interactions between plant growth-promoting rhizobacteria

187
1 INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA (PGPR) AND SQUASH PLANTS THE ROLE OF PGPR-MEDIATED SUPPRESSION OF PHYTOPHTHORA BLIGHT By XIAODAN MO A DISSERTATION PRESENTED TO THE GRADUATE SCHOOL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY UNIVERSITY OF FLORIDA 2013

Upload: others

Post on 02-Oct-2021

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

1

INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA (PGPR) AND SQUASH PLANTS – THE ROLE OF PGPR-MEDIATED SUPPRESSION

OF PHYTOPHTHORA BLIGHT

By

XIAODAN MO

A DISSERTATION PRESENTED TO THE GRADUATE SCHOOL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT

OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY

UNIVERSITY OF FLORIDA

2013

Page 2: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

2

© 2013 Xiaodan Mo

Page 3: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

3

This dissertation work is dedicated to my husband, Jiebin, for his endless love, support and encouragement during the challenges in research and life. This work is also

dedicated to my parents and my baby, who have always loved me unconditionally.

Page 4: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

4

ACKNOWLEDGMENTS

I am very fortunate to have performed my graduate work at University of Florida;

and there are many people to thank for their part in my dissertation work. I would like to

express my deepest appreciation to my advisor, Dr. Shouan Zhang, for giving me

extraordinary support on my research and to assist me in the dissertation writing. I am

grateful for his guidance and the opportunities he has afforded me. He is incredibly

organized and a great problem solver, both of these qualities were immensely helpful in

moving my project forward. I would also like to thank my dissertation committee

members, Dr. Bruce Schaffer, Dr. Pamela D. Roberts, Dr. Jeffrey B. Jones, and Dr.

Wenyuan Song, for serving on my committee and contribution to my dissertation. I am

very appreciated to Dr. Jodie V. Johnson in Chemistry Department, University of Florida

for his assistance in quantitative analysis of surfactins, and Dr. Joseph W. Kloepper in

the Department of Entomology and Plant Pathology, Auburn University for providing

Bacillus PGPR strains.

I would like to thank all members of Dr. Zhang’s lab who have contributed to the

progress I have made. Yuqing Fu and Dr. Xiaohui Fan have given their time to teach me

techniques and aid in experiments since the first day I joined the lab. I would like to

thank Dr. Zelalem Marsha and Dr. Jaimin S. Patel for sharing their own expertise and

enthusiasm for science.

Finally, I would like to thank my friends and family for their continued support and

encouragement. Thanks for listening to my problems, providing perspective, being

patient with me, and being there whenever I need you.

Page 5: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

5

TABLE OF CONTENTS page

ACKNOWLEDGMENTS .................................................................................................. 4

LIST OF TABLES ............................................................................................................ 9

LIST OF FIGURES ........................................................................................................ 11

LIST OF ABBREVIATIONS ........................................................................................... 13

ABSTRACT ................................................................................................................... 14

CHAPTER

1 LITERATURE REVIEW .......................................................................................... 16

Summer Squash ..................................................................................................... 16 Squash Crown and Root Rot Disease Caused by Phytophthora capsici ................ 18

Plant Growth-Promoting Rhizobacteria ................................................................... 22 Rhizosphere ..................................................................................................... 22

PGPR as Biocontrol Agents (BCAs) on Vegetable Crops ................................ 23 Root Colonization of PGPR .............................................................................. 26

Mode of Action of Biocontrol Agents ....................................................................... 28 Antagonism Mediated by Metabolites ............................................................... 28

Competition for Nutrient and Niches ................................................................. 30 Defense Mechanisms in Plants ........................................................................ 30

2 BIOLOGICAL MANAGEMENT OF PHYTOPHTHORA BLIGHT AND PLANT GROWTH PROMOTION BY SELECTED PGPR STRAINS UNDER GREENHOUSE CONDITIONS ............................................................................... 36

Introduction ............................................................................................................. 36 Materials and Methods............................................................................................ 40

PGPR Strains and Their Growth Conditions ..................................................... 40

Media Used in the This Study ........................................................................... 41 Growth of Squash Plants .................................................................................. 41

Application of PGPR Treatments in the Greenhouse ....................................... 42 Inoculum Preparation and Inoculation .............................................................. 42

Disease Rating ................................................................................................. 43 Evaluation of PGPR for Suppressing Phytophthora blight ................................ 43 Plant Growth Promotion by PGPR ................................................................... 44 Effect of PGPR Concentration on Disease Suppression .................................. 44 Effect of Timing of PGPR Application on Disease Suppression ....................... 44

Effect of PGPR on Chlorophyll Content in Squash Leaves .............................. 45 Statistical Analyses .......................................................................................... 45

Results .................................................................................................................... 46

Page 6: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

6

Susceptibility of Squash to Different P. capsici Isolates ................................... 46

Effect of PGPR on Phytophthora Blight in the Greenhouse.............................. 46 Effect of PGPR on Plant Growth in the Greenhouse ........................................ 47

Effect of Concentration of Selected PGPR on Disease Suppression ............... 48 Effect of Application Timing of PGPR on Disease Development and Age-

related Resistance to Phytophthora Crown and Root Rot on Squash Treated with IN937b ...................................................................................... 49

Effect of PGPR on Chlorophyll Content in Squash Leaves .............................. 49

Discussion .............................................................................................................. 50 Virulence Differentiation of P. capsici Isolates .................................................. 50 Control of Phytophthora Crown and Root Rot by Bacillus spp. ........................ 51 Enhanced Plant Growth by Bacillus spp. .......................................................... 53

Relation of Leaf Chlorophyll Content and Plant Health .................................... 53 Optimization of Application Rate and Timing .................................................... 54

3 THE ROLE OF BIOFILM FORMATION AND COLONIZATION IN THE RHIZOSPHERE BY BACILLUS SUBTILIS IN937B WITH SYNTHESIS OF SURFACTINS ......................................................................................................... 75

Introduction ............................................................................................................. 75 Materials and Methods............................................................................................ 79

PGPR Strains and Media ................................................................................. 79 Pellicle Formation ............................................................................................. 79

Colony Morphology Assay ................................................................................ 80 Biofilm Production Assay in Microtiter Plates ................................................... 80

DNA Extraction and Polymerase chain reaction (PCR) Analysis ...................... 80 Surfactin Extraction .......................................................................................... 82 HPLC/ESI-MSn Assay of Surfactin Production ................................................. 82

Screening IN937b Mutant for Rifampicin Resistance ....................................... 83 Root Treatment by IN937b Rifampicin-Resistance Mutant ............................... 83

Rhizosphere Sampling ..................................................................................... 84 Recovery of IN937b Mutant with Rifampicin-Resistance from Roots ............... 84

Results .................................................................................................................... 85 Architecture of Pellicles and Colonies of Bacillus spp. ..................................... 85 In Vitro Test for Biofilm formation ..................................................................... 86

Effect of Bacterial Density on Biofilm Formation ............................................... 86 PCR Detection of Lipopeptide Genes in Bacillus PGPR Strains in Relation

to Biofilm Formation ...................................................................................... 87 HPLC/MS Analysis of Surfactins Production in B. subtilis IN937b .................... 88

Survival of IN937b in the Rhizosphere ............................................................. 88 Survival of IN937b in Squash Roots ................................................................. 89

Discussion .............................................................................................................. 89

4 THE ROLE OF SUBTILOSIN A-LIKE ANTIMICROBIAL PEPTIDE IN SUPPRESSION OF PHYTOPHTHORA CAPSICI BY BACILLUS SUBTILIS IN937B .................................................................................................................. 110

Page 7: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

7

Introduction ........................................................................................................... 110 Materials and methods.......................................................................................... 113

Dual Culture Assay ......................................................................................... 113

Antibiotic Extraction and Analysis ................................................................... 113 Antifungal Activity of Methanol Extracts Against Phytophthora capsici .......... 114

Detached Leaf Assays ................................................................................... 115 HPLC/MS Assays of Surfactin Production ...................................................... 116 DNA Extraction and PCR Analysis ................................................................. 116

Results .................................................................................................................. 117 Dual Culture Assays ....................................................................................... 117 Methanol Extracts of IN937b Culture .............................................................. 117 Chromatographic Separation by HPLC .......................................................... 117 PCR Detection of sbo Genes in Bacillus PGPR IN937b ................................. 118

Assays of Hyphal Growth Inhibition ................................................................ 118 Effect of Methanol Extracts on Mycelium Morphology .................................... 119

The EC50 of the Extracts ................................................................................. 119

Effect of Methanol Extracts on Zoospore Germination ................................... 119 Effect of Methanol Extracts on Sporangial Formation .................................... 119 Effect of Methanol Extracts on Detached Leaf Infection ................................. 120

Discussion ............................................................................................................ 120 Hyphal Growth Inhibition ................................................................................ 121

Zoospore Germination and Sporangial Formation .......................................... 122 Detached Leaf Assays ................................................................................... 122 Identification of Subtilosin A ........................................................................... 123

5 ELICITATION OF PLANT SYSTEMIC DEFENSE BY BACILLUS PGPR STRAINS .............................................................................................................. 135

Introduction ........................................................................................................... 135 Materials and Methods.......................................................................................... 139

Inoculum Preparation and Inoculation ............................................................ 139 Application of Bacillus PGPR Treatment in the Greenhouse .......................... 139

Split-Root Assay ............................................................................................. 140

Enzyme Activity Assays.................................................................................. 140 Enzyme Activity Assays.................................................................................. 141

Results .................................................................................................................. 141 Induction of Systemic Resistance by PGPR with P. capsici Challenge .......... 141

Increased Activities of PO and PAL Induced Locally by PGPR Treatment with P. capsici Challenge ............................................................................ 142

Systemically Induced Enzyme Activity as Determined by Split-Root System without P. capsici Challenge ....................................................................... 143

Increased PAL and PO Activities Systemically in Response to P. capsici in Different Seedling Stages............................................................................ 143

Discussion ............................................................................................................ 144 Defense-Related Enzymes ............................................................................. 144 Locally Induced Enzyme Activity .................................................................... 146 Priming: Getting Ready for Pathogen Invasion ............................................... 147

Page 8: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

8

6 SUMMARIES AND FUTURE PERSPECTIVES .................................................... 158

Summaries ............................................................................................................ 158 Future Perspectives .............................................................................................. 160

LIST OF REFERENCES ............................................................................................. 163

BIOGRAPHICAL SKETCH .......................................................................................... 187

Page 9: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

9

LIST OF TABLES

Table page 1-1 Bacillus-based biofungicides on vegetables ....................................................... 35

2-1 Relevant information for PGPR strains tested in this study ................................ 58

2-2 Disease development in squash inoculated with different isolates of P. capsici ................................................................................................................ 59

2-3 Effect of the concentration of P. capsici isolate 151 on disease development .... 60

2-4 Efficacy of PGPR (1-8 PGPR strains) on suppression of Phytophthora crown and root rot under greenhouse conditions .......................................................... 61

2-5 Efficacy of PGPR (9-16 PGPR strains) on suppression of Phytophthora crown and root rot under greenhouse conditions................................................ 62

2-6 Effect of PGPR on plant growth under greenhouse conditions ........................... 63

2-7 Effect of IN937b concentration on crown and root rot disease ........................... 64

2-8 Effect of SE76 concentration on crown and root rot disease .............................. 65

2-9 Effect of IN937a concentration on crown and root rot disease ........................... 66

2-10 Effect of Bacillus spp. application timing (stage 1: seed) on crown and root rot disease .......................................................................................................... 67

2-11 Effect of Bacillus spp. application timing (stage 2: 10 days after emergence) on crown and root rot disease ............................................................................ 68

2-12 Effect of Bacillus spp. application timing (Stage 3: 17 days after emergence) on crown and root rot disease ............................................................................ 69

2-13 Effect of PGPR on chlorophyll content in squash leaves under greenhouse conditions ........................................................................................................... 70

3-1 List of primer sequences used for biosynthetic genes of surfactins, bacillomycin, iturin A, Fengycin, and Zwittermicin A. .......................................... 96

3-2 Biofilm formation by Bacillus strains with initial bacterial populations ................. 96

3-3 Detection of the genes related to biofilm formation ............................................. 96

3-4 Identification of surfactin ions by HPLC/ESI-MS ................................................. 97

Page 10: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

10

3-5 HPLC/(+)ESI-MS peak areas and relative areas for the detected surfactins in the crude extracts of IN937b............................................................................... 97

4-1 Antagonism of Bacillus PGPR strains against P. capsici on PDA plates .......... 125

4-2 Effect of Methanol Extracts of IN937b on hyphal growth of P. capsici .............. 126

4-3 Effect of Methanol Extracts of IN937b in Detached Leaf assays ...................... 127

5-1 Effect of Bacillus PGPR on number and size of lesions in the detached leave assay under the lab conditions ......................................................................... 149

5-2 Effect of different concentrations of IN937b on the number and size of lesions in the detached leave assay under the lab conditions ...................................... 149

Page 11: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

11

LIST OF FIGURES

Figure page 2-1 Regression analysis of the concentration of P. capsici isolate 151 on disease

development ....................................................................................................... 71

2-2 Regression analysis of Effect of PGPR strains concentration on crown and root rot disease ................................................................................................... 72

2-3 Effect of IN937b on squash root and shoot dry weight. ...................................... 73

2-4 Chlorophyll content in leaves of squash plants treated with PGPR. ................... 74

3-1 Architecture of pellicles and colony morphology of PGPR strains. ..................... 98

3-2 Pellicle development of Bacillus strains. ............................................................. 99

3-3 Biofilm development properties of Bacillus strains. ............................................. 99

3-4 Biofilm formation rate of Bacillus strains. .......................................................... 100

3-5 HPLC/MS analysis of the fraction from crude extraction from B. subtilis IN937b culture. ................................................................................................. 101

3-6 Colonization of the rhizosphere of squash plants by a mutant of IN937b with rifampicin resistance. ........................................................................................ 105

3-7 Colonization of the squash roots by a mutant of IN937b with rifampicin resistance ......................................................................................................... 106

3-8 PCR product profiles of the sfp gene ................................................................ 107

3-9 Comparison of the predicted amino acid sequence for the sfp gene present in several Bacillus strains. .................................................................................... 109

3-10 Phylogram of sfp homologs from several Bacillus strains ................................. 109

4-1 Antagonism of Bacillus spp. against P. capsici on PDA plates ......................... 128

4-2 Antagonism of methanol extracts of IN937b against P. capsici on CM agar plates ................................................................................................................ 129

4-3 Effect of methanol extracts of IN937b on morphology of P. capsici .................. 130

4-4 Effect of methanol extract concentrations of IN937b on zoospore germination 131

4-5 Effect of methanol extracts of IN937b on sporangial formationmg/mL) ............ 132

Page 12: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

12

4-6 Effect of methanol extracts on detached leaf infection ..................................... 134

5-1 Induction of PAL by PGPR in squash roots ...................................................... 150

5-2 Induction of PO by PGPR in squash roots ........................................................ 151

5-3 Induction of PAL by PGPR in roots against P. capsici ...................................... 152

5-4 Induction of PO by PGPR in roots against P. capsici ....................................... 153

5-5 Induction of PAL by IN937b in split root system ............................................... 154

5-6 Induction of PO by IN937b in split root system ................................................. 155

5-7 Induction of PAL by PGPR against P. capsici in different growth stages .......... 156

5-8 Induction of PO by PGPR against P. capsici in different growth stages ........... 157

Page 13: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

13

LIST OF ABBREVIATIONS

ASM Acibenzolar-S-methyl

BCA Biological control agent

CFU Colony forming units

ET Ethylene

ETI Effector-triggered immunity

HR Hypersensitive response

ISR Induced systemic resistance

JA Jasmonic acid

LPS Lipopolysaccharides

MAMP Microbe associated molecular pattern

PAL Phenylalanine ammonia lyase

PCR Polymerase chain reaction

PGPR Plant growth-promoting rhizobacteria

PO Peroxidase

PR Pathogenesis related

PTI PAMP-triggered immunity

R protein Resistance protein

SA Salicylic acid

SAR Systemic acquired resistance

Page 14: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

14

Abstract of Dissertation Presented to the Graduate School of the University of Florida in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy

INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

(PGPR) AND SQUASH PLANTS – THE ROLE OF PGPR-MEDIATED SUPPRESSION OF PHYTOPHTHORA BLIGHT

By

Xiaodan Mo

December 2013

Chair: Shouan Zhang Major: Plant Pathology

Phytophthora blight is a devastating disease in vegetable production worldwide.

This disease is caused by the soil-borne Oomycete Phytophthora capsici. The objective

of this study was to investigate the suppression of Phytophthora blight on squash and

plant defense induced by plant growth-promoting rhizobacteria (PGPR). The biocontrol

efficacy of 16 Bacillus PGPR strains to reduce disease severity was evaluated under

greenhouse conditions. A rifampicin-resistant mutant of Bacillus PGPR strain IN937b

was obtained and used to monitor its colonization on squash roots. The underlying

mechanisms of disease suppression by PGPR were determined by analyses of

secondary metabolites involving P.capsici inhibition and induced systemic resistance on

squash.

In this study, Bacillus PGPR strain IN937b were consistent in reducing

Phytophthora blight on squash when applied at the concentration of 108 CFU/mL and on

older plants, i.e. the first PGPR application was made at 17 days after emergence. The

growth promotion effect of IN937b only occurred under conditions without fertilization.

IN937b was capable of producing surfactins and forming biofilms. IN937b colonized the

Page 15: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

15

rhizosphere of squash and also occupied the internal roots. At the early stage of

colonization, the population of IN937b remained relatively high in the rhizosphere. The

bacterial population in the rhizosphere decreased 14 days after bacterization, while

bacterial cells in root tissues survived longer and accumulated gradually. IN937b

showed in vitro inhibition activities against P. capsici. Subtilosin A, an antimicrobial

peptide produced by B. subtilis IN937b was found to be involved in reduction in

zoospore germination, sporangial formation and hyphal growth of P. capsici, and the

suppression of lesion expansion on the detached leaves after being challenged with P.

capsici. Greater levels of Peroxidase and phenylalanine ammonia-lyase induced by

IN937b may have contributed collectively to induced resistance - a plant innate

resistance in squash plants against P. capsici, which eventually resulted in lower

disease severity compared to the non-treated control. This study provides future

researchers with insights into understanding biocontrol mode of action that can be used

to develop effective biocontrol agents and to manipulate rhizosphere ecosystems for

improved crop production.

Page 16: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

16

CHAPTER 1 LITERATURE REVIEW

Summer Squash

Summer squash (Cucurbita pepo L.), originating from the Americas, is one of

widely grown and consumed vegetable crops in the world. As a member of the highly

diverse species in the gourd family Cucurbitaceae, is closely related to both muskmelon

(Cucumis melo L.) and cucumber (Cucumis sativus L.) (Paris, 2008). The initial

domestication of squash occurred in Mexico and Central America about 10,000 years

ago, where it spread quickly to Asia, Europe, and Africa, and now it is distributed

throughout the world (Smith, 1997).

Summer squash can be used in a wide variety of dishes both as fresh and

processing products. The word “squash” comes from a Massachusetts Native American

word “askutasquash”, meaning eaten raw or uncooked (Kemble et al., 2005). The fruits,

seeds, flowers, shoots and roots of squash are eatable. Consumption of summer

squash increases every year (Locke et al., 2009). Summer squash is used as folk

medicine because of its nutritional benefits. Since squash contains calcium,

magnesium, potassium, vitamins B6 and E, the fruit may contribute to health of the

heart, lungs and bones. Squash is also a good source of fiber and protein with a low

sugar and sodium concentrations (NASS, 2011).

According to the Food and Agriculture Organization of the United Nations (FAO),

more than 1.8 million hectares (ha) of pumpkin, squash, and gourds are harvested

annually with a total yield of 24 million tons in 2010 (FAO, 2011). As an easy to grow,

short season crop, summer squash is adapted to temperate and subtropical climates

and is grown in many regions (Paris, 1996). Today, the largest producers of summer

Page 17: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

17

squash include China, India, Russia, Iran, USA, Egypt and Mexico (NASS, 2012).

According to the United States Department of Agriculture (USDA), the United States

grew 0.38 million tons of squash for fresh market valued at 248.7 million US dollars in

2012 (NASS, 2013). During 2011-2012, squash growers planted 18,210 ha for fresh

market in the United States. The average harvested yield of summer squash is 19.3

tons/ha. The cultivation of squash can be found in a wide latitudinal range from the

North (New York) to the South (Florida).

Florida ranks first nationally for the area harvested and value of production on

squash processing and fresh market. It is one of the few crops that is shipped every

month of the year throughout the United States (Mossler and Nesheim, 2001).The

growing acreage was 4,047 ha and the value of the production reached 66 million US

dollars in 2012 (NASS, 2012). In 2006, 60% of the squash acreage was in Miami-Dade

County (Li et al., 2006). In the Homestead area, where the majority of the state's squash

is produced, squash is grown from August to the following April (Mossler and Nesheim,

2001).

Squash grows best under conditions of high humidity, and temperature between

18 and 24 ℃, and properly fertilized soil. Squash plants are largely direct seeded in

Florida (Hochmuth and Hanlon, 2010). Sandy loam soil is suitable for squash due to the

shallow root system of this crop (Li et al., 2006). The water requirement for squash of

approximately one inch of water per week, is slightly lower than that for other vegetable

crops requiring (Ertek et al., 2004). Due to hot and dry weather in the winter, more than

90% of squash-producing farms are irrigated to increase production (Mossler and

Nesheim, 2001). Under favorable environmental and fertilized conditions, the squash

Page 18: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

18

plants grow rapidly and produce fruits continuously. Squash requires more phosphorus

and potassium than nitrogen (Ozoreshampton et al., 1994). Timely and appropriate

application of fertilizers can make a significant difference in the quality and quantity of

fruits, and may promote early fruit maturation, allowing the fruit to be harvest early.

In Florida, common diseases of squash in the greenhouse and field include

pathogens Fusarium spp., Rhizoctonia spp., Phytophthora spp., and Pythium spp

(Zhang et al., 2007). Diseases cause substantial losses of the yield, up to 20% annually

and also decrease the fruit quality (Agrios, 2005). Some diseases occur infrequently

such as Phytophthora blight (Phytophthora capsici), downy mildew

(Pseudoperonospora cubensis), powdery mildew (Podosphaera xanthii), angular leaf

spot (Pseudomonas syringae pv. lachrymans), gummy stem blight (Mycosphaerella

citrullina) and Alternaria leaf spot (Alternaria cucumerina) (Zhang et al., 2007).

However, when the weather is favorable for disease development, these diseases may

cause up to 100% yield loss (Babadoost, 2000).

Squash Crown and Root Rot Disease Caused by Phytophthora capsici

Phytophthora blight is a devastating disease of vegetable crops in Florida (Ploetz

et al., 2002). This disease is caused by the soil-borne Oomycete Phytophthora capsici.

Phytophthora capsici was first described on chili pepper at New Mexico in 1922

(Leonian, 1922). Phytophthora capsici which is widespread throughout the United

States has been reported to infect a wide range of plant taxa involving a total of more

than 50 species, including major vegetable crops such as cucurbits, peppers, eggplants,

tomatoes and snap beans (Granke et al., 2012). In recent years, vegetable producers in

eastern U.S. have experienced severe losses of vegetable crops due to P. capsici

infection (Hausbeck and Lamour, 2004).

Page 19: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

19

Summer squash is highly susceptible to P. capsici. Diseased plants are easily

pulled from the soil due to the rotted roots system (Padley et al., 2008). Water-soaked

lesions on the stem result in girdling and plant death because water and nutrient

transport are reduced (Gevens et al., 2008). Affected fruit can be completely covered by

sporangia causing fruit to rots, and can occur after harvest during shipment (Ando et al.,

2009).

Phytophthora capsici releases zoospores from sporangia, or sporangia

germinate directly to initiate infection on squash. The zoospores swim through water in

the soil and are chemotactically attracted by the exudates released by the host roots

(Granke and Hausbeck, 2010). After the zoospores have adhered to the root surface,

they encyst and produce a precisely oriented germ tube that grows into adjacent host

plant tissue (Raftoyannis and Dick, 2006). The development of zoospores, cysts, and

germ tubes is triggered by chemical signals in the environment, some which are

produced by the plant roots (Bishop-Hurley et al., 2002).

Phytophthora capsici has caused longstanding problems for agriculture. In the

field, the disease can spread rapidly in the presence of heavy rainfall and excessive

irrigation water. When surface moisture is present, zoospores landing on host plants

can invade plant tissues through natural openings such as stomata and hydathodes

(gas or water pores, respectively) (Granke and Hausbeck, 2010). The sexual spore of

P. capsici is the oospore. The oospore’s thick walls help it survive unfavorable periods

in the soil. It can also be the overwintering propagule in the soil and the primary source

of inoculum (Granke et al., 2009). Dissemination may also occur by contaminated

Page 20: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

20

equipment and through movement of plant debris or other hosts (Lamour and

Hausbeck, 2000).

Practices for managing of soil-borne pathogens include crop rotation, fungicide

applications, and the use of resistant or tolerant varieties. Field production of vegetable

crops in Florida is based on high-input, intensively managed production systems that

utilize a broad spectrum fumigants such as methyl bromide to manage soil-borne pests

(French-Monar et al., 2007). The fumigant methyl bromide has been used extensively to

control soil-borne pathogens, and is effective against the mycelia and the long-term

persistent oospores of P. capsici in the soil (Zhang et al., 2010). However, agricultural

emissions of methyl bromide is a significant cause of ozone depletion (Portmann et al.,

2012). Alternatives to methyl bromide either are not as effective as methyl bromide or

not economically practical, since squash is not a high value crop (Zhang et al., 2010).

With the phase-out of methyl bromide, the ban on this fumigant could potentially cost

about 1.5 billion US dollars in lost crop production annually in the U. S. (Sande et al.,

2011).

Crop rotation is an important component of integrated disease management;

however, the long-term survival of P. capsici oospores, even in the absence of a host,

limits the effectiveness of this strategy. Crop rotation is needed for more than 3 years to

avoid increasing inoculum levels (Hausbeck and Lamour, 2004). In Florida, the

oospores survive in the soil up to 5 years. It is very difficult to eliminate the oospores

from an infested field (French-Monar et al., 2007).

There are some fungicides registered for use on cucurbits, but none are highly

effective against P. capsici under hot and humid conditions of South Florida (Mossler

Page 21: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

21

and Nesheim, 2001). Among more than 50 fungicides tested, only six were effective for

controlling P. capsici (Babadoost et al., 2008). No single fungicide has consistently and

effectively suppressed losses caused by P. capsici. More importantly, P. capsici

developed resistance to some fungicides used for Phytophthora blight control. For

example, Phytophthora blight often has been managed with application of fungicides

containing the active ingredient mefenoxam. Since the mefenoxam-resistant isolates

have developed, the usefulness of fungicides containing mefenoxam in the field trials

has diminished (Lamour and Hausbeck, 2000).

Plant breeding for resistance is considered an economically viable control

method for Phytophthora blight. Commercial cucurbit varieties vary with respect to their

Phytophthora blight resistance, but highly resistant varieties with ideal horticultural traits

are still not available (Babadoost et al., 2008). Phytophthora capsici is a heterothallic

species with two mating types, designated as A1 and A2, which interact to produce

sexual oospores. The presence of both mating types in the same area provides more

potentially virulent phenotypes of the pathogen through sexual recombination. Variation

in genetics and virulence among isolates of P. capsici lead to costly and time-

consuming breeding endeavors (Granke et al., 2012).

At present, there is no single method which can provide adequate control of P.

capsici. Since the pathogen is difficult to eradicate, induction of plant resistance against

pathogen is considered as a potential method in reduction and control of the disease

(Shoresh et al., 2010). Plants interact continuously with biotic factors (mycorrhiza,

endophytes, insects, nematodes, plant growth-promoting rhizobacteria, and other root-

associated microbes) in the field, which might lead to the inhibit the growth of plant

Page 22: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

22

pathogens (Ghorbani et al., 2008), known as biological control or biocontrol. Biocontrol

can be developed as a component for sustainably integrated disease management with

reduced adverse impact to the environment.

Plant Growth-Promoting Rhizobacteria

Rhizosphere

The term, “rhizosphere” comes from rhizo or rhiza, a Greek word for root, and

“sphere” which means an environment or area influenced by the root (Morgan et al.,

2005). In the rhizosphere, plant roots can interact with soil microorganisms, and soil

microorganisms can, in turn, affect plant growth and health during the entire growth

season (Hartmann et al., 2008).

The rhizosphere is the location where beneficial microorganisms and soil-borne

pathogens compete to establish a relationship with the plant (Raaijmakers et al.,

2009a). Microorganisms compete with each other to acquire nutrients to grow and

multiply (Lugtenberg and Kamilova, 2009). During these processes, metabolic

processes of microorganisms result in rapid nutrient cycles and movement of

microorganisms in the soil (Raaijmakers et al., 2009a). Some beneficial rhizosphere

microorganisms can reduce population densities and activities of soil-borne pathogens

by competition, antagonism, and induced plant defense (Weller et al., 2002). Beneficial

microorganisms such as plant growth-promoting rhizobacteria (PGPR), positively affect

different crops including rice, pepper, tomato, tobacco, and bean (Kloepper and

Schroth, 1978). PGPR are a group of free-living rhizospheric bacteria that enhance

plant health. PGPR can promote plant health not only directly by facilitating mineral

nutrient uptake or phytohormone production, but also indirectly by protecting plants

against pathogens (Persello-Cartieaux et al., 2003; Zahir et al., 2004). There have been

Page 23: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

23

many reports that PGPR prevent diseases caused by Oomycetes and fungal pathogens

including root diseases (damping-off of tomato), foliar diseases (powdery mildews on

cucurbit and strawberry), and postharvest diseases (green, grey and blue molds)

(Kloepper et al., 2004).

