolates were used for analysis. Each pie represents a state of

sampling of the year, with the size of pie being the number of samples, and the number

associated with each color being a sample size of a specific genotype. Genotypes included blue

circle: Type I, orange circle: Type II, and green circle: Type III.

Page 66: Characterization of Dickeya dianthicola and Pectobacterium

51

2.4 DISCUSSION

The 2015 blackleg epidemic was first observed in potato crops planted with seed potatoes

originating in Maine (Johnson, 2016; 2017). As such, it was suspected that the original

contamination in other states initiated from Maine. Initial results suggested multiple origins for

this outbreak since a strain isolated from potatoes in Wisconsin and Texas in 2016 is genetically

distinct from a strain isolated in Maine (Ma et al. 2019; Nasaruddin et al. 2019). The Maine

hypothesis was based on the fact that blackleg is a tuber borne disease, that Maine is the primary

seed potato supplier to states in the Northeastern U. S., and that at least some of the infected

crops in these states had been planted with seed potatoes from Maine. The outbreak led to

comprehensive testing of seed potatoes for Dickeya in Maine and a large reduction in losses due

to this pathogen within just a few years. This affected the current study by reducing the number

of samples that could be collected, and also demonstrated the efficacy that pathogen testing in

seed potato certification has in disease management.

In this study that blackleg disease in this outbreak was not caused by a single genotype of

D. dianthicola strain, although it is not clear whether these strains all came from the same

location. This contrasts with the D. solani outbreak in Europe and the Middle East, which was

caused by a single genotype that was probably a recent introduction to the potato industry

(Parkinson et al. 2015). By fingerprinting with BOX PCR, three D. dianthicola genotypes were

identified, namely Type I, II and III. Although they were discriminated in different groups, there

were no differences of pathogenicity and virulence among the groups. As expected, these types

were strongly related to geographic locations. The vast majority of samples collected from

Maine, and most of the samples from other states, had Type I D. dianthicola. This result is

expected if contaminated seed tubers from Maine were subsequently planted in other states.

Page 67: Characterization of Dickeya dianthicola and Pectobacterium

52

However, from these results, it was not clear if the states outside Maine had D. dianthicola

before or during the outbreak. Genome-wide analysis will be performed in the following work to

assist on further investigation of D. dianthicola causing potato blackleg in USA and other

countries that export agricultural products to USA.

Interestingly, a small portion of samples showed Type II D. dianthicola. In Maine, Type

II was only found in few samples in 2015 and 2018, but not 2016 and 2017. This showed a

chronic disconnection and random distribution. In contrast, Type II was consistently detected

every year outside Maine. I hypothesize that Type I was likely associated with Maine seed

origination, but Type II may have been indigenous across all examined states or that Type II also

initiated from Maine but was more adaptive in the southern region outside Maine. More

importantly, Type III D. dianthicola was found in Pennsylvania, New Jersey and Massachusetts,

but never detected in Maine isolates. Information on where the seed potatoes originated from

was lacked for these samples, but our data suggests that the Type III strain may have spread from

another seed potato producing state or province since it has yet to be found in Maine. Dickeya

dianthicola has been reported in Wisconsin, which also exports seed to states on the east coast,

and strains from Wisconsin differ from those in Maine (Nasaruddin et al. 2019). Dickeya

chrysanthemi was also recently reported in potato, showing that additional Dickeya species are

present in the potato production system and suggesting that other species could cause future

epidemics (McNally et al. 2018).

I suspect that the pathogen has occurred in most potato growing areas at low inoculum

levels in either potato or in alternate hosts. Although D. dianthicola is a seedborne pathogen, it

does survive up to 6 months in soil and it can become established in surface water (Pérombelon

1992). In addition, the pathogen can be harbored and transported by other plants, such as weeds

Page 68: Characterization of Dickeya dianthicola and Pectobacterium

53

and ornamentals and it can initiate an epidemic by spreading from other plant species (Parkinson

et al. 2015). Dickeya is well-known for causing asymptomatic infections that lead to severe

outbreaks when environmental conditions change. The weather conditions in 2014, which had

exceptional levels of rainfall, could have exacerbated this outbreak. Predicted climate change

effects, including warmer temperatures and more severe rain events will provide conditions

favorable to D. dianthicola, so the potato industry is likely to see future epidemics of this

pathogen.

Unlike some other important plant pathogens, there is no suggested center of origin for

soft rot Pectobacteriaceae. These data that were present here show that D. dianthicola strains in

the United States are more diverse than those reported from Europe or Africa. The Type I strain

was not found in collections other than in isolates from the United States, although the limited

number of strains available from other locations does not allow us to determine whether it is

present on other continents. The Type II and Type III strains were found in the United States,

Europe, and Africa. There has been no legal potato import from Europe or Africa into the United

States since 1913 and exports of potato from the United States to Europe or Africa are

uncommon. This suggests that Type II and Type III strains have either been long established in

all three regions or that they were spread by other hosts, such as ornamental plants. A large

number of ornamentals are imported annually from Europe and Africa into the United States and

from the United States throughout the world, providing an invasion route for D. dianthicola in

either direction (Wocial 2013).

Irrigation water has been investigated as one of the dispersal media of D. dianthicola

(Elphinstone et al. 2014). D. dianthicola isolated from river water in the United Kingdom has

helped explain the cross-infection between potato and ornamental hosts (Elphinstone et al. 2014).

Page 69: Characterization of Dickeya dianthicola and Pectobacterium

54

The only D. dianthicola strain, 59W, isolated from water in Maine belonged to Type II. Type II

D. dianthicola was consistently distributed in the southern states, like Pennsylvania, New Jersey

and Virginia, but not in Maine. This suggests that either the seed potatoes originated in a state

where Type II is more prevalent or that this strain is more aggressive in southern states than it is

in Maine, possibly because it is better adapted to higher temperature.

In conclusion, the outbreak of blackleg of potato in the Northeastern United States was

caused by multiple strains of D. dianthicola, which have most likely existed in various potato

production areas for many years. Contaminated potato seed tubers have exacerbated the disease

epidemics in multiple ways, but the disease may also have been enhanced by recent changes in

weather conditions. In managing the disease, certified seed and pathogen testing are always keys

to reduce primary infection, but the change of environmental factors should be considered in

disease forecasting.

Page 70: Characterization of Dickeya dianthicola and Pectobacterium

55

CHAPTER 3

PANGENOMIC ANALYSIS OF DICKEYA DIANTHICOLA STRAINS REVEALS THE

OUTBREAK OF BLACKLEG AND SOFT ROT ON POTATO IN USA

ABSTRACT

Dickeya dianthicola has caused an outbreak of blackleg and soft rot of potato in the

eastern half of the United States since 2015. To investigate genetic diversity of the pathogen, a

comparative analysis was conducted on genomes of D. dianthicola strains. Whole genomes of 16

strains from the US outbreak were fully assembled and compared to 16 reference genomes

retrieved from the NCBI GenBank. Among the 32 strains, eight distinct clades were

distinguished based on phylogenomic analysis. The outbreak strains were grouped into three

clades, with the majority of the strains in clade I. Clade I strains were unique and homogeneous,

suggesting a recent incursion of this strain into potato production from alternative hosts or

environmental sources. Pangenome of the 32 strains contained 6693 genes, 3377 of which were

core genes. By screening primary protein subunits associated with virulence from all USA

strains, it was found that virulence-related gene clusters, such as plant cell wall degrading

enzyme genes, flagellar and chemotaxis related genes, two-component regulatory genes, and

type I/II/III secretion system genes were highly conserved but type IV and type VI secretion

system genes varied. The virulent clade I strains encoded two clusters of type IV secretion

systems, while clade II and III strains encoded only one cluster. Clade I and II strains encoded

one more VgrG/PAAR spike protein than clade III. Thus, it was predicted that the presence of

additional virulence-related genes may have enabled the unique clade I strain to become the

predominant source of the US outbreak.

Page 71: Characterization of Dickeya dianthicola and Pectobacterium

56

3.1 INTRODUCTION

The Pectobacteriaceae pathogens Dickeya and Pectobacterium are pectolytic bacteria

that cause blackleg and soft rot (BSR) on potato. These pathogens also infect numerous other

economically important vegetables, grain and ornamental crops (Adeolu et al. 2016; Dees et al.

2017; Duarte et al. 2004; Ekbataniamiri 2020; Gardan et al. 2003; Khayi et al. 2016; Ma et al.

2007; Parkinson et al. 2014; Perombelon and Kelman 1980; Samson et al. 2005; Sławiak et al.

2009; Tian et al. 2016; van der Wolf et al. 2014).

Dickeya dianthicola has been reported in more than 10 countries in the regions of North

America, Europe, North Africa, and the Middle East on potato, vegetable or ornamental plants

(Jiang et al. 2016; Marković et al. 2020; Nasaruddin et al. 2019; Oulghazi et al. 2017; Patel et al.

2019; Rosenzweig et al. 2016; Sarfraz et al. 2018; Toth et al. 2011). Dickeya dianthicola and D.

solani have emerged in European countries in the past decades and have severely affected the

European potato industries (Toth et al. 2011; van der Wolf et al. 2014). Dickeya dianthicola was

first detected causing stunting and slow wilting of Dianthus in the early 1950s in Denmark, then

in the Netherlands and UK (Hellmers 1958; Parkinson et al. 2015). Although D. solani has not

been found in the United States, D. dianthicola has caused an outbreak in the northeast U.S.

since 2015 (Jiang et al. 2016; Johnson 2016; Ma et al. 2018). Recently, Ge et al. (2020b) have

discovered that strains collected from outbreak samples can be divided into three distinct

genotypes, with one predominant genotype. Furthermore, D. dianthicola populations from

multiple hosts in European regions can be divided into additional genotypes (Parkinson et al.

2015).

Dickeya is a pectinolytic genus that produces large quantities of secreted plant cell wall

degrading enzymes (PCWDE) that are required for pathogenicity and that contribute to bacterial

Page 72: Characterization of Dickeya dianthicola and Pectobacterium

57

growth in plants (Salmond 1994). The PCWDE and other pathogenesis- and survival-related

proteins or effectors are secreted or transported to the target cells via various secretion systems

(Russell et al. 2014; Wallden et al. 2010). For example, PCWDEs are secreted through the type

II secretion system (T2SS) under the control of quorum sensing (QS) system (Charkowski et al.

2012; Salmond 1994; Toth et al. 2006). Type III secretion system (T3SS) contributes to

pathogenesis of some pectinolytic bacteria, including D. dadantii, P. carotovora, P. atroseptica,

and D. zeae (Bell et al. 2002; Ham et al. 1998; Li et al. 2012; Rantakari et al. 2001; Yang et al.

2002). Other genes/gene clusters related to pathogenicity include genes required for flagella

motility and chemotaxis, iron acquisition, polysaccharides and biofilm synthesis, quorum

sensing, and CRISPR-Cas systems (Blin et al. 2020; Li et al. 2018; Motyka-Pomagruk et al.

2020).

Type IV secretion system (T4SS) promotes genetic exchange and/or effector

translocation with consequent impacts on virulence and genome plasticity. The T4SS is

classified into three main types: conjugative, DNA release and uptake, and effector-delivery

(Alvarez-Martinez and Christie 2009). The conjugative and DNA release and uptake systems

play important role in horizontal gene transfer, which allows prokaryotes to adapt to their

environment, and is a powerful driver in the acquisition of novel functions by bacterial

pathogens (Cascales and Christie 2003). In Gram-negative bacteria, the structure of T4SS is

composed of 12 protein subunits (virB1-virB11 and virD4). In general, DNA or effectors are

recruited by APTase analogues (consist of VirB, VirB11, and VirD4), then translocated between

extracellular space and the cytoplasm through the translocation channel (contains VirB6-VirB10)

and the pilus (composed of VirB2 and VirB5) (Chandran et al. 2009; Fronzes et al. 2009;

Wallden et al. 2010).

Page 73: Characterization of Dickeya dianthicola and Pectobacterium

58

The Type VI secretion system (T6SS) is involved in mediating antagonistic or synergistic

communications against adjacent eukaryotes or bacteria (Russell et al. 2014; Wallden et al.

2010). The T6SS apparatus is a tubular structure consisting of a sheath of TssB and TssC, which

surround a tube built from stacked hexameric rings of the haemolysin co-regulated protein (Hcp)

(Brunet et al. 2014; Bönemann et al. 2009). The apparatus assembles on a baseplate complex

(TssE, TssF, TssG, TssK) and is attached to the cell envelope by the membrane complex (TssJ,

TssM, TssL). The trimer VgrG spike protein sits at the tip of the Hcp tube, which is sometimes

capped with a pointed PAAR (Proline-Alanine-Alanine-aRginine) domain-containing protein

(Bönemann et al. 2009). VgrG or VgrG/PAAR spike facilitates the secretion of substrates (or

effectors) to adjacent bacterial or eukaryotic cells by penetrating of the membrane of the

neighboring cell. In addition to VgrG or VgrG/PAAR spike, substrates of T6SS can also directly

bind within the pore of an Hcp hexamer and are secreted together as a substrate-Hcp complex

(Silverman et al. 2013). All characterized bacterial-targeting T6SS proteins function as toxins,

either by killing or preventing the growth of target cells. Transcriptome and comparative

genomic analyses conducted by Bellieny-Rabelo et al. found the T6SS associated with a

toxin/immunity pair in Pectobacterium strains (Bellieny-Rabelo et al. 2019).

Whole genome sequence (WGS) is a powerful tool for describing bacterial species

evolution and pangenome (Medini et al. 2005). Genes in the pangenome that are shared by all

members of the taxa are the core-genome, while genes observed in a subset of strains are

considered as the dispensable genome and genes present just in a single strain form a unique

genome (Lapierre and Gogarten 2009). From the genome-related research, genomic diversity and

evolution of the taxa of interest can be described by analyzing their pangenome. Putative genetic

Page 74: Characterization of Dickeya dianthicola and Pectobacterium

59

differences can be screened out from dispensable or unique genomes to explain the differences in

virulence, morphology or other survival-related characteristics.

In this study, I reveal genetic variations of D. dianthicola and examine the variations

found in recent outbreak strains compared to each other and to previously collected strains in

order to describe biodiversity and emergence of new strains of D. dianthicola. Specific

objectives were to compare putative biological functions required for virulence or interbacterial

competition and to determine if D. dianthicola strains causing the current potato blackleg

outbreak in the United States differ from previously collected strains.

3.2 MATERIALS AND METHODS

3.2.1 Bacterial strains

A total of 32 D. dianthicola strains were analyzed, including 19 from USA and 13 from

elsewhere in the world (Table 3-1). Among the strains from USA, 16 were isolated from potato

(Ge et al. 2020b) and were whole-genome sequenced. Genome sequence data of three additional

strains from USA, one strain from Pakistan, two from Morocco, and 10 from Europe were

retrieved from the NCBI database (https://www.ncbi.nlm.nih.gov/genbank/).

Page 75: Characterization of Dickeya dianthicola and Pectobacterium

60

Table 3-1. Dickeya dianthicola strains used in this study.

Strain name Assembly number Source Country USA State Year Host

ME23 GCA_003403135.1 NCBI a USA Maine 2015 Potato

DE440 GCA_002906725.1 NCBI USA Delaware 2016 Potato

WV516 GCA_002906735.1 NCBI USA West Virginia 2016 Potato

CFBP2015 GCA_009873535.1 NCBI France NA b 1975 Potato

SS70 GCA_003121805.1 NCBI Pakistan NA 2017 Potato

CFBP1888 GCA_009873515.1 NCBI France NA 1978 Potato

RNS 11-47-1-1A GCA_009874325.1 NCBI France NA 2011 Potato

CFBP2982 GCA_009873565.1 NCBI France NA 1988 Potato

S4.16.03.P2.4 GCA_003595075.1 NCBI Morocco NA 2016 Potato

MIE34 GCA_009874285.1 NCBI Switzerland NA 2013 Potato

S4.16.03.LID GCA_003595005.1 NCBI Morocco NA 2016 Potato

GBBC 2039 GCA_000365365.1 NCBI Belgium NA NR c Potato

IPO980 GCA_000430955.1 NCBI Netherlands NA NR Potato

NCPPB 3534 GCA_000365405.2 NCBI Netherlands NA 1987 Potato

NCPPB 453 GCA_000365305.1 NCBI UK NA 1957 Carnation

RNS04.9 GCA_002706485.1 NCBI France NA 2004 Potato

PA24 CP069595 This study USA Pennsylvania 2015 Potato

ME30 CP060830 This study USA Maine 2015 Potato

16SBJ16 CP069596 This study USA Maine 2016 Potato

16PA07 CP060833 This study USA Pennsylvania 2016 Potato

16NJ12 CP060832 This study USA New Jersey 2016 Potato

16NJ11 CP060829 This study USA New Jersey 2016 Potato

16ME21T CP069598 This study USA Maine 2016 Potato

16ME22T CP069597 This study USA Maine 2016 Potato

16MB01 CP069599 This study USA New Jersey 2016 Potato

16MA15T CP069600 This study USA Massachusetts 2016 Potato

16LI04 CP069601 This study USA New York 2016 Potato

16LI02 CP069602 This study USA New York 2016 Potato

16LI01 CP069603 This study USA New York 2016 Potato

16JP05 CP069604 This study USA Virginia 2016 Potato

16JP03 CP069605 This study USA Virginia 2016 Potato

59W CP060831 This study USA Maine 2016 water

a NCBI: National Center for Biotechnology Information. b NA: not available. c NR: not recorded.

Page 76: Characterization of Dickeya dianthicola and Pectobacterium

61

3.2.2 Whole genome sequencing (WGS) and assembly

Bacterial DNA was extracted and prepared using the QIAamp DNA Mini Kit (QIAGEN,

Hilden, Germany) according to the manufacturer’s instructions. WGS was performed with index-

tagged libraries for each strain by paired-end sequencing at the CD Genomic (New York, USA)

on an Illumina HiSeq (Illumina Inc., San Diego, California, United States). The coverage was

greater than 100. Short reads were assembled using the Geneious Primer 2020.2.2 software

(Biomatters, Ltd., San Diego, CA, USA) by mapping to the reference genome of strain ME23

(CP031560). Draft genomes were then finished after closing gaps using the abundant pair

relationships of short reads through GapCloser (Luo et al. 2012). Raw reads used for gap-closing

were filtered by removing low quality, adapters and duplications with the use of FASTP (Chen et

al. 2018).

3.2.3 Genome annotation

Additional genome sequences of three USA strains and 13 non-USA reference strains

were downloaded from the NCBI genome database in a FASTA format. The reference genomes

along with the newly sequenced genomes were evaluated using the quality assessment tool

QUAST (Mikheenko et al. 2018). Gene features of the 32 genomes were then predicted using

Prokka v1.12 (Seemann 2014) using default settings. The newly assembled genome sequences

were deposited in the BioProject at NCBI with accession number PRJNA659753.

3.2.4 Pangenome analysis

Following the annotation using Prokka, pangenome analysis of all 32 strains were

performed using the Roary software package with its default settings (Page et al. 2015). Files

created by Roary were used for generating pan and core genome curves in R (Team 2013). Lists

Page 77: Characterization of Dickeya dianthicola and Pectobacterium

62

containing all the predicted gene clusters that generated from Roary were uploaded to eggnog-

Mapper2 to identify the cluster of orthologs groups (COG) on the basis of an all-against-all

sequence comparison of the proteins (Huerta-Cepas et al. 2019). Subtyping of the 32 strains was

performed using RxAML (Huerta-Cepas et al. 2019) to identify single-nucleotide

polymorphisms (SNP) in the core genomes and to reconstruct a parsimony phylogenomic tree.

Phylogenetic trees were visualized using the online tool iTOL (Letunic and Bork 2019).

3.2.5 Comparison among USA strains of D. dianthicola genomes

Dickeya dianthicola strains from USA were clustered into clades I, II and III. Whole

genome comparison of strains in the three clades were computed with the use of the BLAST

Ring Image Generator (BRIG) v0.95 (Alikhan et al. 2011) to illustrate the similarity among the

genomes. Complete genome sequence of strain ME23 was used as the reference genome for

analysis. Similarity was determined using BLASTn.

To predict the genetic determinants for pathogenesis or survival related differences,

genomes from each clade were evaluated on the evolutionary distance using the Mauve software

(Darling et al. 2004). A set of common genes and a set of unique genes among the three clades

were analyzed using files generated from Roary. Venn diagrams were created using VennPainter

(Lin et al. 2016). The comparative analysis of the secretion system, two-component regulatory

systems, flagellar and chemotaxis, and PCWDEs were conducted using GenBank files generated

by Prokka, and diagrams were drawn using Illustrator for Biological Sequences (IBS) (Liu et al.

2015).

Page 78: Characterization of Dickeya dianthicola and Pectobacterium

63

3.3 RESULTS

3.3.1 Genomic assemblies

Thirty-two genomes of D. dianthicola were used for comparative analysis, with 16 USA

strains being sequenced in this study (Table 3-1). Of the 16 genomes from USA strains, 12 were

assembled into one scaffold, and four were assembled into 6 to 12 scaffolds with no consecutive

gap longer than 10 bases (Table 3-2). The average size of the assembled genome was 4.88 Mb

with the range from 4,775,037 bp (16NJ12) to 4,909,063 bp (16MB01) (Table 3-2). In

measuring the quality of assembled genomes, N50 (minimum length of contigs in which 50% of

the genome are covered) ranged from 752,789 (16NJ12) to 4,909,063 (16MB01) (Table 3-2).

The calculated GC content fell within the range of 55.72 to 55.99 % (Table 3-2). An average of

4,279 predicted protein sequences (CDS) were annotated for each genome, ranging from 4,151

(16NJ12) to 4,388 (RNS-11-47-1-1A) (Table 3-2). For all 16 strains, the quantity of the

annotated rRNA were 22 and number of tRNA ranged from 72-81, which were nearly identical

to the two complete genomes ME23 (CP031560.1) and RNS04.9 (CP017638.1) (Table 3-2).

Intact sequences of rRNA indicated that the assembling pipeline was of good quality.

Except the two complete genomes ME23 and RNS04.9, the number of scaffolds present in the

remaining 14 reference genomes obtained from NCBI database was considerably higher (11-

108) than in the drafted genomes of the 16 USA strains newly obtained for this study (Table 3-

2). Another quality metric of reference assembly, the N50, was relatively lower for these strains.

Moreover, a significantly higher variation (55.60-56.32) in the %GC was observed among the

reference genomes than the newly sequenced ones (Table 3-2). To compare all 32 genomes, the

16 retrieved reference genomes were also annotated with same Prokka-based pipeline. The stated

quantities of tRNA (39-77) were often lower in the reference genomes, while the number of

Page 79: Characterization of Dickeya dianthicola and Pectobacterium

64

rRNA was 1 to 17 in contrast to 22 detected in herein reported newly sequenced genomes (Table

3-2).

Table 3-2. Genome statistics of Dickeya dianthicola strains.