PGPR as Biocontrol Agents (BCAs) on Vegetable Crops

PGPR isolated from different plant species belong to many genera including

Pseudomonas, Rhizobia, and Bacillus (Lugtenberg and Kamilova, 2009). Multiple

Bacillus and Paenibacillus spp. can be readily cultured from both bulk (free of root) and

rhizosphere (influenced by root) soils. The population of PGPR is affected by plant

species and root exudates in the rhizosphere. Plant roots release organic compounds

such as amino acids and sugars into the rhizosphere (Lugtenberg and Kamilova, 2009),

which have a positive effect on the population of many PGPR. The plant rhizosphere is

the preferred ecological niche for PGPR. Due to nutrient availability in the rhizosphere,

there are about 2.5 to 1,260 times more microorganisms in the rhizosphere than the

bulk soil (Rouatt et al., 1960). The population density of Bacillus spp. in the rhizosphere

generally range from 103 to 107 cells per gram of rhizosphere soil (Pandey and Palni,

1997).

PGPR are effective in controlling plant diseases caused either by soil-borne,

foliar or post-harvest pathogens (Raupach and Kloepper, 1998b; Kloepper et al., 2004).

PGPR were isolated from different crops (tomato, pepper, cucumber, etc.) based on

their multiple plant growth promoting and antifungal capabilities. These PGPR isolates

are well adapted when they are introduced to particular rhizospheric soil environment

(vanVeen et al., 1997). The target pathogens include soil-borne fungal pathogens

including Phytophthora, Fusarium, and Rhizoctonia, as well as foliar diseases caused

Page 24: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

24

by bacterial pathogens Xanthomonas and Ralstonia. Several commercial PGPR

products are currently available for use on vegetables and new products are constantly

being developed and commercialized in the market (Kumar et al., 2011b). The details of

commonly used commercial PGPR products for managing important diseases of

vegetables are listed in Table 1-1.

PGPR as biocontrol agents have the following advantages: (1) PGPR which were

isolated from the rhizosphere or plants are environmentally friendly and nontoxic

(Beneduzi et al., 2012), (2) PGPR utilize root exudates leading to their rapid growth and

mass production (Lugtenberg and Kamilova, 2009), (3) PGPR adapt to the environment

by colonizing and multiplying in the rhizosphere and interior of the plant (Compant et al.,

2010), and (4) PGPR undertake a defined range of functions including antibiosis, growth

promotion, and induced systemic resistance (Compant et al., 2005).

The success of biocontrol agents mainly depends on the existing seed, soil, or

foliar application technologies that help integrate PGPR into the cropping system. Seed

inoculation with the P. fluorescens isolate CW2 positively affects tomato growth and

protects seedlings from damping-off disease caused by Pythium ultimum (Salman and

Abuamsha, 2012). A foliar spray of P. fluorescens Pf1, Py15, and B. subtilis Bs16

significantly reduced the incidence of early blight caused by Alternaria solani (Latha et

al., 2009). The most commonly used application method for PGPR is as a soil drench.

Soil application of Bacillus subtilis QST713 consistently reduced infection by

Plasmodiophora brassicae on canola roots. Antibiosis and induced resistance are

involved in clubroot suppression by QST713 (Lahlali et al., 2013). The PGPR strains

Page 25: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

25

aggressively colonized plant roots, thus helping to inhibit pathogens in the rhizosphere

by producing antibiotics and inducing host resistance.

The dosage of the BCAs significantly influence the biocontrol efficacy through

their effect on establishment of pathogens and the development of diseases. To achieve

a beneficial effect on a target plant, PGPR have to be implemented into the field in a

timely manner and in high numbers. In the field, a cell concentration at 109 CFU per mL

of Pseudomonas brassicacearum MA250 was required for effective biocontrol of

Fusarium spp. on winter cereals (Levenfors et al., 2008). The implemented PGPR strain

had to reach a threshold population density in order to effectively suppress pathogens in

the rhizosphere. For some PGPR strains, the maximum root colonization was less than

106 CFU per gram of root in the nutrient-rich environment (Ahmad et al., 2011).

Pseudomonas putida strain WCS358 reached a threshold population density of

approximately 105 CFU per gram of root, which was required for a significant

suppression of Fusarium wilt on radish (Raaijmakers et al., 1995b).

Host age influences the effectiveness of PGPR against pathogens. Plants vary in

their susceptibility to certain pathogens due to their age or stage of development, known

as age-related resistance (Garcia-Ruiz and Murphy, 2001). In general, young plants are

susceptible to pathogens while adult plants have a certain degree of resistance. Age-

related resistance to P. capsici is found in cucumber. Immature fruits are more

susceptible to P. capsici than mature fruits (Meyer and Hausbeck, 2013). Plant age also

affects PGPR populations through a modification in the root exudation (Roesti et al.,

2006). At a specific age of the host plant and the appropriate timing of PGPR

application, a plant may respond to PGPR and pathogens by activating its resistance

Page 26: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

26

locally and systemically (Katagiri and Tsuda, 2010b). Some proteins and metabolites

form during this particular period. A resistance protein (R protein) - mediated recognition

of P. syringae is involved in age-related resistance in Arabidopsis (Kus et al., 2002).

Salicylic acid (SA) signaling was involved in age-related resistance of Nicotiana

benthamiana against P. infestans (Shibata et al., 2010). Young and susceptible plants

failed to induce the salicylic acid (SA) - mediated signaling pathway. Pretreating young

plants with acibenzolar-S-methyl (ASM, Actigard®, Syngenta Crop Protection,

Greensboro, NC, USA) or SA analogs induced resistance to P. infestans.

Root Colonization of PGPR

The first step in all interactions between plants and PGPR is bacterial

colonization of plant root systems. The steps of colonization include recognition,

adherence, invasion, occupation and multiplication, and signals to establish the

interactions (Compant et al., 2010). Effective root colonization by PGPR depends on

their ability to survive and proliferate in the presence of indigenous microorganisms over

a period of time (Maheshwari, 2011). For example, P. fluorescens 2112 consistently

lasts for 18 weeks in the rhizosphere of pea (Kim et al., 2012).

Chemotaxis is one way that plants attract PGPR to the rhizosphere through the

release of carbohydrates and amino acids (Somers et al., 2004). The major chemo-

attractants in tomato root exudates for P. fluorescens WCS365 were identified as

leucine and dicarboxylic acids (Lugtenberg et al., 1999). Another organic acid, malate,

appeared to be the major chemo-attractant for B. subtilis FB17 in root exudates of

Arabidopsis thaliana (Rudrappa et al., 2008). Biocontrol agents may be distributed in

the specific crop or the different niches of the hosts (Beneduzi et al., 2008). For

example, B. amyloliquefaciens FZB42 primarily colonized in root hair and tips on

Page 27: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

27

Arabidopsis. However, when FZB42 was applied to maize roots, the bacterial cells

concentrated at the base of primary roots indicating that FZB42 colonized plant roots in

different niches (Fan et al., 2011).

Some bacteria present in the rhizosphere are capable of entering the plants as

endophytes that do not cause harm and can establish a mutualistic association with

plants. Endophytic bacteria have been isolated from a wide range of plant species. In

general, endophytic bacteria occur at lower population densities than rhizospheric

bacteria (Compant et al., 2010). The population density of endophytes is greatly

variable, depending mainly on the bacterial species and genotypes of the host plant,

also influences by the developmental stage of the host, inoculum density, and

environmental conditions (Sturz et al., 2005).

Populations of PGPR usually develop and form biofilms in their natural

environments (Seneviratne et al., 2010). Biofilms are a group of highly structured cells

capable of sticking together on the surface of liquid or solid substances in a self-

produced extracellular matrix (Costerton, 1995). Biofilms change the growth rate of cells

and transform from free-living cell to sessile biofilm cells through nutrient limitation

(Morikawa, 2006). Biofilm formation is important for PGPR to act as a biocontrol agent

against plant pathogens and to respond quickly to environmental conditions

(Seneviratne et al., 2010). Paenibacillus polymyxa protected Arabidopsis thaliana

against Phytophthora palmivora and Pythium aphanidermatum by establishing biofilm

on plant roots to protect the colonization sites such as root tips. Competition for

nutritional elements in the root exudate, biofilm-forming PGPR act as a sink for the

nutrients can inhibit pathogen growth in the rhizosphere (Timmusk et al., 2005). In

Page 28: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

28

addition, biofilm-forming PGPR can produce a variety of antimicrobial metabolites which

include broad spectrum lipopeptides such as surfactins that are potent biosurfactants for

maintaining structure of biofilms and are important for suppressing pathogens (Bais et

al., 2004b).

Although beneficial aspects of PGPR have been well documented, there are

some problems about variability in colonization efficiency, field performance, and

rhizosphere competence (Babalola, 2010). In the field, PGPR behave unpredictably in

different environments (Spadaro and Gullino, 2005). Variability in rhizosphere

colonization has been proposed to be a major obstacle to agronomic application of

promising PGPR strains (Cooper and Rao, 2006). Research into the mechanisms of

action has provided a comprehensive understanding of PGPR as biocontrol agents.

Research on the mechanism that is responsible for how the PGPR interact with the

hosts and pathogens would help to improve the efficacy of growth promotion and

biocontrol. Detailed studies of the whole system need to be undertaken in order to

reduce the variability of PGPR performance (Compant et al., 2005).

Mode of Action of Biocontrol Agents

The process of biocontrol is the action of suppressing pathogens by using PGPR

with a variety of mechanisms such as antagonism mediated by metabolites, competition

for nutrients and niches in the rhizosphere, and/or activation of plant defenses. Most of

successful commercial BCAs are based on the thorough knowledge of their modes of

action, not only under laboratory conditions, but also in greenhouse and field situations.

Antagonism Mediated by Metabolites

Antagonism is usually mediated by the production of secondary antimicrobial

metabolites (antibiosis), lytic enzymes, and/or effectors (Raaijmakers et al., 2009a). In

Page 29: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

29

the case of PGPR, antibiotics are secondary metabolites with low-molecular weight that

are toxic to other microorganisms (Mannanov and Sattarova, 2001). The metabolites

released by PGPR are involved in directly inhibiting plant pathogens locally before they

reach the plant (Kessler, 2011). Only a small fraction of pathogen populations is

exposed to the antimicrobial compounds in the rhizosphere during a short period of its

life cycle (Raaijmakers et al., 2009b). Therefore, resistance in pathogens to

antimicrobial compounds produced by antagonistic PGPR is presumed not to develop,

or at least develops relatively slowly compared to resistance to chemicals in pathogen

populations (Ishii, 2006).

Pseudomonas and Bacillus species are known to produce antibiotics that are

inhibitory to the growth and/or activities of fungal pathogens of plants. Well-known and

characterized compounds produced by Pseudomonas are phenazine-1-carboxylic acid

(PCA) and 2, 4-diacetylphloroglucinol (DAPG), while cyclic lipopeptides surfactants,

zwittermycin A, and bacteriocins can be produced by Bacillus (Hayat et al., 2010). To

demonstrate the role of antibiotics in biocontrol, the mutants defective in structural

genes for synthesis of the antibiotic must be negative in biocontrol. In a previous study,

knock-out mutants of lipopeptides surfactin genes were severely defective in biofilm

formation in vitro and in colonization of tomato roots. At the same time, the mutant

decreased the efficacy of biocontrol against Ralstonia solanacearum on tomato under

greenhouse conditions (Chen et al., 2013). DAPG is a primary factor contributing to

biological control of plant disease by P. fluorescens strains CHA0 and several

Pseudomonas strains. The 2, 4-DAPG biosynthetic and regulatory mutants

demonstrated that DAPG has a broad spectrum of antifungal activities that inhibit

Page 30: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

30

pathogens directly, and also induce systemic resistance in the host plants (Weller et al.,

2012).

Competition for Nutrient and Niches

An introduced PGPR strain can be defined as an effective root colonizer if it is

able to propagate and survive in the rhizosphere for several weeks. PGPR face the

challenge of carbon starvation in most soils (Knee et al., 2001). Competition for different

carbon sources generally occurs in the soil between the introduced PGPR strains and

the native inhabitants in the rhizosphere (Nihorimbere et al., 2011). For this purpose,

PGPR directly inhibited pathogens under highly competitive conditions (Cunniffe and

Gilligan, 2011). For example, the lack of carbon sources is assumed to inhibit fungal

spore germination in the soil providing a broad spectrum of protection on plants (Tang

et al., 2010).

PGPR can compete with pathogens for niches when they colonize the same host

tissues and develop in different root cortical cells, indicating some sort of competition for

space (Ongena and Jacques, 2008). A number of bacterial functions such as motility,

attachment, growth, stress resistance and the production of secondary metabolites have

been associated with competence for space occupation in the rhizosphere (Lugtenberg

and Kamilova, 2009).

Defense Mechanisms in Plants

Plants possess morphological barriers, secondary metabolites, and antimicrobial

proteins that prevent or overcome a pathogen’s ability invading a plant (van Loon et al.,

2006). Plants possess different strategies to recognize and counteract pathogen

attacks. As a first line of defense, the plant cell surface contains pattern recognition

receptors (PRRs) that recognize potential pathogens through flagella, outer membrane

Page 31: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

31

lipopolysaccharides (LPS), and other cell wall or secreted components (Zipfel, 2008). In

addition to flagella and LPS, there are various bacterial compounds that can be

recognized by the plant, including N-acyl-homoserine lactones, biosurfactants,

siderophores, and the antibiotics DAPG and pyocyanin (Doornbos et al., 2012).

General elicitors that are designated as pathogen-associated molecular patterns

(PAMPs) are isolated from pathogens (Boller and He, 2009). Non-pathogenic

microorganism elicitors are designated as one of microbe-associated molecular patterns

(MAMPs), which are recognized in a similar way (Bittel and Robatzek, 2007). The

perception of general elicitors triggers a broad array of reactions, known as basal

resistance or PAMP-Triggered Immunity (PTI). However, some successful pathogenic

microorganisms may overcome basal resistance by inhibiting signaling pathways, thus

suppressing this first type of immunity (Zamioudis and Pieterse, 2012). In response,

plants have evolved a second line of defense through specific disease resistance (R)

genes, the so-called effector triggered immunity (ETI) (Katagiri and Tsuda, 2010a). R

gene-mediated resistance is expressed through similar defense responses as those that

are active in basal resistance, but on a much greater scale (Henry et al., 2012).

Therefore, PTI and ETI are considered as primary and secondary innate immunity,

respectively.

The phenomenon of induced resistance to pathogens in plants has been studied

intensively in recent years (Hammerschmidt, 1999). Induced resistance can be divided

into systemic acquired resistance (SAR) and induced systemic resistance (ISR). In

SAR, resistance can be activated by chemicals, avirulent pathogens, or attenuated

pathogens. In contrast to SAR, ISR is induced in roots by PGPR. During ISR, gene

Page 32: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

32

expression is induced locally in roots or systemically in leaves upon colonization by

PGPR, and the treated plants are in primed conditions (i.e. defense genes respond

more rapidly and strongly after pathogen attack).

ISR Mediated by PGPR

To fully understand the specific mechanism of ISR, research on ISR in certain

crop species mediated by specific PGPR strains is required. Globally, local and

systemic defense responses triggered by microorganisms are controlled by a signaling

network in which the plant hormones salicylic acid (SA), jasmonic acid (JA), and

ethylene (ET) play important roles (Henry et al., 2012).

Induction of SA-mediated responses has been demonstrated to occur in the

rhizosphere (Doornbos et al., 2012). Several phytohormones have been demonstrated

to negatively cross-communicate with the SA signaling pathway and affect the outcome

of the immune response (Verhagen et al., 2004; Pieterse et al., 2009). Local JA and ET

signaling play a role in the control of beneficial interactions (Lopez-Raez et al., 2010).

ET also appears to have a crucial role in regulating the progress of root colonization

(Choudhary and Johri, 2009). For example, a significant reduction in the expression of

genes encoding ET-related transcription factors has been reported for Arabidopsis roots

colonized by P. fluorescens WCS417 and FPT9601-T5, which supports the hypothesis

that PGPR modulate host immune response by interfering with the ET signaling

pathway (Verhagen et al., 2004).

Most pathogenicity-related (PR) proteins are induced through the action of the

signaling compounds and possess antimicrobial activities (van Loon et al., 2006). B.

cereus strain BS107 against Xanthomonas axonopodis pv. vesicatoria systemically

induced plant defense genes CaPR1 and PR4 in pepper leaves (Yang et al., 2009). In

Page 33: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

33

tobacco (Nicotiana tabacum cv. Samsun NN), at least 16 PR-1-type genes appear to be

present after being induced upon TMV infection (Riviere et al., 2008). These findings

indicate that PR proteins classified in the same family on the basis of sequence

homology can have different properties and hence may differ substantially in biological

activity. Examples of PR proteins include phenylalanine ammonia-lyase (PAL) (Mariutto

et al., 2011), peroxidase (Sarwar et al., 2011), and polyphenoloxidase (Mayer, 2006).

They are generally present constitutively and only increased the expression level during

most infections.

Plants recognize certain ISR determinants to elicit ISR. Volatile organic

compounds and more particularly 2,3-butendiol were the determinants for elicitation

identified from Bacillus spp. (Ryu et al., 2003). Massetolide A produced by P.

fluorescens retains ISR-eliciting activity in tomato plants for the control of Phytophthora

infestans, the causal agent of late blight (Tran et al., 2007). The compound 2,4-

diacetylphloroglucinol (DAPG) may also act as an elicitor of systemic resistance in

some Pseudomonas strains (Bakker et al., 2003). Another class of compounds

emerged recently as ISR elicitors is biosurfactant such as rhamnolipids and

lipopeptides. Pure fengycins and surfactins from B. subtilis provided a significant

protective effect that is similar to the one induced by living cells of the producing strain.

Experiments conducted on bean and tomato showed that overexpression of both

surfactin and fengycin biosynthetic genes in B. subtilis strain 168 (the strain is impaired

in lipopeptide production) was associated with a significant increase in the potential of

the derivatives to induce resistance (Ongena et al., 2007).

Page 34: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

34

B. subtilis mutant strain M1 was constructed with a deletion in a surfactin

synthase gene and thus deficient in surfactin production (Bais et al., 2004b). Bacillus

subtilis M1 was ineffective as a biocontrol agent against P. syringae in Arabidopsis and

also failed to form robust biofilms on either roots or inert surfaces (Bais et al., 2004b).

The understanding of the physiological and biological basis of these induced immunity

mechanisms has greatly advanced the development of novel strategies for disease

control (Henry et al., 2012). The importance of these considerations is expected to have

practical applications. PGPR have the ability to control pathogens through complex

processes of colonization and antibiotic production. Studies of the population dynamics

and physiology of PGPR can be important and useful for applying PGPR for disease

management in vegetable production.

The aim of this dissertation was to investigate PGPR strains for controlling

Phytophthora blight on squash. The first objective was to determine the effectiveness of

Bacillus PGPR strains in the biocontrol of crown and root rot disease on squash caused

by P. capsici in soilless potting mixes under the greenhouse conditions. The second

objective was to focus on understanding of the ecology of Bacillus PGPR and pathogen

to enhance the effective use of Bacillus PGPR in the suppression of plant diseases. The

third objective was to gain insights into the role of antifungal antibiotics in disease

suppression via antagonism. The fourth objective was to gain an understanding of

mechanism in induced resistance by monitoring the defense-related enzyme activities

changes on squash to impair pathogen progression locally and systemically.

Page 35: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

35

Table 1-1. Bacillus-based biofungicides on vegetables

Brands® Bacillus species

Target pathogen

Crop Mode of action

Reference

Kodiak B. subtilis GB03

Colletotrichum orbiculare

Cucumber Antagonism ISR Plant growth promotion

(Raupach and Kloepper, 1998a)

Rhapsody B. subtilis QST 708

Alternaria cucumerina

Muskmelon

Antagonism (Keinath et al., 2007)

Serenade B. subtilis QST713

Pseudomonas syringae

Tomato Antagonism ISR Plant growth promotion

(Gilardi et al., 2010)

Sclerotinia sclerotiorum

Soybean (Zeng et al., 2012b)

Sonata

B. pumilus QST 2808

Phytophthora infestans, Fusarium spp., Pythium ultimum

Potato ISR (Gachango et al., 2012)

Subtilex B. subtilis MBI 600

Phoma valerianellae

Lettuce ISR (Schmitt et al., 2009)

Yield Shield B. pumilus GB34

Rhizoctonia Fusarium

Legumes (Bradley, 2008)

Companion B. subtilis GB03, B. lichenformis and B. megaterium

Fusarium oxysporum

Tomato Antagonism ISR Plant growth promotion

(Myresiotis et al., 2012)

Page 36: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

36

CHAPTER 2 BIOLOGICAL MANAGEMENT OF PHYTOPHTHORA BLIGHT AND PLANT GROWTH PROMOTION BY SELECTED PGPR STRAINS UNDER GREENHOUSE CONDITIONS

Introduction

Squash (Cucurbita pepo L.), in the family Cucurbitaceae, is a widely grown and

consumed vegetable crop in the United States and many other countries as well. In

2012, growers in the United States harvested 340,060 tons of squash for fresh market

valued at $248.7 million (NASS, 2011). Ranked No. 1 nationally in terms of the acreage

harvested and the value of processing and fresh market production, squash became a

valuable vegetable crop for Florida growers. Squash is vulnerable to bacterial and

fungal pathogens. The most devastating squash disease is Phytophthora blight

(Babadoost, 2000), which can cause yield losses up to 100% in warm and humid

conditions (Hausbeck and Lamour, 2004). Phytophthora blight poses major challenges

to growers by reducing yield and adversely affecting the fruit quality.

Phytophthora blight is a soil-borne disease which commonly affects numerous

crops in the United States including all cucurbits, peppers, tomatoes, snap beans and

several weed species (Ploetz and Haynes, 2000). The pathogen that causes this

disease is the Oomycete Phytophthora capsici Leonian (Leonian, 1922), which

produces motile zoospores. When it rains, or irrigation water splashes in the field,

zoospores can be spread with water to reach healthy squash tissues (Granke et al.,

2012). Zoospores and sporangia can infect all parts of squash plants including leaves,

roots, stems and fruits (Hausbeck and Lamour, 2004). The affected plants quickly die in

a week due to break down of the root and/or stem systems.

Symptoms of Phytophthora blight are first visible as water-soaking lesions, which

then turn dark over time (Kousik et al., 2012). Phytophthora capsici is easy to identify on

Page 37: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

37

the fruit or stem as white mass of mycelia, sporangiophores and sporangia on the

surface of the tissue (Roberts et al., 2005). This diploid Oomycete is heterothallic with

two different mating types to complete sexual cycle (Erwin and Ribeiro, 1996). When

both mating types A1 and A2 are found in the same field, the sexual propagule

oospores are produced. Oospores can tolerate adverse environments and survive in the

soil for up to five years (French-Monar et al., 2007). In the spring, oospores are the

primary source for initial infection in the field (Pavon et al., 2008). Therefore, controlling

this disease has become increasingly challenging because oospores can survive a long

time in the soil, even without host tissues. Oospores have the ability to infect all parts of

plants, and more aggressive types of this pathogen may potentially cause frequent

outbreaks (Sanogo and Ji, 2012). So far, there is no single strategy that provides

adequate protection against this disease.

In an effort to prevent or control Phytophthora blight, resistant cultivars remain

the most economical and long-term approach. Although genetic resistance was

identified from some cucurbit species, there is no cucurbit cultivar commercially

available with measurable resistance to P. capsici (Hausbeck and Lamour, 2004).

Mapping genes for resistance to Phytophthora blight is challenging, and breeding for

resistant cucurbits is complex due to continuous changes in the genetics of P. capsici

(Babadoost et al., 2008).

In Florida, intensively managed production systems for vegetable production rely

on methyl bromide, a fumigant commonly used to manage soil-borne pests (French-

Monar et al., 2007). With the impending phase-out of methyl bromide (Noling and

Becker, 1994), growers must apply fungicides frequently to maintain their productivity.

Page 38: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

38

However, overusing chemicals increases the risk of the pathogens developing

resistance to fungicides, resulting in the reduction of chemical efficacy. Chemicals

adversely impact non-target species and the environment, causing potential public

health concerns (Pimentel et al., 1992). As a result, there is an increasing demand for

the development of biologically based, environmentally safe alternatives to chemical

pest-control products.

The use of microorganisms to control plant disease is a promising strategy and

has potential for improving plant health (Cook, 1993). As an eco-friendly and cost-

effective strategy, biological control agents (BCAs) are usually isolated from natural

fields or the inside of plant tissues (Zhang et al., 2009). These BCAs are capable of

colonizing the root surface efficiently when they are reintroduced into the soil. Once

established, BCAs are capable of providing long-lasting and long-term defenses against

soil-borne pathogens attack.

Many members of the genus Bacillus have the ability to effectively control plant

diseases. Bacillus subtilis is an ideal example of a bacterial biological control agent,

which is able to control damping-off caused by Pythium ultimum (Asaka and Shoda,

1996), crown rot caused by Aspergillus niger (Podile and Prakash, 1996), and vascular

wilt caused by Fusarium oxysporum (El-Hassan and Gowen, 2006). Bacillus subtilis has

also been developed commercially for bio-control products including BioYield®

(Kloepper et al., 2007), Subtilex® (Schisler et al., 2004) and Serenade® (Lahlali et al.,

2013).

Bacillus strains have several characteristics making them good candidates for

developing competitive BCAs. First, Bacillus spp. have the potential to produce more

Page 39: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

39

than 20 antibiotics with strong antimicrobial activity (Stein, 2005). Antibiotics help

Bacillus strains successfully survive and multiply in the soil and to compete with other

microorganisms for nutrients and space (Nihorimbere et al., 2011). Bacillus PGPR have

been shown to increase seed emergence, plant weight, and yield (Kloepper et al.,

2004). Bacillus spp. secrete plant hormones such as indole-3-acetic acid (IAA),

cytokinin and gibberellin which enhance plant growth (Spaepen et al., 2007). Bacillus

strains also increase plant uptake of available minerals, nitrogen, and phosphorus from

the soil (Toro et al., 1998). Bacillus strains form heat- and desiccation- resistant

endospores that promote survival in stressful environments (Lim and Kim, 2010).

Commercialization and the use of most bio-control products remains limited due

to inconsistency in their efficacy, particularly under field conditions (Ramaekers et al.,

2010). To exert their beneficial effects, the application method and timing is critical for

avoiding an outbreak of the pathogen prior to the introduction of Bacillus spp. into the

soil (Spadaro and Gullino, 2005). To solve this problem, the appropriate methods and

timing for applying BCAs need to be determined. Application timing is important

because BCAs need to be applied prior to a pathogen establishing itself on a host plant

in order to achieve the best control. Host may regulate its defense response to a

pathogen during its different growth stage, known as age-related resistance (Kus et al.,

2002). Appling BCAs to a plant host prior to the plant’s maturation is imperative to

impede the contact of the pathogen with the host plant. A positive correlation has been

shown between a plant age and resistance to Phytophthora megasperma var sojae in

soybean(Hanley et al., 1995). Thus, the application timing of BCAs may directly

influenced by age-related resistance.

Page 40: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

40

It is generally assumed that the biocontrol efficiency and yield are higher when

BCAs are used in a high dosage. Application rate can be altered to maximize BCAs

efficacy. However, it is probable that introducing too many of microbes could negatively

affect the plant or the rhizosphere (Strigul and Kravchenko, 2006). For example,

population growth of PGPR is suppressed when PGPR population density is high (over

109 CFU/g soil) due to competitiveness for resources and production of secondary

metabolites. Additional work is needed to determine the optimal application rate of

BCAs in for efficient control.

In previous research, PGPR strains from Auburn University (Dr. Joseph

Kloepper’s lab) isolated originally from the plant rhizosphere were efficient in plant

growth promotion and disease control against multiple pathogens on different vegetable

crops under greenhouse and field conditions (Zhang et al., 2010). The overall goal of

the present study was to evaluate bio-control agents including PGPR for suppressing

the Phytophthora blight on squash. The specific objectives were to (1) investigate the

potential of Bacillus spp. for bio-control of Phytophthora blight on squash; (2) optimize

the application rate and timing for better control of this disease.

Materials and Methods

PGPR Strains and Their Growth Conditions

Sixteen bacilli PGPR strains provided by Dr. Joseph W. Kloepper at Auburn

University, AL were isolated from the soil rhizosphere, root surfaces, and roots of

different crops. Information on these PGPR strains are listed in the Table 2-1. PGPR

strains used in this study were stored in Nutrient Broth (NB, pH=7.0) containing 20%

glycerol at -80 ℃.