Strain

No. of

scaffolds

Genome

size (bp) N50

%GC CDS rRNA

repeat

region tRNA tmRNA

ME23 1 4,909,058 4,909,058 55.72 4317 22 3 76 2

DE440 55 4,867,975 246,586 55.72 4321 8 3 68 2

WV516 103 4,906,266 214,389 55.60 4318 6 3 66 2

CFBP2015 55 4,785,831 107,754 55.74 4207 1 3 63 2

SS70 62 4,797,592 177,218 55.94 4217 5 3 61 2

CFBP1888 67 4,844,898 161,968 55.73 4276 1 5 64 2

RNS-11-47-1-1A 78 4,948,067 145,496 55.77 4388 1 4 61 2

CFBP2982 90 4,704,394 94,450 55.86 4171 1 1 62 2

S4.16.03.P2.4 101 4,865,147 92,028 55.67 4306 1 3 39 2

MIE34 94 4,677,677 89,071 55.90 4118 1 2 64 2

S4.16.03.LID 108 4,863,939 71,707 55.68 4300 1 3 39 2

GBBC2039 32 4,798,242 4,649,420 56.32 4235 4 6 59 2

IPO980 27 4,840,931 4,662,278 55.69 4227 5 5 66 2

NCPPB3534 100 4,867,392 4,775,960 55.71 4248 17 4 76 2

NCPPB453 11 4,676,060 462,956 55.99 4101 4 9 64 2

RNS04.9 1 4,720,132 4,720,132 55.96 4110 22 4 77 2

PA24 1 4,909,043 4,909,043 55.72 4318 22 3 76 2

ME30 1 4,909,051 4,909,051 55.73 4306 22 3 76 2

16SBJ16 1 4,909,059 4,909,059 55.72 4318 22 3 76 2

16PA07 9 4,800,845 1,060,962 55.99 4175 22 3 79 2

16NJ12 12 4,775,037 752,789 55.97 4151 22 4 81 3

16NJ11 7 4,813,906 1,130,583 55.97 4173 22 4 80 2

16ME21T 1 4,909,059 4,909,059 55.72 4318 22 3 76 2

16ME22T 1 4,909,042 4,909,042 55.72 4318 22 3 76 2

16MB01 1 4,909,063 4,909,063 55.72 4317 22 3 76 2

16MA15T 1 4,909,059 4,909,059 55.72 4317 22 3 76 2

16LI04 1 4,909,059 4,909,059 55.72 4318 22 3 76 2

16LI02 1 4,909,059 4,909,059 55.72 4318 22 3 76 2

16LI01 1 4,909,059 4,909,059 55.72 4317 22 3 76 2

16JP05 1 4,909,059 4,909,059 55.72 4318 22 3 76 2

16JP03 1 4,909,059 4,909,059 55.72 4318 22 3 76 2

59W 6 4,860,981 1,340,508 55.96 4209 22 3 72 2

Page 80: Characterization of Dickeya dianthicola and Pectobacterium

65

3.3.2 Pangenome

Gene accumulation curves showed that the pangenome plateaued at 6,693 genes, and core

genome kept at 3,205 genes, reflecting a stable core and an almost closed pangenome (Figure 3-

1A). The pangenome of the 32 D. dianthicola strains comprised 6693 CDSs, including 3205 core

genes (genes present in more than 31 strains), 172 soft core genes (genes present in more than 30

and less than 31 strains), 1288 shell genes (genes present in more than 4 and less than 30 strains),

and 2028 cloud genes (genes present in less than 4 strains) (Figure 3-1B).

By combining core and soft-core genes into one core category, genes in clades I, II and

III were sorted into COG; the outcomes of the attribution are depicted in Figure 3-1C. When

sorting the annotated genes into COG, 22.8% were categorized as unknown function which was

excluded from the graphical representation of the COG categories. The most abundant categories

in the core genome were genes involved in storage housekeeping functions, such as transcription

and translation (8.6%, 5.5%), and different metabolism categories (inorganic ion transport and

metabolism 10.5%, amino acid metabolism and transport 9.8%, energy production and

conversion 7.8%). The shell genomes had a larger portion of genes associated with mobile

genetic elements (MGE) such as transposons and bacteriophages (transcription (11.0%),

replication, recombination and repair (16.3%)), which was also the most enriched category

amongst the cloud genomes (11.6%, 19.4% respectively). Genes involved in functions other than

the above two were considerably more prevalent in the core gene pool, compared to the shell and

cloud genes.

Page 81: Characterization of Dickeya dianthicola and Pectobacterium

66

Figure 3-1. Pangenome statistics of 32 Dickeya dianthicola strains. (A) Pangenome curve

generated by plotting the number of genomes (X axis) and number of conserved genes (dotted

line) or total genes (solid line) (Y axis). (B) Size and distribution of genes within subgroups

including core (in blue color) and soft-core genes (in orange) shared by more than 30 strains,

shell genes (in grey) shared by 5 to 30 strains and cloud genes (in yellow) found in less than 5

strains. (C) Percentage of each cluster of orthologs group that distributed in core (in blue), shell

(in grey) and cloud genes (in yellow), Cluster of orthologs groups include a: metabolism, b:

storage and processing, and c: cellular processes and signaling.

Page 82: Characterization of Dickeya dianthicola and Pectobacterium

67

3.3.3 SNPs-based phylogeny

The 32 strains were classified into eight clades (I-VIII) using phylogenetic reconstruction

based on SNPs (Figure 3-2). Lineages of D. dianthicola contained differential levels of

observable diversity. Particularly, there was a distinct distribution of USA strains into three

different clades (I, II, III), which was consistent in matching the three genotypes described by Ge

et al. (2020b). Clade I consisted of the majority of strains from USA (n = 15), while the

remaining four of the USA strains were clustered into clade II and III. Strains from European

countries were more diversified into five clades (III, IV, V, VI and VIII). One strain from Asia

(Pakistan) was solely clustered in clade VII. One strain from the USA was clustered with two

European strains, and two strains from Africa (Morocco) were also clustered with two different

European strains. Two strains out of the 32 were isolated from non-potato sources, with one in

clade II and one in clade III.

Page 83: Characterization of Dickeya dianthicola and Pectobacterium

68

Figure 3-2. Geographic and phylogenetic relationship of Dickeya dianthicola strains used in this

study. (A) Pie chart showing the geographic origin of Dickeya dianthicola strains included in this

study. Strains are colored based on country of origin, USA in orange, Belgium in aquamarine,

France in purple, Morocco in neon green, the Netherlands in yellow, Pakistan in vivid blue,

Switzerland in Emerald green, and the UK in purple in Irish. (B) Phylogenetic tree of 32 strains

established based on single nucleotide polymorphisms (SNPs) of the core genome, with color

codes indicating the country of origin described in (A). Fifteen out of 19 strains from the USA

were clustered within clade I. Three strains from the USA were clustered within clade II. One

strain from USA was clustered with two other European strain into clade III. Two strains out of

the 32 were isolated from non-potato sources, with one in clade II and one in clade III.

Page 84: Characterization of Dickeya dianthicola and Pectobacterium

69

3.3.4 Analyses of USA strains

To identify the region of interest in clades I, II and III containing USA strains of D.

dianthicola, one strain from each clade was selected for comparative analysis. Strains with high

integrity genomes (ME23, 16PA01, RNS04.9), each with 3760 conserved genes were selected as

representative. Clade I by strain ME23 shared 116 genes with clade II strain 16PA07 genome

and shared 50 genes with clade III strain RNS04.9 genome (Figure 3-3). Furthermore, 391

unique genes were present in the genome of ME23.

Similarities among the genomes was illustrated by comparing genomic nucleotide

sequences using BLAST. Genomic nucleotide sequences of strains were compared using

BLASTn to reference strain ME23. Graphical output of BRIG analysis displayed an overview of

identity and the major features related with virulence. The features included secretion systems,

two-component regulatory systems, PCWDE and flagella motility and chemotaxis (Figure 3-4).

Genomes of D. dianthicola in clade I possessed nearly identical genomic structures to that of

ME23, while multiple regions of substantial variation were observed in genomes of strains in

clade II and III (Figure 3-4). By combining the results from COG clusters, shell and cloud genes

were associated with bacteriophage or transposon-elements.

Page 85: Characterization of Dickeya dianthicola and Pectobacterium

70

Figure 3-3. Whole genome alignment and comparison of orthologous genes from clades I, II and III of Dickeya dianthicola from

USA. (A) Mauve progressive alignment of RNS04.9 (clade III) genome, 16PA07 (clade II) genome, and ME23 (clade I) genome. (B)

Venn diagram showing the number of clusters of orthologous genes shared and unique among clades I, II and III.

Page 86: Characterization of Dickeya dianthicola and Pectobacterium

71

Figure 3-4. BLAST Ring Image Generator (BRIG) analysis of genomes of Dickeya dianthicola

strains clustered in clades I, II and III against reference strain ME23. Reference features related

with virulence were labeled in the outermost layer, including secretion system in red, with types

I, II, III, IV and VI, two-component regulatory system in light blue, plant cell wall degrade

enzymes (PCWDE) and flagellar and chemotaxis related coding sequences (CDS) in black.

Page 87: Characterization of Dickeya dianthicola and Pectobacterium

72

Screening for PCWDE showed that all strains shared a set of extracellular enzymes with

same number of copies, including pectate lyase, pectate disaccharidelyase, pectinesterase,

polysaccharide lyase, rhamnogalacturonate lyase, and cellulase family glycosylhydrolase. In

addition, D. dianthicola strains in clades I, II, and III all contained conserved T1SS, T2SS and

T3SS gene clusters, which are used by many Gram-negative pathogenic bacteria in transporting

effectors or some other proteins. Flagellar- and chemotaxis-related genes were also conserved.

The T4SS and T6SS gene clusters differed among the three clades of USA strains. Two

clusters of T4SS system, containing 18 core components were found in clade I, while only one

cluster of T4SS system was found in clade II, consisting of 8 core components, and one cluster in

clade III, consisting of 9 core components (Figure 3-5). The T4SS DNA binding domain

containing protein was encoded downstream of the T4SS system cluster 1 in all clades. Twelve

core T6SS components were highly conserved in all three clades (Figure 3-6). Four VgrG, three

PAAR and four hcp genes were found in clade I and clade II, while only three VgrG, two PAAR

and three hcp genes were discovered in clade III (Figure3-6). Additionally, there was one

additional VgrG/PAAR spike genes in clade I and clade II with Type II Toxin-antitoxin system

(CcdA/CcdB) upstream of this cluster (Figure 3-6).

Page 88: Characterization of Dickeya dianthicola and Pectobacterium

73

Figure 3-5. Schematic map of type IV secretion system in clades I, II and III of Dickeya dianthicola from USA. Arrows indicate

putative transcriptional units with directions. Long distant intervals were disconnected by cut lines. T4SS components depicted in blue

color and hypothetical proteins depicted in grey color. T4SS components include two clusters (cluster 1: virB1/ virB2/ virB3/ virB5/

virB6/ virB8/ virB9/ virB11; cluster 2: virB1/ virB2/ virB3/ virB5/ virB6/ virB8/ virB9/ virB10/ virB11). Both cluster 1 and cluster 2

were found in clade I, while clade II had cluster 1 only, and clade III had cluster2 only. T4SS DNA binding domain containing protein

(coded by DDILJCIA_01468, AGKNFJGG_01395, EAMBJKPE_01341) were found in all three clades.

Page 89: Characterization of Dickeya dianthicola and Pectobacterium

74

Figure 3-6. Schematic map of type VI secretion system in clades I, II and III of Dickeya dianthicola, including T6SS components

(coded by tssA/ tssB/ tssC/ tssE/ tssF/ tssG/ tssJ/ tssK/ tssL /tssH and tssM) in light blue color, PAAR in green, VgrG in vivid blue, hcp

in red, T6SS-related rshB loci in red octagon, and hypothetical proteins coding genes in grey color. Arrows indicate directions of

transcription.

Page 90: Characterization of Dickeya dianthicola and Pectobacterium

75

3.4 DISCUSSION

The newly sequenced genomes in this study were assembled by mapping to the reference

genome ME23 with Geneious Primer 2020.2.2, which has an intuitive interface equipped with

essential genomic tools for data analysis. De novo assembling through SOAPdenovo2 (Luo et al.

2012) was utilized to compare with the Geneious Primer. After closing the gaps generated during

the GapCloser assembly, longer and more complete sequences were obtained from Geneious

Primer plus GapCloser. Therefore, I took the advantage of the Geneious Primer for assembling

and GapCloser for gap closing, then switched to other tools for further analysis. In view of a high

demand of genomic sequences and the significant threat of D. dianthicola to potato production in

USA (Charkowski 2018; Hao et al. 2016; Johnson 2016), whole genomes of D. dianthicola

strains isolated during the blackleg outbreak and from representative regions in the USA were

sequenced. Through the above workflow, the 16 D. dianthicola genomes that were re-sequenced

and assembled in this study, 12 have been nearly closed to a full chromosome, and the remaining

4 contained only 6 to 12 scaffolds. Furthermore, they contained a complete number of rRNA.

Shortly after the start of the recent D. dianthicola outbreak in the United States, multiple

D. dianthicola genotypes were identified (Ge et al. 2020b). The extent of genetic variability in

this important pathogen needed to be clarified to understand disease epidemiology and pathogen

evolution and to aid in disease management. Therefore, pangenomic analyses of 32 D.

dianthicola strains were completed to obtain a comprehensive representation of the D.

dianthicola core- (3377 proteins) and pangenomes (6693 proteins). The number and the

representativeness of the genomes affect the status of pangenomes of a certain species (Rouli et

al. 2015). In a closed pangenome, new coding capabilities will not be discovered with the

continuous adding of new genomes (Bosi et al. 2015). Importantly, the number of sequences in

Page 91: Characterization of Dickeya dianthicola and Pectobacterium

76

this study was representative for the geographic origins and are sufficient for a nearly closed

pangenome of D. dianthicola. Attribution of the genes from the studied D. dianthicola

pangenome to functional COG categories showed overrepresentation of transcription-associated

groups and MGE in the cloud and shell pangenome parts. In this context, MGE promote bacterial

evolution through horizontal gene transfer, which frequently function through bacterial secretion

systems (Gyles & Boerlin 2014; Rodríguez-Beltrán et al. 2020).

PBSR can be caused by a number of bacterial species, but each particular outbreak of this

disease is typically caused by one predominant genotype probably due in part to contamination

and spread of that genotype in one or a few widely used seed potato sources. In the United

States, seed potatoes from a small number of sources in the northeastern quarter of the United

States appeared to be important during the early stages of the outbreak. In Europe, D. dianthicola

was isolated from potato in 1970s (Parkinson et al. 2009; Toth et al. 2011), and became a

predominant species, until it was supplanted by D. solani (Blin et al. 2020; Parkinson et al. 2009;

Patel et al. 2019; Sarfraz et al. 2018; Toth et al. 2011). The predominant species must have

highly competitive traits in fitness and pathogenicity. Dickeya solani has some advantages over

D. dianthicola. Dickeya solani is more efficient and effective in infecting and rotting potato

tubers, giving it a higher level of fitness and quick proliferation in infected tissues, primarily due

to a higher gene expression of pectate lyases and balanced frequencies of two alleles in the

VfmB protein involved in quorum sensing (Blin et al. 2020), as well as a quick colonization

(Toth et al. 2011).

Dickeya solani emerged as a potato pathogen recently and was quickly spread from seed

potato fields in the Netherlands throughout Europe, the Middle East, Africa, and China through

the seed trade (Chen et al. 2015; Toth et al. 2011). Currently, D. solani isolates in potato are

Page 92: Characterization of Dickeya dianthicola and Pectobacterium

77

highly homogeneous, reflecting its recent emergence (Motyka-Pomagruk et al. 2020). Dickeya

dianthicola strains from potato show a stark contrast to D. solani. By comparing genomes of all

available D. dianthicola strains, it was found that it appears to have a clonal distribution across at

least three continents, reflecting ecological evolution and possible interference by trade of

agricultural products (Carvajal-Yepes et al. 2019; Shahbandeh 2020).

An exceptionally high level of distinction was found among the eight clades of D.

dianthicola genomes through a SNP analysis of the core genomes. The predominant USA strains

were distinct from the European strains (Ge et al. 2020b; Parkinson et al. 2015). Clade III was

found in other countries, but clade I and II were unique to USA and clade I had extraordinarily

high homogeneity within the collected bacterial isolates. Interestingly, three genotypes identified

from our previous work consistently fall into three clades (I, II, and III) obtained in this study,

namely the former genotypes I, II, and III corresponded to clades I, II, and III, respectively.

Clade I was a predominant population in the US outbreak as previously reported (Ge et al.

2020b; Parkinson et al. 2015). The characteristic genome of clade I possibly enables D.

dianthicola to be more efficient in adapting to the ecological niches, such as capabilities for

temperature adaption, competition or survival.

The homogeneity and predominance of the clade I population indicated the outbreak of

PBSR in the United States was primarily caused by one strain. The primary pathogenicity genes

in D. dianthicola, such as the PCWDE and quorum-sensing genes, are conserved. Some genes

that may contribute to virulence differ among clades. For example, Clade I encode two copies of

the T4SS, whereas clades II and III only encode one copy and the clades differed in the secreted

proteins associated with the T6SS. Either or both of the T4SS and T6SS differences may affect

strain fitness. It is possible from pangenomic analysis that clade I strains have pathogenic

Page 93: Characterization of Dickeya dianthicola and Pectobacterium

78

advantages over other clades. However, these differences do not fully explain clade I strains as

being predominant over others, so further experimental studies should be conducted. One

supporting evidence from previous studies (Ge et al. 2020b) is that clade I (or genotype I)

consistently showed up in Maine every year in the five years of data collection, while clade II

only showed up in 2015 as a small subset in Maine and disappeared since then.

Clade I and II strains possessed an additional VgrG/PAAR spike genes than clade III

strain, with Type II Toxin-antitoxin system (CcdA/CcdB) upstream of the VgrG/PAAR genes,

although whether the adjacent toxin can be secreted by T6SS and act on neighboring cells

remains unknown. According to other studies, rhsB gene (coding for deoxyribonuclease RhsB) is

delivered into target cells in a VgrG-dependent process in D. dadantii, and also closely

associated with VgrG spike in D. dianthicola (Bellieny-Rabelo et al. 2019; Koskiniemi et al.

2013), but their functions need to be further confirmed. Nonetheless, the differences of T4SS and

T6SS might partially explain why clade I population was dominant in the PBSR outbreak and

these gene clusters are important targets for future fitness and pathogenicity studies (Ge et al.

2020b).

Currently, there is no enough evidence to explain how the clade I strain emerged and

caused the outbreak of PBSR. Since it had additional copies of secretion system or pathogenicity

related factors, it might be a variant from clade II or other clades by horizontal gene transfer

(Arber 2014; De la Cruz & Davies 2000). Although clade I of D. dianthicola has been

predominant in the outbreak, it does not mean we should ignore the other clades. Since other

clades are present in the environment, additional outbreaks with these strains could occur and

therefore screening protocols need to be able to detect all clades. Dickeya dianthicola has a wide

host range and the clades not currently detected in the United States could enter through trade in

Page 94: Characterization of Dickeya dianthicola and Pectobacterium

79

ornamental plants or seed and could have a large impact on agricultural production (Toth et al.

2011).

In summary, genomic constructions and compositions determine the distribution and

dominance of a unique strain of Dickeya dianthicola that was the cause of the recent PBSR

outbreak in the northeastern USA in recent years. I infer that the presence of additional copies of

virulence-related genes allowed clade I (or Type I as previously described by Ge et al. 2020b)

population to be predominant over the other strains. This can be a foundation in understanding

the epidemic of PBSR in potato production.

Page 95: Characterization of Dickeya dianthicola and Pectobacterium

80

CHAPTER 4

PATHOGEN COMPLEX AND MICROBIAL ASSOCIATION OF POTATO

BLACKLEG AND SOFT ROT IN NORTHEASTERN AMERICA

ABSTRACT

Potato blackleg and soft rot (PBSR) which can be caused by multiple bacterial species

has become a serious problem in the Northeastern US since 2015. The aim of this study was to

examine the taxa of causal agents and their microbial associations. Potato plants showing

blackleg symptoms were collected from northeastern states. Bacteria were isolated from plant

samples. Polymerase chain reaction (PCR) was conducted with primers specific to Dickeya or

Pectobacterium spp. Illumina sequencing was conducted on the V3-V4 region of 16S rDNA

from the samples, which were analyzed for bacterial communities based on operational

taxonomic units. Samples were divided into four groups: D: samples containing Dickeya only; P:

Pectobacterium only; DP: both Dickeya and Pectobacterium; and N: neither of the two genera.

Pectobacteriaceae was the predominate family in sample groups D, P, and DP. Genera Dickeya

and Pectobacterium prevailed in the microbial community when they each existed alone, while

Dickeya surpassed Pectobacterium when they coexisted. Pseudomonas, Acinetobacter, and

Providencia were always associated with potato samples as endophytes. Among the pathogens,

D. dianthicola was the only species present within the Dickeya genus, while species varied

within Pectobacterium genus, with P. parmentieri, P. polaris, P. carotovora subsp. carotovora,

P. carotovora subsp. odoriferum being the most prevalent presumptive species. Isolates of the

four presumptive species and P. carotovora subsp. brasiliensis were identified by sequencing the

gapA and confirmed to be pathogenic on potatoes. Thus, PBSR was caused by intergeneric or

Page 96: Characterization of Dickeya dianthicola and Pectobacterium

81

intrageneric species of Dickeya and Pectobacterium that contribute collectively to the disease

complex.

4.1 INTRODUCTION

Potato (Solanum tuberosum) is the third most important food crop in the US, and its total

production is estimated at 25.2 million US tons per year (U.S. Department of Agriculture 2019).

Production has been threatened by blackleg and soft rot (BSR) of potato in northeastern U.S.

since the 2015 outbreak, resulting in millions of dollars lost to the potato industry (Charkowski

2018). Although multiple bacterial species cause PBSR, Dickeya dianthicola has been identified

as the most prevalent species causing PBSR in the region (Hao et al. 2016; Jiang et al. 2016;

Johnson 2016; Ma et al. 2018; Patel et al. 2019). It is believed the bacterium may contaminate

and stay in/on dormant tubers during storage periods and cause field symptoms when these

tubers are planted.

Besides D. dianthicola, other species in genera Dickeya, such as D. chrysanthemi, D.

solani, and D. dadantii have been found potato samples and confirmed to be pathogenic to potato

(McNally et al. 2018; Ngadze et al. 2010; van der Wolf et al. 2014). Dickeya zeae and D.

aquatica were isolated from an irrigation pond adjunct to a potato field, but never found in

potato samples (Parkinson et al. 2014). Species in the genus of Pectobacterium are soil

inhabitants, which are well adapted to survive as saprophytes in the soil until a host plant is

reachable (Aizawa 2014). Owing to their existence in soil and wide host range, Pectobacterium

spp. cause soft rot diseases on potato and a wide range of vegetable crops, including carrot,

cabbage, maize, and sugar beet (Ma et al. 2007). Pectobacterium carotovora, in the family of

Pectobacteriaceae, is a diverse species consisting of P. carotovora subsp. carotovora, P.

carotovora subsp. odoriferum and P. carotovora subsp. brasiliensis, which are all causal agents

Page 97: Characterization of Dickeya dianthicola and Pectobacterium

82

of PBSR (Duarte et al. 2004; Gallois et al. 1992; Gardan et al. 2003; Zhang et al. 2016). In

addition, P. parmentieri (formerly P. wasabiae) from potato, which was once classified within P.

wasabiae from Japanese horseradish (Goto and Matsumoto 1987; Khayi et al. 2016), has been

found in potato plants and other crops all over the world (Baghaee-Ravari et al. 2011; De Boer

et al. 2012; Ma et al. 2007; Ngadze et al. 2012; Pitman et al. 2008; Rahmanifar et al. 2012).