Page 41: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

41

PGPR strains were streaked on nutrient agar (NA) and incubated at 28 ℃ for 24

h. Single colonies were inoculated in NB and cultured for 24 h in a shaker incubator

(SteadyShake 757, Amerex Instruments, Inc., Lafayette, USA) at 150 revolutions per

minute (rpm) at 28 ℃. The bacterial culture was centrifuged at 3,000 rpm for 10 minutes

(min) and the cells were suspended in 20 milliliter (mL) of sterilized deionized water (DI

water). The concentrations of the bacterial suspensions were adjusted to 108 cells/mL

based on the optical density (OD600) with a spectrophotometer (UNICO SpectroQuest

TM 4802 UV/VIS Double Beam, United Products & Instruments Inc., NJ).

Media Used in the This Study

Pimaricin–ampicillin–rifampicin–pentachloronitrobenzene with hymexazol

cornmeal agar (PARPH) medium was modified (Mitchell and Kannwischer-Mitchell,

1992): corn meal agar (Difco, 17 g/L) amended with: ampicillin (100 mg/L), hymexazol

(50 mg/L), pimaricin (10 mg/L), PCNB (100 mg/L), rifampicin (10 mg/L).

Clarified V8 broth was prepared with 250 mL V8 juice by mixing with 0.2 g

CaCO3 and centrifuging at 3,000 rpm for 20 min at room temperature. The supernatant

was filtered through Whatman® No. 1 filter paper. The final volume of supernatant was

200 mL and mixed with 800 mL DI water.

V8 agar (unclarified) was prepared with 10% V8 juice, 1% CaCO3 and 15 g/L

Agar. The mixture was heated and stirred prior to autoclaving in order to dissolve the

solutes completely.

Growth of Squash Plants

Zucchini (Cucurbita pepo cv. Senator) seeds were obtained from S&B Inc.

(Homestead, FL). The seeds were washed in DI water (pH 2.0, adjusted with 6M HCl) to

remove fungicides in the coating, and grown in the plastic pots (7.62 cm high and 10.16

Page 42: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

42

cm at the opening with 8 drain holes) containing potting mix (Fafard® Lightweight Mix 2,

Conrad Fafard, Inc., Aquawam, MA) in the greenhouse (24 ± 10 ℃) during the

experiments. After the squash plants had two completely expanded leaves, each

seedling was fertilized with 50 mL (3.5g/L) of a liquid soluble fertilizer (Peter’s 20:20:20

NPK, Peters Fertilizer Products, Fogelsville, PA) every two weeks.

Application of PGPR Treatments in the Greenhouse

To evaluate the effect of PGPR on Phytophthora blight and plant growth, PGPR

were applied two times in the greenhouse. For the first PGPR application, 20 mL of

each PGPR suspension at 1 × 108 CFU/mL was applied as a soil drench around the

base of the stem. One week later, the plants were treated with the same PGPR a

second time by soil drench to enhance the PGPR population and potential of disease

control.

Inoculum Preparation and Inoculation

Isolates (#121, #146, and #151) of P. capsici were provided by Dr. Pamela D.

Roberts, Southwest Florida Research and Education Center, University of Florida,

Immokalee, FL. The inoculum of P. capsici was prepared according to (Ploetz et al.,

2002) with slight modifications. Phytophthora capsici was grown on the selective

medium, PARPH, containing a mixture of fungicide and antibiotics to prevent

contamination by bacteria and other soil-borne pathogens. A disc (5-mm diameter) was

cut from the edge of the medium with P. capsici, transferred to V8 agar and grown for

one week at 25 ℃. Ten discs (5 mm) with P. capsici were cut from the V8 agar plate,

transferred to a Petri dish (90-mm diameter) containing 20 mL of V8 broth and

incubated for one week at 25 ℃ in the dark. The mycelium in the V8 broth was washed

twice with DI water, and 15 mL of DI water was added to merge the mycelium. Then the

Page 43: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

43

plates were placed under light at room temperature for 2 days to induce sporangial

formation. The plates were chilled at 4 ℃ for 1 hour, before they were moved into an

incubator at 28 ℃ for 30 minutes to enhance the release of zoospores. The

concentration of zoospores was determined under an Olympus BH2-RFCA microscope

(Olympus, Tokyo, Japan) using a hemocytometer after the sporangial solution was

stirred in a vortex mixer for 1 min to break flagella of the zoospores.

For inoculation with P. capsici, the concentration of the inoculum was adjusted to

2×104 zoospores/mL using DI water. Three days after the second PGPR treatment, 5

mL of the zoospore inoculum was carefully added to the soil around the base of plant

stems. Inoculated plants were maintained on benches in the greenhouse with high soil

moisture by hand watering daily.

Disease Rating

Symptoms of Phytophthora blight usually started at the base of the stem about 4-

5 days after inoculation. The disease was rated according to a 0-5 scale (Zhang et al.,

2010), where 0=healthy plant, 1= water soaked lesions at the bottom of the stem, 2=

stem was girdled and had lesions or started to exhibit white mycelium on the surface of

diseased tissue, 3= all leaves turned yellow and the whole plant started to collapse, 4=

plants wilted and fell down, and 5= dead plant.

Evaluation of PGPR for Suppressing Phytophthora blight

Sixteen PGPR strains were tested in two groups. Each group included eight

PGPR strains, one standard fungicide (Presidio®) and a non-treated control. There were

ten plants in each treatment with each plant as a replicate. The experiment was

designed as a randomized complete block. Each experiment was conducted three

Page 44: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

44

times. The replications and duration of experiments were similar in all trials unless

specified.

Plant Growth Promotion by PGPR

To evaluate the effect of PGPR on plant growth promotion, soil was drenched

with PGPR suspensions (1×108 CFU/mL) twice after two plant leaves were completely

expanded. In order to test the influence of fertilizers on plant growth promotion by

PGPR treatment, each pot received 20 mL (3.5 g/L) of the liquid fertilizer 20:20:20 NPK

mixture (Peters General Purpose 20-20-20; Peters Fertilizer Products, Fogelsville, PA)

3 days after the first PGPR application. In another experiment, plants were grown in the

potting mix without fertilization. Ten days after the second PGPR treatment, plant

height, stem diameter, fresh and dry weight of shoot and root, root length were

measured and recorded. The experiment was designed as a randomized complete

block, and ten plants were used in each treatment with each plant as a replicate.

Effect of PGPR Concentration on Disease Suppression

The effect of PGPR concentration on disease control efficacy was determined by

soil drench with PGPR suspensions at 105, 106, 107 and 108 CFU/mL. The different

PGPR concentrations were prepared by serial dilution from 108 CFU/mL with DI water.

The standard plant resistance inducer Actigard® was included in this experiment. Each

pot was treated with 10 mL of Actigard (30 mg/L) at the same time that the PGPR

treatment was applied.

Effect of Timing of PGPR Application on Disease Suppression

The effect of timing of PGPR application on Phytophthora blight in squash was

evaluated at different plant growth stages. The first PGPR application was applied 10

(Stage 1), 17 (Stage 2) and 24 (Stage 3) days after emergence in each individual trial.

Page 45: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

45

The second PGPR application was applied 7 days after the first application.

Phytophthora capsici inoculation was the same as described previously.

Effect of PGPR on Chlorophyll Content in Squash Leaves

The chlorophyll content in squash leaves was estimated with a Minolta SPAD

502 meter (Minolta Camera Co., Osaka, Japan) to determine the effect of PGPR

treatment on chlorophyll contents. The data were collected one day before inoculation,

and five days after inoculation when plants started to show symptoms. The readings

were taken from the 3rd, 4th, 5th leaf from apex in all treatments. Data were averaged

from three different leaves for each plants.

Statistical Analyses

Data of disease severity was collected using the rating system from 0 (no

disease) to 5 (dead plant). The area under the disease progress curve (AUDPC) was

calculated for each replicate plant in each treatment using the formula

∑ [𝑦𝑖+𝑦𝑖+1

2] (𝑡𝑖+1 − 𝑡𝑖)

𝑛−1𝑖 , where yi = disease severity at the ith day, ti = time at the ith day,

and n=total number of evaluation times (Ristaino, 1991).

All analyses were performed by using the statistical analysis system (SAS

version9.1, SAS Institute, Cary, NC). Data of AUDPC and growth variables in response

to different PGPR strains in the greenhouse tests were analyzed by a one-way analysis

of variance (ANOVA). Significant differences among means were determined with

Fisher’s protected LSD test (p = 0.05).

Page 46: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

46

Results

Susceptibility of Squash to Different P. capsici Isolates

Isolate 121 was the most aggressive among three P. capsici isolates (121, 146,

and 151) tested, resulting in highest disease severity and AUDPC value at 103, 104 and

105 zoospores/mL 8 days after inoculation (Table 2-2). Isolate 146 was had a slight

effect on inducing disease symptoms at 105 zoospores/ mL, and isolate 151 had a

moderate effect based on the fact that Phytophthora symptoms were the lowest at 104

and 105 zoospores/mL. Based on the result, isolate 151 was chosen to use in the

following experiments. The isolate 151 at 1x 105 zoospores/mL incited significantly

greater disease than 104 zoospores/mL, indicated as AUDPC values (Table 2-3). There

was no significant difference in AUDPC values by P. capsici from 2×104 to 8×104

zoospores/mL. Next, we analyzed the correlation between zoospores levels and

disease development, there would have a positive relationship between zoospores

concentrations and AUDPC values (Figure 2-1). Based on the results above, 2×104

zoospores/mL of isolate 151 was selected as the P. capsici inoculum concentration for

testing the effect of PGPR on suppression of Phytophthora crown and root rot, unless

indicated otherwise.

Effect of PGPR on Phytophthora Blight in the Greenhouse

Sixteen selected Bacillus strains of PGPR were evaluated in the greenhouse for

suppressing Phytophthora blight - in this case, Phytophthora crown and root rot. In the

progression of disease development, some PGPR strains significantly (P≤0.05) reduced

AUDPC value of Phytophthora blight compared to the non-treated control. PGPR strains

IN937b and SE76 consistently provided significant (P ≤ 0.05) protection against P.

capsici in all three trials (Table 2-4). When plants were treated with IN937b and SE76,

Page 47: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

47

symptoms of crown rot caused by P. capsici developed more slowly than it did on non-

treated control plants. PGPR strains SE56, 1PC-11, T4 and GB03 significantly reduced

the AUDPC value in two of the three trials, PGPR strains SE52, INR7, IN937a, SE49,

1PN-19, and Companion significantly reduced the disease AUDPC in one of the three

trials, whereas the strains from commercial products Actinovate AG, BU EXP 1216S,

BU EXP 1216C, and SE34 were the least effective treatments for suppressing

Phytophthora blight on squash in this greenhouse study (Table 2-4 and Table 2-5).

However, no PGPR completely protected the plants from infection by P. capsici. At the

end of the evaluation, plants were dead in all treatments except for the standard

fungicide Presidio®.

Effect of PGPR on Plant Growth in the Greenhouse

When fertilization was adequate for plant growth, there was no difference in all

treatments concerning growth parameters (shoot and root dry weights, shoot height,

root length, etc.) (Table 2-6). Strain IN937a was added as a positive control because it

was shown to improve tomato plants growth (Adesemoye et al., 2009a). Mean plant

height (33.0 cm), stem caliper growth (6.6 mm), fresh shoot weight (34.8 g) and dry

shoot weight (4.2 g) observed in IN937a treated plants were highest in all treatments

(Table 2-6). IN937b numerically increased plant height by 4.5% when compared with

the non-treated control, albeit not significantly. Dry root weight in IN937b and SE76

increased by 50% compared with the non-treated control. Interestingly, IN937b and

SE76 significantly reduced root length by 12% and 7%, respectively. This indicated that

IN937b and SE76 did not result in the longest roots, but produced more lateral roots.

The ability of plant growth promotion by IN937b appears to be, at least partially, due to

the change of root architecture such as lateral roots to increase nutrient and water

Page 48: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

48

uptake by the roots. The ASM treatment did not increase dry root weight or root length

indicating that ASM did not improve lateral root production.

When no fertilizer was added, squash treated with IN937b had significantly

greater dry shoot and root weight compared to the non-treated control (Figure 2-3). In

the experiment to evaluate the effect of fertilization with PGPR treatment on plant

growth promotion, no significant difference in plant growth was observed among strains

and the non-treated control after fertilizer applied, which are consistent with the result in

the Table 2-6.

Effect of Concentration of Selected PGPR on Disease Suppression

Application of IN937b at 106, 107 and 108 CFU/mL significantly reduced

Phytophthora blight severity compared to the non-treated control. IN937b applied at 108

CFU/mL significantly suppressed disease development by 45% compared to the non-

treated control based on the AUDPC values (Table 2-7). When the concentration was

reduced to 105 CFU/mL, disease development was similar to that in non-treated control

plants. A significant positive correlation (r2=0.92) was found between the IN937b

inoculum concentrations and AUDPC value (Figure 2-2).

SE76 at 108 CFU/mL had significantly lower AUDPC values when compared with

the untreated control (Table 2-8). Similarly as IN937b, treatment with SE76 at the

highest concentration (108 CFU/mL) had the highest efficacy to reduce disease severity

by 42% 6 days after inoculation, which was the last day of the investigation.

IN937a had significant disease control in one of the three trials (Table 2-4).

IN937a at high concentrations (108 and 107 CFU/mL) showed disease symptoms earlier

and had greater AUDPC values than the non-treated control (Table 2-9). In this trial, the

low concentration (105 and 106 CFU/mL) significantly reduced disease severity by 25%

Page 49: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

49

and 27%, respectively, compared to the non-treated control. In this study, there is no

significant correlation (r2=0.022) between the IN937a inoculum concentrations and

AUDPC value (Figure 2-2).

Effect of Application Timing of PGPR on Disease Development and Age-related Resistance to Phytophthora Crown and Root Rot on Squash Treated with IN937b

Symptom development of crown and root rot in squash treated with IN937b

varied with the timing of application. When IN937b was applied at growth stage 1 (10

days after plant emergence) and stage 2 (17 days after plant emergence), symptoms

developed as early as 3 days after inoculation. Water-soaked lesions formed initially on

the young seedlings causing the whole plant to wilt quickly (Table 2-10 and 2-11);

whereas when applied at the growth stage 3 (24 days after emergence), symptom

development was delayed, and symptoms were first visible 7 days after inoculation

(Table 2-12). Lesions had visible signs of P. capsici, including mycelia and

sporangiophores with sporangia being easily found in the late stage of development.

IN937b and SE76 significantly reduced disease severity compared to the

untreated control when they were applied at all growth stages. In the trial with best

control efficacy, IN937b and SE76 significantly suppressed the disease by 54% and

57%, respectively, when they were applied at the stage 2 i.e. the first PGPR application

was made at 17 days after seeding (Table 2-11).

Effect of PGPR on Chlorophyll Content in Squash Leaves

The chlorophyll content (as estimated with a SPAD meter) is an indirect indicator

of plant health. The chlorophyll content was significantly greater in leaves of IN937b and

ASM (Actigard) treated plants when compared to the untreated (Figure 2-4). There was

Page 50: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

50

no difference in chlorophyll content among the treatments of SE76, IN937a and the non-

treated control.

Five days after inoculation, significant differences in chlorophyll content were

observed in P. capsici infected plants in the treatments SE76 and ASM compared with

the untreated control (Table 2-13). Damage of roots and stems led to a drastic decrease

in SPAD values in the leaves of IN937a treated plants and the non-treated control. On

the other hand, plant treated with IN937b showed only a moderate decrease compared

with SE76 and ASM treatments, although IN937b treated plant did not show any

symptoms at the time of data collection.

Discussion

In the present study, select Bacillus PGPR strains (IN937b, SE76, SE56, SE49,

1PC-11, and GB03) showed effective biological control activity against P. capsici which

causes crown and root rot on squash in this greenhouse. We also investigated the

effect of Bacillus PGPR strains IN937b, SE76 and IN937a on squash growth promotion.

However, the growth promotion effect only occurred under conditions without

fertilization. Moreover, in our system, PGPR strains IN937b and SE76 were consistent

in improving the performance in managing of Phytophthora crown and root rot on

squash, when applied at the concentration of 108 CFU/mL and a longer time period i.e.

the first PGPR application was made at 17 days after seeding.

Virulence Differentiation of P. capsici Isolates

In this study, the susceptibility of the squash cultivar, Senator, to Phytophthora

crown and root rot differed significantly among three isolates 121, 146, and 151. Based

on the results of disease severity and AUDPC value, there appeared to be differences

in virulence among these P. capsici isolates. Isolate 121 was highly virulent, causing

Page 51: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

51

crown and root rot. Isolate 146 was the least virulent based on slow symptom

development and lower disease incidence compared to plants treated with the other two

isolates. Different levels of virulence among P. capsici isolates led to similar conclusions

in previous studies with pepper (Oelke et al., 2003). Knowing that different levels of

virulence exist among P. capsici isolates infecting squash, we used isolate 151 with

moderate virulence caused symptoms to develop and tissue to collapse more than one

week after inoculation in the greenhouse.

In our tests, zoospores of P. capsici were used for inoculation. The development

of Phytophthora crown and root rot was affected by inoculum density (CafeFilho and

Duniway, 1996). When P. capsici was tested at 1, 2, 4, 6, 8, and 10 ×104 zoospores/mL

to determine the optimum concentration for the future experiments, high levels of crown

and root rot developed, and plant disease increased very quickly the first week after

inoculation when the concentrations were 6, 8, and 10 ×104 zoospores/mL. However,

only low levels of crown and root rot developed, and disease did not develop in some

plants inoculated with 1×104 zoospores/mL. Optimum disease levels were observed in

plants inoculated with P. capsici at the concentrations of 2 to 8 ×104 zoospores/mL.

Consequently, 2 ×104 zoospores/mL was chosen for further greenhouse experiments to

guaranty optimum plant infection and symptom development.

Control of Phytophthora Crown and Root Rot by Bacillus spp.

The development of biological control agents (BCAs) effective for control of

Phytophthora blight depends on the selection of an efficient strain, the introduction of

adequate bacterial cells into the soil, the interaction between the pathogen and the bio-

control strain, and the timing and method of application. In general, introducing large

populations of BCAs might provide a competitive advantage over residential

Page 52: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

52

microorganisms. Sometimes, the bacterial strains that showed the greatest potential for

plant growth promotion were not necessarily the best for inhibiting pathogens. Bacillus

plays an important role in the control of soil-borne plant pathogens such as in

suppressing of crown and root rot caused by P. capsici (Zhang et al., 2010).

Competition for nutrients and space between PGPR and the pathogen is considered to

be one of the mechanisms whereby PGPR biologically control plant diseases.

In our greenhouse trials, some PGPR strains provided strong suppression of

Phytophthora crown and root rot, whereas others showed only moderate activity or no

activity at all. Two PGPR strains IN937b and SE76 consistently showed significant

suppression of Phytophthora blight on squash, which agrees with earlier studies

(Zhang et al., 2010). These Bacillus strains were most effective among the PGPR

strains tested based on data from our greenhouse trials. PGPR strains IN937b and

SE76 showed strong disease suppression, which may make them promising candidates

for the development of BCAs that manage Phytophthora blight on squash. In particular,

B. subtilis IN937b has shown potential to suppress disease and to promote plant

growth, which makes it a potentially promising BCAs for management of Phytophthora

blight on squash.

The effect of disease suppression by IN937b and SE76 was greater at high

concentrations (108 CFU/mL) than low concentrations (107 and 106 CFU/mL), although

these two PGPR strains applied at low concentrations were also effective in some of the

greenhouse trials. The efficacy of the lowest concentration (105 CFU/mL) was

decreased due to the lack of adequate viable cells for optimal control of crown rot.

Page 53: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

53

Similar results were reported when 105 CFU/g root was present in the case of

Pseudomonas spp. (Raaijmakers et al., 1995a).

Enhanced Plant Growth by Bacillus spp.

The definition of PGPR comes from the fact that PGPR are able to increase plant

growth. Improvements of plant growth by PGPR on many different crops are well

documented (Kloepper et al., 2004). The ability of Bacillus strains IN937b, SE76 and

IN937a to increase squash plant biomass only occurred in the present study only when

plants were not fertilized, indicating that plant growth was strongly influenced by

fertilization. When the nutrient concentrations in the soil were sufficient, it was difficult to

promote plant growth by application of Bacillus PGPR. On the other hand, Bacillus spp.

could reduce fertilizer application rates due to its ability to increase nutrients available to

crops without any significant reduction of yield (Adesemoye et al., 2009b).

IN937b had a positive effect on plant growth, especially for lateral root growth.

Root biomass is commonly used to indicate growth promotion by PGPR and root

functions such as water and nutrient uptake (Lopez-Bucio et al., 2003). The increase in

lateral root formation by PGPR strains helps to explain the observed growth promotions.

Nutrient uptake by plant generally depends on the amount of root surface area

(Simunek and Hopmans, 2009).

Relation of Leaf Chlorophyll Content and Plant Health

The present study showed that chlorophyll content increased in leaves of IN937b

treated plants under greenhouse conditions, which agrees with the previous report

indicating that treating lettuce leaves with PGPR increased leaf chlorophyll content

(Arkhipova et al., 2005). In this present study, the increase in chlorophyll content in

IN937b treated squash plants indicated that foliar nitrogen content in plant leaves was

Page 54: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

54

increased by IN937b, which might indirectly influence plant growth improvement. The

increase in chlorophyll content might be due to the changes in the plant’s metabolism.

Five days after inoculation with P. capsici, squash plants began to show

symptoms of water-soaked lesions on the stem, especially the leaves turned yellow.

Treatments with PGPR strain SE76 significantly reduced disease severity compared

with the untreated control, while leaf chlorophyll content was also significantly higher

than in the untreated control. IN937a had no effect on disease suppression and similar

leaf chlorophyll content as the untreated control.

Optimization of Application Rate and Timing

Loss of biocontrol efficiency observed after 2 weeks is associated with a

decrease in the populations of PGPR (Strigul and Kravchenko, 2006). Maintain a high

number of PGPR as well as the timing of applications is possible to obtain an optimal

long-term control of diseases. This requires to assess the behavior of a specific BCA.

Understanding the behavior of a BCA in the environment may lead to improved

performance of the BCA, including nutrient use (Ji et al., 1997) and site preference (Fan

et al., 2011). If certain behaviors can be correlated with better disease control, then it

may be possible to promote these behaviors and, hence, improve disease control.

BCAs can be made into more reliable and predictable products by promoting the

behaviors associated with improved disease control and by optimizing application

methods.

Our results indicated that 108 CFU/mL is the optimum concentration for PGPR

strains IN937b and SE76 to consistently promote plant growth and suppress disease

development. Bacterial concentrations of PGPR above 108 CFU/mL were not tested

Page 55: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

55

since the highest vegetative cells of PGPR strains cultured within 24 h in NB are

approximately 1 × 108 CFU/mL.

It is important to determine a threshold of inoculum density for IN937b and SE76

at which they significantly suppress the Phytophthora blight on squash. Compared to

the low concentration of 105 CFU/mL, treatments with PGPR at high concentrations

demonstrated greater control efficacy. This is because symptom development was

delayed due to the high population of Bacillus spp. in the potting mix which may play a

long-term role in disease suppression. Some studies have suggested that specific

PGPR strains secreted secondary metabolites to prevent pathogen invasion and spread

(Compant et al., 2013). Thus, the low severity of disease in PGPR treatments was

probably due to low frequency of infection by the pathogen, reducing the potential for

pathogens to attack and colonize the root tissues of plants (Nihorimbere et al., 2010). In

general, two applications of PGPR were generally more effective than one treatment,

since the seedling stage highlighted the importance of maintaining a high population of

PGPR in the soil. This high population may reduce infection by pathogens (Poleatewich

et al., 2012). This requirement, however, poses a challenge to maintain high PGPR

populations in the soil because vegetative cells of PGPR strains lose efficacy quickly

under dry soil conditions due to poor colonization (West et al., 1985). Repeated

applications of PGPR may enhance the survival of PGPR in the soil (Schmidt et al.,

2004). In the present study, treatments of two applications likely helped to maintain high

populations of PGPR throughout the experiments.

The effect of age-related resistance in squash plants treated with PGPR remains

unknown (Sunwoo et al., 1996). Results from our study demonstrated that the timing of

Page 56: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

56

applying PGPR can be crucial in the efficacy of PGPR against infection by P. capsici.

Application to the stage 2 (stage 1 was the youngest stage) seedlings was remarkably

more efficient in disease suppression than when it was applied at the two older stages,

suggesting that the PGPR strains IN937b and SE76 exhibited strong reactions starting

from early growth stage. It also suggested that PGPR strains need time to establish a

relationship with the host plant. In particular, the younger seedlings developed more

crown rot disease, similar to that described by Morin for gorse (Ulex europaeus L.)

plants (Morin et al., 1998). The gorse plants tested were susceptible to pathogens, but

younger plants were more easily killed. The age of plants influenced the capability of

pathogens to infect and spread due to thickening of the cuticle and inducing resistance

response. However, the effectiveness of Bacillus strains in suppressing pathogens

decreases as plants mature.

During the past decade, several studies have focused on bio-control activities of

B. subtilis in controlling plant diseases. Antimicrobial metabolites are often associated

with BCAs in the genus Bacillus, which may act directly on the pathogen. Indigenous

bacteria that are antagonistic to plant pathogens in the soil could make a substantial

contribution to preventing plant diseases, and therefore represent an alternative to using

chemical pesticides. It is clear that BCAs are still too variable in their performance to be

successfully used as a common practice for disease management. This inconsistency

has been attributed to a number of factors including fluctuations in the environment and

the variable expression of genes involved in disease suppression and root colonization

by the BCA. The present suggests that further investigations are needed regarding the

role of environmental factors in modulating performance of Bacillus PGPR strains in

Page 57: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

57

order to improve their efficacy and performance in suppressing phytopathogens under

field conditions. A better understanding of the bio-control mechanisms may be helpful

for guiding future efforts in BCA development.

Page 58: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

58

Table 2-1. Relevant information for PGPR strains tested in this study

Tested PGPR strains a

Identity

Protected plant

Target pathogen Reference

SE34

B. safensis Rice Xanthomonas oryzae pv. oryzae

(Chithrashree et al., 2011)

SE49

B. safensis Cucumber Colletotrichum orbiculare

(Wei et al., 1996)

SE52

B. safensis Loblolly pine Cronartium quercuum f. sp. fusiforme

(Enebak and Carey, 2000)

SE56 Lysinibacillus

boronitolerans Pepper

Colletotrichum gloeosporioides

(Jetiyanon and Kloepper, 2002)

SE76

B. safensis Tomato Colletotrichum gloeosporioides

(Jetiyanon and Kloepper, 2002)

IN937a

B. subtilis Tomato Ralstonia solanacearum

(Jetiyanon, 2007)

IN937b

B. subtilis Pepper Sclerotium rolfsii (Jetiyanon, 2007)

INR7

B. pumilus Pepper Xanthomonas axonopodis pv. vesicatoria

(Yi et al., 2013)

T4

B. safensis Tobacco Peronospora tabacina

(Zhang et al., 2002)

GB03

B. subtilis Arabidopsis thaliana

Pseudomonas syringae pv. tomato DC3000

(Kumar et al., 2012)

1PC-11

B. macauensis Squash Phytophthora capsici

(Zhang et al., 2010)

1PN-19

B. subtilis Squash Phytophthora capsici

(Zhang et al., 2010)

Companion (GB03)

B. subtilis

Arabidopsis thaliana

Pseudomonas syringae pv. tomato DC3000

(Kumar et al., 2012)

BU EXP 1216S (MBI600)

B. subtilis Rice

Rhizoctonia solani

(Kumar et al., 2010)

BU EXP 1216C (MBI600)

B. subtilis Squash

Podosphaera xanthii

(Zhang et al., 2011)

Actinovate AG

Streptomyces lydicus WYEC 108

Soybean Sclerotinia sclerotiorum

(Zeng et al., 2012a)

a PGPR strains were provided by Dr. Joseph W. Kloepper.

Page 59: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

59

Table 2-2. Disease development in squash inoculated with different isolates of P. capsici

Zoospore concentration a

Disease severity b AUDPC c

121 146 151 121 146 151

102 0.0 b d 0.0 b 0.0 b 0.0 e 0.0 e 0.0 e

103 3.3 a 0.0 b 0.0 b 1.6 d 0.0 e 0.0 e

104 5.0 a 0.0 b 3.8 a 5.3 b 0.0 e 3.8 c

105 5.0 a 3.8 a 5.0 a 7.5 a 1.9 d 7.3 a

a Zoospore solutions of three isolates (121, 146, 151) were applied in the pots as a soil drench when plants had four completely expanded leaves. Each plant was inoculated with 5 mL of P. capsici inoculum. b. Disease severity was rated based on the scale from 0 (healthy) to 5 (dead). Data were collected 7 days after pathogen inoculation. c AUDPC was calculated based on the disease severity rated 5, 6, and 7 days after pathogen inoculation d In each column in disease severity, each value represents the mean of four replicates. Values followed by the same letter are not significantly different according to Fisher’s protected Least Significant Difference (LSD) test (P = 0.05).