Pectobacterium atrosepticum (formerly P. carotovora subsp. atrosepticum) causes blackleg in

the field and soft rot of tubers during storage (Pérombelon and van der Wolf 2002), which is very

host specific to potato. These bacteria, however, may live in the rhizosphere of cultivated plants

and weeds without causing disease. One newly named bacterium, P. polaris, was always

associated with PBSR, but requires a wound on plants or high concentration of pathogen to cause

disease (Dees et al. 2017). The distribution of these species within the genus Pectobacterium

vary, depending on geographical location of samples.

With such diversified species and their distribution, regardless of whether they are soil

inhabitants or invaders, potato plants may be exposed to multiple species of bacteria that are

potentially causal agents of PBSR. How these bacteria interact in plant infection needs further

investigation. In this study, the interested subject was the microbes associated with PBSR that

occurred in the US outbreak. If multiple species of Dickeya and Pectobacterium can cause

PBSR, was PBSR outbreak in the US caused by a single species or multiple species? Therefore,

the objectives of this study were to analyze the structure of bacterial communities in potato

disease samples in circumstances with or without the presence of PBSR pathogens and

characterize the associated pathogenic bacterial species.

Page 98: Characterization of Dickeya dianthicola and Pectobacterium

83

4.2 MATERIALS AND METHODS

4.2.1. Disease samples

Potato stems and tubers showing PBSR symptoms were collected from fields in most of

the Northeastern US states from 2015 to 2020, where PBSR outbreak occurred. DNA was

extracted using FastDNA SPIN Kit (MP Biomedical, Santa Ana, CA, USA). Disease was

confirmed by PCR with primer pairs pelADE targeting on Dickeya spp. and PEC targeting on

both Dickeya and Pectobacterium spp. (Ge et al. 2020b). Potato samples collected in 2015 to

2017 have been described in previous studies and were used in this study for microbial analysis

and pathogen isolation. Seventy-six samples, including 73 with PBSR symptoms and 3 without

PBSR symptoms, were randomly subsampled for microbial analysis. In the 73 symptomatic

samples, 61 tested positive for Dickeya spp. and possibly positive for Pectobacterium spp. and

12 only Pectobacterium spp. positive. The three non-BSR-symptomatic samples were Dickeya

and Pectobacterium negative (Table 4-1).

Table 4-1. Bacterial detection in potato samples using polymerase chain reaction (PCR) with

primers pelADE targeting on Dickeya spp. and PEC targeting on both Dickeya and

Pectobacterium spp.

State State

code

Total

samples

Dickeya

+ z

Dickeya

-

Pectobacterium

+

Pectobacterium

-

Maine ME 16 12 2 14 2

Massachusetts MA 11 11 0 11 0

Pennsylvania PA 10 8 1 9 1

Rhode Island RI 9 7 2 9 0

New Jersey NJ 10 9 1 10 0

New York NY 5 4 1 5 0

Virginia VA 5 4 1 5 0

Michigan MI 4 1 3 4 0

Delaware DE 5 5 0 5 0

Florida FL 1 0 1 1 0

Total 76 61 12 73 3 z “+” indicate a positive PCR reaction and “-” indicates a negative reaction.

Page 99: Characterization of Dickeya dianthicola and Pectobacterium

84

4.2.2 Metagenomic sequencing

Concentration of DNA from the 76 samples was adjusted to 10 ng/l and submitted to

Michigan State University Genomic Core Facility (East Lansing, MI, USA) for metagenomic

sequencing in 2018 spring. Illumina compatible amplicon libraries of the V3-V4 hypervariable

region of the 16S rRNA gene were amplified using barcoded Illumina compatible primer pair

b341/806r (Caporaso et al. 2012; Juck et al. 2000). This generated a PCR amplicon

encompassing both the V3 and V4 regions. Sequencing was performed in a 2x250 bp paired end

format using a 500 cycle v2 reagent cartridge on Illumina MiSeq platform (Illumina Inc., San

Diego, CA, USA). The raw reads are deposited in the NCBI Sequence Read Archive (SRA) and

can be found with BioProject accession no PRJNA664821.

For community analysis, 1% relative abundance (RA) was used as a cut-off point. Any

species with RA > 1% were counted, otherwise was considered as negative. By combining PCR

assay and RA values, the samples were classified into four groups, which included group N: no

Dickeya nor Pectobacterium spp.; group P: containing Pectobacterium spp.; group D: containing

Dickeya spp.; and group DP: containing both Dickeya and Pectobacterium spp.

4.2.3 Bacterial isolation and identification

Sequencing analysis with QIIME. The standard operating pipeline QIIME was used to

process and analyze Illumina sequencing data (Bolyen et al. 2019). Generated sequences were

modified by trimming off primers and adapters. The quality of the sequences in FASTQ format

was analyzed after pooling all forward or reverse sequences. Overlapping paired-end reads were

denoised and stitched together. Operational taxonomic unit (OTU) was used to classify bacteria.

Alpha diversity, referring to the diversity within a sample, was expressed by the species richness

and species diversity. Observed OTUs in certain sampling depth was analyzed to express the

Page 100: Characterization of Dickeya dianthicola and Pectobacterium

85

species richness. Shannon diversity index was also calculated, which measures both species

richness and inequality between species abundances. Significance of observed OTUs and

Shannon diversity index among the four groups were analyzed using Kruskal-Wallis H test at

= 0.05. The 341f/805r region of 16S rRNA gene in Silva 132 QIIME-compatible database was

used for reference. Abundance in levels kingdom, phylum, class, order, family and genus was

calculated. RA was used to describe the percent composition of each OTU.

Sequencing data processing with MOTHUR. The MOTHUR software pack (version

1.39.5) was used only for the analysis at species level by analyzing OTUs (Schloss et al. 2009).

Sequences were processed according to the MiSeq SOP (https://mothur.org/wiki/miseq_sop/),

including reducing sequences and PCR errors, finding unique sequences, aligning sequences to

the reference alignment (reference database: silva.bacteria.fasta), and assessing error rates.

FASTA format of sequences of each OTU were printed out, and the output was BLASTed to 16S

rRNA database which was downloaded from NCBI using Python (Rossum 1995). Parameters

were applied to filter out five hits with the top five scores. The hit with the top score was

designated as the final species in excel. Only sequences of Dickeya spp. or Pectobacterium spp.

were selected for analysis with RA.

4.2.4 Bacterial isolation and identification

Potato stems showing PBSR symptoms were cut from the base and dipped into sterile

water for 5 min to let bacteria discharge. A portion of tissue was cut from symptomatic tubers

and ground in a 50 mL centrifuge tube containing sterile water. Bacteria were isolated by

streaking the sap onto crystal violate pectate (CVP) plates (Hélias et al. 2012), and incubated at

28 C for 24 hours in darkness. Purified colonies formed in cavities on the plates were

Page 101: Characterization of Dickeya dianthicola and Pectobacterium

86

transferred to tryptic soy broth (TSB) and incubated at 28 C overnight. Bacterial cells in TSB

were mixed with glycerol and stored at -80C.

Isolated bacteria were revived from -80C freezer by streaking on tryptic soy agar (TSA)

plates. Pure-cultured colonies were transferred to TSB and incubated at 28C overnight. One

milliliter of cell suspension was used for DNA extraction with QIAamp DNA Mini Kit

(QIAGEN, Germantown, MD, USA). DNA concentration was measured with NanoDrop 2000

(Thermo-Fisher, Waltham, MA USA) and adjusted to 10 ng/L. Metabolism gene gapA (coding

for protein glyceraldehyde-3-phosphate dehydrogenase A) was amplified by PCR using primer

pair gapA-7F/gapA-938R (Cigna et al. 2017). The thermal cycler was programmed for an initial

denaturation of 5 min at 94°C, followed by 40 cycles of 30 sec at 94°C, 30 sec at 54°C and 30

sec at 72°C, with a final elongation of 10 min at 72°C. Amplicons from PCR were cleaned using

UltraClean 15 DNA purification kit (MO BIO, Carlsbad, CA, USA), and then sequenced. The

sequences are deposited in NCBI Genbank; study accession numbers are MW027689 –

MW027797. Phylogenetic analysis was performed using MEGA7 software (Kumar et al. 2016).

The maximum likelihood phylogenetic tree was constructed with 500 bootstrap replicates.

4.2.5 Pathogenicity assay

The pathogenicity of Pectobacterium isolates was conducted using potato ‘Lamoka’ with

two inoculation methods. In the first method, washed potato tubers were stabbed to a depth of 1

cm with a pipette tip. Ten microliters of bacterial suspension at ~109 CFU/ml in TSB was

transferred into the hole which was sealed with Parafilm. Inoculated tubers were incubated at

room temperature (23ºC ± 2) for 5 days. Then the tubers were cut through the inoculation point.

Width and length of the lesion were measured and lesion size (width length) was calculated.

Ten P. parmentieri isolates, 11 P. carotovora subsp. brasiliensis isolates, 41 P. carotovora

Page 102: Characterization of Dickeya dianthicola and Pectobacterium

87

subsp. odoriferum isolates and 33 P. Polaris isolates were selected for pathogenicity test on

tubers. Each isolate was randomly assigned to 4 tubers as replicates. The second method

consisted of stem injection. Four isolates of each species were randomly selected for

pathogenicity test on stems. Potato tubers were planted in potting mix in a plastic pot and

maintained at 25°C (day) and 15°C (night) in a greenhouse. Six-week-old potato stems were

injected with 100 l of 108 CFU/ml bacterial suspension at 2 cm above the soil line using a

syringe. Four stems were used for each isolate. The wound was sealed with Parafilm. Symptoms

were visually inspected. Both trials were arranged using completely randomized design with four

replicates. Three weeks after inoculation, stems were cut along the stem for disease evaluation.

For both tuber and stem inoculation assay, TSB was used as a negative control. This experiment

was conducted twice. Means of lesion sizes were compared by Tukey HSD multiple range test at

= 0.05 using the SAS software (SAS Institute Inc. 2015, Cary, NC, USA).

4.2.6 Bacterial growth on CVP medium

To compare the growth of Pectobacterium and Dickeya spp., 9 standard strains were

selected for this experiment (Table 4-2). Bacterial cultures of each strain from -80C freezer

were revived as described. After incubation in TSB overnight, bacterial concentration was

approximately 109 CFU/ml. Cell suspension was prepared with three concentrations including

109, 106 and 102 CFU/ml. Ten microliters of cell suspension of each concentration were

transferred onto CVP plates and incubated at 28C. Each isolate was replicated three times on

two plates with six replicates in total. Diameter of cavity was measured at 12, 24, 36 and 48 h

after incubation. The experiment was conducted twice.

Page 103: Characterization of Dickeya dianthicola and Pectobacterium

88

Table 4-2. Dickeya and Pectobacterium spp. strains used for pathogen-bacteria interactions.

Pathogen Isolate Source

D. dianthicola ME30 University of Maine

P. parmentieri ME175 University of Maine

D. zeae 1686 A. Charkowski, Colorado State University

D. dadantii 3937 A. Charkowski, Colorado State University

D. solani IPO2222 A. Charkowski, Colorado State University

P. carotovora subsp.

brasiliensis

16MA04 University of Maine

P. carotovora subsp.

odoriferum

16MA30 University of Maine

P. polaris 16MA27 University of Maine

P. atroseptica ATCC33260 A. Charkowski, Colorado State University

4.3 RESULTS

4.3.1 Bacterial communities associated with PBSR

To understand the microbial community within PBSR, DNA of 76 samples were

analyzed for microbiome based on OTUs at 99% sequence similarity. After removing low-

quality and non-specific reads, a total of 10,362,620 reads were obtained from the 76 potato

samples from ten states, ranging from 57,643 to 207,222 per sample, with the median of

134,792. The number of reads was normalized to the smallest number of reads for alpha diversity

comparisons. Shannon diversity and observed OTUs (Figure 4-1) demonstrated that sample

group D had significantly (P < 0.05) lower bacterial diversity and evenness than groups N, P and

DP. Bacterial communities within sample groups DP, D and P were dominated by the family

Pectobacteriaceae, with RA of 45.9%, 58.4% and 45.8% accordingly (Figure 4-2A). At genus

level, Dickeya (RA = 52.9%) and Pectobacterium (RA = 36.0%) respectively dominated in

groups D and P (Figure 4-2B). Comparing to groups D and P, RA of Dickeya (22.0%) was much

higher than Pectobacterium (8.9%) in group DP.

Page 104: Characterization of Dickeya dianthicola and Pectobacterium

89

Figure 4-1. Number of operational taxonomic unit (OTU) and Shannon evenness index of

bacterial community in potato tissues. Significance among groups was analyzed using Kruskal-

Wallis H test ( = 0.05). N: no Dickeya nor Pectobacterium spp.; P: containing Pectobacterium

spp.; D: containing Dickeya spp.; DP: containing both Dickeya and Pectobacterium spp.

Page 105: Characterization of Dickeya dianthicola and Pectobacterium

90

Figure 4-2. Relative abundances of bacteria in potato samples at family level (A) and genus

level (B). Taxa with relative abundance < 0.001 were excluded). Samples were grouped

according to polymerase chain reaction (PCR). N: no Dickeya nor Pectobacterium spp.; P:

containing Pectobacterium spp.; D: containing Dickeya spp.; DP: containing both Dickeya and

Pectobacterium spp.

Page 106: Characterization of Dickeya dianthicola and Pectobacterium

91

Genera including Acinetobacter, Pseudomonas, Providencia, Stenotrophomonas,

Sphingobacterium and others were found in samples containing either Dickeya/ Pectobacterium

or not (Table 4-3). Arthrobacter, Kouboulia, Sporosarcina and Anaerosinus were only found in

the group N which comprised lower levels of Dickeya and Pectobacterium (Table 4-3). With the

exception of Dickeya and Pectobacterium, a few genera were only detected in sample groups

comprising higher levels of Dickeya and Pectobacterium (DP, D, P), including Pantoea, Delftia,

Flavobacterium, Serratia, Ralstonia and Aliidiomarina (Table 4-3).

4.3.2 Distribution of Dickeya and Pectobacterium spp.

As causal agents of PBSR, genera Genera Dickeya and Pectobacterium were retrieved

and analyzed individually, as they are causal agents of PBSR. The 76 samples were also grouped

by location (state). States showed a significant difference in RA of Dickeya and Pectobacterium

(Figure 4-3A, C). States with samples containing Dickeya spp. were New York, Massachusetts,

New Jersey and Pennsylvania, and the highest RA for each state was 98.4%, 91.5% 96.6% and

94.8%, respectively. In Michigan, five samples showed extremely low RA of Dickeya (<0.4%),

but a wide range in RA of Pectobacterium (<76.0%). Sixteen samples from Maine and nine

samples from Rhode Island contained lower RA of Dickeya with lower range of RA, which

showed similar distribution on Pectobacterium. The highest RA of Dickeya was 78.7% in group

DP and 98.3% in group D. Meanwhile, the highest RA of Pectobacterium was 39.2% in group

DP and 88.1% in group P. RA of both Dickeya and Pectobacterium was substantially reduced

when they co-existed (Figure 4-2B, D).

Page 107: Characterization of Dickeya dianthicola and Pectobacterium

92

Table 4-3. Relative abundance (%) of shared and exclusive genera of bacteria in four groups (D,

DP, P, and N) of potato samples showing symptoms of blackleg and soft rot.

Genus DPy (n = 22) D (n =22) P (n = 22) N (n = 10)

Ax

Acinetobacter 13.2 4.8 9.7 14.1

Providencia 7.8 1.5 2.8 12.6

Pseudomonas 13.8 18.1 11.2 22.0

Sphingobacterium 1.8 2.7 1.6 1.2

Stenotrophomonas 1.9 3.3 2.6 2.0

Unassigned 6.2 4.1 6.5 8.1

Comamonas 3.9 0.0 2.4 2.5

Myroides 1.8 0.0 1.3 3.0

Alcaligenes 0.0 0.0 1.9 5.1

Citrobacter 0.0 1.6 0.0 1.6

B

Anaerosinus 0.0 0.0 0.0 1.8

Arthrobacter 0.0 0.0 0.0 2.2

Koukoulia 0.0 0.0 0.0 2.0

Lachnoclostridium 5 0.0 0.0 0.0 1.4

Morganella 0.0 0.0 0.0 1.4

Paenalcaligenes 0.0 0.0 0.0 1.1

Enterobacter 0.0 0.0 0.0 1.4

Raoultella 0.0 0.0 0.0 1.1

Sporosarcina 0.0 0.0 0.0 1.8

C

Delftia 1.8 1.0 0.0 0.0

Dickeya 22.0 52.9 0.0 0.0

Flavobacterium 1.6 0.0 0.0 0.0

Pantoea 2.7 0.0 3.0 0.0

Pectobacterium 8.9 0.0 36.0 0.0

Serratia 1.5 0.0 0.0 0.0

Ralstonia 0.0 0.0 1.5 0.0

Aliidiomarina 0.0 0.0 1.5 0.0 x A, B and C were determined by the presence in types of samples N, DP, D or P. Group A contains the

genus only found in all samples; group B contains the genus only found in sample N; group C contains

the genus only found in sample DP, D, or P. y DP: containing both Dickeya and Pectobacterium spp.; D:

containing Dickeya spp.; P: containing Pectobacterium spp.; N: no Dickeya nor Pectobacterium spp.

Page 108: Characterization of Dickeya dianthicola and Pectobacterium

93

Figure 4-3. Relative abundances of Dickeya and Pectobacterium spp. by geographic locations or

states (A and C) and sample groups (B and D) of PBSR. Samples were grouped (x-axis)

according to polymerase chain reaction (PCR) results. N: no Dickeya nor Pectobacterium spp.;

P: containing Pectobacterium spp.; D: containing Dickeya spp.; DP: containing both Dickeya and

Pectobacterium spp.

DE MA ME NJ PA RI VA NY FL MI

Rel

ativ

eab

und

ance

0

0.2

0.4

0.6

0.8

1

1.2

Dickeya

DP D P None

Rel

ativ

eab

und

ance

0

0.2

0.4

0.6

0.8

1

1.2

Dickeya

DE MA ME NJ PA RI VA NY FL MI

Rel

ativ

eab

un

dan

ce

0

0.2

0.4

0.6

0.8

1

1.2

Pectobacterium

DP D P None

Rel

ativ

eab

un

dan

ce

0

0.2

0.4

0.6

0.8

1

1.2

Pectobacterium

A B

C D

DP (n=22) D (n=22) P (n=22) N (n=10)

DP (n=22) D (n=22) P (n=22) N (n=10)

Page 109: Characterization of Dickeya dianthicola and Pectobacterium

94

4.3.3 Presumptive bacterial species in genera Dickeya and Pectobacterium

To better understand the species community associated with PBSR, processed sequences

of each OTU from metagenomic sequencing were BLASTed into NCBI database of 16S rRNA

gene. Length of trimmed sequences ranged from 261 nt to 432 nt, with a mean of 423 nt.

Samples in group N were excluded here. All the species present in the alignment results with the

highest bit score (identical similarity) are included (Table 4-4). Due to a small fragment of 16S

rRNA gene sequence used in the blast, limited variable regions among species resulted in more

than one alignment with the highest bits score from one sequence. Percent composition of each

species relative to the total reads of Dickeya and Pectobacterium was calculated (Table 4-4).

Dickeya dadantii and D. dianthicola were the two species with almost same identical similarity

in V3-V4 region, resulting in same percentage of composition. However, only D. dianthicola

was detected by PCR and confirmed to be pathogenic to potato, while D. dadantii was not

detected in another study (Ge et al. 2020b). Compared to Dickeya spp., species community in

Pectobacterium were more complicated and differed greatly in each sample. Overall, P.

parmentieri, P. polaris, P. carotovora subsp. carotovora, P. carotovora subsp. odoriferum, were

the most prevalent species which were viewed as the presumptive causal agent of PBSR.

Pectobacterium carorovorum subsp. actinidiae, P. aroideanum, P. atrosepticum and P.

carotovora subsp. brasiliensis were found in few samples. Samples in group DP showed both

high percentage of Dickeya and Pectobacterium species.

Page 110: Characterization of Dickeya dianthicola and Pectobacterium

95

Table 4-4. Relative abundance (%) of shared and exclusive genera of bacteria in four groups (D,

DP, P, and N) of potato samples showing symptoms of blackleg and soft rot.