Page 60: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

60

Table 2-3. Effect of the concentration of P. capsici isolate 151 on disease development

Zoospore concentration (×104) a

Disease severity b Time after inoculation (Day)

AUDPC c 5 7 9

1 0.6 a d 3.0 a 3.0 b 9.6 b

2 0.6 a 3.0 a 3.8 ab 10.4 ab

4 0.8 a 3.0 a 3.5 ab 10.3 b

6 1.1 a 2.8 a 3.3 b 9.9 b

8 1.8 a 3.0 a 4.5 a 12.3 ab

10 1.8 a 4.0 a 4.5 a 14.3 a

a The concentration of P. capsici inoculum was diluted to 1, 2, 4, 6, 8, and 10 × 104 zoospores/mL. Squash plants were inoculated with 5 mL of P. capsici inoculum. b Disease severity was rated based on the scale from 0 (healthy) to 5 (dead). c AUDPC was calculated based on the disease severity rated 5, 7, and 9 days after pathogen inoculation. d Each value represents the mean of four replicates. Means followed by the same letter within a column are not significantly different according to Fisher’s protected LSD test (p = 0.05).

Page 61: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

61

Table 2-4. Efficacy of PGPR (1-8 PGPR strains) on suppression of Phytophthora crown and root rot under greenhouse conditions

Treatment a AUDPC b

Trial 1 Trial 2 Trial 3

SE34 14.4 a c 17.4 ab 24.8 a

SE52 10.0 b 15.6 b 24.4 a

SE56 12.0 ab 14.7 b 20.1 b

SE76 5.8 c 15.6 b 17.1 c

INR7 13.6 a 15.3 b 22.9 a

1PN-19 13.6 a 15.3 b 23.7 a

IN937a 13.6 a 15.9 b 23.9 a

IN937b 5.6 c 15.3 b 19.6 b

Presidio® 1.0 d 0.0 c 0.0 d Non-treated control 12.8 ab 19.2 a 23.9 a

a PGPR strains were applied twice (1.0 ×108 CFU/mL; 20mL) as a soil drench. Presidio® (1.2 mL/L; 20mL) was applied as a standard fungicide which provide control of Phytophthora blight. b P. capsici inoculum on crown and roots was diluted to 2 × 104 zoospores/mL. Plants were inoculated by 5 mL of P. capsici inoculum. Disease severity was collected based on the rating scale from 0 (healthy) to 5 (dead). AUDPC calculated based on the disease severity. Trial 1: 11, 13, and 15 days after pathogen inoculation; Trial 2: 9, 12, and 15 days after pathogen inoculation; Trial 3: 7, 9, and 14 days after pathogen inoculation. c Each value represents the mean of five replicates. Means followed by the same letter within a column are not significantly different according to Fisher’s protected LSD test (p = 0.05).

Page 62: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

62

Table 2-5. Efficacy of PGPR (9-16 PGPR strains) on suppression of Phytophthora crown and root rot under greenhouse conditions

Treatment a AUDPC b

Trial 1 Trial 2 Trial 3

SE49 15.6 ab c 10.3 bc 5.8 b

Com 9.6 cd 12.4 abcd 8.6 ab

AG 14.4 ab 16.1 a 8.4 ab

GB03 8.6 d 7.8 c 8.8 ab

1216S 14.4 ab 13.2 abc 8.4 ab

1216C 14.8 ab 13.8 ab 9.6 ab

T4 12.6 bcd 9.8 bc 11.4 a

1PC-11 13.6 bcd 11.7 bcd 8.2 ab

Presidio® 0.0 e 0.0 d 0.0 c Non-treated control 18.2 a 16.8 a 10.2 ab

a PGPR strains were applied twice (1.0 ×108 CFU/mL; 20mL) as a soil drench. Presidio® (1.2 mL/L; 20mL) was applied as a standard fungicide which provide control of Phytophthora blight. b P. capsici inoculum on crown and roots was diluted to 2 × 104 zoospores/mL. Plants were inoculated by 5 mL of P. capsici inoculum. Disease severity was collected based on the rating scale from 0 (healthy) to 5 (dead). AUDPC calculated based on the disease severity. Trial 1: 11, 15, and 17 days after pathogen inoculation; Trial 2: 12, 14, and 17 days after pathogen inoculation; Trial 3: 7, 9, and 11 days after pathogen inoculation. c Each value represents the mean of five replicates. Means followed by the same letter within a column are not significantly different according to Fisher’s protected LSD test (p = 0.05).

Page 63: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

63

Table 2-6. Effect of PGPR on plant growth under greenhouse conditions

Treatment a

Parameters of plant growth b

Plant height cm

Stem diameter mm

Fresh shoot g

Dry shoot g

Fresh root g

Dry root g

Root elongation cm

IN937b 32.3 ab c 6.4 a 33.6 ab 3.9 a 4.6 a 0.3 a 14.3 c

SE76 31.6 ab 6.2 a 34.0 ab 3.9 a 4.3 a 0.3 a 15.0 bc

IN937a 33.0 a 6.6 a 34.8 a 4.2 a 4.3 a 0.2 a 15.3 b

Actigard® 32.9 a 6.5 a 32.6 ab 4.1 a 4.3 a 0.2 a 14.4 c Non-treated control

30.9 ab 6.3 a 34.1 ab 4.0 a 4.4 a 0.2 a 16.2 a

Non-treated control (-F)

30.3 b 6.2 a 30.7 b 2.9 b 3.3 b 0.1 b 13.3 d

a PGPR strains were applied twice (1.0 ×108 CFU/mL; 20 mL) as a soil drench. Actigard® (30 mg/L; 20 mL) was applied as a positive SAR inducer. Water soluble fertilizer 20:20:20 once a week (3.5 g/L; 20 mL per plant) was applied to all treatments except for the untreated control (-F). The untreated control (-F) represents the plants grown without fertilization. b Plants were investigated at 40 days after planting, the root and shoot were separated, and the biomass was measured. The length of root tissues was recorded to determine the root elongation. c Each value represents the mean of ten replicates. Means followed by the same letter within a column are not significantly different according to Fisher’s protected LSD test (p = 0.05).

Page 64: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

64

Table 2-7. Effect of IN937b concentration on crown and root rot disease

Treatment a

Disease severity b Time after inoculation (Day) AUDPC c

3 4 5 6

105 CFU/mL 0.3 a d 1.4 b 2.1 ab 3.5 ab 5.4 ab

106 CFU/mL 0.1 b 1.2 b 1.9 bc 3.2 b 4.8 bc

107 CFU/mL 0.0 b 1.0 b 2.0 bc 3.7 ab 4.8 bc

108 CFU/mL 0.0 b 1.0 b 1.5 c 2.4 c 3.7 c

Non-treated control 0.0 b 2.1 a 2.6 a 4.0 a 6.7 a

Actigard® 0.0 b 0.0 c 0.0 d 0.0 d 0.0 d

Presidio® 0.0 b 0.0 c 0.0 d 0.0 d 0.0 d a PGPR were applied twice as a soil drench. The concentration of IN937b was adjusted to 105, 106, 107, and 108 CFU/mL, and 20 mL was applied to each plant. Presidio® (1.2 mL/L; 20mL/plant) and Actigard® (30 mg/L; 20mL/plant) were applied as a positive fungicide and SAR control, respectively, of Phytophthora blight. b P. capsici inoculum on crown and roots was diluted to 2 × 104 zoospores/mL. Plants were inoculated by 5 mL of P. capsici inoculum. Disease severity was collected at 3, 4, 5, and 6 days after pathogen inoculation based on the rating scale from 0 (healthy) to 5 (dead). c AUDPC calculated based on the disease severity 3, 4, 5, and 6 days after pathogen inoculation. d Each value represents the mean of ten replicates. Means followed by the same letter within a column are not significantly different according to Fisher’s protected LSD test (p = 0.05).

Page 65: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

65

Table 2-8. Effect of SE76 concentration on crown and root rot disease

Treatment a

Disease severity b Time after inoculation (Day) AUDPC c

3 4 5 6

105 CFU/mL 0.0 b d 1.0 a 1.4 ab 2.9 b 3.9 ab

106 CFU/mL 0.5 a 1.1 a 1.7 ab 3.8 a 4.9 ab

107 CFU/mL 0.0 b 1.5 a 1.0 b 2.7 bc 3.8 ab

108 CFU/mL 0.0 b 1.1 a 1.1 b 2.2 c 3.2 b

Non-treated control 0.0 b 1.8 a 2.1 a 3.8 a 5.8 a

Actigard® 0.0 b 0.0 b 0.0 c 0.0 d 0.0 c

Presidio® 0.0 b 0.0 b 0.0 c 0.0 d 0.0 c a PGPR were applied twice as a soil drench. The concentration of SE76 was adjusted to 105, 106, 107, and 108 CFU/mL, and 20 mL was applied to each plant. Presidio® (1.2 mL/L; 20mL/plant) and Actigard® (30 mg/L; 20mL/plant) were applied as a positive fungicide and SAR control, respectively, of Phytophthora blight. b P. capsici inoculum on crown and roots was diluted to 2 × 104 zoospores/mL. Plants were inoculated by 5 mL of P. capsici inoculum. Disease severity was collected at 3, 4, 5, and 6 days after pathogen inoculation based on the rating scale from 0 (healthy) to 5 (dead). c AUDPC calculated based on the disease severity 3, 4, 5, and 6 days after pathogen inoculation. d Each value represents the mean of ten replicates. Means followed by the same letter within a column are not significantly different according to Fisher’s protected LSD test (p = 0.05).

Page 66: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

66

Table 2-9. Effect of IN937a concentration on crown and root rot disease

Treatment a

Disease severity b Time after inoculation (Day) AUDPC c

3 4 5 6

105 CFU/mL 0.0 b d 1.3 c 1.6 b 3.6 bc 4.7 b

106 CFU/mL 0.0 b 1.6 bc 1.6 b 2.9 d 4.6 b

107 CFU/mL 0.4 a 2.2 a 2.8 a 4.0 a 7.2 a

108 CFU/mL 0.5 a 2.3 a 2.6 a 3.4 c 6.9 a

Non-treated control 0.0 b 1.9 ab 2.4 a 3.9 ab 6.3 a

Actigard® 0.0 b 0.0 d 0.0 c 0.0 e 0.0 c

Presidio® 0.0 b 0.0 d 0.0 c 0.0 e 0.0 c

a PGPR were applied twice as a soil drench. The concentration of IN937a was adjusted to 105, 106, 107, and 108 CFU/mL, and 20 mL was applied to each plant. Presidio® (1.2 mL/L; 20mL/plant) and Actigard® (30 mg/L; 20mL/plant) were applied as a positive fungicide and SAR control, respectively, of Phytophthora blight. b P. capsici inoculum on crown and roots was diluted to 2 × 104 zoospores/mL. Plants were inoculated by 5 mL of P. capsici inoculum. Disease severity was collected at 3, 4, 5, and 6 days after pathogen inoculation based on the rating scale from 0 (healthy) to 5 (dead). c AUDPC calculated based on the disease severity 3, 4, 5, and 6 days after pathogen inoculation. d Each value represents the mean of ten replicates. Means followed by the same letter within a column are not significantly different according to Fisher’s protected LSD test (p = 0.05).

Page 67: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

67

Table 2-10. Effect of Bacillus spp. application timing (stage 1: seed) on crown and root rot disease

Treatment a

Disease severity b Time after inoculation (Day) AUDPC c

3 5 6 7

IN937b 0.4 b d 1.4 b 2.8 bc 3.6 bc 7.2 bc

SE76 0.0 b 1.1 b 2.5 c 3.0 c 5.3 c

IN937a 0.5 b 2.5 a 3.6 ab 4.6 ab 9.5 ab

Non-treated control 1.3 a 3.0 a 3.9 a 5.0 a 11.6 a

Actigard® 0.0 b 0.0 c 0.0 d 0.0 d 0.0 d

Presidio® 0.0 b 0.0 c 0.0 d 0.0 d 0.0 d a PGPR strains were applied twice (seed soaking and 10 days after emergence) as a soil drench (1.0 ×108 CFU/mL; 20mL). Presidio® (1.2 mL/L; 20mL) and Actigard® (30 mg/L; 20mL/plant) was applied as the positive control of Phytophthora blight. b P. capsici inoculum on crown and roots was diluted to 2 × 104 zoospores/mL. Plants were inoculated by 5 mL of P. capsici inoculum. Disease severity was collected at 3, 5, 6, and 7 days after pathogen inoculation based on the rating scale from 0 (healthy) to 5 (dead). c AUDPC calculated based on the disease severity 3, 5, 6, and 7 days after pathogen inoculation. d Each value represents the mean of ten replicates. Means followed by the same letter within a column are not significantly different according to Fisher’s protected LSD test (p = 0.05).

Page 68: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

68

Table 2-11. Effect of Bacillus spp. application timing (stage 2: 10 days after emergence) on crown and root rot disease

Treatment a

Disease severity b Time after inoculation (Day) AUDPC c

3 5 7 9

IN937b 0.8 ab d 1.2 b 1.2 b 3.1 b 8.6 b

SE76 0.7 bc 1.2 b 1.3 b 2.5 b 8.0 b

IN937a 1.6 a 2.5 a 2.7 a 3.0 b 14.9 a

Non-treated control 1.2 ab 3.0 a 3.5 a 4.7 a 18.8 a

Actigard® 0.0 c 0.0 c 0.0 c 0.0 c 0.0 c

Presidio® 0.0 c 0.0 c 0.0 c 0.0 c 0.0 c a PGPR strains were applied twice (10 days and 17 days after emergence) as a soil drench (1.0 ×108 CFU/mL; 20mL). Presidio® (1.2 mL/L; 20mL) and Actigard® (30 mg/L; 20mL/plant) was applied as the positive control of Phytophthora blight. b P. capsici inoculum on crown and roots was diluted to 2 × 104 zoospores/mL. Plants were inoculated by 5 mL of P. capsici inoculum. Disease severity was collected at 3, 5, 7, and 9 days after pathogen inoculation based on the rating scale from 0 (healthy) to 5 (dead). c AUDPC calculated based on the disease severity 3, 5, 7, and 9 days after pathogen inoculation. d Each value represents the mean of ten replicates. Means followed by the same letter within a column are not significantly different according to Fisher’s protected LSD test (p = 0.05).

Page 69: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

69

Table 2-12. Effect of Bacillus spp. application timing (Stage 3: 17 days after emergence) on crown and root rot disease

Treatment a

Disease severity b Time after inoculation (Day) AUDPC c

7 9 11 13

IN937b 1.4 b d 1.4 c 2.6 b 3.0 c 11.3 c

SE76 1.8 ab 2.1 b 2.7 b 3.4 bc 14.1 b

IN937a 2.0 a 2.1 b 2.9 b 3.9 b 15.9 ab

Non-treated control 1.9 a 2.8 a 3.6 a 5.0 a 18.6 a

Actigard® 0.5 c 0.7 d 1.0 c 1.1 d 4.9 d

Presidio® 0.0 d 0.0 e 0.0 d 0.0 e 0.0 e a PGPR strains were applied twice (17 days and 24 days after emergence) as a soil drench (1.0 ×108 CFU/mL; 20mL). Presidio® (1.2 mL/L; 20mL) and Actigard® (30 mg/L; 20mL/plant) was applied as the positive control of Phytophthora blight. b P. capsici inoculum on crown and roots was diluted to 2 × 104 zoospores/mL. Plants were inoculated by 5 mL of P. capsici inoculum. Disease severity was collected at 7, 9, 11, and 13 days after pathogen inoculation based on the rating scale from 0 (healthy) to 5 (dead). c AUDPC calculated based on the disease severity 7, 9, 11, and 13 days after pathogen inoculation. d Each value represents the mean of ten replicates. Means followed by the same letter within a column are not significantly different according to Fisher’s protected LSD test (p = 0.05).

Page 70: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

70

Table 2-13. Effect of PGPR on chlorophyll content in squash leaves under greenhouse conditions

Treatment a Disease severity b SPAD reading c

IN937b 0.0 b d 21.7 b

SE76 0.3 b 29.3 a

IN937a 1.0 a 16.6 b

Actigard® 0.0 b 32.9 a

Non-treated Control 1.0 a 16.8 b

a PGPR were applied twice as a soil drench. The concentration of IN937b was adjusted to 108 CFU/mL, and 20 mL was applied to each plant. Actigard® (30 mg/L; 20mL/plant) was applied as a positive SAR control of Phytophthora blight. b Phytophthora capsici inoculum was diluted to 2 × 104 zoospores/mL. Plants were inoculated with 5 mL of P. capsici inoculum. Disease severity was rated at 5 days after pathogen inoculation based on the rating scale from 0 (healthy) to 5 (dead). c The SPAD measurement were carried out 5days after inoculation. SPAD values were measured on three leaves at the same position on each plant. d Each value represents the mean of 5 replicates. Means followed by the same letter within a column are not significantly different according to Fisher’s protected LSD test (p = 0.05).

Page 71: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

71

Figure 2-1. Regression analysis of the concentration of P. capsici isolate 151 on

disease development. AUDPC was calculated based on the disease severity rated 5, 7, and 9 days after pathogen inoculation. Each value represents the mean of four replicates. Means followed by the same letter within a column are not significantly different according to Fisher’s protected LSD test (p = 0.05).

y = 0.4482x + 8.8175R² = 0.741

0

2

4

6

8

10

12

14

16

0 2 4 6 8 10 12

AU

DP

C

Zoospore concentration (×104 /mL)

Page 72: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

72

A.

B.

C.

Figure 2-2. Regression analysis of Effect of PGPR strains concentration on crown and root rot disease. A: IN937b, B: SE76, C: IN937a. PGPR were applied twice as a soil drench. The concentration of PGPR was adjusted to 105, 106, 107, and 108 CFU/mL, and 20 mL was applied to each plant. Presidio® (1.2 mL/L; 20mL/plant) and Actigard® (30 mg/L; 20mL/plant) were applied as a positive fungicide and SAR control, respectively, of Phytophthora blight. P. capsici inoculum on crown and roots was diluted to 2 × 104 zoospores/mL. Plants were inoculated by 5 mL of P. capsici inoculum. Disease severity was collected at 3, 4, 5, and 6 days after pathogen inoculation based on the rating scale from 0 (healthy) to 5 (dead). AUDPC calculated based on the disease severity 3, 4, 5, and 6 days after pathogen inoculation. Values with same letter within a column are not significantly different according to Fisher’s protected LSD test (p = 0.05). Vertical bars represent mean ± S.D. (n = 5).

y = -0.3371x + 6.833R² = 0.9209

0.0

5.0

10.0

0.0 2.0 4.0 6.0 8.0 10.0A

UD

PC

IN937b concentration (log10 CFU/mL)

y = -0.2892x + 5.8237R² = 0.7674

0.0

5.0

10.0

0.0 2.0 4.0 6.0 8.0 10.0

AU

DP

C

SE76 concentration (log10 CFU/mL)

y = 0.0582x + 5.6371R² = 0.0220.0

5.0

10.0

0.0 2.0 4.0 6.0 8.0 10.0

AU

DP

C

IN937a concentration (log10 CFU/mL)

Page 73: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

73

A

B

Figure 2-3. Effect of IN937b on squash root and shoot dry weight. A, Average dry weight of squash shoots measured 30 days after emergence. B, Average dry weight of squash roots measured 30 days after emergence. In the fertilization experiment, plants were fertilized with water soluble fertilizer (20:20:20, 3.5 g/mL, 20 mL/plant) once each week. Values with same letter within a column are not significantly different according to Fisher’s protected LSD test (p = 0.05). Vertical bars represent mean ± S.D. (n = 10).

a aa

b

0.00

0.50

1.00

1.50

2.00

2.50

3.00

3.50

4.00

4.50

With fertilization Without fertilization

Dry shoot

Dry

sho

ot

wei

ght

(g)

IN937b Non-treated control

a

aa

b

0.00

0.05

0.10

0.15

0.20

0.25

With fertilization Without fertilization

Dry root

Dry

ro

ot

wei

ght

(g)

IN937b Non-treated control

Page 74: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

74

CholorophyII

IN937b SE76 IN937a ASM CK

SP

AD

0

10

20

30

40

50

a

c

bc

ab

c

Figure 2-4. Chlorophyll content in leaves of squash plants treated with PGPR. PGPR

strains were applied twice as a soil drench to squash roots. The concentration of IN937b, SE76, and IN937a was adjusted to 108 CFU/mL, and 20 mL was applied to each plant. Actigard® (30 mg/L; 20mL/plant) was applied as a positive SAR control of Phytophthora blight. The SPAD measurement were carried out 5days after inoculation. SPAD values were measured on three leaves at same position on each plant. Values with same letter within a column are not significantly different according to Fisher’s protected LSD test (p = 0.05). Vertical bars represent mean ± S.D. (n = 5).

Page 75: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

75

CHAPTER 3 THE ROLE OF BIOFILM FORMATION AND COLONIZATION IN THE RHIZOSPHERE

BY BACILLUS SUBTILIS IN937B WITH SYNTHESIS OF SURFACTINS

Introduction

Plants attract beneficial rhizobacteria such as PGPR by exudates released from

plant roots (Raaijmakers et al., 2009b). PGPR are directly involved in the nutrition,

health and quality of plants. They have a role in cycling carbon, nitrogen, and

phosphorus, as well as suppressing soil-borne plant pathogens (Garbeva et al., 2004).

The PGPR-mediated nutrient cycling is related to increase nitrogen-fixation and the

availability of soluble phosphates in the soil. The mechanism by biocontrol-PGPR

protects plant roots from pathogens include microbial antagonism, biofilm formation and

the stimulation of induced systemic resistance in the host plants. The functional

activities and community diversities of PGPR are positively affected by plant roots.

Roots are able to release a wide range of organic and inorganic compounds into the

rhizosphere including sugars, amino acids, phenolics, enzymes, siderophores, etc.

(Pinton et al., 2001). The root exudates create specific habitat conditions suitable for

microbial development, and play an important role in affecting microbial populations,

availability of nutrients,

PGPR colonize the plant roots steadily by chemotaxis attraction and biofilm

formation (Yaryura et al., 2008). The process includes bacterial attachment to the root,

movement and proliferation of PGPR along the elongating root, and penetration of the

epidermis (Liu et al., 2011). PGPR are likely to locate roots through particular signals

emanating from the roots. Carbohydrates and amino acids also stimulate PGPR to

move towards roots (Bais et al., 2006). PGPR can also produce biofilms to initiate

beneficial interactions with host plants. Biofilm foster increased resistance to certain

Page 76: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

76

environmental stresses as well as antimicrobial tolerance (Danhorn and Fuqua, 2007).

Biofilm is a multicellular structure which adheres to the root surface as a protective

mechanism against adverse environmental conditions. The formation of biofilm is

associated with root colonization by bacteria such as PGPR (Ahmad et al., 2011).

Root colonization by PGPR is considered integral for PGPR to be successful as

biocontrol agents (Whipps, 2001). The biocontrol effect depends on population densities

of PGPR applied for suppressing pathogens (Kumar et al., 2011a). For example, P.

fluorescens Pf1TZ could protect grape vine (Vitis vinifera L. cv. Chardonmy) plantlets

against Botrytis cinerea after 3 weeks of bacterial inoculation (Kilani-Feki et al., 2010).

The population density of Pf1TZ colonized roots required for the bacteria to protect

plants from the pathogen was 104 CFU/mg of fresh roots. However, shortly after PGPR

inoculation, the population declines progressively over time (vanVeen et al., 1997).

Natural complexity of the rhizosphere (soil pH, temperature, nutrients, competitors, and

predators) poses challenges to PGPR survival in different soil environments (Strigul and

Kravchenko, 2006). As effective biocontrol agents, PGPR must establish themselves in

the rhizosphere, which is affected by many abiotic and biotic factors.

Bacillus PGPR strains are widely considered to be capable of colonizing roots

(Compant et al., 2010). For example, B. pumilus SE34 grew on the root surface and

intercellularly penetrated roots of pea plants, allowing B. pumilus SE34 to get nutrients

from the plant and compete for root colonization (Benhamou et al., 1996). Many Bacillus

strains are capable of adopting and switching between free-living (planktonic) and

biofilm (surface-attached) forms. The planktonic form can rapidly proliferate as free-

floating individual cells, whereas the biofilm form are a unit of well-organized bacterial

Page 77: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

77

cells which helps to tolerate adverse environments (O'Toole et al., 2000). As highly

structured microbial communities, biofilm helps bacterial strains adhere to a soild

surface where nutrients in an aqueous environment tent to concentrate. It is a survival

strategy that designed to anchor PGPR in a nutritionally advantageous environment and

permit their escape when growth factors have been exhausted.

Bacillus strains have the ability to form biofilm both in the laboratory and in nature

(O'Toole et al., 2000). Under laboratory conditions, B. subtilis 3610 produced

architecturally complex colonies on solid medium and formed floating biofilm called

pellicles at the interface of the liquid medium and air (Lopez et al., 2009). Bacillus

subtilis 3610 forms robust biofilms on the root surfaces of tomato plants (Chen et al.,

2013). Biofilm formation by Bacillus plays an important role in protecting plants from

pathogen attacks. B. subtilis ATCC 6051 suppress Pseudomonas syringae on

Arabidopsis by competition on colonization (Bais et al., 2004). Paenibacillus polymyxa

B1 and B2 protect Arabidopsis thaliana from infections by Pythium aphanidermatum by

biofilm formation on root tips (Timmusk et al., 2005).

Biofilm formation by Bacillus spp. is associated with production of surfactins,

which are well-known secondary metabolites with antifungal activities. Surfactins

produced by Bacillus spp. are required for biofilm formation and for the successful bio-

control of plant pathogens. A knock-out mutant of surfactin synthase resulted in the loss

of swarming motility of B. subtilis which involves the differentiation of vegetative cells

and population migration across solid root surfaces (Kinsinger et al., 2003; Koumoutsi et

al., 2004). Mutants deficient in surfactin synthesis were unable to suppress pathogens

(Bais et al., 2004b). B. amyloliquefaciens FZB 42 produces surfactins. Knock-out

Page 78: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

78

mutants of surfactin production were unable to repress Fusarium oxysporum grown on

agar plates (Koumoutsi et al., 2004). Moreover, surfactins obtained from B.

amyloliquefaciens strain S499 were capable of triggering an oxidative burst which is

mediated perception by tobacco root cells that induce a defensive state in the plant

(Henry et al., 2011). These results indicate that surfactins play an important role in root

colonization by Bacillus spp. and in disease suppression via direct inhibition and

stimulation of the host plant’s immune system (Ongena and Jacques, 2008).

The possible role that surfactins play in biofilm formation includes functioning as

bio-surfactants, which have hydrophobic properties for water-insoluble substrates (Ron

and Rosenberg, 2001), and influences on attachment and detachment of

microorganisms to roots (Branda et al., 2001). Bacillus subtilis has long served as the

powerful model organism for laboratory studies the mechanisms of biofilm formation

(Vlamakis et al., 2013). However, studies need to be done to expand our understanding

of Bacillus PGPR biofilm formation on plant roots, or whether biofilm influences root

colonization by Bacillus PGPR.

In the present study, we focused on in vitro assays of biofilm formation and

colonization of selected Bacillus PGPR strains in the rhizosphere of squash. The

selected strains of Bacillus PGPR were able to suppress Phytophthora blight disease

caused by P. capsici through soil inoculation (see Chapter 2). The aim of this work were

the following: (1) to evaluate the ability of Bacillus PGPR to form biofilm in vitro, (2) to

investigate the timing and concentration of B. subtilis IN937b required for biofilm

formation, (3) to determine biosynthesis of surfactin characteristics of PGPR strains

Page 79: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

79

corresponding to biofilm formation, (4) to assess whether IN937b would colonize and

form biofilm on squash roots.

Materials and Methods

PGPR Strains and Media

All Bacillus PGPR strains used in this study were stored in nutrient broth (NB)

containing 15% glycerol at -80 ℃ prior to use. Bacterial cultures were prepared first by

streaking each PGPR strain taken from ultra-cold storage onto nutrient agar (NA; Difco)

plates, then incubating them at 28 ℃ for 24 h.

For pellicle formation experiments, a minimal medium MSgg (5 mM potassium

phosphate, pH 7, 100 mM MOPS, pH 7, 2 mM MgCl2, 700 µM CaCl2, 50 µM MnCl2, 50

µM FeCl3, 1 µM ZnCl2, 2 µM thiamine, 0.5% glycerol, 0.5% glutamate, 50 µg ml-1

tryptophan, 50 µg ml-1 phenylalanine and 50 µg ml-1 threonine) was used (Branda et al.,

2001).

Surfactin production by PGPR was elicited using No. 3S medium containing 10 g

of soybean peptone, 10 g of glucose, 1 g of KH2PO4, and 0.5 g of MgSO4·7H2O per liter

of the medium (pH 7.0) (Tsuge et al., 2005).

Bacteria were grown in Luria broth (LB) medium for DNA extraction: 10 g of

bacto tryptone (Difco), 5 g of yeast extract (Difco), 10 g of NaCl per liter of the medium

(pH 7.0) (Gellert et al., 1976).

Pellicle Formation

PGPR strains were grown in 5 mL Nutrient Broth (NB; Difco) medium on a rotary

shaker incubator at 150 rpm overnight at 28 ℃. Five microliters of cell cultures were

transferred into 1 mL liquid MSgg medium in 10 mL plastic tubes (or 30 mL liquid MSgg

medium in 100 mL beakers). The plastic tubes and beakers were placed in the

Page 80: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

80

incubator at 28 ℃ for 3 days when the pellicle, a biofilm formed at the air-liquid

interface, was observed and photographed with a Canon EOS Rebel T1i digital camera

(Canon, Tokyo, Japan).