Groupy Ddaz Ddi Par Pat Pca Pcb Pcc Pco Ppa Ppo

P (n = 22)

0 0 0 0 0 0 0 0 98 99

0 0 0 1 0 1 57 63 41 41

1 0 0 0 0 0 0 0 98 98

1 1 1 32 2 32 63 64 33 3

1 1 1 2 1 5 35 37 60 59

1 1 0 0 0 0 0 0 97 97

1 1 51 1 43 1 2 2 3 3

1 1 1 1 0 1 6 16 69 82

1 1 0 1 2 0 1 2 93 94

2 2 2 1 70 1 25 26 2 2

4 4 0 1 0 0 3 4 90 91

0 0 1 0 0 1 97 98 1 1

1 0 1 0 0 1 91 93 7 7

1 1 1 0 0 1 94 95 3 3

2 1 12 0 16 1 36 38 35 36

2 2 18 1 6 3 60 61 11 12

2 2 0 1 0 0 25 26 71 71

4 4 3 0 13 6 71 72 3 3

4 4 1 1 3 2 82 84 9 9

4 4 2 3 1 2 2 2 88 86

15 14 1 1 1 1 8 9 72 72

30 30 1 2 1 3 31 60 8 6

DP (n = 22)

17 16 1 1 1 2 77 78 3 3

17 16 22 1 33 29 14 14 14 15

20 20 2 2 1 2 4 10 63 61

20 20 1 1 1 1 17 39 37 37

22 22 1 0 1 0 29 62 14 15

26 26 1 0 71 0 0 0 2 2

33 32 1 1 1 2 43 44 19 19

38 38 9 1 25 1 16 16 9 9

40 40 0 1 1 1 1 1 45 53

40 40 0 0 0 0 0 0 54 55

42 42 1 1 1 1 54 54 1 1

48 47 0 1 0 0 0 0 42 43

55 55 0 1 1 1 19 38 4 4

56 56 0 0 0 0 0 0 38 38

59 59 0 1 0 0 10 10 26 26

72 71 0 1 0 1 23 24 3 3

79 78 0 0 1 0 1 3 8 17

Page 111: Characterization of Dickeya dianthicola and Pectobacterium

96

Groupy Ddaz Ddi Par Pat Pca Pcb Pcc Pco Ppa Ppo

80 80 0 0 0 0 0 1 17 17

88 88 0 0 0 0 5 9 2 2

89 88 1 1 0 1 3 6 3 2

94 94 0 0 1 0 0 0 3 3

96 95 0 0 1 0 2 3 0 0

D (n = 22)

96 96 0 1 0 1 1 1 2 1

97 97 0 1 0 0 0 1 2 2

98 97 0 0 0 0 0 1 1 1

99 99 0 0 0 0 0 0 0 0

99 99 0 0 0 0 0 0 0 0

99 99 0 0 0 0 0 0 0 0

99 99 0 0 0 0 0 0 0 0

99 99 0 0 0 0 0 0 0 0

94 94 0 0 0 1 0 1 1 1

95 94 0 0 0 1 0 0 2 2

95 94 0 0 0 1 2 2 1 1

95 95 0 0 0 0 1 1 1 1

96 95 0 1 0 1 1 1 1 1

97 97 0 0 0 0 0 1 1 1

98 97 0 0 0 0 1 1 1 1

98 98 0 0 0 0 0 1 1 1

99 98 0 0 0 0 0 0 0 0

81 81 9 0 1 1 2 3 4 4

82 82 1 2 0 3 9 9 4 2

87 86 1 1 2 1 4 4 4 4

73 72 1 6 3 10 8 11 14 8

92 91 0 1 0 1 2 2 3 3

x Values of relative abundance were calculated as the percentage of reads related to each species on the total number

of reads related to genus Dickeya and Pectobacterium. y Bacterial groups include P containing Pectobacterium spp.,

D containing Dickeya spp., and DP containing both Dickeya and Pectobacterium spp. z Presumptive species in

genera Dickeya and Pectobacterium by blasting all V3-V4 sequences against the NCBI 16S rRNA database,

including D. dadantii (Dda), D. dianthicola (Ddi), P. aroidearum (Par), P. carotovora subsp. actinidiae (Pca), P.

carotovora subsp. brasiliensis (Pcb), P. carotovora subsp. carotovora (Pcc), P. carotovora subsp. odoriferum (Pco),

P. parmentieri (Ppa), and P. polaris (Ppo).

Page 112: Characterization of Dickeya dianthicola and Pectobacterium

97

4.3.4 Pathogenicity and growth of Pectobacterium spp.

Pectobacterium species were preliminarily confirmed to present in the PBSR samples by

metagenomic analysis. To confirm the presence of these presumptive species, Pectobacterium

species were isolated as many as possible and 110 pure cultures were obtained. Phylogenetic

analysis of the gapA gene showed that all isolates were classified into five Pectobacterium

species that were associated with PBSR in Northeastern US, including P. parmentieri, P.

carotovora subsp. carotovora, P. carotovora subsp. brasiliensis, P. carotovora subsp.

odoriferum and P. Polaris (Figure 4-4). Pectobacterium carotovora subsp. carotovora was

excluded for pathogenicity test as only one isolate was obtained.

Results from stem injection assay showed that P. carotovora subsp. brasiliensis and P.

parmentieri caused the stem to be completely rotted and collapsed (Figure 4-5). However,

symptoms caused by P. polaris and P. carotovora subsp. odoriferum caused mild or non-obvious

symptoms outside stems, but internal degradation only observed after dissection (Figure 4-5).

This difference might be related with the ability of pectic enzyme secretion and pectate

degradation. Results from tuber inoculation assay showed that all isolates of Pectobacterium

species could cause soft rot on potato tubers (Figure 4-6A). There was no significant difference

of pathogenicity among species (Figure 4-6B).

Page 113: Characterization of Dickeya dianthicola and Pectobacterium

98

Figure 4-4. Phylogenetic tree constructed based on partial sequences the gapA gene (850 bp)

using maximum-likelihood algorithm with strains of Pectobacterium spp., including P.

carotovora subsp. odoriferum (PCO), P. polaris (PPO), P. parmentieri (PPA), P. carotovora

subsp. carotovora (PCC), and P. carotovora subsp. brasiliensis (PCB), which were collected in

years 2015 (n = 37), 2016 (n = 51), and 2017 (n = 22). Reference strains (3-2, BC S7,

NBIO1006, BZA12, 67, IFB5485) and Escherichia coli O83:H1 (outgroup) were obtained from

the GenBank. Numbers next to the branching points indicate the relative support from 500

bootstrap replicates (only scores above 50 are shown). Colors in outer layer represent the five

Pectobacterium spp, and colors (legend) highlighted the sample ID in the middle layer represent

location (state) of potato samples.

Page 114: Characterization of Dickeya dianthicola and Pectobacterium

99

Figure 4-5. Pathogenicity test of Pectobacterium spp. Species include P. carotovora subsp.

brasiliensis (PCB), P. carotovora subsp. odoriferum (PCO), P. parmentieri (PPA), and P.

polaris (PPO) on potato ‘Lamoka’ stems. Bacterial inoculum (100 l of 109 CFU/ml) was

injected using a syringe on a stem at around 5 cm from soil line and incubated in a greenhouse at

25C (day) and 15C (night). Symptoms included stem collapse (PCB and PPA), blackened

vascular (PCO and PPO).

Page 115: Characterization of Dickeya dianthicola and Pectobacterium

100

Figure 4-6. Pathogenicity test of Pectobacterium spp. Species include P. carotovora subsp.

brasiliensis (PCB), P. carotovora subsp. odoriferum (PCO), P. parmentieri (PPA), and P.

polaris (PPO) on potato ‘Lamoka’ tubers. Potato tubers were inoculated adding 10 l bacterial

suspension at ~108 CFU/ml into a poked hole on a tuber, sealed with Parafilm, and were

incubated at 22C for 3 days. A) Decayed tissue was observed on inoculation point; B): Lesion

size was measured as width and length of a tuber dissected through the inoculation point. Filled

boxes represent range of lesion area (width length) and error bars represent standard deviation.

Page 116: Characterization of Dickeya dianthicola and Pectobacterium

101

In forming cavities on CVP plates, P. polaris, P. carotovora subsp. odoriferum, P.

carotovora subsp. brasiliensis showed cavities in less than 24 h, while P. parmentieri and P.

atroseptica required longer than 24 h. Dickeya dianthicola and D. solani required around 48 h,

while other Dickeya species formed cavities within 24 h (Table 4-5). Furthermore, cavity sizes

of P. polaris, P. carotovora subsp. odoriferum, P. carotovora subsp. brasiliensis were

significantly larger than that of D. dianthicola and P. atroseptica under any concentration and

time point (Table 4-5).

Table 4-5. Cavity diameter (cm) of bacterial species belonging to genera Dickeya and

Pectobacterium estimated by measuring the colony diameter on crystal violet pectate agar plates

incubated at 28 C in darkness for 24 and 48 hours.

Pathoge

n

24 h 48 h

109

CFU/mL

106

CFU/mL

102

CFU/mL

109

CFU/mL

106

CFU/mL

102

CFU/mL

PCBx 1.66 a 1.28 a 1.14 a 2.63 a 2.60 a 2.45 a

PPO 1.48 abcd 1.21 ab 0.94 ab 2.23 ab 2.23 b 2.05 ab

DZE 1.55 ab 1.21 ab 1.06 a 2.15 b 2.05 bc 1.85 bc

DDA 1.50 abc 1.23 a 0.60 bc 1.99 b 1.95 bc 1.59 cd

PPA 1.36 bcd 0.65 d 0.50 c 2.10 b 1.85 c 1.33 de

PCO 1.28 d 1.03 c 1.01 a 2.08 b 1.89 bc 1.68 bcd

DSO 1.32 cd 1.08 bc 0.78 abc 1.93 b 1.73 c 1.98 bc

DDI 0.60 e 0.50 e 0.50 c 1.23 c 0.65 d 0.90 ef

PAT 0.64 e 0.49 e 0.53 c 1.35 c 0.54 d 0.50 f x Abbreviation of full name of pathogens, PCB: P. carotovora subsp. brasiliensis, PPO: P. polaris, DZA: D.

zeae, DDA: D. dadantii, PPA: P. parmentieri, PCO: P. carotovora subsp. odoriferum, DSO: Dickeya solani,

DDI: D. dianthicola, PAT: P. atrosepticum.

Page 117: Characterization of Dickeya dianthicola and Pectobacterium

102

4.4 DISCUSSION

PBSR is a complex disease that can potentially be caused by more than one species of

bacteria. Dickeya dianthicola has been frequently and almost exclusively reported to be a

predominant causal agent of PBSR during the recent outbreak. In this study, PCR and

metagenomic approaches were employed to analyze pathogen composition and microbiota

associated with PBSR. It has been found that in a majority of plant samples, D. dianthicola was

the only species in the genus Dickeya. However, PBSR was not only caused by D. dianthicola

but caused or accompanied by several Pectobacterium spp. and maybe other potential pathogenic

bacteria. Therefore, PBSR could be a complex.

Dickeya and Pectobacterium are within the family Pectobacteriaceae. Thus, it was not

surprising that Pectobacteriaceae were the most abundant bacteria in the samples from diseased

plants, and that Dickeya spp. or Pectobacterium spp. was the most abundant genera present.

From our observations, it was found that Dickeya was more aggressive than Pectobacterium in

potato infection, and when they coexisted in potato tissues, PBSR was more severe than when

they occurred individually. This synergistic phenomenon was confirmed by subsequent

experimentation (chapter 5). Bacteria in families Pseudomonadaceae, Moraxellaceae and

Comamonadaceae, or the genera Pseudomonas and Acinetobacter were also common in many of

the potato samples, but their potential roles in disease development are not clear.

Plants are a driving force for the composition and diversity of microbiomes. Soils,

particularly rhizospheric soils provide unique environments for diversified endophytes and

epiphytes (Hardoim et al. 2015; Mendes et al. 2013; Hartmann et al. 2008; Buchholz et al. 2019).

In addition, PBSR pathogens may have strong associations and interactions with these

microorganisms. Comparing taxa in the four groups, genera such as Acinetobacter,

Page 118: Characterization of Dickeya dianthicola and Pectobacterium

103

Pseudomonas, Providencia, Stenotrophomonas and more were found in all four groups, which

might be endophytes in rhizosphere or phyllosphere that spread systemically via the xylem.

Arthrobacter, Sporosarcina and Anaerosinus were also commonly found in soil, but were only

detected in samples without Dickeya or Pectobacterium. Interestingly, several genera, such as

Alcaligenes, Pantoea and Ralstonia with very low level of abundance were only present in

diseased samples along with the pathogens Dickeya spp and Pectobacterium spp. These bacteria

may share common substrate with Dickeya spp. Pantoea and Ralstonia are worthy of further

investigation on their association with PBSR, since they were reported as unconventional or

important plant pathogens (Hayward 1991; Coutinho & Venter, 2009).

The results from conventional PCR and next generation sequencing were consistent in

diagnostic results, with few exceptions. For example, in samples that were detected by PCR as

Dickeya spp. or Pectobacterium spp. positive, exhibited higher relative abundance for Dickeya

and Pectobacterium spp. by next generation sequencing. Limitations for most 16S-sequence-

based microbiome studies are that they are short reads and do not give a fine resolution for

classification at species level. In this study, V3-V4 regions were used, which contain the most

variable regions and made the final sequence length of around 435 nt. A longer fragment favors

the analysis at species level. However, there was still a high similarity among species within

genera Dickeya or Pectobacterium, thus species-level results in this study were presumptive.

Dickeya dianthicola and D. dadantii were the two species that could not be clearly distinguished.

As such, using PCR, D. dianthicola has been detected in all samples with species-specific

primers to confirm its presence and isolated from a small percentage of samples, while D.

dadantii and D. chrysanthemi were not detected in the samples by PCR (chapter 2). This may be

Page 119: Characterization of Dickeya dianthicola and Pectobacterium

104

the reason why D. dianthicola has been reported as the primary causal agent in the northeastern

region, which was based on bacterial isolation and PCR detection.

Species in genera Dickeya and Pectobacterium were persistently reported worldwide but

without coexistence. In this study, it was proved that Dickeya and Pectobacterium existed

together and played important roles in causing PBSR. Pectobacterium within samples contained

multiple presumptive species, including P. parmentieri, P. polaris, P. carotovora subsp.

odoriferum, P. carotovora subsp. carotovora and P. carotovora subsp. brasiliensis. Dickeya.

dianthicola, P. parmentieri and P. carotovora subsp. brasiliensis resulted in complete collapse of

potato stems, so that they were regarded as the most aggressive pathogens to potato stems and

tubers (Figure 5, Ge et al. 2021). As most of Pectobacterium spp. are soil inhabitants, their

presence may not necessarily mean that they are pathogens. With such a rich community of

pathogenic bacteria, the severe soft rot in potato storage can be readily explained. The

importance of interacted and molecular behavior within a bacterial community has been

highlighted in recent research, and a metabolic complementarity is required for a stable bacterial

consortium (Egland et al. 2004; Kim et al. 2008).

In conclusion, D. dianthicola, P. parmentieri, P. carotovora subsp. brasiliensis, P.

carotovora subsp. odoriferum, and P. polaris were frequently found in the disease complex of

PBSR. The existence of D. dianthicola restricted the growth of Pectobacterium spp. to some

extent. In addition, appearance of the above pathogens reduced the microbial diversity in potato

plants.

Page 120: Characterization of Dickeya dianthicola and Pectobacterium

105

CHAPTER 5

INTERACTION BETWEEN DICKEYA DIANTHICOLA AND PECTOBACTERIUM

PARMENTIERI IN POTATO INFECTION OF POTATO UNDER FIELD CONDITIONS

This chapter has been published in Microorganisms:

Ge, T.L., Ekbataniamiri, F., Johnson, S.B., Larkin, R.P., and Hao, J.J. Interaction between Dickeya

dianthicola and Pectobacterium parmentieri in potato infection under field conditions. Microorganisms

2021, 9, 316.

ABSTRACT

Dickeya and Pectobacterium spp. both cause blackleg and soft rot of potato, which can be

a yield-reducing factor for potato production. The purpose of this study was to examine the

interaction between these two bacterial genera causing potato infection, and subsequent disease

development and yield responses under field conditions. Analysis of 883 potato samples

collected in the Northeastern USA using polymerase chain reaction determined that Dickeya

dianthicola and P. parmentieri were found in 38.1% and 53.3% of all samples, respectively, and

that 20.6% of samples contained both D. dianthicola and P. parmentieri. To further investigate

the relationship between the two bacterial species and their interaction, field trials were

established. Potato seed pieces of ‘Russet Burbank’, ‘Lamoka’, and ‘Atlantic’ were inoculated

with bacterial suspensions of D. dianthicola at 107 colony-forming units (CFU)/ml using a

vacuum-infiltration method, air dried, and then planted in the field. Two-year results showed that

there was a high correlation (P < 0.01) between relative yield loss and percentage of inoculated

seed pieces. In a secondary field trial conducted in 2018 and 2019, seed pieces of potato

‘Shepody’, ‘Lamoka’ and ‘Atlantic’ were inoculated with D. dianthicola, P. parmentieri, or

mixtures of both species, and then planted. In 2019, disease severity index as measured by the

most sensitive variety ‘Lamoka’ was 16.2 with D. dianthicola inoculation, 10.4 with P.

Page 121: Characterization of Dickeya dianthicola and Pectobacterium

106

parmentieri, 25.4 with inoculation with both bacteria. Two-year data had a similar trend. Thus,

D. dianthicola was more virulent than P. parmentieri, but the co-inoculation of the two species

resulted in increased disease severity compared to single-species inoculation with either

pathogen.

5.1 INTRODUCTION

Blackleg and soft rot (BSR) of potato (Solanum tuberosum) are caused by many bacterial

species in the genera Dickeya and Pectobacterium. The predominant species of bacteria

responsible for the disease varies depending on geographical location. For example,

Pectobacterium atroseptica was dominant in Europe before 1970, but Dickeya dianthicola and

followed by D. solani have become dominant in some European countries in recent decades

(Toth et al. 2011; van der Wolf et al. 2014). An outbreak of PBSR in the Northeastern USA was

caused by D. dianthicola in 2015 and the following years, while D. chrysanthemi was found in

the central USA, and Pectobacterium spp. were found everywhere in the USA (Hao et al. 2016;

Jiang et al. 2016; Johnson 2016; Ma et al. 2018; McNally et al. 2017, 2018; Patel et al. 2019).

Symptoms of PBSR were expressed as rot and blackened stems (blackleg), wilting plants and

decayed tubers (soft rot), which has threatened potato production, resulting in millions of dollars

lost by the potato industry (Johnson 2016)

BSR caused by a pathogen complex adds an additional layer of difficulty in the

development of effective control measures. One plant may be infected by more than one

pathogen (Fitt et al. 2006; Lamichhane and Venturi 2015). Such a complex disease structure is

determined on the interactions between pathogen species, host plant, biotic and abiotic

environmental factors (Dung et al. 2014; Fitt et al. 2006; Hibbing et al. 2010; Mazzaglia et al.

2012; Short et al. 2014).

Page 122: Characterization of Dickeya dianthicola and Pectobacterium

107

It is now generally believed that D. dianthicola is primarily a seed-borne pathogen, and

most Pectobacterium spp. are soil inhabitants but can be seed borne as well. Therefore, stored

tubers are the most important source in harboring bacterial pathogens and initiating disease.

Pathogens surviving in soil can colonize roots then penetrate and move into the xylem

(Czajkowski et al. 2009). Some studies indicate that P. parmentieri is less virulent than P.

carotovora, owing to the lack of Type III secretion system (T3SS) (Kim et al. 2009; Nykyri et

al. 2012). However, I have demonstrated that P. parmentieri isolates are highly pathogenic on

both tubers and stems of potato (Ge et al. 2018b).

In the USA PBSR outbreak that happened in 2015 and following years, Dickeya

dianthicola was determined to be the predominant pathogen involved, based on its primary and

predominant detection and isolation from field samples. However, the detection frequency of this

species progressively declined in the following years. In contrast, some Pectobacterium spp.

were found on postharvest tubers and later identified in symptomatic potato plants in increasing

amounts in subsequent years. Common species of PBSR causal agents included P. parmentieri,

P. polaris, P. carotovora subsp. carotovora, P. c. subsp. oderiferium, and other Pectobacterium

spp. (Ge et al. 2018a; Ge et al. 2018b; Jiang et al. 2016; Ma et al. 2018). The above species

were classified as one species, Erwinia carotovora before 1981, but were further differentiated

into several subspecies soon afterwards (Gardan et al. 2003; Perombelon and Kelman 1980;

Samson et al. 2005). In addition, P. parmentieri extended its distribution beyond the

Northeastern USA causing blackleg and soft rot of potato (Dung et al. 2012).

Whether the co-existence of Dickeya and Pectobacterium spp. has advantages for

pathogen infection is a key subject that impacts both biological studies and disease management.

Our questions were: Is the predominance of bacterial species influenced by environmental

Page 123: Characterization of Dickeya dianthicola and Pectobacterium

108

factors or bacterial virulence? Is there an enhanced response when both species co-occur in the

same infected region? The aims of this study were to understand how bacterial infection affected

potato growth and the corresponding yield responses. Specifically, I wanted to determine how

multiple species of pathogens interact, and how they affect disease development and yield

responses.

5.2 MATERIALS AND METHODS

5.2.1 Frequency and distribution of Dickeya dianthicola and Pectobacterium parmentieri in

naturally infected potato

From 2015 to 2020, symptomatic potato stems and tubers showing PBSR were collected

from fields in most of the northeastern states by growers, consultants, or plant pathologists. All

samples were processed and assayed at the University of Maine. Segments of diseased stem

tissue or peels of tubers were obtained from the samples. Bacteria were extracted from the tissue

sap by incubating them in sterile distilled water (Ge et al. 2020b). Genomic DNA was extracted

from the stem- or tuber-derived sap using the FastDNA SPIN Kit (MP Biomedical, Santa Ana,

CA, USA). Concentration of DNA was estimated using NanoDrop 2000c spectrophotometer

(Thermo Fisher Scientific Inc, Wilmington, DE, USA), and diluted to 10 ng/µL for polymerase

chain reaction (PCR). A total of 883 DNA samples were selected and amplified by polymerase

chain reaction (PCR) with primer pairs of PW7011 specific for P. parmentieri, and DIA-A

specific to D. dianthicola (Kim et al. 2012; Pritchard et al. 2013). PCR amplifications were

performed in a 25 L total volume reaction system containing 17.9 L of sterile water, 5 L of

5x PCR buffer, with MgCl2 at a final concentration of 0.25 mM dNTPs, 0.1 M of each pair of

primers, 0.5 U GoTaq DNA Polymerase and 1 L of DNA. All PCR reagents were from

Promega Corporation (Madison, MI). The thermal cycler for primer DIA-A was programmed for

Page 124: Characterization of Dickeya dianthicola and Pectobacterium

109

an initial denaturation of 5 min at 94°C, followed by 35 cycles of 1 min at 94°C, 30 s at 53°C

and 1 min at 72°C, with a final elongation of 10 min at 72°C, while the annealing temperature of

primer PW7011 was set at 67°C for 30 s.

5.2.2 Tuber inoculation using vacuum infiltration

Dickeya dianthicola strain ME30 and P. parmentieri strain ME175 were cultured on

crystal violate pectate (CVP) plates at 28 C for 48 h. The culture was transferred to tryptic soy

broth (TSB) and incubated on an incubator shaker overnight at 28 C and 120 rpm. The chamber

was disinfected with 10% bleach every time after inoculation of each treatment. At planting

inoculation, the prepared bacterial culture was diluted with non-autoclaved tap water and

adjusted to 107 CFU/mL as working inoculum. Well-suberized seed pieces were washed and

submerged into the bacteria suspension in a vacuum chamber (BVV10GL, BEST VALUE

VACS, USA) and tightly sealed. The chamber was held between -60 to -80 kPa for 15 min

(Czajkowski et al. 2010; Ge et al. 2020a). A control was set up by treating seed pieces with tap

water (pathogen free) as ‘non-inoculated’ through the same procedure in the vacuum chamber.

The treated tubers were air dried and planted one week after inoculation.

5.2.3 Effects of bacterial inoculation on potato growth

Field studies were conducted in 2020 at Presque Isle, Maine. A randomized complete

random block design (RCBD) was used with four replicate blocks. Potato varieties ‘Lamoka’,

‘Atlantic’, and ‘Russet Burbank’ were inoculated with D. dianthicola strain ME30 using the

vacuum-infiltration method as described above. For each variety, 50 tubers were planted on May

2020, and each treatment contained a mixture of bacteria-inoculated and bacteria-free tubers at

different percentages, including 0, 10%, 20%, 40% and 80%, which were arranged in a RCBD.

Fertilizer (N:P:K = 14:14:14) was applied at time of planting at a rate of 1,100 lb/A. Nuprid 2SC

Page 125: Characterization of Dickeya dianthicola and Pectobacterium

110

(a.i. 21.4% imidacloprid) was also applied at time of planting at 20 oz/A to control insects. Plant

stand was evaluated on 13 July 2020. Plants were visually inspected for the development of

PBSR symptoms. Disease severity was evaluated weekly or every other week after symptoms

appeared and rated from 0 to 5 with 0 being healthy and 5 being dead plants. Potato vines were

killed with two applications of Reglone at 1.5 pt/A. Potatoes were harvested on 2 October 2020.