Colony Morphology Assay

For colony architecture analysis on the solid media, PGPR strains were streaked

on NA plates to get the single colonies. The plates were incubated at 28 ℃ for at least

24 h to observe morphology of the single colonies. The architecture was clearly

observable under a dissecting microscope (Olympus DF Planapo Ⅸ, Olympus, Tokoya,

Japan). Colonies were imaged using a Canon EOS Rebel T1i digital camera.

Biofilm Production Assay in Microtiter Plates

Bacillus PGPR strains were grown in 10 mL NB medium at 28 ℃ on a rotary

shaker incubator at 150 rpm overnight. Biofilm production assays were performed with

MSgg medium (Branda et al., 2001). The inoculum for the microtiter plate assays was

obtained by diluting the Bacillus cells to 106 CFU/mL. Final inocula were diluted to an

identical density and calculated based on OD600. One hundred microliters of the

inoculum were grown in each well of the 6-well polyvinylchloride (PVC) microtiter plates,

each containing 10 mL liquid MSgg medium. The bacterial cells were grown at room

temperature for 7 days when they reached the stationary growth phase. Biofilm

production and morphologies were imaged using a Canon EOS Rebel T1i digital

camera.

DNA Extraction and Polymerase chain reaction (PCR) Analysis

Bacillus strains were inoculated into 10 mL LB medium and incubated on a rotary

shaker incubator at 150 rpm overnight at 28 ℃. DNA extraction was carried out using

Page 81: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

81

the Qiagen kit (Qiagen QIAamp DNA mini kit, Qiagen, Valencia, CA) according to the

manufacturer’s instructions. Polymerase chain reaction (PCR) reactions were carried

out based on the methodology and specific primers established by Hsie et al. (2004).

Primers used are listed in Table 3-1, that amplify genes encoding the antibiotics

synthetases of iturin A, bacillomycin D, surfactin, zwittermicin A, and fengycin (Hsieh et

al., 2004; Ramarathnam et al., 2007).

The PCR reaction was performed in a final volume of 25 µL containing 2.5 µL of

10X PCR buffer, 0.5 µL of dNTP mix, 1.0 µL each of PCR primers, 0.5 µL of Taq DNA

polymerases, and 1 µL each of template genomic DNA. Amplification was done with a

conventional thermal cycler (BioRad, Laboratories, Hercules, CA) with an initial

denaturing at 94 ℃ for 1 min, annealing at 46 ℃ for 30s and extension at 72 ℃ for 1

min, for a total of 30 cycles, and followed by a final extension of 72 ℃ for 10 min.

Following amplification, the PCR reaction mix was separated through 1% agarose gel

electrophoresis.

PCR products were purified using a Qiagen PCR purification kit (Qiagen,

Valencia, CA). DNA sequencing was conducted at the Interdisciplinary Center for

Biotechnology Research, University of Florida, Gainesville, FL. Sequence data were

processed and analyzed with Basic Logal Alignment Search Tool (BLAST) software of

National Center for Biotechnology Information (NCBI)

(http://www.ncbi.nlm.nih.gov/BLAST). The amino acid sequence of sfp in IN937b was

determined using the translation tool in the software DNAMAN version 7 (Lynnon

Biosoft, Quebec, Canada). The amino acid sequences for B. amyloliquefaciens FZB42,

B. cereus E33L, B. megaterium QM B1551, B. polymyxa SC2, B. pumilus ATCC 7061,

Page 82: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

82

B. subtlis 168, B. subtilis IN937b, Brevibacillus brevis NBRC 100599, and B.

licheniformis ATCC14580 were obtained from GenBank (http://www.ncbi.nlm.nih.gov/).

Multiple sequence alignment and Phylogenetic analysis were done with DNAMAN

version 7. Phylogenetic analysis was performed using a non-rooted neightbor-joining

tree with Poisson-correction (Bootstrap values: 1000 replicates) distances.

Surfactin Extraction

Surfactin was extracted from PGPR cultures according to Yakimov et al. (1995).

Bacterial cells were grown in No. 3S medium on a rotary shaker incubator at 150 rpm

for 48h at 28 ℃. Cells were separated from No.3S medium by centrifugation at 3,000

rpm for 30 min at room temperature. The supernatant was subjected to acid

precipitation by adding 6 M HCl to achieve a final pH of 2.0 and allowing the precipitate

to form at 4 ℃ overnight (Yakimov et al., 1995). The pellet was collected by

centrifugation at 3,000 rpm for 30 min at room temperature, and washed two times with

DI water (pH 2.0, adjusted by 6M HCl). The pellet was dissolved in methanol (pH 7.0,

adjusted by 1 M NaOH) by stirring for 2 h, and was placed in the refrigerator at 4 ℃

overnight. The crude material was collected for further analyses.

HPLC/ESI-MSn Assay of Surfactin Production

The crude material extracted as described above and the surfactin standard

(Sigma-Aldrich, St. Louis, USA) were sent to Dr. Jodie V. Johnson’s laboratory in the

Chemistry Department at the University of Florida, Gainesville, FL for detecting the

presence of surfactins and for further quantitative analyses. The extracts and surfactin

standard were analyzed via high performance liquid chromatography combined with

electrospray ionization tendem mass spectrometry (HPLC/ESI-MSn).

Page 83: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

83

Both (+) and (-) ESI-MSn collision-induced dissociations (CID) were conducted to

quantify surfactins produced in the testing samples. Both positive and negative ion

modes are provided in the Table 3-4. The gradient was changed to allow better

separation in (+)ESI-MS compared to (-)ESI-MS. (+)ESI-MS was chosen for further

analyses. The relative amount of different surfactin isoforms was detected in the crude

extracts of IN937b expressed in terms of HPLC/(+)ESI-MS peak areas and relative

areas. The sum of the intensities of the [M+H]+ and [M+Na]+ were used to produce the

ion-peaks for the surfactins. The peak area was obtained by integration of the

appropriate peak in the relative abundance of surfactins in the ion chromatograms

(Table 3-5).

Screening IN937b Mutant for Rifampicin Resistance

Spontaneous mutants with rifampicin-resistance were selected from nutrient agar

amended with rifampicin (NAR) plates according to Bolstridge et al. (2009) with

modifications. Wild type of B. subtilis IN937b was plated on NA amended with 10 µg/mL

rifampicin (NAR) and incubated at 28 ℃ for 48 h. The mutants with the same

morphology as the wild type were transferred to NAR with the higher concentration of

rifampicin (25 µg/mL). And the above step was repeated until the mutants had the ability

to resist 75 µg/mL rifampicin. Stability of the rifampicin resistance was confirmed by

sub-culturing ten times on NA and then on NAR (75 µg/mL). All mutants were stored in

20% glycerol at -80 ℃ (Bolstridge et al., 2009).

Root Treatment by IN937b Rifampicin-Resistance Mutant

A spontaneous mutant of IN937b resistant to 75 µg/mL rifampicin was used to

monitor changes in the IN937b population introduced into the rhizosphere. A single

colony of IN937b with rifampicin resistance was inoculated in 10 mL NB containing 50

Page 84: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

84

µg/mL rifampicin in 50 mL centrifuge tubes. After the tube was incubated at 28 ℃ on a

rotary shaker incubator at 150 rpm overnight, 1 mL of the bacterial suspension was

inoculated into 200 mL NB in 500 mL flasks, and incubated for 24 h under the same

conditions as mentioned above. The cell pellet was collected by centrifugation at 3000

rpm for 15 min and washed twice with sterile DI water. The cell concentration used in

this experiment was adjusted to 108 CFU/mL, determined by serial dilutions and plating.

Squash seeds were planted in plastic pots (7.62 cm high and 10.16 cm at the

opening with 8 drain holes) filled with potting mix. One seed was placed in each pot. In

the greenhouse, two-week-old squash plants were treated with rifampicin resistant

IN937b by soil drench at 107 CFU/g soil per plant.

Rhizosphere Sampling

Samples of the squash rhizosphere were taken from each of the pots 1, 3, 5, 7,

and 14 days after IN937b treatment. The roots of the squash plants were shaken

vigorously to get rid of bulk soil that did not tightly adhere to the roots. Four replicates in

each treatment (IN937b and non-treated control) were obtained at each sampling time.

The rhizosphere soil is defined as the soil still adhering to the roots after vigorous

shaking. The roots and rhizosphere soil were collected together, and 50 g of soil and

the roots were placed in a 500 mL plastic container with 100 mL sterile DI water. Total

bacterial populations were assessed by serial dilutions and plating of the resulting soil

suspension on NAR medium after the container was shaken on a rotary shaker

incubator at 150 rpm for 30 min at room temperature.

Recovery of IN937b Mutant with Rifampicin-Resistance from Roots

For measuring internal root colonization, roots were separated from the

rhizosphere samples mentioned above and were collected by rinsing the roots three

Page 85: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

85

times with DI water. Root samples were surface disinfected with sodium hypochlorite

(1%) for 5 min followed by ethanol (70%) for 30s. Then the roots were chopped and

ground with 2 mL sterile DI water in a sterile mortar using a pestle. In order to make

sure that no surface bacteria remained after surface-disinfection, one intact root for

each sample before chopping was placed on the NAR medium for surface bacteria

growing. The population of inoculated bacteria in internal root tissues was determined

by serial dilutions and plating of the homogenate on the NAR plates and incubated at 28

℃ for 48 h.

Results

Architecture of Pellicles and Colonies of Bacillus spp.

Tubes made of PVC which stimulate biofilm formation (Pedersen, 1990), were

used to test if selected Bacillus PGPR had the ability to form biofilm on an abiotic

surface. The Bacillus strains were allowed to grow on the surface of the liquid MSgg

medium. Five days after inoculation, the biofilm was observed as a flat layer at the

interface between the air and liquid medium (Figure 3-1 A.). PGPR strains IN937b,

1216S, GB03, IN937a, INR7 and 1PC-19 had the ability to form a thick, floating biofilm

composed of a large group of bacterial cells. However, strains SE76, 1PC-11, SE34,

SE49, SE52 and T4 exhibited poor growth of the layer, and they were unable to form a

biofilm initially.

When spotted on NA plates, the Bacillus strains produced colonies with

morphological characteristics of their corresponding pellicles (Figure 3-1 B.). Colonies

varied widely in morphology among Bacillus strains. Five days after inoculation, PGPR

strains IN937b, 1216S, GB03, IN937a, INR7 and 1PC-19 had wrinkled (because of its

rough appearance) colonies which were consistent with the result of pellicle formation in

Page 86: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

86

the liquid medium test (Figure 3-1 A.). SE76, 1PC-11, SE34, SE49, SE52 and T4

exhibited smooth (because of its small and flat appearance) colonies on NA plates

which were not able to form biofilm in the liquid medium. The wrinkled bacterial colonies

grew wider on the solid medium surface than the smooth colonies did.

Closer examination of the colonies under a dissecting microscopy (×10

magnification) revealed a highly structured bacterial community, with clearly observable

architecture. As shown in Figure 3-2, the wrinkled pattern of IPC-19 generated denser

layers compared to the other test PGPR strains. IN937b and 1216S consisted of many

wave-like columns that projected upward from the agar surface. IN937a, INR7 and

GB03 had aerial structures composed of cells bundled together in a flat form.

In Vitro Test for Biofilm formation

The highly structured biofilm was formed by the Bacillus strains on the surface of

a liquid MSgg medium in a beaker. Bacillus PGPR strains SE34, SE52, SE76, and T4

failed to form biofilm (Figure 3-3). GB03 and 1216S formed an extremely thin, fragile,

and smooth pellicle that lacked a distinctive architecture. Contrastingly, PGPR strains

IN937b and 1PC-19 formed a thick pellicle with an intricate vein-like appearance on its

surface, and a thick and highly folded structure of biofilm was clearly visible through the

beaker.

Effect of Bacterial Density on Biofilm Formation

Based on the observation of the in vitro test for biofilm formation, wherein

bacterial cells move in the liquid without a solid surface, we found that the process of

self-organization in biofilm has been linked to bacterial density and bacterial growth

speed in the liquid-air interface. In order to determine the influence of PGPR densities

on biofilm formation, we assayed the selected PGPR strains in liquid MSgg medium in

Page 87: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

87

polystyrene plates that has 6 wells each. Biofilm formation and progression were

visually inspected daily for 7 days. The initial concentration in liquid MSgg medium was

calculated as 104, 103, and 102 CFU/mL. At days 1 and 2 after bacterial inoculation,

IN937b had a high concentration (104 CFU/mL) and the positive strain 1PC-19 (104

CFU/mL) formed thin, loosely associated aggregations of bacterial cells at the surface

of the liquid medium in the well. At days 2 and 3, a thin pellicle comprised of both live

cells and extracellular materials formed on the air-surface interface in the wells

containing IN937b at 104 and 103 CFU/mL and 1PC-19 (Figure 3-4). From days 3 to 5,

the pellicle at the interface thickened and developed a thick mat of cording and

reticulations. In addition to bacterial cells, extracellular materials from biofilms produced

by the strain 1PC-19 adhered to and moved up the walls of the well during days 2 to 5

(Figure 3-4 and Table 3-2). IN937b at the lowest concentration (102 CFU/mL) and the

negative strain SE52 failed to form biofilm and the reticulation was incomplete, bacterial

cells were unable to disperse in the well and could not self-organize.

PCR Detection of Lipopeptide Genes in Bacillus PGPR Strains in Relation to Biofilm Formation

A 675-bp PCR product was amplified from IN937b using the primers specific to

the sfp gene which is responsible for surfactin production (Figure 3-8). Amplification of

the PCR products indicates that IN937b and IN937a harbor the sfp gene and may

produce surfactins. No PCR product was amplified from SE76 and SE52 using any of

the primers tested. No product was amplified from IN937b with primers specific to the

genes encoding the antibiotics iturin A, fengycin, zwittermicin A, and bacillomycin D

(Table 3-3).

Page 88: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

88

DNA sequences analyzed by BLAST software revealed that this PCR fragment in

IN937b had 100% similarity to B. subtilis 168 (Accession No. AL009126) whose

genome has been fully sequenced. The results indicated that IN937b is positive for the

presence of the 4’-phosphopantetheinyl transferase gene sfp for surfactin synthesis.

Alignment of the predicted amino acid sequences revealed that all the 5 amino acid

residues are conserved in all Bacillus for surfactin synthetase (Figure 3-9). The

phylogenetic tree based on amino acid sequences of sfp revealed a close relationship

among B. subtilis IN937b, B. subtilis 168, and B. amyloliquefaciens FZB42 (Figure 3-

10).

HPLC/MS Analysis of Surfactins Production in B. subtilis IN937b

The methanol extracts of No. 3S culture of strain IN937b were separated by

HPLC/ESI-MS, and a clear peak pattern of surfactins was eluted based on the

molecular weight range of the standard sample. The molecular weight (MW) of surfactin

isoforms was determined by ESI-MS (Figure 3-5 A-G), and the HPLC fractions that had

more than one peak were found to represent a cluster of isoforms at MW 993, 1007,

1021, 1035, 1049, and 1063 Da. The fatty acids in surfactins were determined (Table 3-

4). The structure of standard surfactins consists of a seven amino acid cyclopeptide with

a saturated fatty acid chain (R) attached. The chain length of R varies from C9 to C14

leads to detection of the major molecular weights. Surfactin isoforms with MW 1035 and

1021 Da contained the highest levels of relative abundance of 45.76% and 39.84%,

respectively (Table 3-5).

Survival of IN937b in the Rhizosphere

Recovery of IN937b from the rhizosphere began at 1 day after treatment to two

weeks after soil drench with IN937b (Figure 3-6). There was a high bacterial density at

Page 89: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

89

106 CFU/g soil per plant in the potting mix at 1 day after application. The population

declined dramatically to 104 CFU/g soil per plant at 5 days after treatment. And then the

populations slowly decreased to 103 CFU/g soil per plant at 14 days after treatment.

Survival of IN937b in Squash Roots

Internal root colonization by IN937b was evaluated by serial dilutions and plating

on rifampicin-amended NA plates. No IN937b cells were detected on plates from the

non-treated squash roots after surface-sterilization. IN937b was recovered in squash

roots at 1 day after treatment. However, the endophytic population of IN937b recovered

from the roots fluctuated from 101 to 103 CFU/g fresh root (Figure 3-7). At 2 weeks after

IN937b treatment, an increasing trend in the population of IN937b was observed with

level up to 103 CFU/g fresh root. Bacterial densities of IN937b colonized in the roots

were as high as 103 CFU/g fresh root 14 days after treatment.

IN937b colonized the rhizosphere and also occupied the internal tissues of

squash roots after application. At the early stage of colonization, the population of

IN937b survived in the rhizosphere. Over a 14 days period, the bacterial population in

the rhizosphere decreased, while endophytic population in root tissue survived longer

and accumulated gradually.

Discussion

Bacillus strains had the ability to form biofilms in liquid or solid medium. Roots of

some plants contain tightly bound Bacillus strains that form abundant biofilms. Both in

vitro and in vivo assays have showed that the rapid formation of biofilm under laboratory

conditions reflect the ability of biofilm formation in the natural environment (Bais et al.,

2004b). Interestingly, all strains which morphologically vary in colony morphology, are

Page 90: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

90

able to form biofilms, implying that Bacillus strains developed the same major adaptive

strategy to survive in the soil environment.

Under laboratory conditions, the bacterial cells produced extracellular polymers

that bind cells together and provide structural support for biofilm that resided at the

liquid-air interface of the medium in glass beakers. The lack of solid surface at this

interface imposes a greater requirement for self-organization on the constituent cells.

When grown in beakers filled partly with a liquid medium, bacterial strains form large

groups of cells on the liquid surface. During this process, bacterial strains produce

extracellular polysaccharide (EPS), which affords better access to oxygen (Flemming

and Wingender, 2010). EPS is effective in binding cells together and in anchoring them

to the beaker walls.

The architecture of the colony and the thickness of the pellicle formed at the air-

liquid interface are both influenced by surfactin produced from different Bacillus strains

(Leclere et al., 2006). Among these, B. subtilis is believed to form a robust biofilm when

it attains a certain population density. The motile cells of B. subtilis migrate to the air-

medium interface and biofilm formation rate is associated with cell populations. B.

subtilis cells assimilate sufficient oxygen to produce submerged fermentation that is

subjected to colony morphology and growth rate. Faster growth of bacterial cells causes

nutrient depletion in the culture, resulting in faster sporulation of Bacillus spp.

Sporulation of B. subtilis is delayed in biofilm when compared to that of B. subtilis at

submerged fermentation (Lindsay et al., 2005). Therefore, Bacillus cells in the biofilm

remain in their metabolically active state for a longer period of time compared to those

in their sporulation stage.

Page 91: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

91

Bacillus subtilis often found in the rhizosphere might be closely associated with

plant roots. Recently, the ability of Bacillus in the rhizosphere to form beneficial biofilm

has been well documented (O'Toole et al., 2000; Vlamakis et al., 2013). The behavior of

bacterial cells tending to attach and aggregate near plant root surfaces or internal roots

by self-organization can be considered a type of biofilm (Lugtenberg et al., 2001). These

communities have spatial and temporal organization due to cellular differentiation

(Ramey et al., 2004).

Biofilm helps microbial populations to swarm over the root surface and reach new

niches to attain more nutrients (Branda et al., 2001). On the well-colonized root surface,

approximately 10 - 40% of root surfaces can be occupied by Bacillus that formed biofilm

(Vlamakis et al., 2013). Motility of the bacterial cells is important for biofilm formation

and colonization by the bacteria. The motility that performed as swimming and

swarming was achieved through the cell movement of colonies. The strains with fast

motility usually spread or occupy large area in the medium, indicating that motility is an

important mechanism for Bacillus colonization, especially for those introduced to a new

environment (Kinsinger et al., 2003). As a first step of biofilm formation, the motility that

helps bacterial strains to access favorable colonization sites in the rhizosphere involves

rapid and coordinated population migration. As a second step, bacterial cells must

attach to highly structured biofilms on solid surfaces or pellicles and adhere to roots and

soil particle surfaces (Danhorn and Fuqua, 2007).

The population of IN937b recovered from potting mix was 103 CFU/g of fresh root

10 days after IN937b introduced into the rhizosphere of squash plants. The results from

the survival assay of IN937b showed poor colonization in the rhizosphere, although

Page 92: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

92

IN937b formed robust biofilms in defined medium (and likely on plant roots as well).

Also, results from the microtiter assays showed that biofilm formation strongly depended

on the initial population of IN937b inoculated in the medium, indicating that initial

population might be critical for biofilm formation.

In this experiment, the initial populations of IN937b were high and dramatically

decreased in the potting mix 2 weeks after treatment. Potting mix is a relatively simple

environment compared to the soil, which is a mixture of mineral compounds and

microorganisms. Bacterial behavior may be influenced by environmental factors

including the nutrients, water content in the soil and bacterial movement in the root

systems. On the other hand, the rapid colonization of IN937b in the internal roots

indicated that IN937b had potential to predominate in the root competition. These also

enable Bacillus strains to find colonization sites and acquire nutrients for multiplication

over a long period.

Different types of distribution were found in Bacillus strains. Some bacteria were

localized at the root cap while others were evenly distributed along the root surface.

Fluorescence in situ hybridization (FISH) revealed that the favorable site for Bacillus

strains to form biofilm could be extended 2-30 µm in the rhizosphere outward from root

surface of wheat (Watt et al., 2006). Effective colonization leads to plant growth

promotion (Ongena et al., 2005). Bacillus strains secrete a large number of compounds

into the rhizosphere which are beneficial to plant roots and shoot growth. Bacillus works

as biofertilizer for many plant species due to its interaction with plant roots through

biofilms on root surfaces (Seneviratne et al., 2010).

Page 93: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

93

Bacillus strains in this study were selected from different crops in diverse

environmental niches. These bacteria showed distinct features, especially in colony

morphology, biofilm formation and biocontrol activity. However, the ability to form

biofilms is still highly conserved in these test strains. These results imply that different

Bacillus spp. have developed the same strategy for adapting to different environments

(Chen et al., 2013).

B. subtilis IN937b produced surfactins. Surfactins are well known to stimulate

biofilm formation (Bais et al., 2004a). For example, surfactins are detected in the

rhizosphere of cucumber and tomato inoculated with B. subtilis strains (Ongena et al.,

2007; Kinsella et al., 2009). The movement of bacterial cells is largely dependent on the

production of surfactins, which help cells to spread over the root surfaces (Raaijmakers

et al., 2006). Surfactin production is widespread in Bacillus genus (Price et al., 2007).

Surfactin-producing Bacillus species can be easily isolated from soils. However,

surfactins are susceptible to degradation by other microorganisms in nature or in the

laboratory. In sterile soils without interference from other microorganisms, levels of

surfactins remained stable for 25 days (Asaka and Shoda, 1996).

Surfactins are important for microorganisms because they lower the surface

tension of water. In addition, surfactins have antimicrobial activities due to their abilities

to penetrate cellular membranes (Heerklotz and Seelig, 2007). Moreover, surfactins act

as a signal in the quorum-sensing to regulate bacterial population density in response to

environmental changes (Lopez et al., 2009). Surfactins are reported to act as a

stimulus in biofilm formation. High levels of surfactins were observed during colonization

and biofilm formation by B. amyloliquefaciens S499 on tomato roots. Recovery of

Page 94: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

94

surfactins from the rhizosphere indicated that bacterial colonization on the root leads to

high levels of surfactins that stimulate biofilm formation (Nihorimbere et al., 2012).

Surfactin biosynthesis requires enzymes called surfactin synthetases. The gene

sfp serves a regulatory role as well as a more direct role in surfactin synthesis. In this

study, PCR primers specific to the regions within sfp were used to screen a number of

Bacillus spp. strains for presence of the sfp. Only IN937b was confirmed to carry the sfp

gene responsible for the production of surfactins.

To confirm the PCR result for surfactin production, HPLC and quantitative

analysis were adopted to detect the amount of surfactins produced from B. subtilis

strain IN937b. Surfactins were extracted from No. 3S culture broth of IN937b by classic

methods including acid precipitation with HCl, recrystallization and extraction by organic

solvents such as methanol. Results of HPLC were consistent with the PCR detection.

Naturally occurring surfactins are a mixture of different molecules based on chain

length, β-hydroxy fatty acid, and amino-acid sequence (Lin and Jiang, 1997). In the

present study, several isomers of surfactins with different molecular weights were

detected.

Biofilm is important for bacteria to colonize in the rhizosphere and to act as a

biocontrol agent (Bais et al., 2004b). The biocontrol efficacy of B. subtilis against

Rhizoctonia solani has been shown to be achieved by the production of surfactins and

iturin A. It is possible that biofilm and surfactin production allowed B. subtilis and its

close relatives, such as B. amyloliquefaciens, to efficiently colonize plant roots and also

to provide protection to their hosts. A knockout of the sfp gene resulted in flat and small

colonies of the bacterium. This knockout mutant spread very little on the agar surface

Page 95: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

95

(Branda et al., 2001) revealing that the mutant failed to form mature aerial structures

related to biofilm formation. Loss of surfactin biosynthesis led to a deficiency in motility

and reduced colonization of roots and surfaces. Biocontrol of P. syringae infection of

Arabidopsis roots by B. subtilis 6051 is related to surfactin production on the surface of

the root (Bais et al., 2004b). B. subtilis 6051 can produce a variety of lipopeptides

including surfactins that are important for maintaining the aerial structure of biofilms.

Upon root colonization, B. subtilis 6051 formed a stable, extensive biofilm and secreted

surfactins, which act together to protect plants against infection by pathogenic bacteria

(Morikawa, 2006). It is strongly suggested that the production of surfactins is essential

for biofilm formation and colonization of plant roots.

In conclusion, our results suggest that many Bacillus PGPR strains are capable

of forming biofilms. It will be important to construct knock-out mutants of surfactin

synthesis genes. Future work is needed to determine the physiological relevance of the

regulation of surfactin synthesis genes and biofilm formation. The surfactin mutants will

continue to be useful in future studies on colonization models in the squash

rhizosphere, and this capability may be important for disease suppression. Biofilm

formation by IN937b may increase its colonization to the squash root. However, the

exact mechanisms in which biofilm formation contributes to plant biocontrol still remain

unknown. Successful colonization on root tissue is a first step critical in eliciting plant

biocontrol activities against soil borne pathogens. In the future, it will be imperative to

characterize the signaling molecules released by the plant host that improve Bacillus

spp. ability to colonize and form biofilm on the root surfaces.

Page 96: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

96

Table 3-1. List of primer sequences used for biosynthetic genes of surfactins, bacillomycin, iturin A, Fengycin, and Zwittermicin A.

Antibiotic Primer sequence

Surfactin F: ATGAAGATTTACGGAATTTA

(Hsieh et al., 2004) R: TTATAAAAGCTCTTCCTACG

Bacillomycin D

F:GAAGGACACGGGCAGAGAGTC

(Athukorala et al., 2009) R: CGCTGATGACTGTTCATGCT

Iturin A F: GATGCGATCTCCTTGGATGT R: ATCGCATGTGCTGCTTGAG

Fengycin F: TTTGGCAGCAGGACAAGTTT

(Ramarathnam et al., 2007)

RGCTGTCCGTTCTGCTTTTTC

Zwittermicin A F: TTGGGAGTTTATACAGCTCT

R: GACCTTTTGAAATGGGCGTA

Table 3-2. Biofilm formation by Bacillus strains with initial bacterial populations

Concentration (CFU/mL)a Biofilm formation time (Days)b

IN937b (104) 2 IN937b (103) 3 IN937b (102) 5 IN937b (101) None IN-19 (104, positive strain) 3 SE52 (104, negative strain) None

a The inocula for the microtiter plate assays were obtained by diluting the Bacillus cells to approximately 104, 103, 102 and 10 CFU/mL. One hundred µL of inocula were grown in 6-well PVC microtiter plates containing 10 mL of liquid MSgg medium at room temperature for 5 days before imaging. b Biofilm growth were observed by visual inspection. The biofilm formation was checked daily for the timing of biofilm formation. Table 3-3. Detection of the genes related to biofilm formation

Antibiotica

PCR results

IN937b SE76 SE52

Surfactin +b - - Zwittermicin A - - - Bacillomycin D - - - Iturin A - - - Fengycin - - -

a Five selected antibiotics are antimicrobial peptides that are associated with biosurfactins and in the fitness of Bacillus strains in the rhizosphere. b Polymerase chain reaction (PCR) screening using specific primers (Table 3-1) for antibiotic biosynthetic genes in strains IN937b, SE76 and SE52. + represents detection of the genes. – represents no detection of the genes.

Page 97: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

97

Table 3-4. Identification of surfactin ions by HPLC/ESI-MS

Surfactin ionsa

(+)ESI-MSb (-)ESI-MS

R mass R, u

MW, u

[M+H]+ [M+Na]+ [M-H]- [M-2H+Na]-

C9H19 127.1 993.6 994.3 1016.5 992.5 1014.6

C10H21 141.2 1007.7 1008.4 1030.6 1006.6 1028.7

C11H23 155.2 1021.7 1022.5 1044.7 1020.7 1042.8

C12H25 169.2 1035.7 1036.4 1058.6 1034.6 1056.7

C13H27 183.2 1049.7 1050.4 1072.6 1048.6 1070.7

C14H29 197.2 1063.7 1064.5 1086.7 1062.7 1084.8 a The surfactin consists of a seven-amino acid cyclopeptide with a saturated fatty acid chain (R) attached. The chain length of R varies from C9 to C14 to provide the major molecular weights (u) detected. Mass R represents the molecular weight of saturated fatty acid chain. MW represents the molecular weight of surfactins. b HPLC-ESI-MS ions assay for surfactins. Both positive and negative ion modes were given in the table. (+)ESI-MS represents positive ion mode; (-)ESI-MS represents negative ion mode. Table 3-5. HPLC/(+)ESI-MS peak areas and relative areas for the detected surfactins in

the crude extracts of IN937b.

Isomer 1 Isomer 2

MWa RT-1 minb

Area-1c %Total RT-2 min

Area-2 %Total

993 55.94 1,081,996,051 1.63 nd 0 0.00

1007 56.91 1,966,233,219 2.96 58.67 363,655,897 0.55

1021 58.07 26,434,118,167 39.81 59.29 1,430,159,095 2.15

1035 58.80 30,383,407,593 45.76 60.05 1,612,113,268 2.43

1049 59.92 2,308,838,698 3.48 61.02 219,785,102 0.33

1063 60.60 601,132,399 0.91 nd 0 0.00

SubTotal 62,775,726,127 3,625,713,362

Total 66,401,439,489 a (+)ESI-MS mass chromatograms of the [M+Na]+ ions of the surfactins produced by IN937b, which match those of the surfactin standard. b RT represents retention time from 54 to 63 during which surfactins elute through the HPLC column. c The sum of the intensities of the [M+H]+ and [M+Na]+ were used to produce the ion-peaks for the surfactins. The peak area was integrated in the appropriate relative abundance ion chromatograms for determining the abundance of surfactins.

Page 98: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

98

A

B Figure 3-1. Architecture of pellicles and colony morphology of PGPR strains. A. Pellicle

formation at the air-liquid interface in liquid MSgg medium of 12 Bacillus strains. Cells of Bacillus strains were inoculated in the liquid medium in the standing tubes at 28 °C for 3 days. The band at the top of the liquid medium represents the floating pellicle. B. Colony formation of 12 Bacillus strains on the surface of solid MSgg medium. Bacteria were streaked to form single colony on the medium and incubated at 28 °C for 24 hours.

Page 99: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

99

Figure 3-2. Pellicle development of Bacillus strains. Pellicle formation at the air-liquid

interface under a dissecting microscope (10 ×). Strains that formed biofilm in pellicle formation assays (Fig 3.1 A) were selected to observe pellicle morphology under a microscope. The floating pellicles at the top were carefully placed on top of a microscopic glass slide, and photographed (10 ×) for surface structure.

Figure 3-3. Biofilm development properties of Bacillus strains. Biofilm formation from

selected Bacillus strains on the surface of a liquid MSgg medium in a beaker. Bacteria were inoculated in liquid MSgg medium in a beaker for floating pellicle development. The images were taken at room temperature 5 days after inoculation. All four strains (IN937b, IPC-19, 1216S, GB03) formed a biofilm with a wrinkled (rough surface appearance) pattern on the surface as shown in the top set of photo. The strains shown in the bottom set of photos were biofilm-negative phenotype.

Page 100: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

100

Figure 3-4. Biofilm formation rate of Bacillus strains. Biofilm growth was observed by

visual inspection. The inocula for the microtiter plates were obtained by diluting the Bacillus cells to approximately 104, 103, and 102 CFU/mL. One hundred µL of inocula were grown in 6-well PVC microtiter plates containing 10 mL MSgg liquid medium at room temperature for 5 days before imaging. 1: IN937b (104 CFU/mL), 2: IN937b (103 CFU/mL), 3: IN937b (102 CFU/mL), 4: Control, 5: IPC-19 (+) (104 CFU/mL), 6: SE52 (104 CFU/mL) (-).

Page 101: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

101

Figure 3-5. HPLC/MS analysis of the fraction from crude extraction from B. subtilis IN937b culture.

A

B

Page 102: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

102

Figure 3-5. Continued

C

D

Page 103: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

103

Figure 3-5. Continued

E

F

Page 104: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

104

Figure 3-5. Continued

G

A) Chromatogram of surfactins detected by HPLC with (+)ESI-MS. MW 1016, 1030, 1044, and 1058 with black underline, 1072 and 1086 are not shown in the figure, B) Surfactins with MW 993: HPLC/(+)ESI-MS/MS of m/z 1016 [M+Na]+: At least four isomers: two defined by m/z 707 and 772 and two defined by m/z 693 and 786, C) Surfactins with MW 1007: HPLC/(+)ESI-MS/MS of m/z 1030 [M+Na]+: At least four isomers: two defined by m/z 707 and 786 and two defined by m/z 693 and 800 product ions, D) Surfactins with MW 1021: HPLC/(+)ESI-MS/MS of the m/z 1044 [M+Na]+ ions shows the presence of at least four (4) isomers: m/z 707, 800, 693, and 814 product ions, E) Surfactins with MW 1035: HPCL/(+)ESI-MS/MS of m/z 1058 [M+Na]+ At least four isomers with two defined by m/z 707 and 814 and two defined by m/z 693 and 828, F) Surfactins with MW 1049: HPLC/(+)ESI-MS/MS of m/z 1072 [M+Na]+ ions shows more isomers: three defined by m/z 707 and 828, on the top, the m/z 707 shifted to m/z 721, G) Surfactins with MW 1063: HPLC/(+)ESI-MS/MS of m/z 1086 [M+Na]+ ions: The two major isomers were those defined by m/z 707 and 842 and on the bottom, the m/z 707 shifted to m/z 721.

Page 105: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

105

Figure 3-6. Colonization of the rhizosphere of squash plants by a mutant of IN937b with

rifampicin resistance. The data are expressed as log10 CFU per gram of potting mix. Data points are the means and standard errors of four replicates. Data were collected at the 1, 3, 5, 7 and 14 days after treatment. Values with same letter within a column are not significantly different according to Fisher’s protected LSD test (p = 0.05). Vertical bars represent mean ± S.D. (n = 5).

6.75 a

5.56 b

4.44 c

3.18 d 3.17 d

0

2

4

6

8

1 3 5 7 14

Lo

g c

fu/g

so

il

Time (Days)

Page 106: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

106

Figure 3-7. Colonization of the squash roots by a mutant of IN937b with rifampicin

resistance. The data are expressed as log10 CFU per gram of fresh squash root. Data points are the means and standard errors of four replicates. Data were collected at the 1, 3, 5, 7 and 14 days after treatment. Populations of IN937b colonizing squash seedling roots. Values with same letter within a column are not significantly different according to Fisher’s protected LSD test (p = 0.05). Vertical bars represent mean ± S.D. (n = 5).

2.26 d

1.58 e

3.35 b

2.65 c

3.81 a

0

0.5

1

1.5

2

2.5

3

3.5

4

4.5

1 3 5 7 14

Lo

g c

fu/g

ro

ots

Time (Days)

Page 107: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

107

Figure 3-8. PCR product profiles of the sfp gene. Gel displaying bands created by using primers for the sfp gene. Lane 1: Ladder 100-1000 bp makers, 2: SE76, 3: IN937a, 4: T4, 5: IN937b, 6: Negative control.

1 2 3 4 5 6

700bp

500bp

Page 108: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

108

24FZB42

36E33L

0QMB1551

40SC2

27ATCC7061

24168

7IN937b

25NBRC100599

24ATCC14580

Consensus

.

.

.

M

.

.

.

.

.

.

.

.

V

.

.

.

.

.

.

.

.

L

.

.

.

.

.

.

.

.

Y

.

.

.

.

.

.

M

.

T

.

.

.

.

.

.

I

.

E

.

.

.

.

.

.

E

.

Q

.

.

.

.

.

.

S

.

L

.

.

.

.

.

.

K

.

I

.

.

.

.

.

.

V

.

L

.

.

.

.

.

.

V

.

T

.

.

.

.

.

.

D

.

R

.

.

.

.

.

.

S

.

E

.

.

.

.

.

.

I

.

D

M

.

.

.

.

.

P

.

I

K

.

.

.

.

.

T

.

V

I

.

.

M

.

M

L

.

M

F

M

.

I

M

K

N

.

K

A

K

.

A

K

I

E

.

A

I

I

.

I

I

Y

N

.

A

Q

Y

.

Y

Y

G

D

.

V

L

G

.

A

A

V

C

.

S

Q

I

.

L

V

Y

Q

.

F

H

Y

.

K

Y

M

I

.

V

L

M

.

C

M

D

W

.

H

D

D

.

P

D

R

W

.

A

D

R

.

A

Q

P

G

.

V

I

P

.

V

L

L

R

.

C

H

L

.

M

P

S

I

.

H

A

S

.

E

E

A

S

.

A

R

Q

.

K

A

G

D

.

G

K

E

.

D

G

E

L

.

A

Q

E

.

S

Q

E

Q

.

S

I

N

.

F

F

D

S

.

L

D

E

M

N

N

R

W

.

D

Q

R

N

R

R

M

H

.

P

L

F

G

F

M

M

Y

.

E

K

M

S

L

M

A

N

.

I

P

T

C

Q

A

A

L

.

F

F

F

L

A

M

V

L

.

H

V

I

S

L

V

64FZB42

76E33L

0QMB1551

80SC2

67ATCC7061

64168

47IN937b

65NBRC100599

64ATCC14580

Consensus

S

N

.

S

S

S

I

P

S

A

D

.

D

A

P

P

E

E

E

V

.

E

E

E

E

E

E

K

E

.

C

K

K

K

K

K

R

R

.

S

R

R

R

R

R

E

E

.

Y

A

E

E

E

K

K

K

.

F

A

K

K

R

K

C

A

.

E

A

C

C

V

V

R

N

.

S

E

R

R

N

G

R

S

.

L

R

R

R

R

R

F

Y

.

L

F

F

F

F

Y

Y

H

.

F

R

Y

Y

R

R

H

H

.

Q

F

H

H

N

Y

K

S

.

R

L

K

K

P

P

E

A

.

R

I

E

E

A

E

D

D

.

K

D

D

D

D

D

A

R

.

S

A

A

A

S

A

H

A

.

N

R

H

H

Y

S

R

R

.

F

R

R

R

R

R

T

F

.

L

T

T

T

A

T

L

I

.

L

L

L

L

L

L

I

I

.

G

L

L

L

L

I

G

G

.

R

G

G

G

A

G

D

C

.

L

E

D

D

D

E

M

V

.

S

V

V

V

V

V

L

I

.

A

L

L

L

L

L

I

S

.

K

V

V

V

V

A

R

R

.

Q

R

R

R

R

R

T

L

.

A

Q

S

S

S

H

A

V

.

A

M

V

V

L

I

A

L

.

S

I

I

I

I

I

A

G

.

L

H

S

S

C

S

K

K

.

L

D

R

R

E

E

A

I

.

L

M

Q

Q

A

T

Y

L

.

S

Y

Y

Y

Y

F

G

S

.

E

D

Q

Q

E

H

L

M

.

S

L

L

L

I

L

D

S

.

N

P

D

D

S

P

P

P

.

L

F

K

K

N

N

A

V

.

S

D

S

S

D

E

97FZB42

114E33L

0QMB1551

117SC2

100ATCC7061

97168

80IN937b

98NBRC100599

97ATCC14580

Consensus

G

Q

.

N

E

D

D

E

Q

I

V

.

I

I

I

I

I

I

S

P

.

S

V

R

R

E

R

F

I

.

V

F

F

F

Y

F

G

D

.

E

E

S

S

D

T

V

R

.

T

T

T

T

Y

E

Q

M

.

G

E

Q

Q

N

G

E

C

.

V

G

E

E

A

R

Y

P

.

F

N

Y

Y

Y

Y

G

V

.

G

G

G

G

G

G

K

C

.

Q

K

K

K

K

K

P

K

.

P

P

P

P

P

P

Y

L

.

L

V

C

C

F

A

I

Q

.

L

V

I

I

L

A

P

H

.

R

R

P

P

K

E

A

G

.

T

H

D

D

S

G

L

R

.

T

I

L

L

F

L

P

P

.

G

P

P

P

P

N

D

Q

.

A

S

D

D

N

D

M

L

.

P

F

A

A

F

F

H

P

.

G

H

H

H

C

H

F

E

.

Y

F

F

F

F

F

N

G

.

Q

N

N

N

N

N

.

M

.

.

.

.

.

.

.

.

P

.

.

.

.

.

.

.

.

Q

.

.

.

.

.

.

.

.

L

.

V

.

.

.

.

.

.

S

.

S

.

.

.

.

.

I

V

.

I

L

I

I

V

I

S

S

.

A

S

S

S

S

S

H

H

.

H

H

H

H

H

H

S

S

.

C

S

S

S

S

S

G

G

.

D

G

G

G

G

G

R

E

.

D

D

R

R

E

K

W

W

.

I

W

W

W

W

W

I

V

.

G

V

V

V

V

I

.

.

.

T

.

.

.

.

.

.

.

.

A

.

.

.

.

.

V

V

.

V

V

I

I

V

V

C

V

.

V

C

G

C

C

C

134FZB42

154E33L

34QMB1551

154SC2

137ATCC7061

134168

117IN937b

135NBRC100599

134ATCC14580

Consensus

A

A

.

F

A

A

A

A

G

.

F

.

.

.

.

.

.

.

V

T

.

P

I

F

F

T

A

D

K

.

E

D

D

D

H

G

S

F

M

A

D

S

S

D

A

K

A

I

H

A

Q

Q

S

Q

P

P

S

P

P

P

P

Q

P

I

V

V

M

V

I

I

V

I

g

G

G

G

G

G

G

G

G

G

I

V

I

V

I

I

I

I

V

d

D

D

D

D

D

D

D

D

D

I

V

L

I

I

I

I

V

V

E

E

V

E

E

E

E

E

E

K

Q

D

R

E

K

K

Q

K

M

M

I

I

I

T

T

I

I

K

N

E

S

K

K

K

C

K

P

P

R

P

P

P

P

P

P

G

N

F

N

I

I

I

I

I

T

V

K

R

D

S

S

E

N

I

D

L

N

L

L

L

L

F

D

V

A

K

A

E

E

D

D

I

M

L

T

I

I

I

I

I

.

K

.

.

.

.

.

.

.

.

M

.

.

.

.

.

.

.

A

A

S

M

A

A

A

A

A

K

D

R

E

K

K

K

T

K

R

G

S

S

R

R

R

H

R

F

V

G

Q

F

F

F

Y

F

F

L

E

L

F

F

F

F

F

S

T

L

T

S

S

S

A

S

P

D

F

M

A

K

K

P

P

T

S

V

E

D

T

T

A

S

E

E

S

E

E

E

E

E

E

Y

K

R

R

Y

Y

Y

V

H

S

A

I

E

E

S

S

E

D

D

Q

L

R

D

D

D

D

D

L

V

N

I

L

L

L

L

L

Q

M

N

V

L

L

L

L

M

A

K

Y

A

S

A

A

A

E

K

L

E

V

Q

K

K

K

K

173FZB42

193E33L

73QMB1551

194SC2

176ATCC7061

173168

156IN937b

174NBRC100599

173ATCC14580

Consensus

H

P

I

R

P

D

D

H

D

P

N

A

G

A

K

K

P

E

D

E

K

V

E

D

D

N

S

Q

Q

Y

E

R

E

E

E

E

Q

K

D

E

Q

Q

Q

R

R

T

I

S

Y

E

T

T

V

L

D

E

Y

S

A

D

D

S

S

Y

G

Q

V

Y

Y

Y

Y

Y

F

F

W

L

F

F

F

F

F

Y

L

A

L

F

Y

Y

Y

Y

H

T

I

T

H

H

H

D

H

L

Y

L

V

L

L

L

L

L

.

.

.

S

.

.

.

.

.

W

W

F

W

W

W

W

W

W

S

T

S

T

S

S

S

T

T

M

R

A

V

M

M

M

L

M

k

K

K

K

K

K

K

K

K

K

e

E

E

E

E

E

E

E

E

E

S

A

C

A

A

S

S

S

S

F

V

A

I

F

F

F

Y

F

I

L

I

S

I

I

I

I

I

k

K

K

K

K

K

K

K

K

K

Q

A

M

I

L

Q

Q

A

Q

A

T

I

L

T

E

E

R

A

G

G

K

R

G

G

G

G

G

K

E

Q

T

K

K

K

M

K

G

G

I

G

G

G

G

G

G

L

L

P

F

L

L

L

L

L

S

M

K

A

S

S

S

S

S

L

I

D

A

Y

L

L

I

L

P

S

F

S

G

P

P

P

P

L

P

R

L

L

L

L

L

L

D

V

L

Q

S

D

D

Q

D

S

D

R

V

S

S

S

S

S

F

I

D

L

F

F

F

F

F

S

T

I

E

T

S

S

A

S

V

I

Q

I

A

V

V

I

V

R

S

V

N

R

R

R

R

K

L

A

D

E

L

L

L

K

L

K

P

I

I

S

H

H

N

N

208FZB42

233E33L

103QMB1551

229SC2

211ATCC7061

208168

180IN937b

209NBRC100599

208ATCC14580

Consensus

D

N

L

R

E

Q

Q

L

E

D

D

S

F

D

D

D

D

Q

G

S

K

K

G

G

G

Q

G

H

P

D

S

Q

Q

Q

S

R

V

N

E

G

A

V

V

I

A

S

L

L

C

T

S

S

T

S

I

L

L

F

L

I

I

I

V

E

V

V

Q

A

E

E

S

E

L

F

T

S

L

L

L

Q

T

P

K

F

T

P

P

P

S

P

D

D

D

F

D

D

D

D

P

G

R

E

S

H

S

S

S

G

H

Q

S

H

E

H

H

R

Y

E

E

L

F

A

S

S

E

P

P

L

Y

P

P

P

P

A

P

C

V

K

Q

C

C

C

W

C

F

E

Q

Y

F

Y

Y

A

Y

I

N

V

A

V

I

I

F

I

R

T

D

A

Q

K

K

Q

Q

T

V

L

Q

P

T

T

Q

A

Y

M

P

S

Y

Y

Y

Y

Y

E

R

T

Y

S

E

E

E

E

A

D

N

R

L

V

V

I

I

.

L

S

.

.

.

.

.

.

.

R

I

.

.

.

.

.

.

.

P

K

.

.

.

.

.

.

.

S

G

.

.

.

.

.

.

.

V

.

.

.

.

.

.

.

D

D

.

L

D

D

D

D

D

E

Y

.

G

P

P

.

P

P

E

M

.

N

T

G

.

G

G

Y

A

.

Y

Y

Y

.

Y

Y

K

S

.

I

Q

K

.

K

K

L

I

.

F

M

M

.

L

M

A

A

.

T

A

A

.

S

A

V

I

.

I

V

V

.

V

V

C

C

.

V

C

C

.

C

C

A

S

T

A

T

A

.

A

S

A

K

Y

P

R

A

.

T

A

H

E

K

R

K

H

.

T

A

224FZB42

249E33L

115QMB1551

252SC2

230ATCC7061

224168

180IN937b

230NBRC100599

226ATCC14580

Consensus

P

V

V

K

P

P

.

N

P

D

T

F

T

S

D

.

Q

E

F

E

K

E

A

F

.

F

F

C

I

D

W

A

P

.

A

P

D

T

C

S

E

E

.

V

G

G

Q

I

M

R

D

.

Q

R

I

L

V

N

V

I

.

V

I

E

D

T

.

E

T

.

V

V

M

A

H

M

I

M

.

M

M

K

V

L

S

L

V

.

K

K

T

A

I

R

S

S

.

D

S

Y

L

A

L

C

Y

.

L

C

E

L

.

Y

S

E

.

Q

A

E

T

.

E

D

E

.

D

E

L

Y

.

Q

V

L

.

L

L

L

K

.

F

L

L

.

H

S

.

.

.

N

S

.

.

Q

M

.

.

.

S

R

.

.

A

L

.

.

.

I

L

.

.

I

.

.

.

.

F

.

.

.

F

.

.

.

.

H

.

.

.

H

.

.

.

.

P

.

.

.

.

.

.

.

.

S

.

.

.

.

.

.

.

.

C

.

.

.

.

.

Page 109: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

109

Figure 3-9. Comparison of the predicted amino acid sequence for the sfp gene present in several Bacillus strains. Multiple sequence alignment of sfp homologs from B. amyloliquefaciens FZB42 (FZB42), B. cereus E33L (E33L), B. megaterium QM B1551 (QMB1551), B. polymyxa SC2 (SC2), B. pumilus ATCC 7061 (ATCC7061), B. subtlis 168 (168), B. subtilis IN937b (IN937b), Brevibacillus brevis NBRC 100599 (NBRC100599), B. licheniformis ATCC14580 (ATCC14580). Conserved amino acid residues were shaded in the following manner: black = 100% identity or conserved substitution, grey ≥ 75% identity, light grey ≥ 50% identity, white < 50% identity.

Figure 3-10. Phylogram of sfp homologs from several Bacillus strains. Multiple sequence alignment of sfp homologs from B. amyloliquefaciens FZB42 (FZB42), B. cereus E33L (E33L), B. megaterium QM B1551 (QMB1551), B. polymyxa SC2 (SC2), B. pumilus ATCC 7061 (ATCC7061), B. subtlis 168 (168), B. subtilis IN937b (IN937b), Brevibacillus brevis NBRC 100599 (NBRC100599), B. licheniformis ATCC14580 (ATCC14580). An unrooted neighbor-joining phylogenetic tree using Poisson-correction (Booststrp values: 1000 replications) distances was constructed based on the amino acid sequences alignment.

Page 110: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

110

CHAPTER 4 THE ROLE OF SUBTILOSIN A-LIKE ANTIMICROBIAL PEPTIDE IN SUPPRESSION

OF PHYTOPHTHORA CAPSICI BY BACILLUS SUBTILIS IN937B

Introduction

Bacillus spp. are widely recognized as producers of secondary metabolites of

low-molecular-weight compounds that are deleterious to other microorganisms

(Thomashow et al., 2007). Approximately 4% of the B. subtilis 168 genome is devoted

to the biosynthesis of secondary metabolites (Kunst et al., 1997). Similarly,

approximately 8.5% of the whole genome in B. amyloliquefaciens FZB42 is dedicated to

synthesize antibiotics which have significant potential for plant disease management

(Chen et al., 2007).

The classification of antimicrobial metabolites produced by Bacillus spp. is largely

based on non-ribosomal or ribosomal biosynthesis (Montesinos, 2007). Non-ribosomal

peptide antibiotics are often modified posttranslationally, such as lipopeptides (surfactin,

Chapter 3) and polyketides (Stein, 2005). This group of antibiotics possess antimicrobial

capabilities, also involved in biofilm, swarming development, and induction of plant

resistance (Ongena et al., 2005; Ongena et al., 2007). Secondary metabolites through

ribosomal biosynthesis act as pheromones in quorum-sensing, or lead to programmed

cell death. For example, bacteriocins have fungicidal and bactericidal capabilities for

suppressing Alternaria solani and Erwinia carotovora (Hammami et al., 2012).

Antibiotic production by Bacillus spp. may play a major role in maintaining

population density when they are introduced into the soil. The nutrition resources for

microbial proliferation and survival are limited in the rhizosphere. Over time,

consumption of the limiting nutrients will influence the dynamics between Bacillus and

other microbes. Bacillus spp. compete with their neighbors for space and nutrients by

Page 111: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

111

impairing or killing other microbes (Hibbing et al., 2010). As microbial growth inhibitors,

antibiotics result in changes of microbial community structure by inhibiting growth of

different species or related strains of the same species. In general, Bacillus spp.

produce a wide range of antibiotics that are highly specialized for a given habitat may

target different potential competitors (Stein, 2005). On the other hand, the stimulation or

repression of biosynthesis of antimicrobial metabolites helps Bacillus respond

appropriately to the environment such as changes in minerals and carbon levels

(Mannanov and Sattarova, 2001).

Antifungal antibiotic production is involved in disease suppression. Bacillus spp.

produce many antibiotics which directly suppress plant fungal pathogens, such as

Rhizoctonia solani, Botrytis cinerea, Verticillium dahliae, and Sclerotinia sclerotiorum

(Raaijmakers et al., 2002; Ongena and Jacques, 2008). Some commercial Bacillus

strains such as FZB42, GB03, QST713, and MBI600 produce cyclic lipopeptides (CLPs)

including fengycin, surfactin, or iturin. These CLPs have been associated with

suppression of plant pathogens (Joshi and Gardener, 2006). For example, iturins and

surfactin were active against R. solani, and pure fengycin inhibited Pyricularia oryzae

(Asaka and Shoda, 1996; Joshi and Gardener, 2006). Studies on surfactin lipopeptides

showed that surfactins caused hyphal swelling and fungal growth inhibition (Vitullo et

al., 2012). When growing on an agar surface supplemented with surfactins, F.

oxysporum mycelia exhibited retarded growth accompanied by increased branching and

rosette formation as well as hyphal swelling. The production of bacillomycin and

fengycin by B. subtilis UMAF6614, UMAF6639 and UMAF8561 were implicated as the

mechanism of control of powdery mildew on melon caused by Podosphaera fusca

Page 112: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

112

(Romero et al., 2007). Mycosubtilin-overproducing mutants of B. subtilis ATCC6633

were more effective than the wild-type strain in controlling Pythium damping-off on

tomatoes (Leclere et al., 2005).

Disintegration of membranes is well known as the primary function of antibiotics

against bacteria, fungi and oomycetes. Subtilosin A produced by several B. subtilis

strains (Stein et al., 2002) has a macrocyclic structure with three inter-residual linkages

(Marx et al., 2001), which involve permeabilized cell membrane by insertion of the

peptide after interaction with a cell-surface receptor (Moll et al., 1999). The damage

results in irreversible modifications leading to cell death (Aly et al., 2013). For

oomycetes, zoospore germination and sporangial formation represent the important

steps in the asexual life cycle and the spreading of disease, and as a result,

suppressing zoospore and sporangial germination may be an efficient strategy for the

management of diseases caused by oomycetes (Yoshida et al., 2001). A close

correlation between the production of antifungal compounds and the prevention of

zoospore germination has been reported in biocontrol of Pythium by B. cereus (Shang

et al., 1999).

To date, only a small proportion of the antimicrobial compounds produced by soil

bacteria have been studied (El-Hassan and Gowen, 2006). One challenge is that only 1-

5% of soil bacteria can be cultured. Moreover, biosynthesis of antibiotics by

microorganisms is often regulated by nutrition resources (carbon level) and greatly

influenced by environmental factors (pH and temperature) in the soil.

In the context of biocontrol, a successful exploratory attempt to repress

Phytophthora blight by using antagonistic B. subtilis has been reported on pepper (Lim

Page 113: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

113

and Kim, 2010). The mechanisms involving Bacillus spp. against Phytophthora blight on

squash have not yet been adequately explored. The Bacillus PGPR strain IN973b

significantly suppressed Phytophthora blight on squash in the greenhouse and

displayed in vitro inhibitory activities against P. capsici.

The specific objectives of this study were to investigate the mechanisms involved

in biocontrol of P. capsici on squash with B. subtilis strain IN937b. To evaluate the

inhibitory effect of antifungal compounds from the cell-free culture of IN937b, P. capsici

zoospore germination, hyphal growth, and its effect in detached leaf assays were

determined. To evaluate the role of antifungal compounds secreted by B. subtilis

IN937b in the inhibition of P. capsici, we set out to determine the gene that is

responsible for biosynthesis of Subtilosin A like litibiotics and its structural

characteristics.

Materials and methods

Dual Culture Assay

All the bacterial isolates were screened for antagonism against P. capsici on

potato dextrose agar (PDA) plates using the dual culture technique (Yoshida et al.,

2001). One plug of P. capsici isolate #151 mycelium (6-mm diameter) was placed in the

center of PDA medium. The bacterial strain was streaked on the other side of medium 3

cm away from the center. The plates were incubated at 25 °C for 7 days after which the

inhibition zone was measured and recorded.

Antibiotic Extraction and Analysis

For isolation of the antagonistic compounds responsible for inhibiting fungal

growth, strain IN937b was cultured in No.3S medium (Tsuge et al., 2005) and incubated

at 28°C on a rotary shaker incubator at 150 rpm for 2 days. The culture was centrifuged

Page 114: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

114

at 3,000 rpm for 30 min and the supernatant was collected and acid precipitated as

described in Chapter 3. The methanol (HPLC grade) extracts were further analyzed by

HPLC for preliminary characterization.

Antifungal Activity of Methanol Extracts Against Phytophthora capsici

Mycelial Growth. Crude extract of IN937b cell-free culture was dried in a clean

fume hood (Labconco Corporation, Kansas City, Mo). The dry matter was dissolved in

the methanol (40 mg/mL). The effective concentration (EC50) of the extracts was

determined for P. capsici by serial dilution. The EC50 was considered to be the lowest

concentration of the methanol extract that caused 50% inhibition of fungal growth. The

corn meal (CM) agar was amended with methanol extracts at a final concentration of 0,

0.2, 0.4, 0.8, 1.0 mg/mL. An agar plug (6 mm in diameter) was taken from the colony of

P. capsici on PARPH medium and placed at the center of CM agar amended with

methanol extracts. Each treatment had 3 replicates. All the plates were incubated at 25

°C for 9 days. The diameter of the colony was measured every 3 days after incubation.