Total yield was measured.

5.2.4 Interaction between D. dianthicola and P. parmentieri in potato infection

Field studies were performed in 2018 and repeated in 2019 in Presque Isle, Maine.

Dickeya dianthicola strain ME30 and P. parmentieri strain ME175 were used. Seed pieces of

potato cultivars ‘Lamoka’, ‘Atlantic’, and ‘Shepody’ were inoculated with either ME30, ME175,

or a mixture of both bacterial species using the vacuum-infiltration method as described above.

For the co-inoculation, cell suspension of strains ME30 and ME175 were mixed together at a 1:1

ratio with the final concentration of 107 CFU/mL, which was used for vacuum infiltration. The

treated tubers were air dried for at least one week before planting. A RCBD was used with four

replications in the field. Thirty seed pieces were planted in each plot on 25 May 2018 and 28

May 2019. Field management was as described in the above trial. Emergence was assessed as the

number of emerged plants per plot on 2 July 2018 and 1 July 2019. Plants were visually

inspected for the development of blackleg symptoms using a 0 to 5 disease severity scale as

described above. Disease incidence was calculated as DSD = number of symptomatic stems /

total number of stems. Disease severity index (DSI) was calculated as DSI = [sum (disease scale

frequency × score of disease scale)] / [(total number of stems) × (maximal disease severity

index)] × 100. Potatoes were harvested on 5 September 2018 and 19 September 2019. Total yield

was measured.

Page 126: Characterization of Dickeya dianthicola and Pectobacterium

111

5.2.5 Statistical analysis

Statistical analysis was conducted using the SAS statistical program (SAS university

edition, Red Hat (64-bit) version, SAS Institute Inc., Cary, NC, USA). Categorical variables

were analyzed using a chi-square test. Numerical variables, including emergence, disease

incidence, disease severity index and yield were analyzed using the ANOVA procedure with

Tukey’s multiple range test at a significance level α = 0.05. Relative emergence loss and relative

yield loss (Y) were regressed against percentage of seed infection (X) using linear regression

equation Y = b0 + b1X. Relative emergence loss was expressed as (noninfected plot stand count –

Infected plot stand count) / noninfected plot stand count. Relative yield loss = (yield in

noninfected plot – yield in infected plot / yield in noninfected plot).

5.3 RESULTS

5.3.1 Frequency and distribution of Dickeya dianthicola and Pectobacterium parmentieri in

naturally infected potato

Potato samples collected from fields were assayed using PCR to detect both Dickeya and

Pectobacterium spp. Out of the 883 samples, 761 were from potato stems, and 122 were from

potato tubers. Primers DIA-A specific to D. dianthicola and PW7011 specific to P. parmentieri

were used in the detection. Through the six years, 38.1% of samples contained D. dianthicola

and 53.3% of samples contained P. parmentieri. Furthermore, 20.6% of samples contained both

D. dianthicola and P. parmentieri, and 29.2% of samples neither contained D. dianthicola nor P,

parmentieri (Table 5-1). Chi square test showed that the percentage of D. dianthicola positive

stem samples (40.3%) was significantly (P < 0.001) higher than in tubers (23.8%), whereas the

percentage of P. parmentieri positive stems samples (51.8%) was significantly (P < 0.01) lower

than on tubers (63.1%) (Table 5-2).

Page 127: Characterization of Dickeya dianthicola and Pectobacterium

112

There was a trend that the percentage of samples infected by P. parmentieri increased

during the years, from 30.4% in 2015 to 95.3% in 2020 (Table 5-1). The percentage of total

samples infected by D. dianthicola declined from 41.7% to 9.5% through the six years (Table 5-

1).

Table 5-1. Detection of Dickeya and Pectobacterium spp. in symptomatic potato stems or tubers

using polymerase chain reaction (PCR).

Year

Number of

samples

D. dianthicola

(%)

P. parmentieri

(%)

D. dianthicola + P.

parmentieri (%)

2015 319 133 (41.7) 97 (30.4) 39 (12.2)

2016 230 114 (49.6) 134 (58.3) 72 (31.3)

2017 171 60 (35.1) 132 (77.2) 50 (29.2)

2018 81 14 (17.3) 45 (55.6) 9 (11.1)

2019 61 13 (21.3) 43 (70.5) 10 (16.4)

2020 21 2 (9.5) 20 (95.3) 2 (9.5)

Total 883 336 (38.1) 471 (53.3) 182 (20.6)

Table 5-2. Chi-square test on tissue distribution of Dickeya dianthicola (DDi) and

Pectobacterium parmentieri (PPa) on potato tissues showing blackleg and soft rot symptoms.

Number (percentage) of sample

Stem

(observed)

Tuber

(observed) Tuber (expected)a Chi-Sq df P value

DDi positive 307 (40.3%) 29 (23.8%) 49 (40.3%) 13.64 1 < 0.001

DDi negative 454 (59.7%) 93 (76.2%) 73 (59.7%)

Total 761 122 122

PPa positive 394 (51.8%) 77 (63.1%) 63 (51.8%) 6.43 1 < 0.01

PPa negative 367 (48.2%) 45 (36.9%) 59 (48.2%)

Total 761 122 122

a The number of tuber samples showing pathogen (DDi/PPa) positive or negative with the

assumption that there was no association between pathogen and tissue.

Page 128: Characterization of Dickeya dianthicola and Pectobacterium

113

5.3.2 Effects of bacterial inoculation on potato growth

Extremely dry growing conditions in 2020 resulted in an unexpectedly low emergence,

with the non-inoculated treatment having an emergence rate of 33%, 56% and 67% on ‘Lamoka’,

‘Atlantic’, and ‘Russet Burbank’ respectively (Table 5-3). Potato varieties showed different

responses in yield and emergence to bacterial inoculation (Table 5-3). The responses were

evaluated with regression models. Relative emergence loss and inoculation had a linear

relationship with coefficient (r2) of 0.44 on ‘Lamoka’, 0.62 on ‘Atlantic’ and 0.36 on ‘Russet

Burbank’, while yield loss and inoculation on ‘Lamoka’ and ‘Atlantic’ showed linear

relationships with r2 of 0.44 and 0.55, respectively (Figure 5-1). The level of disease

susceptibility was estimated by the slope of the regression equation between relative emergence

loss and inoculation. Slope of the regression equation between relative emergence loss and

inoculation was ‘Lamoka’ (0.687), ‘Atlantic’ (0.475), and ’Russet Burbank’ (0.332) (Figure 5-

1). ‘Lamoka’ was the most susceptible variety, the 100% inoculated seed treatment resulted in

68.7% emergence loss and 68.1% yield loss (Figure 5-1). For the tolerant variety ‘Russet

Burbank’, yield was not significantly affected by bacterial inoculation because emergence was

not significantly reduced (Figure 5-1).

5.3.3 Interaction between D. dianthicola and P. parmentieri in potato infection

Three potato varieties had different responses when inoculated with either D. dianthicola,

P. parmentieri, or mixture of both bacterial species, with ‘Lamoka’ being the most susceptible

variety. ‘Atlantic’ and ‘Shepody’ were less susceptible. This result was consistent in both years.

For example, in 2019, disease symptoms with inoculation of the bacterial mixture

appeared on ‘Lamoka’ at 49 days post planting with the disease severity index of 5.7, while the

disease severity indices of ‘Atlantic’ and ‘Shepody’ at 49 days post planting were 0.75 and 1.6

Page 129: Characterization of Dickeya dianthicola and Pectobacterium

114

respectively (Figure 5-2). At the end of the growing season, disease severity index of ‘Lamoka’

was 25.4, while ‘Atlantic’ and ‘Shepody’ showed relatively lower disease severity indices of 7.0

and 7.1, respectively (Figure 5-2).

Table 5-3. Effects of infection level of seed pieces with Dickeya dianthicola on yield and

emergence of potato in 2020.

Variety infected seeds

(%)

Total yield

(cwt/A)

Emergence

(%)

Total yield/plant

(cwt/A)

Lamoka 0 100.7 33.0 6.8

10 69.1 20.0 6.7

10 77.4 28.6 6.0

40 63.8 18.0 7.1

80 34.7 13.0 6.3

Atlantic 0 148.2 56.0 5.3

10 151.2 44.6 6.7

20 131.0 46.0 5.7

40 136.8 41.6 6.7

80 104.2 31.0 6.7

Russet

Burbank

0 171.3 67.0 5.2

10 148.9 67.6 4.4

20 154.7 65.0 4.8

40 191.0 62.6 6.2

80 162.7 49.0 6.6

Page 130: Characterization of Dickeya dianthicola and Pectobacterium

115

Figure 5-1. Regression analysis between relative emergence loss/relative yield loss and infection

percentage of potato. 1) (left panels) regression models between relative emergence loss (Y) and

infection percentage (X) of potato ‘Lamoka’ (Y = 0.687X+0.062, r2 = 0.44, P = 0.0017),

‘Atlantic’ (Y = 0.475X+0.073, r2 = 0.62, P < 0.0001), ‘Russet Burbank’ (Y = 0.332X-0.038, r2 =

0.36, P = 0.0056) seed pieces inoculated with Dickeya dianthicola; 2) (right panels) regression

models between relative yield loss (Y) and infection percentage (X) of potato ‘Lamoka’ (Y =

0.681X+0.095, r2 = 0.44, P = 0.0014), ‘Atlantic’ (Y = 0.374X-0.019, r2 = 0.55, P = 0.0002),

‘Russet Burbank’ (r2 = 0.02, P = 0.5850) seed pieces inoculated with Dickeya dianthicola.

Predicted data and confident intervals (CI) were shown as the dash lines and orange curves,

respectively, 2020.

Page 131: Characterization of Dickeya dianthicola and Pectobacterium

116

Figure 5-2. Disease progress of blackleg in potato ‘Lamoka’, ‘Atlantic’ and ‘Shepody’ grown

from seed pieces inoculated with either Dickeya dianthicola (DDi), Pectobacterium parmentieri

(PPa), or DDi + PPa under field conditions in 2018 (upper panels) and 2019 (bottom panels).

Non-inoculated (NI) plants were used for control.

Page 132: Characterization of Dickeya dianthicola and Pectobacterium

117

Dickeya dianthicola was more aggressive than P. parmentieri in the field. For example,

in the season of 2019 and on the most susceptible variety ‘Lamoka’, symptoms were first

observed with inoculation of D. dianthicola. The disease severity index was 6.7 for D.

dianthicola inoculation but 0.7 for P. parmentieri (Figure 5-2). Furthermore, by the end of the

season, disease severity index of ‘Lamoka’ caused by D. dianthicola was 16.2 higher than that

caused by P. parmentieri (10.4) (Figure 5-2). However, the mixture of both bacteria showed the

highest disease level, with disease severity index being 25.4 (Figure 5-2). Thus, the mixture of

two bacterial species had advantages in plant infection. The responses of ‘Shepody’ and

‘Atlantic’ were similar to ‘Lamoka’ for the infection by either single or mixture of bacterial

species, although the difference among treatments was not significant (P > 0.05), because both

varieties were less susceptible to blackleg (Figure 5-2). At early growing stages of potato, a

higher disease incidence was observed with D. dianthicola inoculation compared to the treatment

with P. parmentieri inoculation. While inoculation with the mixture of bacterial species started

with disease level slightly lower than that of D. dianthicola, but soon surpassed it. Although not

statistically significant, the disease severity index of treatments in 2018 showed a trend that D.

dianthicola inoculation was higher than P. parmentieri inoculation regardless of potato variety

(Figure 5-2).

All bacterial inoculations significantly reduced potato yield compared to the non-

inoculated treatment in 2019, yield differences in the 2018 trial were not significant (Table 5-4).

The treatment with a mixture of bacterial species did not show a significant difference compared

to D. dianthicola inoculation (Table 5-3). Thus, D. dianthicola was a key to cause higher yield

loss when potatoes were either infected solely by D. dianthicola or coinfected by D. dianthicola

and P. parmentieri.

Page 133: Characterization of Dickeya dianthicola and Pectobacterium

118

Table 5-4. Effects of seed tuber inoculation with Dickeya dianthicola (DDi), Pectobacterium

parmentieri (PPa) and non-inoculated (NI) on emergence and yield of potato varieties in years

2018 and 2019.

Variety Inoculation 2018 2019

Emergence

(%)

Total yield

(cwt/A)

Emergence

(%)

Total yield

(cwt/A)

Lamoka NI 95 118.3 98.3 166.2 ax

DDi 95.0 100.4 85.0 120.8 b

PPa 93.3 117.9 90.8 152.7 b

DDi & PPa 96.7 123.2 89.2 120.4 b

Atlantic NI 96.7 129.8 85.8 147.8 a

DDi 86.7 124.4 76.7 113.6 b

PPa 83.3 131.8 84.2 121.6 b

DDi & PPa 86.7 97.8 82.5 122.0 b

Shepody NI 96.7 127.0 95.0 188.8 a

DDi 96.7 101.4 86.7 138.6 b

PPa 96.7 157.8 87.5 145.9 b

DDi & PPa 91.7 111.2 90.8 136.6 b xMean values followed by different letters were significantly different (P < 0.05).

5.4 DISCUSSION

Dickeya dianthicola was found to be the predominant species causing PBSR during the

outbreak happened in 2015 in the Northeastern USA. There was a trend that the number of

samples infected by P. parmentieri increased during the years, while the percentage of total

samples infected by D. dianthicola declined through the six years. Pectobacterium parmentieri

turned to be another problem along with D. dianthicola, which was persistently detected as the

predominant pathogen in the following years. Furthermore, D. dianthicola was more likely a

pathogen associated with stems, and P. parmentieri was highly associated with potato tubers.

At least 20.6% of infected potato plants contained both D. dianthicola and P. parmentieri

in this study. In addition, several other Pectobacterium spp. have also frequently been found in

Page 134: Characterization of Dickeya dianthicola and Pectobacterium

119

naturally infected seed potatoes. It seemed that the coexistence of multiple bacterial species is

common in blackleg, forming a pathogen complex. Since D. dianthicola was consistently

isolated as the predominant species, it is believed that D. dianthicola was the primary pathogen

that aggressively caused disease in the field. Many Pectobacterium spp. can survive well in soil

or they are soil inhabitants. Since they are weak pathogens, they may not cause infection.

However, their availability in fields allows them to follow the initiation of infection by Dickeya

spp. thus disease severity increased.

It is possible that D. dianthicola, compared to P. parmentieri, has more virulence-related

factors in plant infection, such as enzymes to degrade pectin and cellulose (Barras et al. 1994), or

secretion systems to transfer virulence effectors (Charkowski et al. 2012; Kim et al. 2009).

Regarding pathogenicity, D. dianthicola might have advantages over P. parmentieri owing to the

lack of T3SS in P. parmentieri (Kim et al. 2009; Nykyri et al. 2012), as T3SS contributes to the

virulence of bacteria during the early stages of infection (Ham et al. 2004; Rantakari et al.

2001). Studies have reported that D. dadantii exhibited a high number of pectinase-related

enzymes that would enhance the degradation of pectin (Dittoe et al. 2020), and Dickeya spp.

possess complex regulatory networks in order to express virulence genes (Reverchon et al.

2016). However, detailed studies about virulence-related mechanisms have not been established

for D. dianthicola.

Our two-year data supports that D. dianthicola causes a higher level of PBSR severity

than P. parmentieri, and that inoculation with both bacterial species did increase PBSR severity

later during the growing season (Figure 5-2). This suggested that D. dianthicola should be

considered the primary pathogen in disease control. Meanwhile, due to the wide distribution of

P. parmentieri, it is possible that both bacterial species co-existed in the environment and co-

Page 135: Characterization of Dickeya dianthicola and Pectobacterium

120

infect or -damage potato plants. This could be a consequence of synergy between the two

bacterial species. In production, if D. dianthicola is eliminated, the less-aggressive P.

parmentieri would cause less impact on potato growth and yield. More importantly, it would

remove the complex of co-infection with two or more bacteria. Therefore, it is necessary to

further investigate how Dickeya spp. and Pectobacterium spp. interact with each other inside

potato plants in increasing disease severity.

Competition, cooperation, and coexistence are major means of interactions of bacteria

that co-exist in a same niche (Abdullah et al. 2017). Although no coexistence has been reported

of D. solani and D. dianthicola in fields, D. solani could outcompete D. dianthicola in co-

inoculated plants (Czajkowski et al. 2013b). In this study, there was no antagonistic effect

between D. dianthicola and P. parmentieri when they coexisted but aggravated the disease

epidemic at a higher level.

Surprisingly, even though D. dianthicola or a mixture of the two bacterial species caused

higher disease than the sole infection by P. parmentieri, the total yield did not show a significant

difference (Table 5-4). It was interpreted that yield is constrained by early infection with stand

loss, but less affected by late disease progress. Furthermore, among the three bacterial

treatments, P. parmentieri had less impact on yield compared to the other two. This implied that

D. dianthicola was more aggressive or virulent than P. parmentieri.

Yield losses of potato caused by artificial inoculation of D. dianthicola were evaluated in

this study. There was a high correlation of percentage of seed inoculation/contamination with

lack of emergence, as well as with yield reduction. This was because yield per area of field was

highly correlated with plant emergence or total number of plants. PBSR showed up in the

growing season at a low level and most of the plants survived. Therefore, low percent in-season

Page 136: Characterization of Dickeya dianthicola and Pectobacterium

121

disease did not significantly affect potato yield. In addition to disease severity, yield is also

dependent on weather conditions, such as precipitation and soil temperature, soil properties,

pathogen aggressiveness, and the ability of the crop to compensate for stand loss. Based on our

results, ‘Lamoka’ was the most susceptible variety to PBSR, while ‘Russet Burbank’ was more

tolerant, ‘Atlantic’ and ‘Shepody’ were moderate. Such information may be incorporated in

breeding a program for BRS resistance.

In conclusion, tuber contamination with either D. dianthicola or P. parmentieri caused

direct stand loss that translated to yield losses, and the percent of tuber infection was negatively

correlated with emergence. When a single species of pathogen infected potato, D. dianthicola

caused more symptoms than P. parmentieri. However, co-infection of the two species showed

even higher disease severity of PBSR, and a synergistic effect of bacterial species in plant

infection.

Page 137: Characterization of Dickeya dianthicola and Pectobacterium

122

DISSERTATION SUMMARY

This research was conducted on potato blackleg and soft rot (PBSR) and its recent

outbreak in the northeastern U.S. and attempted to address four key questions related to

understanding the pathogen, disease, and the outbreak. These questions were: 1) What was the

origin and cause of the outbreak? 2) What are the genetic characteristics of the pathogen

population responsible for the outbreak compared to those occurring throughout the world? 3)

Were multiple bacteria species involved in outbreak infections? And 4) How do multiple bacteria

species in a pathogen complex interact and affect disease development and yield? By analyzing

thousands of diseased potatoes samples collected from the Northeastern states using molecular

and genetic tools, a clear image has been drawn of the outbreak. The disease epidemic was solely

caused by domestic USA strains of Dickeya dianthicola, which contained three genotypes.

Genotype I was the predominant population and its incursion into potato production could be the

main reason for the PBSR outbreak. It was found at a high percentage in Dickeya positive

samples in all states every year for the five years of sample collection and was distinctly different

from all known genotypes from other parts of the world. By analyzing the whole genome of the

bacteria, genotype I was confirmed to have genetic advantages by having extra copies of types

IV and VI secretion systems, which are highly correlated to bacterial pathogenesis and virulence.

This genetic arrangement might have enhanced the genotype I strain allowing in to become more

competitive. Genotype III was mostly common in the more southerly states. It was a causal agent

of PBSR but not a reason for the outbreak since it was not observed in Maine where much of the

region’s potato seed is produced. Seed was the primary carrier of PBSR pathogens for long-

distance movement.

Page 138: Characterization of Dickeya dianthicola and Pectobacterium

123

PBSR can be a complex as many other bacteria were involved. This was discovered

using high throughput DNA sequencing. At least four species of Pectobacterium have been

retrieved from symptomatic potato tissues and confirmed to be virulent to potato. More

importantly, D. dianthicola and P. parmentieri showed a synergy in plant infection. The trend of

the stem-associated pathogen D. dianthicola declining and the tuber-related pathogen P.

parmentieri increasing in the past 5 years may indicate another challenge to potato production,

which should be addressed in the future. This work has provided answers to the fundamental

questions surrounding the origin and pathogens involved in the recent outbreak of PBSR, as well

as insight into more effective management of these diseases. Thus, this work significantly

contributed to the productivity and sustainability of potato production, the most important

vegetable crop in the world.

Page 139: Characterization of Dickeya dianthicola and Pectobacterium

124

REFERENCES

Abdullah, A. S., Moffat, C. S., Lopez-Ruiz, F. J., Gibberd, M. R., Hamblin, J., and Zerihun, A.

2017. Host–multi-pathogen warfare: pathogen interactions in co-infected plants. Frontiers

in Plant Science 8:1806.

Adams, M. J. 1975a. Potato tuber lenticels: susceptibility to infection by Erwinia carotovora var.

atroseptica and Phytophthora infestans. Annals of Applied Biology 79:275-282.

Adams, M. J. 1975b. Potato tuber lenticels: development and structure. Annals of Applied

Biology 79:265-273.

Adams, M. J., and Lapwood, D. H. 1978. Studies on the lenticel development, surface microflora

and infection by common scab (Streptomyces scabies) of potato tubers growing in wet

and dry soils. Annals of Applied Biology 90:335-343.

Adeolu, M., Alnajar, S., Naushad, S., and Gupta, R. S. 2016. Genome-based phylogeny and

taxonomy of the ‘Enterobacteriales’: proposal for Enterobacterales ord. nov. divided

into the families Enterobacteriaceae, Erwiniaceae fam. nov., Pectobacteriaceae fam.

nov., Yersiniaceae fam. nov., Hafniaceae fam. nov., Morganellaceae fam. nov., and

Budviciaceae fam. nov. International Journal of Systematic and Evolutionary

Microbiology 66:5575-5599.

Aizawa, S.-I. 2014. Flagellar World: electron microscopic images of bacterial flagella and

related surface structures. Elsevier.

Alikhan, N.-F., Petty, N. K., Zakour, N. L. B., and Beatson, S. A. 2011. BLAST Ring Image

Generator (BRIG): simple prokaryote genome comparisons. BMC Genomics 12:402.

Alivizatos, A. S. 1985. Bacterial wilt of tomato in Greece caused by Erwinia chrysanthemi. Plant

Pathology 34:638-639.

Alič, Š., Van Gijsegem, F., Pedron, J., Ravnikar, M., and Dreo, T. 2018. Diversity within the

novel Dickeya fangzhongdai sp., isolated from infected orchids, water and pears. Plant

Pathology 67:1612-1620.

Alvarez-Martinez, C. E., and Christie, P. J. 2009. Biological diversity of prokaryotic type IV

secretion systems. Microbiology and Molecular Biology Reviews 73:775-808.