The inhibition rate was defined as 𝐈𝐧𝐡𝐢𝐛𝐢𝐭𝐢𝐨𝐧 𝐫𝐚𝐭𝐞(%) = [(𝐍𝟎 − 𝐍𝟏) / 𝐍𝟎] × 𝟏𝟎𝟎%,

where N0 is the mycelium diameter at a final concentration of 0 mg/mL, and N1

represents the mycelial diameter of colonies on the medium amended with methanol

extracts at different concentrations.

Zoospore Germination. Zoospore suspensions were prepared using the

method described in Chapter 2. One hundred microliters of zoospore suspension were

mixed with 100 µL of methanol extracts at the final concentration of 0, 10, 20, 30, or

40mg/mL (total volume is 200 µL) for the extracts. The mixture was flooded on the CM

agar and placed in the hood to allow to air dry. The plates were incubated at 25 °C for 3

h and the percentage of zoospore germination on each plate was determined by

Page 115: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

115

viewing under a microscope at 400× magnification. The relative proportion of

germinated zoospores was calculated (Keinath, 2007) by dividing the number of

germinated zoospores by the total number (50) of zoospores counted on each plate.

Two plates were used for each concentration as replicates and the experiment was

conducted twice under the same conditions.

Sporangial Formation. Phytophthora capsici was grown on plugs in the DI

water with the methanol extracts added at final concentrations of 0 and 40 mg/mL of

crude extract. Four Petri dishes were used for each concentration and were placed

under the light (one 20-watt General Electric F20T12-CW linear fluorescent lamp) at

room temperature to induce sporangial formation. After 48 h, the plug was examined for

sporangial formation under a dissecting microscope at 70× magnification. The number

of sporangia formed on the plug was counted.

Detached Leaf Assays

Detached leaf assays described by Chiou and Wu (2001) were modified to

investigate the direct and indirect effects of methanol extracts on P. capsici on the

leaves. Young, fully developed leaves were collected from 2-week-old squash plants

and the leaf surfaces were rinsed three times with sterile DI water. Twenty microliters of

methanol extracts were dropped onto the detached leaves, and leaves were placed in

the Petri dishes supplemented with moist filter paper. Twenty microliters of zoospore

suspensions (104 zoospore/mL) were placed on leaves 3 cm away from the site where

methanol extracts were applied. The leaves were placed under the light at room

temperature for 7 days. Disease severity was visually assessed using a scale of 0-5,

where 0 = no symptoms, 1 = less than 10%, 2 = 10-25%, 3 = 50%, 4 = 51-75%, and 5 =

Page 116: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

116

75-100% of leaves covered with brown lesions (Chiou and Wu, 2001). Leaves treated

with methanol served as a non-treated control. Each treatment had three replications.

HPLC/MS Assays of Surfactin Production

Subtilosin A was quantified using HPLC/ESI-MSn by Dr. Jodie V. Johnson as

described in Chapter 3.

DNA Extraction and PCR Analysis

Bacillus PGPR strain IN937b was inoculated into 10 mL of LB medium and

incubated on a rotary shaker incubator at 150 rpm overnight at 28 °C. DNA extraction

was carried out using the Qiagen QIAamp DNA mini kit according to the manufacturer’s

instructions. PCR reactions were carried out based on the methodology and specific

primers established by Velho et al. (2013). The following primers were used for

amplifying the subtilosin A (sboA) gene that are critical for production of the

antimicrobial peptides subtilosin A:

sboA-f: 5’-CATCCTCGATCACAGACTTCACATG-3’

sboA-r: 5’- CGCGCAAGTAGTCGATTTCTAACAC-3’.

The PCR reaction was performed in a final volume of 25 µL containing 2.5 µL of

10X PCR buffer, 0.5 µL of dNTP mix, 1.0 µL each of PCR primers, 0.5 µL of Taq DNA

polymerases, and 1 µL each of template genomic DNA. Amplification was done with a

conventional thermal cycler (BioRad, Laboratories, Hercules, CA) with an initial

denaturing at 94 ℃ for 1 min, annealing at 50 ℃ for 30s and extension at 72 ℃ for 1

min, for a total of 35 cycles, and followed by a final extension of 72 ℃ for 10 min.

Following amplification, the PCR reaction mix was separated through 1% agarose gel

electrophoresis.

Page 117: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

117

Results

Dual Culture Assays

To evaluate the antifungal activities of all selected Bacillus PGPR, bacterial

strains were tested using a dual culture assay. Inhibition zones were clearly observed

between the mycelial growth and bacterial colonies (Figure 4-1). Results of the dual

culture assays for these strains are given in Table 4-1. Strain 1PN-19 exhibited the

greatest inhibition to growth of P. capsici with an inhibition zone of 11 mm. Bacillus

strains Companion, GB03, IN937a and IN937b showed moderately inhibitory activities

with an inhibition zone less than 10 mm, whereas strains SE34, SE49, SE52, SE76,

INR7, 1PC-11 and Actinovate AG had no effects on the growth of P. capsici.

Methanol Extracts of IN937b Culture

The active substances were collected from the cell-free culture of IN937b

(surfactin production No. 3s medium) described in Chapter 3. At pH 2.0, the substances

of antibiotic activity were precipitated from the cell-free supernatant and stabilized in

acidic solution at 4 °C. The active substances were then extracted by methanol under

neutral conditions (adjusted to pH=7.0). After evaporation, the dry weight of the active

substances was approximately 40 mg from 1 liter of liquid medium culture.

Chromatographic Separation by HPLC

The extracts contained numerous compounds. A compound with MW 3400 Da in

the extracts was abundant. At a retention time of 45-49 min, an m/z ratio for an ion had

been selected for fragmentation. The compound produced m/z 1699 [M-2H]2- and m/z

1710 [(M-H+Na)-2H]2- ions in (-)ESI-MS and m/z 1701 [M+2H]2+ and m/z 1712 [(M-

H+Na)+2H]2+ ions in (+)ESI-MS. The molecular mass of the corresponding metabolite

Page 118: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

118

has the same molecular weight as subtilosin A which has been found to be produced by

a great variety of B. subtilis strains (Stein, 2005).

PCR Detection of sbo Genes in Bacillus PGPR IN937b

A 666-bp PCR product was amplified from IN937b using the primers sboA-f and

sboA-r specific to fragment corresponding to the B. subtilis subtilosin gene. DNA

sequences analyzed by BLAST software revealed that this PCR fragment in IN937b had

100% similarity to B. subtilis 168 (Accession No. AL009126) whose genome has been

fully sequenced. The results indicated that IN937b is positive for the presence of the

gene sbo for cyclic peptide antibiotic subtilosin A synthesis. Amplification of the PCR

products indicates that IN937b may produce subtilosin A. The result is consistent with

the observation in the samples tested by HPLC.

Assays of Hyphal Growth Inhibition

The antifungal activity of the methanol extracts was confirmed from the culture of

IN937b. One hundred microliters of methanol extracts (40 mg/mL) were pipetted into

holes (6 mm diameter) 2 cm away from the edge of the CM agar. The methanol extracts

of IN937b strongly inhibited mycelial growth of P. capsici on CM agar 7 days after being

added into the hole (Figure 4-2 A.). The average growth of the colony toward the hole

was less than 5 mm, representing an inhibition of 75% compared to the unaffected

colonies. When the concentration of the methanol extract solution was reduced to 10

mg/mL, the hyphae of P. capsici were able to continue growing toward the inhibition

zone, but the layer of hyphae were thinner and more branches were observed

compared to the unaffected hyphae (Figure 4-2 B.).

Page 119: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

119

Effect of Methanol Extracts on Mycelium Morphology

The hyphal morphology affected by the methanol extracts was examined in detail

by microscopy. The tips of P. capsici hyphae grown on CM agar plants were normal

(Figure 4-3 A.) from the edge of the mycelia, whereas the extract challenged hyphae of

P. capsici were typically induced branching, and the hyphal tips became swollen as

shown in Fig 4-3.

The EC50 of the Extracts

The 0.2, and 0.4 mg/mL methanol extracts had no effects on mycelial growth of

P. capsici (Table 4-1). None of the extracts completely inhibited mycelial growth. The

extracts at 0.8 and 1.0 mg/mL had a stronger effect on the mycelial growth than the

other concentrations. The estimated EC50 of the extracts was 1.0 mg/mL.

Effect of Methanol Extracts on Zoospore Germination

Zoospores began to germinate 2 h after incubation at 28 °C. As the concentration

of methanol extracts increased, a prominent reduction in the percentage of germinated

zoospores was observed (Figure 4-4). The greatest inhibition (46.2%) was achieved

with the 40 mg/mL extract, and none of the tested concentrations completely inhibited

P. capsici.

Effect of Methanol Extracts on Sporangial Formation

For the assessment of the inhibitory effects of methanol extracts on sporangial

formation, P. capsici hyphae were induced under the light for sporangial formation for 2

days at room temperature. Normal sporangia with an intact shape were present in DI

water. The hyphae treated with methanol extracts had deformed and swollen tips, the

same as the observations for mycelial growth inhibition (Figure 4-5). The 40 mg/mL

Page 120: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

120

methanol extract completely inhibited sporangia formation (Table 4-1), and adversely

affected the production of sporangia.

Effect of Methanol Extracts on Detached Leaf Infection

The methanol extracts were evaluated for their potential to suppress

Phytophthora blight using a detached leaf bioassay (Figure 4-6). Fully expanded young

leaves of squash plants from the greenhouse were used in this experiment. The 40

mg/mL extract significantly (p≤0.05) inhibited the development of Phytophthora blight

compared to the methanol control (Table 4-3). No disease developed on leaves treated

with the 40 mg/mL extract.

Discussion

Bacillus subtilis strain IN937b isolated from tomato rhizosphere (Murphy et al.,

2000) showed inhibition activities against P. capsici on PDA plates (Zhang et al., 2010).

Subtilosin A, an antimicrobial peptide produced by B. subtilis IN937b was found to be

involved in reduction in zoospore germination, sporangial formation and hyphal growth

of P. capsici, and in vitro the suppression of lesion expansion on the detached leaves

after being challenged with P. capsici.

B. subtilis is well known for producing various antimicrobial substances as

secondary metabolites to inhibit plant pathogens (Mannanov and Sattarova, 2001).

Antibiotics secreted by B. subtilis include non-ribosomally synthesized peptide

antibiotics (e.g. surfactin and mycobacillin) and ribosomally synthesized precursor

lantibiotics (e.g. subtilin and subtilosin A) (Stein et al., 2002; Bais et al., 2004b; Leclere

et al., 2005; Arguelles-Arias et al., 2009).

The antifungal activity of bacteria might be due to competition for nutrients with

other microorganisms (Janisiewicz et al., 2000). Bacillus strains which contribute to

Page 121: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

121

biologically controlling plant diseases synthesize antibiotic compounds that suppress or

kill other microbes, modify microbial attachment to plant tissue surfaces, and improve

survival of the Bacillus cells in their habitat. The findings in our studies, together with the

fact that biocontrol efficiency of Bacillus strains were closely correlated with antibiotics

production, strongly support the relevance of antibiosis as an important factor involved

in the protection of squash against P. capsici by these PGPR strains.

Hyphal Growth Inhibition

Sixteen Bacillus strains were tested for their ability to inhibit hyphal growth of P.

capsici on PDA plates. These strains varied in their capacities against P. capsici. Strain

IN937b showed moderate levels of fungal inhibition (inhibition zone was approximately

5 mm). Thus, the antibiotics produced by IN937b were presumably partially responsible

for suppressing the hyphal growth of P. capsici (Ahmed et al., 2003; Zhang et al., 2010).

Methanol extract of strain IN937b was increased inhibition of hyphal growth

compared to IN937b cells did on PDA plates. In the present study, antibiotics produced

by IN 937b were condensed by precipitation with HCl from the culture supernatants and

were extracted from the pellet with methanol (Vater et al., 2002). The condensed

extracts had a 40-fold increase in inhibition of hyphal growth compared with the culture

of IN937b. The very weak antifungal activity displayed by the culture of IN937b

suggested that antibiotics are produced in very small quantities. The methanol extracts

could have resulted in the enrichment of antibiotics and therefore led to the increased

antifungal activity. Application of antibiotics rather than bacterial cells into agricultural

systems may improve disease control (McManus et al., 2002).

Page 122: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

122

Zoospore Germination and Sporangial Formation

We demonstrated that zoospore germination and sporangial formation of P.

capsici were significantly inhibited by the methanol extracts of IN937b. Zoospore

germination is believed to be the most important phase in the asexual life cycle and

spread of P. capsici (Judelson and Blanco, 2005). In the present study, the extracts

were able to reduce P. capsici zoospore germination with results similar to those

previously reported (Kone et al., 2009). Microscopic observations revealed that

sporangial formation of P. capsici was completely inhibited in DI water pretreated with

the extracts. These findings support the hypothesis that secretion of antifungal

compounds is the primary mechanism contributing to the suppression of B. subtilis

strain IN937b against P. capsici. Antibiotics are capable of modifying the fungal

membrane permeability causing enlargement of the mycelium and distortion of

sporangia and zoospores (Leelasuphakul et al., 2008).

Detached Leaf Assays

In this study, we used detached leaves to assess disease severity on squash

leaves treated with the extracts produced by IN937b. Although conditions in this assay

were highly favorable to P. capsici infection, the extracts of IN937b, applied before

pathogen challenge, completely suppressed the disease development on the attached

leaves. However, if treated with the extracts simultaneously or after the pathogen

inoculation, the extracts failed to suppress the disease development. Therefore,

antibiosis by the strain IN937b plays a major role in the suppression of Phytophthora

blight in squash. Inhibition of P. capsici on squash leaves by the extracts observed in

this study suggests that the extracts do not have a therapeutic effect but exert

preventive activities against the Phytophthora blight disease (Curtis et al., 2004). Once

Page 123: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

123

the pathogen is established in the plant tissues, the effect of suppression on mycelial

growth diminished.

Identification of Subtilosin A

Subtilosin A was isolated from IN937b culture, and identified by HPLC/ESI-MS.

Subtilosin A, with a molecular mass of 3399.7 Da (Marx et al., 2001), is a cyclic

polypeptide that is ribosomally synthesized in Bacillus spp.. HPLC analysis showed that

there were large quantities of subtilosin A in the methanol extracts. The initial pH and

culture temperature are critical factors for microbial growth and metabolic biosynthesis,

and they substantially affect antibiotic production (Gong et al., 2006).

Subtilosin A targets cell membranes to form transient pores inducing lipid-

perturbation and contributing to its antibiotic activity. It has been reported that subtilosin

A involves permeabilization of target cell membranes (Thennarasu et al., 2005).

Subtilosin A promotes internal osmotic imbalance and wide cytoplasmic disorganization

through abundant vacuoles formation. The aggregation of cytoplasm increases along

with a loss of characteristic organelles. These changes result in the plasmolysis of

fungal cells and ultimately cause morphological damage visible in sporangia and

zoospore germination.

The exploitation of antibiotics as active ingredients of pesticides is promising in

developing novel fungicides to control crop diseases. Development of compounds

suitable for agricultural use as pesticide ingredients requires intrinsic toxicity and

stability. Therefore, exploring less toxic and more stable compounds may be one of the

most promising strategies for developing novel biocontrol agents. B. subtilis strains

could be important candidates for biocontrol agents because they are known to produce

more than two dozen antibiotics (Stein, 2005).

Page 124: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

124

In conclusion, we have demonstrated that secondary metabolites secreted by B.

subtilis IN937b strongly inhibited the growth of P. capsici. Although antibiosis seems to

be an important mechanism involved in biocontrol activity of the active compounds.

Further studies of the modes of action of these secondary metabolites are currently

underway in order to improve their efficacy.

Page 125: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

125

Table 4-1. Antagonism of Bacillus PGPR strains against P. capsici on PDA plates

Bacillus strains Isolates of P. capsici

#121 #146 #151

SE34 - - -

SE49 - - -

SE52 - - -

SE56 - - -

SE76 - - -

INR7 - - -

1PC-11 - - -

Actinovate AG - - -

1PN-19 +++ +++ +++

BUEX 1216S ++ ++ ++

BUEX 1216C ++ ++ ++

Companion + + +

GB03 + + ++

IN937b + + +

IN937a ++ ++ ++

- Represents no inhibition zone. + Represents 1–5 mm wide inhibition zone, ++ represents 6–10 mm wide zone, +++ represents more than 10 mm wide zone. Values are average of three replications.

Page 126: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

126

Table 4-2. Effect of Methanol Extracts of IN937b on hyphal growth of P. capsici

a All the plates were incubated at 25 °C for 9 days. The diameter of mycelial colony was measured 3, 6 and 9 days after incubation. b Each value represents the mean of three replicates. Means followed by the same letter within each column were not significantly different according to Fisher’s protected LSD test (p = 0.05).

Concentration

(mg/mL)

Diameter of mycelial colony (cm) and inhibition rate (%) a

Day 3 (%) Day 6 (%) Day 9 (%)

0.0 3.9 a b 0.0 e 6.7 a 0.0 c 8.0 a 0.0 e

0.2 3.7 b 5.3 d 6.6 a 1.3 c 7.5 b 6.3 d

0.4 2.9 c 25.8 c 4.4 b 34.3 b 7.1 c 11.3 c

0.8 2.7 d 30.8 b 4.4 b 34.3 b 5.5 d 31.2 b

1.0 2.0 e 48.8 a 3.1 c 53.8 a 5.1 e 36.3 a

Page 127: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

127

Table 4-3. Effect of Methanol Extracts of IN937b in Detached Leaf assays

Treatment

Disease severity a

Day 5 Day 7 Day 9

Methanol (CK) 1.1±0.3 b a 2.7±0.5 a 4.8±0.4 a

Methanol extracts

(40 mg/mL) 0 b 0 b 0 b

a Disease severity was evaluated based on the scale of 0-5: where 0 = no symptoms; 1 = less than 10%; 2 = 10-25%; 3 = 50%; 4 = 51-75%; 5 = 75-100% of leaves covered with lesions. b Each value represents the mean of four replicates. Means followed by the same letter within each column were not significantly different according to Fisher’s protected LSD test (p = 0.05).

Page 128: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

128

A

B

Figure 4-1. Antagonism of Bacillus spp. against P. capsici on PDA plates. In vitro assays of Bacillus strains to suppress P. capsici were conducted on PDA plates. A 6 mm-diameter mycelial plug of P. capsici was placed in the center of PDA. The bacterial strain was streaked a 4 cm long line 3 cm away from the center of plug. The plates were incubated at 25 °C for 7 days. A) strains IN937b and IN937a, CK: only mycelial plug of P. capsici on the PDA, B) strains SE76 and IPC-11, CK: only mycelial plug of P. capsici on the PDA.

Page 129: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

129

A

B

Figure 4-2. Antagonism of methanol extracts of IN937b against P. capsici on CM agar plates. One hundred microliters of methanol extracts were pipetted into each of the holes (6 mm diameter) 2 cm away from the edge of the CM agar. A 6 mm-diameter mycelial plug of P. capsici was placed in the center of CM agar plates. A) 100 µL methanol extracts (40 mg/mL), B) 100 µL methanol extracts (10 mg/mL). A thin layer and more branches were observed close to the hole (black arrow).

Page 130: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

130

A

B

Figure 4-3. Effect of methanol extracts of IN937b on morphology of P. capsici. Effect of methanol extracts on morphology of P. capsici under a light microscope. A) normal mycelia of P. capsici observed at the edge of the plate away from the extracts, B) abnormal mycelial swollen and branching of P. capsici towards to the extracts. Distorted morphology in the presence of methanol extracts (white arrows).

Page 131: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

131

Figure 4-4. Effect of methanol extract concentrations of IN937b on zoospore germination. One hundred microliters of zoospore suspensions (105 zoospores/mL) mixed with 100 µL methanol extracts at the final concentration of 0, 10, 20, 30, or 40mg/mL. Data are the percentage of germinated zoospores after 3 h incubation at 25 °C. The equation of regression analysis represents a dose response of germinated zoospores to the extract concentration, where x is the extract concentration and y is the zoospore germination rate. Each value represents the mean of four replicates. Means followed by the same letter within each column were not significantly different according to Fisher’s protected LSD test (p = 0.05).

83.8 a

73.1 a

54.6 b47.3 bc

45.1 bc

35.4 c

y = -1.1041x + 75.871R² = 0.86

0

10

20

30

40

50

60

70

80

90

0 10 20 30 40

Ge

rmin

ati

on

Rate

(%

)

Concentrations of methanol extracts (mg/mL)

Page 132: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

132

A

B

Figure 4-5. Effect of methanol extracts of IN937b on sporangial formation. The mycelial plug was placed in 15 mL DI water with 100 µL methanol extracts (40 mg/mL). The plates were placed under the light at room temperature to induce sporangial formation. After 48 h, the number of sporangia was examined under a dissecting microscope. A) Sporangia were observed without methanol extracts, B) No sporangia were found and swollen tips of hypha existed in treatment with methanol extracts.

Page 133: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

133

A

B

C

D

Page 134: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

134

Figure 4-6. Effect of methanol extracts on detached leaf infection. Squash leaves (less than 9 cm) were collected from 2 week-old plants. Twenty microliters of methanol extracts were dropped on to the detached leaves and placed in the Petri dishes supplemented with moist filter paper. One day later, 20 µL zoospore suspensions (104 zoospore/mL) were dropped on to the leaves 3 cm away from the site that treat with methanol extracts. A) Rating scale of disease severity. Disease severity was evaluated using a scale of 0-5: where 0 = no symptoms; 1 = less than 10%; 2 = 10-25%; 3 = 50%; 4 = 51-75%; 5 = 75-100% of leaves covered with brown lesions, B) Symptom development recorded 5 days after pathogen inoculation, C) Symptom development recorded 7 days after inoculation, D) Symptom development recorded 9 days after inoculation.

Page 135: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

135

CHAPTER 5 ELICITATION OF PLANT SYSTEMIC DEFENSE BY BACILLUS PGPR STRAINS

Introduction

Plants are constantly exposed to various pathogens. In order to prevent potential

plant infection by pathogens, plants possess an immune system that recognizes the

pathogens and activate defenses that protect plants from infectious fungi, oomycetes,

bacteria, and nematodes, etc. (Kuc, 1982). A first line of defense is called pathogen-

associated molecular pattern (PAMP) - triggered immunity (PTI), resulting in non-host

resistance that prevents further pathogen invade (Jones and Dangl, 2006). PTI leads to

fast defense responses via synthesis of toxins, antifungal proteins, and chemical signals

during initial stages of pathogen infection (Nicaise et al., 2009). However, some

pathogens develop strategies to suppress PTI. Therefore, plants have to develop a

second line of defense called effector-triggered immunity (ETI), which relies on systemic

signals produced at and dispersed from infection sites (Amil-Ruiz et al., 2011). In

general, ETI is associated with a hypersensitive response (HR) and a systemic

activation of plant defenses at the site of signal perception with pathogenesis-related

(PR) proteins, phytoalexins and reactive oxygen species.

Plant defenses are not only activated in response to pathogen infection, but also

can be stimulated by beneficial microbes. After exposure to biotic stimuli provided by

PGPR, plants can acquire enhanced levels of resistance to subsequent pathogen

attack. This phenomenon is commonly referred to as induced systemic resistance (ISR)

(Kloepper et al., 2004). Biological control using introduced PGPR with the capacity to

elicit ISR against plant pathogens has been extensively studied over the past two

decades in greenhouse and field trials (Yan et al., 2002). In many cases, ISR elicits

Page 136: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

136

plant immunity against multiple plant pathogens (Kloepper et al., 2004). For example,

select mixtures of compatible Bacillus spp. with the capacity to elicit ISR protect the

hosts against infection by pathogens Ralstonia solanacearum on tamato, Colletotrichum

gloeosporioides on pepper, and Rhizoctonia solani on green kuang futsoi (Brassica

chinensis var. parachinensis) (Jetiyanon and Kloepper, 2002).

A common feature of resistance responses induced by PGPR is “priming”, a

state in which plants treated with PGPR display fast and strong activation of defense

responses upon pathogen challenge. A priming effect is when PGPR strains cause the

plant to be in an “alert” state to detect pathogens with the response occurring faster

and/or stronger compared to plants not previously exposed to the priming stimulus

(Yang et al., 2009). Primed plants display a unique physiological status that can

respond rapidly to pathogen attacks (Conrath et al., 2006). For example, Bacillus sp.

CHEP5 and Pseudomonas sp BREN6 strains reduced root and stem wilt disease

severity caused by Sclerotium rolfsii on peanut (Arachis hypogaea L) (Laura Tonelli et

al., 2011). PGPR-mediated ISR caused a faster increase in phenylalanine ammonia-

lyase and peroxidase levels in peanut plants compared to non-treated control plants

after pathogen challenge, indicating that a state of priming was activated by treatment

with CHEP5 and BREN6. However, ISR usually does not confer adequate control, but

provides long-lasting increased resistance against a broad range of pathogens (Akram

et al., 2008).

ISR occurs when plants respond to an appropriate elicitor from the tissues

surrounding the initial infection site, which spreads throughout the plant via the emission

of molecular signals that reach distant tissues triggering increased resistance to the

Page 137: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

137

pathogens. The signaling of ISR is controlled by salicylic acid (SA), jasmonic acid (JA)

and ethylene (ET) (Pieterse et al., 2002). Paenibacillus alvei K165 primed for enhanced

SA-dependent defenses (Tjamos et al., 2005), while some other Bacillus spp. such as

Bacillus cereus AR156, were able to prime both the SA and the JA/ET pathways (Niu et

al., 2011). Recently, lipopolysaccharides, 2,4-diacetylphloroglucinol, and siderophores

produced by PGPR have been found to be responsible for ISR (Henry et al., 2011). All

these signaling molecules can distribute from local to far sites.

To defend against pathogen attack, there are two defense pathways that are

generally associated with an enhanced level of resistance conferred by ISR: the

phenylpropanoid pathway and the oxylipin pathway. Phytoalexin synthesis induction

through stimulation of the expression of genes involve phenylalanine ammonia-lyase

(PAL). The first step of the oxylipin pathway is catalyzed by PAL, which activates and

catalyzes phenylalanine to cinnamic acid, the precursor of lignin, salicyclic acid, and

phytoalexin phenylpropanoids, leading to secondary metabolites such as phytoalexins

that are toxic to fungi (Silva et al., 2004a). PAL is also involved in lignin biosynthesis

and accumulation of phenolics in response to fungal infection. Enzyme stimulation as

part of ISR is generally activated after pathogen infection (Choudhary et al., 2007).

ISR is associated with an oxidative burst (Bargabus et al., 2003) and cell wall

reinforcement (van Loon, 2007) by accumulation of a defense related enzyme

peroxidase (PO). PO shows affinity to cellular lignification and has direct antimicrobial

activity in the presence of hydrogen peroxide (Silva et al., 2004b). The oxidative burst is

the earliest event in plant defense response in both compatible and incompatible plant-

pathogen interactions as well as in rhizobacteria - plant interactions (Jourdan et al.,

Page 138: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

138

2009). In general, the transient primary burst is nonspecific and has no effect on

disease progression. Only during incompatible interactions, more prolonged peaks of

hydrogen peroxide production lead to specific recognition (Van Breusegem et al., 2001).

The enforcement of cell walls by lignification involves blocking of fungal pathogen

penetration. Upon challenge infection with the fungus, the root cell walls are rapidly

strengthened at the sites of attempted fungal penetration through appositions that

contain callose as well as phenolic materials and fungal ingress is effectively prevented

(Yoshida et al., 2005). Tobacco plants treated with B. pumilus SE34 exhibited newly

formed barriers beyond the site of fungal infection. These barriers were cell wall

appositions that contained large amounts of callose and were infiltrated with phenolic

compounds (Zhang et al., 2002). Phenolic compounds accumulated in host cell walls,

intercellular spaces, and on the surface and inside of the invading pathogen’s hyphae

(Nakkeeran et al., 2006).

ISR induced by Bacillus PGPR was shown to be generated in whole squash

plants and a split-root system. In studies of protection of cucumber against diseases

caused by soil-borne fungi, a split-root assay was developed in which the inducing

bacteria and the pathogen were simultaneously inoculated on separate halves of roots,

and then planted in separate pots (Liu et al., 1995). ISR was expressed as a delay in

symptom development, reduced disease severity, and reduced disease incidence

compared to the non-bacterized controls. In a previous study, the compatibility between

P. capsici and susceptible pepper plants (Capsicum chinense Jacq) could be a non-host

compatible interaction due to a consistent increase activity of defense enzymes a few

hours after inoculation with P. capsici (Goretty Caamal-Chan et al., 2011). It appears

Page 139: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

139

that P. capsici has the ability to block the defense response and to manipulate the

metabolism of the host by its effectors for successful infection (van Loon et al., 2006).

The objectives of this study were: (1) to understand the nature of ISR by Bacillus

PGPR IN937b and SE76 in squash plants with split-root assays by measuring the levels

of PAL and PO, (2) to evaluate the potential host defense response in squash plants

against P. capsici by monitoring PAL and PO activities, and (3) to analyze the changes

in enzyme activity of PAL and PO by treatment with Bacillus PGPR applied at different

squash growth stages.