Page 140: Characterization of Dickeya dianthicola and Pectobacterium

125

Ansermet, M., Schaerer, S., Kellenberger, I., Tallant, M., and Dupuis, B. 2016. Influence of

seed-borne and soil-carried inocula of Dickeya spp. on potato plant transpiration and

symptom expression. European Journal of Plant Pathology 145:459-467.

Arber, W. 2014. Horizontal gene transfer among bacteria and its role in biological evolution.

Life 4:217-224.

Armbruster, C. E., Hong, W., Pang, B., Weimer, K. E. D., Juneau, R. A., Turner, J., and Swords,

W. E. 2010. Indirect pathogenicity of Haemophilus influenzae and Moraxella catarrhalis

in polymicrobial otitis media occurs via interspecies quorum signaling. MBio 1:e00102-

10.

Baghaee-Ravari, S., Rahimian, H., Shams-Bakhsh, M., Lopez-Solanilla, E., Antúnez-Lamas, M.,

and Rodríguez-Palenzuela, P. 2011. Characterization of Pectobacterium species from

Iran using biochemical and molecular methods. European Journal of Plant Pathology

129:413-425.

Barras, F., van Gijsegem, F., and Chatterjee, A. K. 1994. Extracellular enzymes and

pathogenesis of soft-rot Erwinia. Annual Review of Phytopathology 32:201-234.

Bartz, J. A., and Kelman, A. 1985. Effect of air-drying on soft rot potential of potato tubers

inoculated by immersion in suspensions of Erwinia carotovora. Plant Disease 69:128-

131.

Bdliya, B. S., and Haruna, H. U. 2007. Efficacy of solar heat in the control of bacterial soft rot of

potato tubers caused by Erwinia carotovora ssp. carotovora. Journal of Plant Protection

Research 47:11-18.

Bell, K. S., Avrova, A. O., Holeva, M. C., Cardle, L., Morris, W., De Jong, W., Toth, I. K.,

Waugh, R., Bryan, G. J., and Birch, P. R. J. 2002. Sample sequencing of a selected region

of the genome of Erwinia carotovora subsp. atroseptica reveals candidate

phytopathogenicity genes and allows comparison with Escherichia coli The GenBank

accession numbers for the 424 sequences determined in this work are BH614193 to

BH614616. Microbiology 148:1367-1378.

Bell, K. S., Sebaihia, M., Pritchard, L., Holden, M. T. G., Hyman, L. J., Holeva, M. C.,

Thomson, N. R., Bentley, S. D., Churcher, L. J. C., and Mungall, K. 2004. Genome

sequence of the enterobacterial phytopathogen Erwinia carotovora subsp. atroseptica and

characterization of virulence factors. Proceedings of the National Academy of Sciences

101:11105-11110.

Page 141: Characterization of Dickeya dianthicola and Pectobacterium

126

Bellieny-Rabelo, D., Tanui, C. K., Miguel, N., Kwenda, S., Shyntum, D. Y., and Moleleki, L. N.

2019. Transcriptome and comparative genomics analyses reveal new functional insights

on key determinants of pathogenesis and interbacterial competition in Pectobacterium

and Dickeya spp. Applied and Environmental Microbiology 85.

Berendsen, R. L., Pieterse, C. M. J., and Bakker, P. A. H. M. 2012. The rhizosphere microbiome

and plant health. Trends in Plant Science 17:478-486.

Berg, G., Rybakova, D., Fischer, D., Cernava, T., Vergès, M.-C. C., Charles, T., Chen, X.,

Cocolin, L., Eversole, K., and Corral, G. H. 2020. Microbiome definition re-visited: old

concepts and new challenges. Microbiome 8:1-22.

Bertin, C., Yang, X., and Weston, L. A. 2003. The role of root exudates and allelochemicals in

the rhizosphere. Plant and Soil 256:67-83.

Binyam, T. 2014. Late blight of potato (Phytophthora infestans) biology, economic importance

and its management approaches. Journal of Biology, Agriculture and Healthcare 4:215-

225.

Bleidorn, C. 2016. Third generation sequencing: technology and its potential impact on

evolutionary biodiversity research. Systematics and Biodiversity 14:1-8.

Blin, P., Robic, K., Khayi, S., Cigna, J., Munier, E., Dewaegeneire, P., Laurent, A., Jaszczyszyn,

Y., Hong, K. W., and Chan, K. G. 2020. Pattern and causes of the establishment of the

invasive bacterial potato pathogen Dickeya solani and of the maintenance of the resident

pathogen D. dianthicola. Molecular Ecology 30:608-624.

Bolyen, E., Rideout, J. R., Dillon, M. R., Bokulich, N. A., Abnet, C. C., Al-Ghalith, G. A.,

Alexander, H., Alm, E. J., Arumugam, M., and Asnicar, F. 2019. Reproducible,

interactive, scalable and extensible microbiome data science using QIIME 2. Nature

Biotechnology 37:852-857.

Bosi, E., Fani, R., and Fondi, M. 2015. Defining orthologs and pangenome size metrics. Pages

191-202 in: Bacterial Pangenomics. Springer.

Bouchez, V., Baines, S. L., Guillot, S., and Brisse, S. 2018. Complete genome sequences of

Bordetella pertussis clinical isolate FR5810 and reference strain Tohama from combined

Oxford Nanopore and Illumina sequencing. Microbiology Resource Announcements 7.

Page 142: Characterization of Dickeya dianthicola and Pectobacterium

127

Brady, C. L., Cleenwerck, I., Denman, S., Venter, S. N., Rodriguez-Palenzuela, P., Coutinho, T.

A., and De Vos, P. 2012. Proposal to reclassify Brenneria quercina (Hildebrand and

Schroth 1967) Hauben et al. 1999 into a new genus, Lonsdalea gen. nov., as Lonsdalea

quercina comb. nov., descriptions of Lonsdalea quercina subsp quercina comb. nov.,

Lonsdalea quercina subsp iberica subsp nov and Lonsdalea quercina subsp britannica

subsp nov., emendation of the description of the genus Brenneria, reclassification of

Dickeya dieffenbachiae as Dickeya dadantii subsp dieffenbachiae comb. nov., and

emendation of the description of Dickeya dadantii. International Journal of Systematic

and Evolutionary Microbiology 62:1592-1602.

Brown, C. R. 1993. Origin and history of the potato. American Potato Journal 70:363-373.

Brown, J. K. M. 2015. Durable resistance of crops to disease: a Darwinian perspective. Annual

Review of Phytopathology 53:513-539.

Callahan, B. J., McMurdie, P. J., and Holmes, S. P. 2017. Exact sequence variants should replace

operational taxonomic units in marker-gene data analysis. The ISME Journal 11:2639-

2643.

Caporaso, J. G., Lauber, C. L., Walters, W. A., Berg-Lyons, D., Huntley, J., Fierer, N., Owens,

S. M., Betley, J., Fraser, L., and Bauer, M. 2012. Ultra-high-throughput microbial

community analysis on the Illumina HiSeq and MiSeq platforms. The ISME Journal

6:1621-1624.

Carvajal-Yepes, M., Cardwell, K., Nelson, A., Garrett, K. A., Giovani, B., Saunders, D. G. O.,

Kamoun, S., Legg, J. P., Verdier, V., and Lessel, J. 2019. A global surveillance system

for crop diseases. Science 364:1237-1239.

Cascales, E., and Christie, P. J. 2003. The versatile bacterial type IV secretion systems. Nature

Reviews Microbiology 1:137-149.

Chakravorty, S., Helb, D., Burday, M., Connell, N., and Alland, D. 2007. A detailed analysis of

16S ribosomal RNA gene segments for the diagnosis of pathogenic bacteria. Journal of

Microbiological Methods 69:330-339.

Chandran, V., Fronzes, R., Duquerroy, S., Cronin, N., Navaza, J., and Waksman, G. 2009.

Structure of the outer membrane complex of a type IV secretion system. Nature

462:1011-1015.

Page 143: Characterization of Dickeya dianthicola and Pectobacterium

128

Chaparro, J. M., Badri, D. V., and Vivanco, J. M. 2014. Rhizosphere microbiome assemblage is

affected by plant development. The ISME Journal 8:790-803.

Charkowski, A., Blanco, C., Condemine, G., Expert, D., Franza, T., Hayes, C., Hugouvieux-

Cotte-Pattat, N., Solanilla, E. L., Low, D., and Moleleki, L. 2012. The role of secretion

systems and small molecules in soft-rot Enterobacteriaceae pathogenicity. Annual

Review of Phytopathology 50:425-449.

Charkowski, A. O. 2018. The changing face of bacterial soft-rot diseases. Annual Review of

Phytopathology 56:269-288.

Chatterjee, A., Cui, Y., Chakrabarty, P., and Chatterjee, A. K. 2010. Regulation of motility in

Erwinia carotovora subsp. carotovora: quorum-sensing signal controls FlhDC, the global

regulator of flagellar and exoprotein genes, by modulating the production of RsmA, an

RNA-binding protein. Molecular Plant-Microbe Interactions 23:1316-1323.

Chen, X. F., Zhang, H. L., & Chen, J. (2015). First report of Dickeya solani causing soft rot in

imported bulbs of Hyacinthus orientalis in China. Plant disease 99, 155.

Chen, S., Zhou, Y., Chen, Y., and Gu, J. 2018. fastp: an ultra-fast all-in-one FASTQ

preprocessor. Bioinformatics 34:i884-i890.

Chujoy, E., and Cabello, R. 2007. The canon of potato science: 29. True potato seed (TPS).

Potato Research 50:323-325.

Cigna, J., Dewaegeneire, P., Beury, A., Gobert, V., and Faure, D. 2017. A gapA PCR-

sequencing assay for identifying the Dickeya and Pectobacterium potato pathogens. Plant

Disease 101:1278-1282.

Collmer, A., and Keen, N. T. 1986. The role of pectic enzymes in plant pathogenesis. Annual

Review of Phytopathology 24:383-409.

Compant, S., Clément, C., and Sessitsch, A. 2010. Plant growth-promoting bacteria in the rhizo-

and endosphere of plants: their role, colonization, mechanisms involved and prospects for

utilization. Soil Biology and Biochemistry 42:669-678.

Compant, S., Samad, A., Faist, H., and Sessitsch, A. 2019. A review on the plant microbiome:

Ecology, functions, and emerging trends in microbial application. Journal of Advanced

Research 19:29-37.

Page 144: Characterization of Dickeya dianthicola and Pectobacterium

129

Cook, L. C., LaSarre, B., and Federle, M. J. 2013. Interspecies communication among

commensal and pathogenic streptococci. MBio 4.

Cother, E. J., Bradley, J. K., Gillings, M. R., and Fahy, P. C. 1992. Characterization of Erwinia

chrysanthemi biovars in alpine water sources by biochemical properties, GLC fatty acid

analysis and genomic DNA fingerprinting. Journal of Applied Bacteriology 73:99-107.

Czajkowski, R. 2016. Bacteriophages of soft rot Enterobacteriaceae—a minireview. FEMS

Microbiology Letters 363.

Czajkowski, R., Grabe, G. J., and van der Wolf, J. M. 2009. Distribution of Dickeya spp. and

Pectobacterium carotovorum subsp. carotovorum in naturally infected seed potatoes.

European Journal of Plant Pathology 125:263-275.

Czajkowski, R., De Boer, W. J., and Van der Wolf, J. M. 2013a. Chemical disinfectants can

reduce potato blackleg caused by ‘Dickeya solani’. European Journal of Plant Pathology

136:419-432.

Czajkowski, R., Ozymko, Z., and Lojkowska, E. 2014. Isolation and characterization of novel

soilborne lytic bacteriophages infecting Dickeya spp. biovar 3 (‘D. solani’). Plant

Pathology 63:758-772.

Czajkowski, R., de Boer, W. J., Velvis, H., and van der Wolf, J. M. 2010. Systemic colonization

of potato plants by a soilborne, green fluorescent protein-tagged strain of Dickeya sp.

biovar 3. Phytopathology 100:134-142.

Czajkowski, R., Perombelon, M. C. M., van Veen, J. A., and van der Wolf, J. M. 2011. Control

of blackleg and tuber soft rot of potato caused by Pectobacterium and Dickeya species: a

review. Plant Pathology 60:999-1013.

Czajkowski, R., De Boer, W. J., Van Veen, J. A., and Van der Wolf, J. M. 2012. Studies on the

interaction between the biocontrol agent, Serratia plymuthica A30, and blackleg‐causing

Dickeya sp.(biovar 3) in potato (Solanum tuberosum). Plant Pathology 61:677-688.

Czajkowski, R., van der Wolf, J. M., Krolicka, A., Ozymko, Z., Narajczyk, M., Kaczynska, N.,

and Lojkowska, E. 2015a. Salicylic acid can reduce infection symptoms caused by

Dickeya solani in tissue culture grown potato (Solanum tuberosum L.) plants. European

Journal of Plant Pathology 141:545-558.

Page 145: Characterization of Dickeya dianthicola and Pectobacterium

130

Czajkowski, R., De Boer, W. J., Van der Zouwen, P. S., Kastelein, P., Jafra, S., De Haan, E. G.,

Van den Bovenkamp, G. W., and Van der Wolf, J. M. 2013b. Virulence of ‘Dickeya

solani’and Dickeya dianthicola biovar‐1 and‐7 strains on potato (Solanum tuberosum).

Plant Pathology 62:597-610.

Czajkowski, R., Ozymko, Z., de Jager, V., Siwinska, J., Smolarska, A., Ossowicki, A.,

Narajczyk, M., and Lojkowska, E. 2015b. Genomic, proteomic and morphological

characterization of two novel broad host lytic bacteriophages ΦPD10. 3 and ΦPD23. 1

infecting pectinolytic Pectobacterium spp. and Dickeya spp. PLoS One 10:e0119812.

da Silva, D. P., Castañeda-Ojeda, M. P., Moretti, C., Buonaurio, R., Ramos, C., and Venturi, V.

2014. Bacterial multispecies studies and microbiome analysis of a plant disease.

Microbiology 160:556-566.

Darling, A. C. E., Mau, B., Blattner, F. R., and Perna, N. T. 2004. Mauve: multiple alignment of

conserved genomic sequence with rearrangements. Genome Research 14:1394-1403.

Davies, H. V. 1984. Mother tuber reserves as factors limiting potato sprout growth. Potato

Research 27:209-218.

De Boer, S. H., and Rubio, I. 2004. Blackleg of potato. The Plant Health Instructor.

De Boer, S. H., Li, X., and Ward, L. J. 2012. Pectobacterium spp. associated with bacterial stem

rot syndrome of potato in Canada. Phytopathology 102:937-947.

De la Cruz, F., and Davies, J. 2000. Horizontal gene transfer and the origin of species: lessons

from bacteria. Trends in Microbiology 8:128-133.

Dean, B. B. 1994. Managing the Potato Production System: 0734. CRC Press.

Decho, A. W., Norman, R. S., and Visscher, P. T. 2010. Quorum sensing in natural

environments: emerging views from microbial mats. Trends in Microbiology 18:73-80.

Dees, M. W., and Wanner, L. A. 2012. In search of better management of potato common scab.

Potato Research 55:249-268.

Page 146: Characterization of Dickeya dianthicola and Pectobacterium

131

Dees, M. W., Lysøe, E., Rossmann, S., Perminow, J., and Brurberg, M. B. 2017. Pectobacterium

polaris sp. nov., isolated from potato (Solanum tuberosum). International Journal of

Systematic and Evolutionary Microbiology 67:5222-5229.

Dennis, P. G., Miller, A. J., and Hirsch, P. R. 2010. Are root exudates more important than other

sources of rhizodeposits in structuring rhizosphere bacterial communities? FEMS

Microbiology Ecology 72:313-327.

Des Essarts, Y. R., Cigna, J., Quêtu-Laurent, A., Caron, A., Munier, E., Beury-Cirou, A., Hélias,

V., and Faure, D. 2016. Biocontrol of the potato blackleg and soft rot diseases caused by

Dickeya dianthicola. Applied and Environmental Microbiology 82:268-278.

Dhanvantari, B. N., and Dirks, V. A. 1987. Bacterial stem rot of greenhouse tomato: etiology,

spatial distribution, and the effect of high humidity. Phytopathology 77:1457-1463.

Dittoe, D. K., Barabote, R. D., Rothrock, M. J., and Ricke, S. C. 2020. Assessment of a Potential

Role of Dickeya dadantii DSM 18020 as a Pectinase Producer for Utilization in Poultry

Diets Based on in silico Analyses. Frontiers in Microbiology 11:751.

Donnelly, J. S. 2002. Great Irish potato famine. The History Press.

Duan, K., Dammel, C., Stein, J., Rabin, H., and Surette, M. G. 2003. Modulation of

Pseudomonas aeruginosa gene expression by host microflora through interspecies

communication. Molecular Microbiology 50:1477-1491.

Duarte, V., De Boer, S. H., Ward, L. J., and De Oliveira, A. M. R. 2004. Characterization of

atypical Erwinia carotovora strains causing blackleg of potato in Brazil. Journal of

Applied Microbiology 96:535-545.

Dung, J. K. S., Johnson, D. A., and Schroeder, B. K. 2012. First report of Pectobacterium

wasabiae causing aerial stem rot of potato in Washington State. Plant Disease 96:1819-

1819.

Dung, J. K. S., Johnson, D. A., and Schroeder, B. K. 2014. Role of co‐infection by

Pectobacterium and Verticillium dahliae in the development of early dying and aerial

stem rot of Russet Burbank potato. Plant Pathology 63:299-307.

Duprey, A., Taib, N., Leonard, S., Garin, T., Flandrois, J. P., Nasser, W., Brochier‐Armanet, C.,

and Reverchon, S. 2019. The phytopathogenic nature of Dickeya aquatica 174/2 and the

Page 147: Characterization of Dickeya dianthicola and Pectobacterium

132

dynamic early evolution of Dickeya pathogenicity. Environmental Microbiology

21:2809-2835.

Egland, P. G., Palmer, R. J., and Kolenbrander, P. E. 2004. Interspecies communication in

Streptococcus gordonii–Veillonella atypica biofilms: signaling in flow conditions

requires juxtaposition. Proceedings of the National Academy of Sciences 101:16917-

16922.

Ekbataniamiri, F. 2020. Potato Resistance and Surface Water Associated with Blackleg and Soft

Rot. The University of Maine, Orono.

Elphinstone, J. G., and Pérombelon, M. C. M. 1986. Contamination of potatoes by Erwinia

carotovora during grading. Plant Pathology 35:25-33.

Elphinstone, J. G., Parkinson, N. M., Saddler, G. S., Cahill, G., Kelly, R., Pritchard, L.,

Humphris, S., and Toth, I. 2014. Biology and control of Dickeya spp. affecting potato in

the UK. Proceedings of the Dundee conference Crop Protection in Northern Britain:235-

241.

F. A. O., T. 2008. International Year of the Potato: FAO Factsheets. Markets Division.

Farran, I., and Mingo-Castel, A. M. 2006. Potato minituber production using aeroponics: effect

of plant density and harvesting intervals. American Journal of Potato Research 83:47-53.

Fierer, N. 2017. Embracing the unknown: disentangling the complexities of the soil microbiome.

Nature Reviews Microbiology 15:579-590.

Fiers, M., Edel-Hermann, V., Chatot, C., Le Hingrat, Y., Alabouvette, C., and Steinberg, C.

2012. Potato soil-borne diseases. A review. Agronomy for Sustainable Development

32:93-132.

Fitt, B. D. L., Huang, Y.-J., Bosch, F. v. d., and West, J. S. 2006. Coexistence of related

pathogen species on arable crops in space and time. Annual Review of Phytopathology

44:163-182.

Frank, J. A., Reich, C. I., Sharma, S., Weisbaum, J. S., Wilson, B. A., and Olsen, G. J. 2008.

Critical evaluation of two primers commonly used for amplification of bacterial 16S

rRNA genes. Applied and Environmental Microbiology 74:2461-2470.

Page 148: Characterization of Dickeya dianthicola and Pectobacterium

133

Fronzes, R., Schäfer, E., Wang, L., Saibil, H. R., Orlova, E. V., and Waksman, G. 2009.

Structure of a type IV secretion system core complex. Science 323:266-268.

Gallois, A., Samson, R., Ageron, E., and Grimont, P. A. D. 1992. Erwinia carotovora subsp.

odorifera subsp. nov., associated with odorous soft rot of chicory (Cichorium intybus L.).

International Journal of Systematic and Evolutionary Microbiology 42:582-588.

Garbaye, J. 1994. Tansley review no. 76 helper bacteria: a new dimension to the mycorrhizal

symbiosis. New Phytologist 128:197-210.

Gardan, L., Gouy, C., Christen, R., and Samson, R. 2003. Elevation of three subspecies of

Pectobacterium carotovorum to species level: Pectobacterium atrosepticum sp. nov.,

Pectobacterium betavasculorum sp. nov. and Pectobacterium wasabiae sp. nov.

International Journal of Systematic and Evolutionary Microbiology 53:381-391.

Garge, S. S., and Nerurkar, A. S. 2017. Evaluation of quorum quenching Bacillus spp. for their

biocontrol traits against Pectobacterium carotovorum subsp. carotovorum causing soft

rot. Biocatalysis and Agricultural Biotechnology 9:48-57.

Ge, T., Marangoni, N. F., Hao, J., Johnson, S., and Larkin, R. P. 2018a. Isolation and

Identification of Bacteria Causing Blackleg and Soft Rot of Potato in Northeastern US.

Phytopathology 108:62.

Ge, T., Johnson, S., Larkin, R. P., Charkowski, A. O., and Hao, J. 2020a. Pathogen synergism of

blackleg disease on potato. Plant Health 2020 Online.

Ge, T., Jiang, H., Johnson, S. B., Larkin, R., Charkowski, A. O., Secor, G., and Hao, J. 2020b.

Genotyping Dickeya dianthicola causing potato blackleg and soft rot outbreak associated

with inoculum geography in the United States. Plant Disease.

https://doi.org/10.1094/PDIS-10-20-2138-RE

Ge, T., Jiang, H., Hao, J., and Johnson, S. B. 2018b. First report of Pectobacterium parmentieri

causing bacterial soft rot and blackleg on potato in Maine. Plant Disease 102:437.

Glasner, J. D., Marquez-Villavicencio, M., Kim, H. S., Jahn, C. E., Ma, B., Biehl, B. S.,

Rissman, A. I., Mole, B., Yi, X., and Yang, C. H. 2008. Niche-specificity and the

variable fraction of the Pectobacterium pan-genome. Molecular Plant-Microbe

Interactions 21:1549-1560.

Page 149: Characterization of Dickeya dianthicola and Pectobacterium

134

Goto, M., and Matsumoto, K. 1987. Erwinia carotovora subsp. wasabiae subsp. nov. isolated

from diseased rhizomes and fibrous roots of Japanese horseradish (Eutrema wasabi

Maxim.). International Journal of Systematic and Evolutionary Microbiology 37:130-135.

Goumans, D. E., and Chatzaki, A. K. 1998. Characterization and host range evaluation of

Pseudomonas viridiflava from melon, blite, tomato, chrysanthemum and eggplant.

European Journal of Plant Pathology 104:181-188.

Greninger, A. L., Naccache, S. N., Federman, S., Yu, G., Mbala, P., Bres, V., Stryke, D.,

Bouquet, J., Somasekar, S., and Linnen, J. M. 2015. Rapid metagenomic identification of

viral pathogens in clinical samples by real-time nanopore sequencing analysis. Genome

Medicine 7:99.

Grube, M., Fürnkranz, M., Zitzenbacher, S., Huss, H., and Berg, G. 2011. Emerging multi-

pathogen disease caused by Didymella bryoniae and pathogenic bacteria on Styrian oil

pumpkin. European Journal of Plant Pathology 131:539.

Gudmestad, N. C., Taylor, R. J., and Pasche, J. S. 2007. Management of soilborne diseases of

potato. Australasian Plant Pathology 36:109-115.

Guttman, D. S., McHardy, A. C., and Schulze-Lefert, P. 2014. Microbial genome-enabled

insights into plant–microorganism interactions. Nature Reviews Genetics 15:797-813.

Gyles, C., and Boerlin, P. 2014. Horizontally transferred genetic elements and their role in

pathogenesis of bacterial disease. Veterinary Pathology 51:328-340.

Ham, J. H., Bauer, D. W., Fouts, D. E., and Collmer, A. 1998. A cloned Erwinia chrysanthemi

Hrp (type III protein secretion) system functions in Escherichia coli to deliver

Pseudomonas syringae Avr signals to plant cells and to secrete Avr proteins in culture.

Proceedings of the National Academy of Sciences 95:10206-10211.

Ham, J. H., Cui, Y., Alfano, J. R., Rodríguez-Palenzuela, P., Rojas, C. M., Chatterjee, A. K., and

Collmer, A. 2004. Analysis of Erwinia chrysanthemi EC16 pelE∷ uidA, pelL∷ uidA, and

hrpN∷ uidA mutants reveals strain-specific atypical regulation of the Hrp type III

secretion system. Molecular Plant-Microbe Interactions 17:184-194.

Haney, C. H., Samuel, B. S., Bush, J., and Ausubel, F. M. 2015. Associations with rhizosphere

bacteria can confer an adaptive advantage to plants. Nature Plants 1:1-9.

Page 150: Characterization of Dickeya dianthicola and Pectobacterium

135

Hao, J., Jiang, H., and Johnson, S. 2016. Detection and characterization of Dickeya species in the

outbreak of blackleg disease of potato in Maine. Phytopathology 106:2.

Harrison, J. G., Searle, R. J., and Williams, N. A. 1997. Powdery scab disease of potato-a

review. Plant Pathology 46:1-25.

Hartmann, A., Schmid, M., Van Tuinen, D., and Berg, G. 2009. Plant-driven selection of

microbes. Plant and Soil 321:235-257.

Hauben, L., Moore, E. R. B., Vauterin, L., Steenackers, M., Mergaert, J., Verdonck, L., and

Swings, J. 1998. Phylogenetic position of phytopathogens within the Enterobacteriaceae.

Systematic and Applied Microbiology 21:384-397.

Hellmers, E. 1958. Four wilt diseases of perpetual-flowering Carnations in Denmark. Dansk

Botanisk Arkiv 18.

Hibbing, M. E., Fuqua, C., Parsek, M. R., and Peterson, S. B. 2010. Bacterial competition:

surviving and thriving in the microbial jungle. Nature Reviews Microbiology 8:15-25.

Hiergeist, A., Gläsner, J., Reischl, U., and Gessner, A. 2015. Analyses of intestinal microbiota:

culture versus sequencing. ILAR journal 56:228-240.

Hijmans, R. J. 2001. Global distribution of the potato crop. American journal of potato research

78:403-412.

Hoek, T. A., Axelrod, K., Biancalani, T., Yurtsev, E. A., Liu, J., and Gore, J. 2016. Resource

availability modulates the cooperative and competitive nature of a microbial cross-

feeding mutualism. PLoS Biology 14:e1002540.

Huerta-Cepas, J., Szklarczyk, D., Heller, D., Hernández-Plaza, A., Forslund, S. K., Cook, H.,

Mende, D. R., Letunic, I., Rattei, T., and Jensen, L. J. 2019. eggNOG 5.0: a hierarchical,

functionally and phylogenetically annotated orthology resource based on 5090 organisms

and 2502 viruses. Nucleic Acids Research 47:D309-D314.

Hélias, V., Hamon, P., Huchet, E., Wolf, J. V. D., and Andrivon, D. 2012. Two new effective

semiselective crystal violet pectate media for isolation of Pectobacterium and Dickeya.

Plant Pathology 61:339-345.

Page 151: Characterization of Dickeya dianthicola and Pectobacterium

136

James, E. K., and Olivares, F. L. 1998. Infection and colonization of sugar cane and other

graminaceous plants by endophytic diazotrophs. Critical Reviews in Plant Sciences

17:77-119.

Jiang, H. H., Hao, J. J., Johnson, S. B., Brueggeman, R. S., and Secor, G. 2016. First report of

Dickeya dianthicola causing blackleg and bacterial soft rot on potato in Maine. Plant

Disease 100:2320-2320.

Johnson, D. A. 2008. Potato health management. American Phytopathological Society (APS

Press).

Johnson, S. 2016. Dickeya, a new potato pathogen in Maine and elsewhere. Phytopathology

106:3.

Johnson, S. B. 2017. Novel Approach for Identification of Potato Seed Lots Suspected to be

Effected by Dickeya. Plant Pathology Journal 16:96-100.

Jones, J. B., Jackson, L. E., Balogh, B., Obradovic, A., Iriarte, F. B., and Momol, M. T. 2007.

Bacteriophages for plant disease control. Annual Review of Phytopathology 45:245-262.

Jones, R. A. C., and Harrison, B. D. 1969. The behaviour of potato mop‐top virus in soil, and

evidence for its transmission by Spongospora subterranea (Wallr.) Lagerh. Annals of

Applied Biology 63:1-17.

Juck, D., Charles, T., Whyte, L. G., and Greer, C. W. 2000. Polyphasic microbial community

analysis of petroleum hydrocarbon-contaminated soils from two northern Canadian

communities. FEMS Microbiology Ecology 33:241-249.

Kastelein, P., Schepel, E. G., Mulder, A., Turkensteen, L. J., and Van Vuurde, J. W. L. 1999.

Preliminary selection of antagonists of Erwinia carotovora subsp. atroseptica (Van Hall)

Dye for application during green crop lifting of seed potato tubers. Potato Research

42:161-171.

Kent, A. D., and Triplett, E. W. 2002. Microbial communities and their interactions in soil and

rhizosphere ecosystems. Annual Reviews in Microbiology 56:211-236.

Khayi, S., Cigna, J., Chong, T. M., Quêtu-Laurent, A., Chan, K.-G., Hélias, V., and Faure, D.

2016. Transfer of the potato plant isolates of Pectobacterium wasabiae to Pectobacterium

Page 152: Characterization of Dickeya dianthicola and Pectobacterium

137

parmentieri sp. nov. International Journal of Systematic and Evolutionary Microbiology

66:5379-5383.

Kim, H.-S., Ma, B., Perna, N. T., and Charkowski, A. O. 2009. Phylogeny and virulence of

naturally occurring type III secretion system-deficient Pectobacterium strains. Applied

and Environmental Microbiology 75:4539-4549.

Kim, H. J., Boedicker, J. Q., Choi, J. W., and Ismagilov, R. F. 2008. Defined spatial structure

stabilizes a synthetic multispecies bacterial community. Proceedings of the National

Academy of Sciences 105:18188-18193.

Kim, M. H., Cho, M. S., Kim, B. K., Choi, H. J., Hahn, J. H., Kim, C., Kang, M. J., Kim, S. H.,

and Park, D. S. 2012. Quantitative real-time polymerase chain reaction assay for

detection of Pectobacterium wasabiae using YD repeat protein gene-based primers. Plant

Disease 96:253-257.

Koeuth, T., Versalovic, J., and Lupski, J. R. 1995. Differential subsequence conservation of

interspersed repetitive Streptococcus pneumoniae BOX elements in diverse bacteria.

Genome Research 5:408-418.

Koskiniemi, S., Lamoureux, J. G., Nikolakakis, K. C., de Roodenbeke, C. t. K., Kaplan, M. D.,

Low, D. A., and Hayes, C. S. 2013. Rhs proteins from diverse bacteria mediate

intercellular competition. Proceedings of the National Academy of Sciences 110:7032-

7037.

Kranz, A., Vogel, A., Degner, U., Kiefler, I., Bott, M., Usadel, B., and Polen, T. 2017. High

precision genome sequencing of engineered Gluconobacter oxydans 621H by combining

long nanopore and short accurate Illumina reads. Journal of Biotechnology 258:197-205.

Kumar, S., Stecher, G., and Tamura, K. 2016. MEGA7: molecular evolutionary genetics analysis

version 7.0 for bigger datasets. Molecular Biology and Evolution 33:1870-1874.

Kůdela, V., Krejzar, V., and PáNkoVá, I. 2011. Pseudomonas corrugata and Pseudomonas

marginalis associated with the collapse of tomato plants in rockwool slab hydroponic

culture. Plant Protection Science 46:1-11.

Ladygina, N., and Hedlund, K. 2010. Plant species influence microbial diversity and carbon

allocation in the rhizosphere. Soil Biology and Biochemistry 42:162-168.

Page 153: Characterization of Dickeya dianthicola and Pectobacterium

138

Lamichhane, J. R., and Venturi, V. 2015. Synergisms between microbial pathogens in plant

disease complexes: a growing trend. Frontiers in Plant Science 6:385.

Lanoot, B., Vancanneyt, M., Dawyndt, P., Cnockaert, M., Zhang, J., Huang, Y., Liu, Z., and

Swings, J. 2004. BOX-PCR fingerprinting as a powerful tool to reveal synonymous

names in the genus Streptomyces. Emended descriptions are proposed for the species

Streptomyces cinereorectus, S. fradiae, S. tricolor, S. colombiensis, S. filamentosus, S.

vinaceus and S. phaeopurpureus. Systematic and Applied Microbiology 27:84-92.

Lapierre, P., and Gogarten, J. P. 2009. Estimating the size of the bacterial pan-genome. Trends in

Genetics 25:107-110.

Lelliott, R. A., and Dickey, R. S. 1984. Genus VII, Erwinia spp.

Letunic, I., and Bork, P. 2019. Interactive Tree Of Life (iTOL) v4: recent updates and new

developments. Nucleic Acids Research 47:W256-W259.

Li, B., Shi, Y., Ibrahim, M., Liu, H., Shan, C., Wang, Y., Kube, M., Xie, G.-L., and Sun, G.

2012. Genome sequence of the rice pathogen Dickeya zeae strain ZJU1202. Journal of

Bacteriology 4452-4453.

Li, X., Ma, Y., Liang, S., Tian, Y., Yin, S., Xie, S., and Xie, H. 2018. Comparative genomics of

84 Pectobacterium genomes reveals the variations related to a pathogenic lifestyle. BMC

Genomics 19:889.

Lin, G., Chai, J., Yuan, S., Mai, C., Cai, L., Murphy, R. W., Zhou, W., and Luo, J. 2016.

VennPainter: a tool for the comparison and identification of candidate genes based on

venn diagrams. PLoS One 11:e0154315.

Liu, W., Xie, Y., Ma, J., Luo, X., Nie, P., Zuo, Z., Lahrmann, U., Zhao, Q., Zheng, Y., and Zhao,

Y. 2015. IBS: an illustrator for the presentation and visualization of biological sequences.

Bioinformatics 31:3359-3361.

Luo, R., Liu, B., Xie, Y., Li, Z., Huang, W., Yuan, J., He, G., Chen, Y., Pan, Q., and Liu, Y.

2012. SOAPdenovo2: an empirically improved memory-efficient short-read de novo

assembler. Gigascience 1:2047-2217X.

López-García, A., Pineda-Quiroga, C., Atxaerandio, R., Pérez, A., Hernández, I., García-

Rodríguez, A., and González-Recio, O. 2018. Comparison of Mothur and QIIME for the

Page 154: Characterization of Dickeya dianthicola and Pectobacterium

139

analysis of rumen microbiota composition based on 16S rRNA amplicon sequences.

Frontiers in Microbiology 9:3010.

Ma, B., Hibbing, M. E., Kim, H.-S., Reedy, R. M., Yedidia, I., Breuer, J., Breuer, J., Glasner, J.

D., Perna, N. T., and Kelman, A. 2007. Host range and molecular phylogenies of the soft

rot enterobacterial genera Pectobacterium and Dickeya. Phytopathology 97:1150-1163.

Ma, X., Schloop, A., Swingle, B., and Perry, K. L. 2018. Pectobacterium and Dickeya

responsible for potato blackleg disease in New York State in 2016. Plant disease

102:1834-1840.

Ma, X., Perna, N. T., Glasner, J. D., Hao, J., Johnson, S., Nasaruddin, A. S., Charkowski, A. O.,

Wu, S., Fei, Z., and Perry, K. L. 2019. Complete Genome Sequence of Dickeya

dianthicola ME23, a Pathogen Causing Blackleg and Soft Rot Diseases of Potato.

Microbiol Resource Announcment 8:e01526-01518.

Malathrakis, N. E., and Goumas, D. E. 1987. Bacterial soft rot of tomato in plastic greenhouses

in Crete. Annals of applied Biology 111:115-123.

Mao, D.-P., Zhou, Q., Chen, C.-Y., and Quan, Z.-X. 2012. Coverage evaluation of universal

bacterial primers using the metagenomic datasets. BMC microbiology 12:66.

Mardis, E. R. 2008. Next-generation DNA sequencing methods. Annu. Rev. Genomics Human

Genetics 9:387-402.

Mardis, E. R. 2017. DNA sequencing technologies: 2006–2016. Nature protocols 12:213.

Marković, S., Stanković, S., Jelušić, A., Ilicic, R., Kosovac, A., Poštić, D., and Popović, T. 2020.

Occurrence and Identification of Pectobacterium carotovorum subsp. brasiliensis and

Dickeya dianthicola causing blackleg in some potato fields in Serbia. Plant Disease.

PDIS-05.

Marquez-Villavicencio, M. d. P., Groves, R. L., and Charkowski, A. O. 2011. Soft rot disease

severity is affected by potato physiology and Pectobacterium taxa. Plant Disease 95:232-

241.

Mattinen, L., Somervuo, P., Nykyri, J., Nissinen, R., Kouvonen, P., Corthals, G., Auvinen, P.,

Aittamaa, M., Valkonen, J. P. T., and Pirhonen, M. 2008. Microarray profiling of host-

Page 155: Characterization of Dickeya dianthicola and Pectobacterium

140

extract-induced genes and characterization of the type VI secretion cluster in the potato

pathogen Pectobacterium atrosepticum. Microbiology 154:2387-2396.

Mazzaglia, A., Studholme, D. J., Taratufolo, M. C., Cai, R., Almeida, N. F., Goodman, T.,

Guttman, D. S., Vinatzer, B. A., and Balestra, G. M. 2012. Pseudomonas syringae pv.

actinidiae (PSA) isolates from recent bacterial canker of kiwifruit outbreaks belong to the

same genetic lineage. PloS One 7: e36518.

McGuire, R. G., and Kelman, A. 1984. Reduced severity of Erwinia soft rot in potato tubers with

increased calcium content. Phytopathology 74:1250-1256.

McNally, R. R., Curland, R. D., Webster, B. T., Robinson, A. P., and Ishimaru, C. A. 2017. First

report of blackleg and tuber soft rot of potato caused by Pectobacterium parmentieri in

Minnesota and North Dakota. Plant Disease 101:2144.

McNally, R. R., Curland, R. D., Webster, B. T., Robinson, A. P., and Ishimaru, C. A. 2018. First

report of stem rot on potato caused by Dickeya chrysanthemi in Minnesota. Plant Disease

102:238.

Mederma, G. J., Schets, F. M., Van de Giessen, A. W., and Havelaar, A. H. 1992. Lack of

colonization of 1 day old chicks by viable, non‐culturable Campylobacter jejuni. Journal

of Applied Bacteriology 72:512-516.

Medini, D., Donati, C., Tettelin, H., Masignani, V., and Rappuoli, R. 2005. The microbial pan-

genome. Current Opinion in Genetics & Development 15:589-594.

Mendes, R., Garbeva, P., and Raaijmakers, J. M. 2013. The rhizosphere microbiome:

significance of plant beneficial, plant pathogenic, and human pathogenic microorganisms.

FEMS Microbiology Reviews 37:634-663.

Mendes, R., Kruijt, M., De Bruijn, I., Dekkers, E., van der Voort, M., Schneider, J. H. M.,

Piceno, Y. M., DeSantis, T. Z., Andersen, G. L., and Bakker, P. A. H. M. 2011.

Deciphering the rhizosphere microbiome for disease-suppressive bacteria. Science

332:1097-1100.

Mikheenko, A., Prjibelski, A., Saveliev, V., Antipov, D., and Gurevich, A. 2018. Versatile

genome assembly evaluation with QUAST-LG. Bioinformatics 34:i142-i150.

Page 156: Characterization of Dickeya dianthicola and Pectobacterium

141

Mina, D., Pereira, J. A., Lino-Neto, T., and Baptista, P. 2020. Impact of plant genotype and plant

habitat in shaping bacterial pathobiome: a comparative study in olive tree. Scientific

Reports 10:1-11.

Monod, J. 1949. The growth of bacterial cultures. Annual Review of Microbiology 3:371-394.

Monteiro, R. A., Balsanelli, E., Wassem, R., Marin, A. M., Brusamarello-Santos, L. C. C.,

Schmidt, M. A., Tadra-Sfeir, M. Z., Pankievicz, V. C. S., Cruz, L. M., and Chubatsu, L.

S. 2012. Herbaspirillum-plant interactions: microscopical, histological and molecular

aspects. Plant and Soil 356:175-196.

Motyka-Pomagruk, A., Zoledowska, S., Misztak, A. E., Sledz, W., Mengoni, A., and Lojkowska,

E. 2020. Comparative genomics and pangenome-oriented studies reveal high

homogeneity of the agronomically relevant enterobacterial plant pathogen Dickeya

solani. BMC Genomics 21:1-18.

Moura, M. L., Jacques, M. A., Brito, L. M., Mourão, I. M., and Duclos, J. 2004. Tomato pith

necrosis (TPN) caused by P. corrugata and P. mediterranea: severity of damages and

crop loss assessment. I International Symposium on Tomato Diseases 695:365-372.

Mousa, W. K., Shearer, C., Limay-Rios, V., Ettinger, C. L., Eisen, J. A., and Raizada, M. N.

2016. Root-hair endophyte stacking in finger millet creates a physicochemical barrier to

trap the fungal pathogen Fusarium graminearum. Nature Microbiology 1:1-12.

Mousa, W. K., Schwan, A., Davidson, J., Strange, P., Liu, H., Zhou, T., Auzanneau, F.-I., and

Raizada, M. N. 2015. An endophytic fungus isolated from finger millet (Eleusine

coracana) produces anti-fungal natural products. Frontiers in Microbiology 6:1157.

Murray, J. L., Connell, J. L., Stacy, A., Turner, K. H., and Whiteley, M. 2014. Mechanisms of

synergy in polymicrobial infections. Journal of Microbiology 52:188-199.

Na, R., and Gijzen, M. 2016. Escaping host immunity: new tricks for plant pathogens. PLoS

Pathogens 12:e1005631.

Nasaruddin, A. S., Charkowski, A. O., Babler, B. N., Perna, N. T., and Glasner, J. D. 2019. First

report of Dickeya dianthicola causing blackleg on potato in Texas. Plant Disease

103:2121.

NASS. 2018. Maine potato size and grade report. New England Field Office.

Page 157: Characterization of Dickeya dianthicola and Pectobacterium

142

Nassar, A., Bertheau, Y., Dervin, C., Narcy, J.-P., and Lemattre, M. 1994. Ribotyping of Erwinia

chrysanthemi strains in relation to their pathogenic and geographic distribution. Applied

and Environmental Microbiology 60:3781-3789.

Nassar, A., Darrasse, A., Lemattre, M., Kotoujansky, A., Dervin, C., Vedel, R., and Bertheau, Y.

1996. Characterization of Erwinia chrysanthemi by pectinolytic isozyme polymorphism

and restriction fragment length polymorphism analysis of PCR-amplified fragments of

pel genes. Applied and Environment Microbiology 62:2228-2235.

Ngadze, E., Coutinho, T. A., and Van der Waals, J. E. 2010. First report of soft rot of potatoes

caused by Dickeya dadantii in Zimbabwe. Plant Disease 94:1263-1263.

Ngadze, E., Brady, C. L., Coutinho, T. A., and Van der Waals, J. E. 2012. Pectinolytic bacteria

associated with potato soft rot and blackleg in South Africa and Zimbabwe. European

Journal of Plant Pathology 134:533-549.

Nykyri, J., Niemi, O., Koskinen, P., Nokso-Koivisto, J., Pasanen, M., Broberg, M., Plyusnin, I.,

Törönen, P., Holm, L., and Pirhonen, M. 2012. Revised phylogeny and novel horizontally

acquired virulence determinants of the model soft rot phytopathogen Pectobacterium

wasabiae SCC3193. PLoS Pathogens 8: e1003013.

Ojala, J. C., Stark, J. C., and Kleinkopf, G. E. 1990. Influence of irrigation and nitrogen

management on potato yield and quality. American Potato Journal 67:29-43.

Oulghazi, S., Khayi, S., Lafkih, N., Massaoudi, Y., El Karkouri, A., El Hassouni, M., Faure, D.,

and Moumni, M. 2017. First report of Dickeya dianthicola causing blackleg disease on

potato in Morocco. Plant Disease 101:1671.

Page, A. J., Cummins, C. A., Hunt, M., Wong, V. K., Reuter, S., Holden, M. T. G., Fookes, M.,

Falush, D., Keane, J. A., and Parkhill, J. 2015. Roary: rapid large-scale prokaryote pan

genome analysis. Bioinformatics 31:3691-3693.

Pagel, W., and Heitefuss, R. 1989. Calcium content and cell wall polygalacturonans in potato

tubers of cultivars with different susceptibilities to Erwinia carotovora subsp.

atroseptica. Physiological and Molecular Plant Pathology 35:11-21.

Parkinson, N., DeVos, P., Pirhonen, M., and Elphinstone, J. 2014. Dickeya aquatica sp. nov.,

isolated from waterways. International Journal of Systematic and Evolutionary

Microbiology 64:2264-2266.

Page 158: Characterization of Dickeya dianthicola and Pectobacterium

143

Parkinson, N., Pritchard, L., Bryant, R., Toth, I., and Elphinstone, J. 2015. Epidemiology of

Dickeya dianthicola and Dickeya solani in ornamental hosts and potato studied using

variable number tandem repeat analysis. European Journal of Plant Pathology 141:63-70.

Parkinson, N., Stead, D., Bew, J., Heeney, J., Tsror, L., and Elphinstone, J. 2009. Dickeya

species relatedness and clade structure determined by comparison of recA sequences.

International Journal of Systematic and Evolutionary Microbiology 59:2388-2393.

Partida-Martinez, L. P., and Hertweck, C. 2005. Pathogenic fungus harbours endosymbiotic

bacteria for toxin production. Nature 437:884-888.

Partida-Martinez, L. P. P., and Heil, M. 2011. The microbe-free plant: fact or artifact? Frontiers

in Plant Science 2:100.

Patel, N., Baldwin, A. C., Patel, R. D., Kobayashi, D. Y., and Wyenandt, C. A. 2019. First report

of Dickeya dianthicola causing blackleg and soft rot on potato (Solanum tuberosum) in

New Jersey, USA. Plant Disease 103:146.

Paula Wilkie, J., and Dye, D. W. 1974. Pseudomonas cichorii causing tomato and celery

diseases in New Zealand. New Zealand Journal of Agricultural Research 17:123-130.

Perombelon, M. C. M. 2000. Blackleg risk potential of seed potatoes determined by

quantification of tuber contamination by the causal agent and Erwinia carotovora subsp.

atroseptica: a critical review. EPPO Bulletin 30:413-420.

Perombelon, M. C. M., and Kelman, A. 1980. Ecology of the soft rot erwinias. Annual Review

of Phytopathology 18:361-387.

Perombelon, M. C. M., and Van Der Wolf, J. M. 2002. Methods for the detection and

quantification of Erwinia carotovora subsp. atroseptica (Pectobacterium carotovorum

subsp. atrosepticum) on potatoes: a laboratory manual. Scottish Crop Research Institute

Annual Report 10.

Petersen, L. M., Martin, I. W., Moschetti, W. E., Kershaw, C. M., and Tsongalis, G. J. 2019.

Third-generation sequencing in the clinical laboratory: exploring the advantages and

challenges of nanopore sequencing. Journal of Clinical Microbiology 58.

Pinto, D., Santos, M. A., and Chambel, L. 2015. Thirty years of viable but nonculturable state

research: unsolved molecular mechanisms. Critical Reviews in Microbiology 41:61-76.

Page 159: Characterization of Dickeya dianthicola and Pectobacterium

144

Pitman, A. R., Wright, P. J., Galbraith, M. D., and Harrow, S. A. 2008. Biochemical and genetic

diversity of pectolytic enterobacteria causing soft rot disease of potatoes in New Zealand.

Australasian Plant Pathology 37:559-568.

Pritchard, L., Humphris, S., Saddler, G. S., Parkinson, N. M., Bertrand, V., Elphinstone, J. G.,

and Toth, I. K. 2013. Detection of phytopathogens of the genus Dickeya using a PCR

primer prediction pipeline for draft bacterial genome sequences. Plant Pathology 62:587-

596.

Pu, X. M., Zhou, J. N., Lin, B. R., and Shen, H. F. 2012. First report of bacterial foot rot of rice

caused by a Dickeya zeae in China. Plant Disease 96:1818.

Pérez‐Mendoza, D., Coulthurst, S. J., Humphris, S., Campbell, E., Welch, M., Toth, I. K., and

Salmond, G. P. C. 2011. A multi‐repeat adhesin of the phytopathogen, Pectobacterium

atrosepticum, is secreted by a Type I pathway and is subject to complex regulation

involving a non‐canonical diguanylate cyclase. Molecular Microbiology 82:719-733.

Pérombelon, M. C. M. 1992. Potato blackleg: epidemiology, host-pathogen interaction and

control. Netherlands Journal of Plant Pathology 98:135-146.

Pérombelon, M. C. M., Fox, R. A., and Lowe, R. 1979. Dispersion of Erwinia carotovora in

aerosols produced by the pulverization of potato haulm prior to harvest.

Phytopathologische Zeitschrift 94:249-260.

Põllumaa, L., Alamäe, T., and Mäe, A. 2012. Quorum sensing and expression of virulence in

pectobacteria. Sensors 12:3327-3349.

Radcliffe, E. B., and Ragsdale, D. W. 2002. Aphid-transmitted potato viruses: the importance of

understanding vector biology. American Journal of Potato Research 79:353-386.

Rahmanifar, B., Hasanzadeh, N., Razmi, J., and Ghasemi, A. 2012. Genetic diversity of Iranian

potato soft rot bacteria based on polymerase chain reaction-restriction fragment length

polymorphism (PCR-RFLP) analysis. African Journal of Biotechnology 11:1314-1320.

Ranganna, B., Kushalappa, A. C., and Raghavan, G. S. V. 1997. Ultraviolet irradiance to control

dry rot and soft rot of potato in storage. Canadian Journal of Plant Pathology 19:30-35.

Rankin, D. J., Bargum, K., and Kokko, H. 2007. The tragedy of the commons in evolutionary

biology. Trends in Ecology & Evolution 22:643-651.

Page 160: Characterization of Dickeya dianthicola and Pectobacterium

145

Rantakari, A., Virtaharju, O., Vähämiko, S., Taira, S., Palva, E. T., Saarilahti, H. T., and

Romantschuk, M. 2001. Type III secretion contributes to the pathogenesis of the soft-rot

pathogen Erwinia carotovora: partial characterization of the hrp gene cluster. Molecular

Plant-Microbe Interactions 14:962-968.

Reinhold-Hurek, B., Bünger, W., Burbano, C. S., Sabale, M., and Hurek, T. 2015. Roots shaping

their microbiome: global hotspots for microbial activity. Annual Review of

Phytopathology 53:403-424.

Reverchon, S., Muskhelisvili, G., and Nasser, W. 2016. Virulence program of a bacterial plant

pathogen: the Dickeya model. Pages 51-92 in: Progress in molecular biology and

translational science. Elsevier 142.

Ritpitakphong, U., Falquet, L., Vimoltust, A., Berger, A., Métraux, J. P., and L'Haridon, F. 2016.

The microbiome of the leaf surface of Arabidopsis protects against a fungal pathogen.

New Phytologist 210:1033-1043.

Robinson, K., and Foster, G. 1987. Control of potato blackleg by tuber pasteurisation: the

determination of time-temperature combinations for the inactivation of pectolytic

erwinia. Potato Research 30:121-125.

Rodríguez-Beltrán, J., Sørum, V., Toll-Riera, M., de la Vega, C., Peña-Miller, R., and San

Millán, Á. 2020. Genetic dominance governs the evolution and spread of mobile genetic

elements in bacteria. Proceedings of the National Academy of Sciences 117:15755-

15762.

Rojas, C. M., Ham, J. H., Deng, W.-L., Doyle, J. J., and Collmer, A. 2002. HecA, a member of a

class of adhesins produced by diverse pathogenic bacteria, contributes to the attachment,

aggregation, epidermal cell killing, and virulence phenotypes of Erwinia chrysanthemi

EC16 on Nicotiana clevelandii seedlings. Proceedings of the National Academy of

Sciences 99:13142-13147.

Rolli, E., Marasco, R., Vigani, G., Ettoumi, B., Mapelli, F., Deangelis, M. L., Gandolfi, C.,

Casati, E., Previtali, F., and Gerbino, R. 2015. Improved plant resistance to drought is

promoted by the root‐associated microbiome as a water stress‐dependent trait.

Environmental Microbiology 17:316-331.

Rolot, J. L., and Seutin, H. 1999. Soilless production of potato minitubers using a hydroponic

technique. Potato Research 42:457-469.

Page 161: Characterization of Dickeya dianthicola and Pectobacterium

146

Rosenzweig, N., Steere, L., Kirk, W. W., Mambetova, S., Long, C., Schafer, R., Dangi, S., and

Byrne, J. 2016. First report of Dickeya dianthicola and Pectobacterium wasabiae causing

aerial stem rot of potato in Michigan, USA. New Disease Report 33:2044-0588.

Rossum, G. 1995. Python reference manual.

Rouli, L., Merhej, V., Fournier, P. E., and Raoult, D. 2015. The bacterial pangenome as a new

tool for analysing pathogenic bacteria. New Microbes and New Infections 7:72-85.

Russell, A. B., Peterson, S. B., and Mougous, J. D. 2014. Type VI secretion system effectors:

poisons with a purpose. Nature Reviews Microbiology 12:137-148.

Salaman, R. N., and Burton, W. G. 1985. The history and social influence of the potato.

Cambridge University Press.

Salmond, G. P. C. 1994. Secretion of extracellular virulence factors by plant pathogenic bacteria.

Annual Review of Phytopathology 32:181-200.

Samson, R., Legendre, J. B., Christen, R., Fischer-Le Saux, M., Achouak, W., and Gardan, L.

2005. Transfer of Pectobacterium chrysanthemi (Burkholder et al. 1953) Brenner et al.

1973 and Brenneria paradisiaca to the genus Dickeya gen. nov as Dickeya chrysanthemi

comb. nov and Dickeya paradisiaca comb. nov and delineation of four novel species,

Dickeya dadantii sp nov., Dickeya dianthicola sp nov., Dickeya dieffenbachiae sp nov

and Dickeya zeae sp nov. International Journal of Systematic and Evolutionary

Microbiology 55:1415-1427.

Sanger, F., and Coulson, A. R. 1975. A rapid method for determining sequences in DNA by

primed synthesis with DNA polymerase. Journal of Molecular Biology 94:441-448.

Sanger, F., Nicklen, S., and Coulson, A. R. 1977. DNA sequencing with chain-terminating

inhibitors. Proceedings of the National Academy of Sciences 74:5463-5467.

Sarfraz, S., Riaz, K., Oulghazi, S., Cigna, J., Alam, M. W., Dessaux, Y., and Faure, D. 2018.

First report of Dickeya dianthicola causing blackleg disease on potato plants in Pakistan.

Plant Disease 102: 2027-2028.

Saygili, H., Aysan, Y., Ustun, N., Mirik, M., and Sahin, F. 2008. Tomato pith necrosis disease

caused by Pseudomonas species in Turkey. Pages 357-366 in: Pseudomonas syringae

Pathovars and Related Pathogens–Identification, Epidemiology and Genomics. Springer.

Page 162: Characterization of Dickeya dianthicola and Pectobacterium

147

Scarlett, C. M., Fletcher, J. T., Roberts, P., and Lelliott, R. A. 1978. Tomato pith necrosis caused

by Pseudomonas corrugata n. sp. Annals of Applied Biology 88:105-114.

Schloss, P. D., Westcott, S. L., Ryabin, T., Hall, J. R., Hartmann, M., Hollister, E. B.,

Lesniewski, R. A., Oakley, B. B., Parks, D. H., and Robinson, C. J. 2009. Introducing

mothur: open-source, platform-independent, community-supported software for

describing and comparing microbial communities. Applied and Environment

Microbiology. 75:7537-7541.

Schoch, C. L., Seifert, K. A., Huhndorf, S., Robert, V., Spouge, J. L., Levesque, C. A., Chen, W.,

and Fungal Barcoding, C. 2012. Nuclear ribosomal internal transcribed spacer (ITS)

region as a universal DNA barcode marker for Fungi. Proceedings of the National

Academy of Sciences 109:6241-6246.

Scholthof, K.-B. G. 2007. The disease triangle: pathogens, the environment and society. Nature

Reviews Microbiology 5:152-156.

Seemann, T. 2014. Prokka: rapid prokaryotic genome annotation. Bioinformatics 30:2068-2069.

Shahbandeh, M. 2020. Trade value of potatoes to the United States from 2015 to 2019. Retrieved

from https://www.statista.com/statistics/816722/us-imports-and-exports-of-potatoes/

Shendure, J., and Ji, H. 2008. Next-generation DNA sequencing. Nature Biotechnology 26:1135-

1145.

Shirsat, S. G., Thomas, P., and Nair, P. M. 1991. Evaluation of treatments with hot water,

chemicals and ventilated containers to reduce microbial spoilage in irradiated potatoes.

Potato research 34:227-231.

Short, F. L., Murdoch, S. L., and Ryan, R. P. 2014. Polybacterial human disease: the ills of social

networking. Trends in Microbiology 22:508-516.

Singer, M. 2010. Pathogen-pathogen interaction: a syndemic model of complex biosocial

processes in disease. Virulence 1:10-18.

Singh, B., Bhardwaj, V., Kaur, K., Kukreja, S., and Goutam, U. 2020. Potato Periderm is the first

layer of defence against biotic and abiotic stresses: a review. Potato Research:1-16.

Page 163: Characterization of Dickeya dianthicola and Pectobacterium

148

Sinha, S. K., and Prasad, M. 1977. Bacterial stalk rot of maize, its symptoms and host-range.

Zentralblatt für Bakteriologie, Parasitenkunde, Infektionskrankheiten und Hygiene.

Zweite Naturwissenschaftliche Abteilung: Allgemeine, Landwirtschaftliche und

Technische Mikrobiologie 132:81-88.

Smolarska, A., Rabalski, L., Narajczyk, M., and Czajkowski, R. 2018. Isolation and phenotypic

and morphological characterization of the first Podoviridae lytic bacteriophages ϕA38

and ϕA41 infecting Pectobacterium parmentieri (former Pectobacterium wasabiae).

European Journal of Plant Pathology 150:413-425.

Soleimani-Delfan, A., Etemadifar, Z., Emtiazi, G., and Bouzari, M. 2015. Isolation of Dickeya

dadantii strains from potato disease and biocontrol by their bacteriophages. Brazilian

Journal of Microbiology 46:791-797.

Steen, A. D., Crits-Christoph, A., Carini, P., DeAngelis, K. M., Fierer, N., Lloyd, K. G., and

Thrash, J. C. 2019. High proportions of bacteria and archaea across most biomes remain

uncultured. The ISME journal 13:3126-3130.

Stevenson, W. R., Loria, R., Franc, G. D., and Weingartner, D. P. 2001. Compendium of potato

diseases. Microbiology Socity for General.

Sławiak, M., van Beckhoven, J. R. C. M., Speksnijder, A. G. C. L., Czajkowski, R., Grabe, G.,

and van der Wolf, J. M. 2009. Biochemical and genetical analysis reveal a new clade of

biovar 3 Dickeya spp. strains isolated from potato in Europe. European Journal of Plant

Pathology 125:245-261.

Team, R. C. 2013. R: A language and environment for statistical computing. Vienna, Austria.

Tekalign, T., and Hammes, P. S. 2005. Growth and productivity of potato as influenced by

cultivar and reproductive growth: II. Growth analysis, tuber yield and quality. Scientia

Horticulturae 105:29-44.

Thomson, S. V., Hildebrand, D. C., and Schroth, M. N. 1981. Identification and nutritional

differentiation of the Erwinia sugar beet pathogen from members of Erwinia carotovora

and Erwinia chrysanthemi. Phytopathology 71:1037-1042.

Thornton, M. 2020. Potato growth and development. Potato Production Systems. Springer: 19-

33.

Page 164: Characterization of Dickeya dianthicola and Pectobacterium

149

Tian, Y., Zhao, Y., Yuan, X., Yi, J., Fan, J., Xu, Z., Hu, B., De Boer, S. H., and Li, X. 2016.

Dickeya fangzhongdai sp. nov., a plant-pathogenic bacterium isolated from pear trees

(Pyrus pyrifolia). International Journal of Systematic and Evolutionary Microbiology

66:2831-2835.

Tollenaere, C., Susi, H., and Laine, A.-L. 2016. Evolutionary and epidemiological implications

of multiple infection in plants. Trends in Plant Science 21:80-90.

Toth, I. K., Pritchard, L., and Birch, P. R. J. 2006. Comparative genomics reveals what makes an

enterobacterial plant pathogen. Annual Review of Phytopathology. 44:305-336.

Toth, I. K., Bell, K. S., Holeva, M. C., and Birch, P. R. J. 2003. Soft rot erwinia: from genes to

genomes. Molecular Plant Pathology 4:17-30.

Toth, I. K., van der Wolf, J. M., Saddler, G., Lojkowska, E., Helias, V., Pirhonen, M., Tsror, L.,

and Elphinstone, J. G. 2011. Dickeya species: an emerging problem for potato production

in Europe. Plant Pathology (Oxford) 60:385-399.

Tseng, T.-T., Tyler, B. M., and Setubal, J. C. 2009. Protein secretion systems in bacterial-host

associations, and their description in the Gene Ontology. BMC Microbiology 9:1-9.

Turner, T. R., James, E. K., and Poole, P. S. 2013. The plant microbiome. Genome Biology 14:1-

10.

USDA. 2019. Potato 2018 summary.

Van de Peer, Y., Chapelle, S., and De Wachter, R. 1996. A quantitative map of nucleotide

substitution rates in bacterial rRNA. Nucleic Acids Research 24:3381-3391.

Van der Ent, S., Van Hulten, M., Pozo, M. J., Czechowski, T., Udvardi, M. K., Pieterse, C. M. J.,

and Ton, J. 2009. Priming of plant innate immunity by rhizobacteria and β‐aminobutyric

acid: differences and similarities in regulation. New Phytologist 183:419-431.

Van Der Heijden, M. G. A., De Bruin, S., Luckerhoff, L., Van Logtestijn, R. S. P., and

Schlaeppi, K. 2016. A widespread plant-fungal-bacterial symbiosis promotes plant

biodiversity, plant nutrition and seedling recruitment. The ISME Journal 10:389-399.

Page 165: Characterization of Dickeya dianthicola and Pectobacterium

150

van der Wolf, J. M., and De Boer, S. H. 2007. Bacterial pathogens of potato. Potato Biology and

Biotechnology. Elsevier: 595-617

van der Wolf, J. M., Nijhuis, E. H., Kowalewska, M. J., Saddler, G. S., Parkinson, N.,

Elphinstone, J. G., Pritchard, L., Toth, I. K., Lojkowska, E., Potrykus, M., Waleron, M.,

de Vos, P., Cleenwerck, I., Pirhonen, M., Garlant, L., Helias, V., Pothier, J. F., Pflueger,

V., Duffy, B., Tsror, L., and Manulis, S. 2014. Dickeya solani sp nov., a pectinolytic

plant-pathogenic bacterium isolated from potato (Solanum tuberosum). International

Journal of Systematic and Evolutionary Microbiology 64:768-774.

Vayssier-Taussat, M., Albina, E., Citti, C., Cosson, J. F., Jacques, M.-A., Lebrun, M.-H., Le

Loir, Y., Ogliastro, M., Petit, M.-A., and Roumagnac, P. 2014. Shifting the paradigm

from pathogens to pathobiome: new concepts in the light of meta-omics. Frontiers in

Cellular and Infection Microbiology 4:29.

Versalovic, J., Koeuth, T., and Lupski, R. 1991. Distribution of repetitive DNA sequences in

eubacteria and application to finerpriting of bacterial enomes. Nucleic Acids Research

19:6823-6831.

von Bodman, S. B., Bauer, W. D., and Coplin, D. L. 2003. Quorum sensing in plant-pathogenic

bacteria. Annual Review of Phytopathology 41:455-482.

Waggoner, P. E., and Berger, R. D. 1987. Defoliation, disease, and growth. Phytopathology

77:393-398.

Wallden, K., Rivera‐Calzada, A., and Waksman, G. 2010. Microreview: Type IV secretion

systems: versatility and diversity in function. Cellular Microbiology 12:1203-1212.

Weirather, J. L., de Cesare, M., Wang, Y., Piazza, P., Sebastiano, V., Wang, X.-J., Buck, D., and

Au, K. F. 2017. Comprehensive comparison of Pacific Biosciences and Oxford Nanopore

Technologies and their applications to transcriptome analysis. F1000Research 6.

West, S. A., Griffin, A. S., Gardner, A., and Diggle, S. P. 2006. Social evolution theory for

microorganisms. Nature Reviews Microbiology 4:597-607.

Wigginton, M. J. 1973. Diffusion of oxygen through lenticels in potato tuber. Potato Research

16:85-87.

Page 166: Characterization of Dickeya dianthicola and Pectobacterium

151

Xu, H.-S., Roberts, N., Singleton, F. L., Attwell, R. W., Grimes, D. J., and Colwell, R. R. 1982.

Survival and viability of nonculturable Escherichia coli and Vibrio cholerae in the

estuarine and marine environment. Microbial Ecology 8:313-323.

Yang, C.-H., Gavilanes-Ruiz, M., Okinaka, Y., Vedel, R., Berthuy, I., Boccara, M., Chen, J. W.-

T., Perna, N. T., and Keen, N. T. 2002. hrp genes of Erwinia chrysanthemi 3937 are

important virulence factors. Molecular Plant-Microbe Interactions 15:472-480.

Young, J. M., Dye, D. W., Bradbury, J. F., Panagopoulos, C. G., and Robbs, C. F. 1978. A

proposed nomenclature and classification for plant pathogenic bacteria. New Zealand

Journal of Agricultural Research 21:153-177.

Yousef, F., and Espinosa-Urgel, M. 2007. In silico analysis of large microbial surface proteins.

Research in Microbiology 158:545-550.

Zamioudis, C., Korteland, J., Van Pelt, J. A., van Hamersveld, M., Dombrowski, N., Bai, Y.,

Hanson, J., Van Verk, M. C., Ling, H. Q., and Schulze‐Lefert, P. 2015. Rhizobacterial

volatiles and photosynthesis‐related signals coordinate MYB 72 expression in

Arabidopsis roots during onset of induced systemic resistance and iron‐deficiency

responses. The Plant Journal 84:309-322.

Zhang, J., Shen, H., Pu, X., Lin, B., and Hu, J. 2014. Identification of Dickeya zeae as a causal

agent of bacterial soft rot in banana in China. Plant Disease 98:436-442.

Zhang, Y., Fan, Q., and Loria, R. 2016. A re-evaluation of the taxonomy of phytopathogenic

genera Dickeya and Pectobacterium using whole-genome sequencing data. Systematic

and Applied Microbiology 39:252-259.

Zwankhuizen, M. J., Govers, F., and Zadoks, J. C. 1998. Development of potato late blight

epidemics: Disease foci, disease gradients, and infection sources. Phytopathology 88:754-

763.

Page 167: Characterization of Dickeya dianthicola and Pectobacterium

152

BIOGRAPHY OF THE AUTHOR

Tongling Ge was born in Heze, Shandong, China on November 13, 1992. She graduated from

Dingtao No.1 High School in 2010. She attended Shandong Agricultural University in 2010 and

graduated in June 2014 with a Bachelor of Science degree in Agronomy. She enrolled in the

Department of Plant Pathology at China Agricultural University in the fall of 2014 and graduated

in June 2016 with a master’s degree in Plant Protection. From July 2016 to June 2017, she

worked in Dr. Hao’s Lab at the University of Maine as a Research Assistant. She continued at

the University of Maine seeking a Ph.D. degree in the program of Ecology and Environmental

Sciences in the fall of 2017. During her Ph.D. program, she won two travel awards for research

conferences and one research award. She contributed six talks and abstracts to American

Phytopathological Society annual meetings and other research conferences. Tongling has three

publications in Plant Disease Journal and 14 in Plant Disease Management Reports as first

author. She also has two publications in peer-reviewed journals and 16 in Plant Disease

Management Reports as a co-author. She is a candidate for the Doctor of Philosophy degree in

Ecology and Environmental Sciences from the University of Maine in May 2021.