Materials and Methods

Inoculum Preparation and Inoculation

Inoculum of P. capsici isolate #151 was prepared according to Chapter 2. The

concentration of zoospores was determined under a microscope using a

hemocytometer.

For foliar inoculation, the concentration of P. capsici inoculum was adjusted to

103 zoospores/mL using DI water. Plant leaves were sprayed with 15 mL of the

inoculum until run off. Each inoculated plant was covered with a plastic bag to maintain

high moisture for 2 days. Inoculated plants were maintained on benches in the

greenhouse with high soil moisture by hand watering daily.

Application of Bacillus PGPR Treatment in the Greenhouse

PGPR strains IN937a, IN937b and SE76 were evaluated for their ability to

systemically induce resistance against P. capsici in the greenhouse. Two-week-old

squash plants were treated with 20 mL of PGPR suspensions prepared according to

Chapter 2 at 1×108 cfu/mL as a soil drench. One week later, the same treatment with

PGPR were applied to the seedlings for a second time.

Page 140: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

140

Split-Root Assay

To set up the split-root systems, germinated seeds were planted in 60-well

plastic trays filled with potting mix. After 10 days, the roots were washed with running

tap water. The roots were then split evenly with a scalpel blade and each half of the root

system of each seedling was quickly moved into two adjacent plastic pots, each

containing potting mix. Each pot was separated to guarantee physical separation of

each half root. The plants were placed on the bench in a greenhouse.

One week after roots were spilt, one half of the root system of each plant was

divided into groups, each treated with a different inoculum. Group 1: one half of the root

system of each plant was treated with 20 mL of IN937b suspensions at 1×108 cfu/mL

applied as a soil drench labeled as IN937b-2. The other half of the root system of each

plant were treated with same amount of DI water labeled as IN937b-1. Group 2: One

half of the root system of each plant was treated with 20 mL of ASM at 30 mg/L applied

as a soil drench labeled as ASM-2. The other half of the root system of each plant were

treated with same amount of DI water labeled as ASM-1. Group 3: Both halves of the

root system of each plant were treated with DI water labeled as non-treated control (CK-

1 and CK-2).

Enzyme Activity Assays

Samples of fresh squash leaves and roots were taken from each of the pots 1, 3,

5, 7, and 10 days after IN937b treatment. The samples were washed with running tap

water, dried and 0.5 g of each was ground in liquid nitrogen in a mortar with a pestle.

The ground tissue was then mixed with 0.1 M sodium phosphate buffer (pH 6.5) and

centrifuged at 12,000 rpm for 10 min at 4 °C. The supernatant used for the enzymatic

activity assay was stored in a 1.5 mL Eppendorf tube at -20 °C. Healthy plants with no

Page 141: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

141

PGPR or pathogen challenge were used as a background. Background PAL and PO

activity was found to be same during experimental period.

Enzyme Activity Assays

Peroxidase (PO) Fifty microliters of roots or leave extracts were added to 2.95

mL of PO reaction buffer consisting of 50 mM sodium phosphate buffer (pH 6.0), 60 mM

guaiacol and 0.6 M H2O2 (Chen et al., 2000). The reaction was measured for 90 s

(recording the value every 15 s) at wavelength of 470 nm with a spectrophotometer

(UNICO SpectroQuest TM 4802 UV/VIS Double Beam, United Products & Instruments

Inc., Dayton, NJ). One unit of enzyme activity was defined by a change in absorbance

of 0.01 for 1 g fresh weight per minute.

Phenylalanine ammonia-lyase (PAL) One hundred µL of crude plant tissue

(roots or leave) extracts were mixed with 1.9 mL of 0.015 M L- phenylalanine and 0.1 M

Tris-HCl buffer solution (pH 8.5) (Chen et al., 2000). The mixture was placed in a water

bath at 40 °C for 30 min. PAL activity was measured spectrophotometrically at a

wavelength of 290 nm. One unit represents the conversion of 1 µmol L-phenylalanine to

cinnamic acid for 1 g fresh weight per minute at 40 °C.

Results

Induction of Systemic Resistance by PGPR with P. capsici Challenge

Soil treatment with IN937b resulted in a significant (P≤0.05) visible reduction in

lesion development (both number and size) compared with the non-treated control 7

days after inoculation (Table 5-1). SE76 failed to elicit ISR because the number and

size of lesions increased as much as in the non-treated control. ASM was the most

effective treatment in this experiment to induce ISR, as it completely stopped the

occurrence and spread of symptoms.

Page 142: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

142

Treatment with IN937b at 108 and 107 CFU/mL significantly (P≤0.05) reduced the

size of lesions compared to the non-treated control 7 days after pathogen inoculation

(Table 5-2). The greatest reduction in lesion size by IN937b was at 108 CFU/mL, where

there was more than 70% reduction compared to the non-treated control plants. IN937b

at 107 and 106 CFU/mL had no effect on disease development compared to the non-

treated control. ASM completely inhibited the disease during the period of this

experiment.

Increased Activities of PO and PAL Induced Locally by PGPR Treatment with P. capsici Challenge

Compared to the non-treated plants, SE76 induced significantly greater levels of

PAL and PO activities in plants roots 3 days after PGPR treatment (Figure 5-1 and

Figure 5-2). When squash roots were challenged with P. capsici, significantly greater

PAL activities were observed in all PGPR treatments and ASM 3 DAI compared to the

non-treated control (Figure 5-3). PAL activities were significantly greater in treatments of

ASM and IN937b than the non-treated control 5 DAI. In general, levels of PAL increased

drastically 1 DAI in all treatments due to the rapid response of the squash plants to the

pathogen. However, PAL activities 3 DAI significantly decreased in all treatments except

for the background control compared to those at 1 DAI.

Bacillus PGPR IN937b and SE76 induced significantly greater levels of PO

activities 5 DAI compared to the non-treated control (Figure 5-4). PO activities reached

the highest levels 3 DAI in all treatments and then decreased 5 DAI. In the non-treated

control plants, PO activities significantly increased 1 and 3 DAI when the pathogen

started to invade squash plants, but dramatically declined 5 DAI when stem wilting

started to occur. Treatment with IN937a had no effect on PO activities, while ASM

Page 143: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

143

treatment resulted in significantly greater levels of PO activities 3 and 5 DAI compared

to the non-treated control plants.

Systemically Induced Enzyme Activity as Determined by Split-Root System without P. capsici Challenge

Data from the split-root system assays showed that IN937b not only locally

increased levels of PO and PAL activities in the root half that was bacterized, but also

systematically increased PO and PAL activities in the other half (distant side) of the root

system which was only treated with water at 10 days after IN937b treatment (Figure 5-5

and 5-6).

Increased PAL and PO Activities Systemically in Response to P. capsici in Different Seedling Stages

Treatment with IN937b and SE76 induced significantly higher PAL activity in

plant leaves compared to the non-treated control 5 DAI for young and older plants (Fig

5-7). When first PGPR treatment was applied 17 days after emergence, significantly

greater PAL activity were detected 5 DAI in all PGPR treatments and ASM compared to

the non-treated control. Similarly, significantly greater PAL activity were detected 5 DAI

in all PGPR treatments compared to the non-treated control when the first PGPR

treatment was applied 10 days after emergence.

Squash plants treated with IN937b, SE76 and ASM induced significantly greater

levels of PO activity in the leaves compared to the non-treated control 5 DAI in the

young plants with the first PGPR treatment 10 days after emergence (Fig 5-8). In older

plants for which the first PGPR treatment was performed 17 days after emergence,

there was no difference of PO activity were observed in all treatment 5 DAI which the

first PGPR treatment was applied 10 days after emergence. PO activity was greater in

leaves treated with ASM for both young and older plants.

Page 144: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

144

Discussion

The purpose of this study was to demonstrate that Bacillus PGPR strains are

able to trigger ISR in squash plants against P. capsici, and to reveal the underlying

mechanisms of PGPR suppression of Phytophthora blight disease. Bacillus PGPR

IN937b applied in the rhizosphere reduced the number and size of lesions on the leaves

infected by P. capsici. Since PGPR and the pathogen were spatially separated, the

protection is not associated with the ability of the PGPR strains to inhibit fungal growth

and, therefore, it may be, at least partially, attributed to ISR-mediated protection. The

results from this study demonstrated that both Bacillus PGPR and the pathogen were

able to affect activities of defense-related enzymes, i.e. PAL and PO.

Actigard® (ASM) is a well-known activator of host resistance in many plants

(Louws et al., 2001; Kone et al., 2009). In this study, ASM induced significantly systemic

resistance in squash, resulting in elimination of P. capsici infection. It suggests that

ASM may induce resistance more effectively than Bacillus PGPR (Barilli et al., 2010).

Defense-Related Enzymes

Plants suppress pathogens by building lines of defense and develop an array of

cellular mechanisms to defend themselves against invading pathogens (Ramamoorthy

et al., 2001). Plants can recognize a pathogen and limit it to the site of the infection

(Jones and Dangl, 2006), and stop pathogen infection and movement (Takemoto et al.,

2003).

In general, increases in enzyme activities related to plant defenses are important

indicators in response to disease development. PO is an oxido-reductive enzyme

involved in the wall binding process (Viswanathan and Samiyappan, 2002). PO has

been implicated in the last enzymatic step of lignin biosynthesis, which is the oxidation

Page 145: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

145

of hydroxyl which is subsequently coupled to lignin polymer. PO is also believed to be

involved in a number of physiological processes that may contribute to resistance

including cross linking of extensin monomers and lignification and they are also

associated with deposition of phenolic compounds into plant cell walls during resistance

reactions (Passardi et al., 2004). To be effective against a pathogen, defense

mechanisms, including increased PO activity, should be rapidly elicited in response to

infection. One of the immediate host responses to infection is the accumulation of

phenolics at the infection site. Similar results were also obtained in a previous study,

where microbial mixture of Bacillus strains IN937a and IN937b were found to induce

maximum SOD and PO activities compared with the non-treated control (Jetiyanon,

2007).

PAL is the first enzyme in the phenylpropanoid biosynthesis pathway leading to

synthesis of phytoalexins or phenols, which have defense functions in plants, such as

reinforcement of plant cell walls, antimicrobial activity and synthesis of signaling

compounds such as salicylic acid (Wen et al., 2005). In the present study, after

challenge with the pathogen P. capsici, PAL activities were positively correlated the

plant’s resistance against the pathogen as indicated by reduced disease severity. PAL

is essential for the synthesis of all protective substances through the phenylpropanoid

pathway. Generally, induction of PAL is correlated with increased resistance in the host

to pathogenic infection. PAL leads to the biosynthesis of phenols, phytoalexins, and

lignins with the involvement of PO.

Increased PAL and PO activities in plants by PGPR enhance the production of

phenolic compounds, and hence induce development of physical and chemical barriers

Page 146: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

146

to pathogen attack. Increased enzyme activities in PGPR-mediated ISR might be one of

the defense mechanisms involved in suppression of P. capsici. Phenolics generally

fungitoxic and increase the mechanical strength of the host cell walls. Lignin is a

phenolic polymer which is difficult for pathogens to breach and has been implicated in

plant defense against pathogens.

In the present study, induced responses in squash plants occurred in a short time

period. Resistance activated by PGPR did affect pathogen infection. However, following

the early and increased expression of defense enzymes, PAL activity dramatically

decreased after P.capsici infection, suggesting that the pathogen had the ability to

suppress induced defense responses in plants by interfering with signaling cascades

and detoxification of phytoalexins (El-Hadrami et al., 2009).

Locally Induced Enzyme Activity

When the squash roots were treated with Bacillus PGPR strains and inoculated

with P. capsici, the differential induction of resistance may have been activated

depending on the induction site. When roots were treated with Bacillus PGPR,

metabolic processes may have produced reactive oxidative oxygen species (ROS) such

as H2O2, O2 and OH (Demidchik, 2012). The production of toxic oxygen derivatives

increased as a result of infection by any foreign microbes. PGPR strains are able to

alleviate ROS via increased activities of antioxidant enzymes such as PO in plants, and

establish a relationship with plant roots without being harmful to roots (Zamioudis and

Pieterse, 2012). In the present study, the observed increased localized PO levels in

squash roots by bacterization are in agreement with the reports that cucumber roots

with PGPR induced localized resistance (Chen et al., 2000; Liang et al., 2011).

Page 147: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

147

Priming: Getting Ready for Pathogen Invasion

PGPR may alter the cellular mechanisms and reprogram metabolism after they

establish a relationship with plants, and induce plant resistance upon subsequent

challenge with a pathogen (Van der Ent et al., 2009). Plants treated with PGPR become

primed to respond faster and show stronger activation of cellular defense responses

upon pathogen challenge compared with unprimed plants (Conrath et al., 2002).

Treatment with Bacillus strains in roots also increased the activities of PO and PAL

enzymes in squash leaves. This is also supported by the observations in the split-root

assays that certain Bacillus strains induced increased PO and PAL activities in the other

half root which was treated with water and distant from the treated half roots.

Older plants tended to develop disease symptoms later compared to younger

plants after they were inoculated with the same amount of P. capsici. PAL and PO

activities were greater in older plants than younger plants in the non-treated control,

IN937b and ASM treatments. In general, the activities of PAL and PO increased in

these treatments with reduced disease severity in squash plants.

In the present study, PAL activity was activated by Bacillus PGPR IN937b and

SE76. It appears that the pathogen blocked the plant defense response development as

previously reported (Cooper and Rao, 2006). PO and PAL activities were promptly

triggered by PGPR treatments, and the levels of PO and PAL activities were faster than

the chemical inducer (ASM) did, and this seems to be a consensus for PGPR treatment.

PO activity is activated faster to contain the reactive oxygen species produced by host

plants that respond to inhibit pathogen growth (Ishida et al., 2008).

In conclusion, higher levels of PO and PAL activities may have contributed

collectively to induced resistance in squash plants against P. capsici, the pathogen of

Page 148: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

148

Phytophthora blight, which eventually resulted in lower disease severity compared to

the non-treated control.

Page 149: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

149

Table 5-1. Effect of Bacillus PGPR on number and size of lesions in the detached leave assay under the lab conditions

lesion size in cm2

Day 5 Day 7

IN937b a 9.0 a c 13.0 b

SE76 12.0 a 32.5 a

ASM 0.0 b 3.9 b

Non-treated control 11.9 a 44.6 a a PGPR were applied twice (1.0 ×108 CFU/mL; 20mL) as a soil drench. ASM (Actigard® 30 mg/L; 20 mL) was applied as a positive SAR inducer for suppressing Phytophthora blight. b P. capsici inoculum was diluted to 1 × 103 zoospores/mL. Plants were sprayed with 20 mL of P. capsici inoculum until ran off. Data of the number and size of lesions were collected at 5 and 7 days after pathogen inoculation. c Each value represents the mean of four replicates. Means followed by the same letter within a column are not significantly different according to Fisher’s protected LSD test (p = 0.05). Table 5-2. Effect of different concentrations of IN937b on the number and size of

lesions in the detached leave assay under the lab conditions

Treatment lesion size in cm2

Day 3 Day 5 Day 7

937b-108 2.5 a 4.0 a 5.0 b

937b-107 3.9 a 5.5 a 8.0 b

937b-106 6.6 a 8.5 a 10.5 a

Non-treated control

6.5 a 7.6 a 16.8 a

a PGPR were applied twice as a soil drench. The concentration of IN937b was adjusted to 106, 107, and 108 CFU/mL, and 20 mL was applied to each plant. ASM (Actigard® 30 mg/L; 20mL/plant) were applied as SAR control for suppressing Phytophthora blight. b P. capsici inoculum was diluted to 1 × 103 zoospores/mL. One leaf of each plant was sprayed with 5 mL of P. capsici inoculum until ran off. Data of the number and size of lesions were collected at 3, 5, and 7 days after pathogen inoculation. c Each value represents the mean of ten replicates. Means followed by the same letter within a column are not significantly different according to Fisher’s protected LSD test (p = 0.05).

Page 150: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

150

Figure 5-1. Induction of PAL by PGPR in squash roots. PAL activities in squash roots treated by PGPR only. Squash roots were sampled 1 day and 3 days after PGPR treatment. Bars represent the standard error of the mean from three replications. Means followed by the same letter are not significantly different according to Fisher’s protected LSD test (p = 0.05).

Time (Day)

PA

L (

U/g

FW

)

0

5

10

15

20

25

30

Non-treated Control

IN937b

SE76

IN937a

ASM

Background Control

1 3

bb

a

ab

ab

b

Page 151: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

151

Figure 5-2. Induction of PO by PGPR in squash roots. PO activities in squash roots treated by PGPR only. Squash roots were sampled 1 day and 3 days after PGPR treatment. Bars represent the standard error of the mean from three replications. Means followed by the same letter are not significantly different according to Fisher’s protected LSD test (p = 0.05)..

Page 152: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

152

Figure 5-3. Induction of PAL by PGPR in roots against P. capsici. PAL activities in squash roots treated by PGPR and challenged with P.capsici (3 days after PGPR treatment). Squash roots were sampled 0, 1, 3 and 5 day after P. capsici inoculation. Bars represent the standard error of the mean from three replications. Means followed by the same letter are not significantly different according to Fisher’s protected LSD test (p = 0.05).

Page 153: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

153

Figure 5-4. Induction of PO by PGPR in roots against P. capsici. PO activities in squash roots treated by PGPR and challenged with P.capsici (3 days after PGPR treatment). Squash roots were sampled 0, 1, 3 and 5 day after P. capsici inoculation. Bars represent the standard error of the mean from three replications. Means followed by the same letter are not significantly different according to Fisher’s protected LSD test (p = 0.05).

Page 154: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

154

Figure 5-5. Induction of PAL by IN937b in split root system. PAL activities on squash roots treated by IN937b only. CK-1: root with water treatment, CK-2: root with water treatment, IN937b-1: root with water treatment, IN937b-2: root with IN937b treatment, ASM-1: root with water treatment, ASM-2: root with ASM treatment. Squash roots were sampled 1, 3, 5 and 10 day after PGPR treatment. Bars represent the standard error of the mean from three replications. Means followed by the same letter are not significantly different according to Fisher’s protected LSD test (p = 0.05).

Page 155: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

155

Figure 5-6. Induction of PO by IN937b in split root system. PO activities in squash roots treated by IN937b only. CK-1: root with water treatment, CK-2: root with water treatment, IN937b-1: root with water treatment, IN937b-2: root with IN937b treatment, ASM-1: root with water treatment, ASM-2: root with ASM treatment. Squash roots were sampled 1, 3, 5 and 10 day after PGPR treatment. Bars represent the standard error of the mean from three replications. Means followed by the same letter are not significantly different according to Fisher’s protected LSD test (p = 0.05).

Page 156: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

156

A

B

Figure 5-7. Induction of PAL by PGPR against P. capsici in different growth stages. PAL activities in squash roots treated by PGPR and challenged with P.capsici (3 days after PGPR treatment). Squash roots were sampled 0, 1, 3 and 5 day after P. capsici inoculation. A) First PGPR treatment started at 10 days after emergence, B) First PGPR treatment started at 17 days after emergence. Bars represent the standard error of the mean from three replications. Means followed by the same letter are not significantly different according to Fisher’s protected LSD test (p = 0.05).

Page 157: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

157

A

B

Figure 5-8. Induction of PO by PGPR against P. capsici in different growth stages. PO activities in squash roots treated by PGPR and challenged with P.capsici (3 days after PGPR treatment). Squash roots were sampled 0, 1, 3 and 5 day after P. capsici inoculation. A) First PGPR treatment started at 10 days after emergence, B) First PGPR treatment started at 17 days after emergence. Bars represent the standard error of the mean from three replications. Means followed by the same letter are not significantly different according to Fisher’s protected LSD test (p = 0.05).

Page 158: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

158

CHAPTER 6 SUMMARIES AND FUTURE PERSPECTIVES

Squash is one of widely grown and consumed vegetable crops worldwide. The

large scale of squash plants suffers heavy losses with pests and diseases. One of the

most economically important diseases on squash is Phytophthora blight, caused by

Phytophthora capsici. Control of this disease, so far, has largely depended on

fungicides. Intensive use of fungicides leads to the appearance of less sensitive strains

and resistant strains. The risk of fungicide resistance causes environmental hazards

and decrease of fungicide efficacy. To prevent the excessive use of fungicides, an

integrated approach is needed for the control of Phytophthora blight, including biological

control along with limited fungicide use. PGPR strains have been demonstrated against

a broad spectrum of pathogens on many vegetable crops. The overall aim of this

dissertation was to investigate Bacillus PGPR strains for controlling Phytophthora blight

on squash and plant defense induced by PGPR.

Summaries

In this dissertation, I was able to show that the application of selected Bacillus

PGPR strains significantly increased plant growth and decreased Phytophthora blight

disease severity on squash plants (Chapter 2). Of the efficient strains tested, Bacillus

subtilis strain IN937b delayed disease development, enhanced the plant growth

parameters, and increased leaf chlorophyll content in leaves under greenhouse

conditions. Also, I established appropriate timing of IN937b applications with different

plant ages to obtain remarkable efficacy of disease control. IN937b applications prior to

the pathogen inoculation, offer adequate time or IN937b to multiply and establish on the

root surface, which prevents early infection by P. capsici.

Page 159: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

159

Another important result of this dissertation is that I have described biofilm

formation of selected Bacillus PGPR strains in the liquid or solid medium, and the

production of surfactins by IN937b (Chapter 3). Use of specific PCR-primers was

applied for the detection of genes involved in the biosynthesis of surfactins in Bacillus

spp. The PCR-based screening has been a useful and efficient approach for the

identification of potential surfactin producing bacteria, especially when the surfactin

primers were designed to detect genes involved in the biosynthesis. The production of

surfactins and biofilm suggests that biofilm may be important for rhizosphere

colonization by IN937b. The population dynamics of IN937b were investigated in the

commercial soilless media using serial dilutions and plating on selective media.

Population densities stayed high at the early stage of inoculation e.g. 1 day after

treatment, but decreased dramatically two weeks after treatment. IN937b also found in

the interior tissues of the roots as they have naturally evolved to compete and survive in

the harsh conditions and compete in space with the pathogen.

The antagonistic effects are mainly due to production of antimicrobial antibiotics,

which play a major role in biological control of plant pathogens. The results from present

study revealed the production of antifungal antibiotic subtilosin A from IN937b. The

efficacy of antibiotics extracted from the culture of IN937b was evaluated for in vitro

suppressing P. capsici in vitro (Chapter 4). The methanol extracts of IN937b inhibited

the germination of the zoospores and sporangia formation. It may stop or delay

zoospores entry into the host tissue at the site of infection, also influenced the

development of P. capsici hypha and accumulated in squash roots. Inhibition of the

pathogen is through the localized direct antifungal activity of IN937b and their cell-free

Page 160: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

160

culture extracts, which points towards the activity of antibiotics as the potential

mechanism of disease control.

Plant resistance against diseases can be systemically induced by plant

pathogens or non-pathogenic rhizobacteria. IN937b was selected for further studies for

its potential to induce systemic resistance in squash against P. capsici (Chapter 5).

Systemic and local induction of resistance by PGPR was assessed through increased

activities of plant defense related enzymes peroxidase (PO) and phenylalanine

ammonia lyase (PAL) in squash plants treated with PGPR and challenged with P.

capsici. To determine the existence of systemic resistance on squash plants, the

pathogen P. capsici was inoculated as inducers on leaves spatially separated from the

PGPR strain IN937b applied as a soil drench. The activities of PAL and PO were

determined in bacterized roots and distant induced squash leaves. The results

demonstrated that systemic resistance against the squash root disease was not only

induced by plant pathogens, but also by non-pathogenic PGPR. When disease

symptoms started to appear on the tissues, the enzyme activates reached a peak. This

reduced pathogen infection, and impaired the disease development on squash roots.

But the pathogen-induced enzyme activities was stronger than the PGPR-mediated ISR

on squash plants. A split-root technique was utilized on squash plants grown in soilless

potting mixes. IN937b stimulated PO activates in the bacterized side of roots locally and

in the distant side of roots systematically. Without P. capsici infection, PAL activities

were not influenced by IN937b locally and systematically.

Future Perspectives

Our data strongly support the hypothesis that PGPR strains provide remarkable

disease suppression as an integrated approach in vegetable crops production. Further

Page 161: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

161

studies need to be performed to evaluate the effect of PGPR in the field and to explain

how plants communicate with PGPR in the soil.

In particular, the following issues might be taken into account in the future:

To extend our results of Chapter 2, I will try to optimize the biocontrol efficacy in

the field including the rate of application, timing, number of applications, and bacterial

survival at specific level of pH, temperature and soil fertility under field conditions.

Furthermore, I will investigate the effect of PGPR combined with commonly used

fungicides in the field to reduce the use of chemicals, which could be integrated into

existing management programs for Phytophthora blight in squash.

Detection and visualization of biofilm formation in situ will be studied (Chapter 3).

Visualization in situ act as a powerful tool that allows us to investigate the spatial

colonization pattern of the pathogen and the biocontrol agent on the root surface, and

also to determine biofilm production pattern of the biocontrol agent coincide with the

spatial colonization of the pathogen. It will be interesting to characterize those signaling

molecules released by the plant host that trigger Bacillus PGPR to colonize and form

biofilm on the root surfaces, and to understand how the PGPR senses and responds to

the signals in stimulating biofilm formation.

To expand the in vitro results in Chapter 4 and to study the hypothesized effects

of subtilosin A, the mode of action of antifungal activity will be studied. As model

systems to study the principles of antibiotics biosynthesis, Bacillus PGPR strains are

natural producers of antibiotics that can be used to enhance antifungal activity. Based

on the role of antibiotics and their action in the biocontrol of pathogens, antibiotics

Page 162: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

162

produced by PGPR are excellent candidates for design and development of novel

fungicides.

PGPR partly induced defense response in the squash plants. It suggested that

signal substances initially form within inoculated sites, and translocate to the entire plant

when plants were induced (Chapter 5). Certainly the signal would play an important role

in connecting bacterized sites and distant sites of the plants. Surfactin, PAL and PO

need to be investigated for their role at the molecular level in regulation of defense

response by using real-time PCR.

It is known that diseases have deleterious effects on crop production that

increasing the cost for growers and food safety. This study provides future researchers

with important tools for understanding biocontrol mode of actions that can be used to

develop powerful and effective biocontrol agents and to manipulate rhizosphere

ecosystems for crop production.

Page 163: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

163

LIST OF REFERENCES

Adesemoye, A.O., Torbert, H.A., and Kloepper, J.W. 2009. Plant growth-promoting rhizobacteria allow reduced application rates of chemical fertilizers. Microbial Ecology 58:921-929.

Agrios, G.N. 2005. Plant Pathology. 5th ed. Elsevier Academic Press.

Ahmad, F., Husain, F.M., and Ahmad, I. 2011. Rhizosphere and root colonization by bacterial inoculants and their monitoring methods: a critical area in PGPR research. microbes and microbial technology. Agricultural and Environmental Applications:363-391.

Ahmed, A.S., Ezziyyani, M., Sanchez, C.P., and Candela, M.E. 2003. Effect of chitin on biological control activity of Bacillus spp. and Trichoderma harzianum against root rot disease in pepper (Capsicum annuum) plants. European Journal of Plant Pathology 109:633-637.

Akram, A., Ongena, M., Duby, F., Dommes, J., and Thonart, P. 2008. Systemic resistance and lipoxygenase-related defence response induced in tomato by Pseudomonas putida strain BTPI. Bmc Plant Biology 8:113.

Aly, A.A., Mohamed, H.I., Mansour, M.T.M., and Omar, M.R. 2013. Suppression of powdery mildew on flax by foliar application of essential oils. Journal of Phytopathology 161:376-381.

Amil-Ruiz, F., Blanco-Portales, R., Munoz-Blanco, J., and Caballero, J.L. 2011. The Strawberry Plant Defense Mechanism: A Molecular Review. Plant and Cell Physiology 52:1873-1903.

Ando, K., Hammar, S., and Grumet, R. 2009. Age-related resistance of diverse cucurbit fruit to infection by Phytophthora capsici. Journal of the American Society for Horticultural Science 134:176-182.

Arguelles-Arias, A., Ongena, M., Halimi, B., Lara, Y., Brans, A., Joris, B., and Fickers, P. 2009. Bacillus amyloliquefaciens GA1 as a source of potent antibiotics and other secondary metabolites for biocontrol of plant pathogens. Microbial Cell Factories 8:63.

Arkhipova, T.N., Veselov, S.U., Melentiev, A.I., Martynenko, E.V., and Kudoyarova, G.R. 2005. Ability of bacterium Bacillus subtilis to produce cytokinins and to influence the growth and endogenous hormone content of lettuce plants. Plant and Soil 272:201-209.

Asaka, O., and Shoda, M. 1996. Biocontrol of Rhizoctonia solani damping-off of tomato with Bacillus subtilis RB14. Applied and Environmental Microbiology 62:4081-4085.

Page 164: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

164

Athukorala, S.N.P., Fernando, W.G.D., and Rashid, K.Y. 2009. Identification of antifungal antibiotics of Bacillus species isolated from different microhabitats using polymerase chain reaction and MALDI-TOF mass spectrometry. Canadian Journal of Microbiology 55:1021-1032.

Babadoost, M. 2000. Outbreak of Phytophthora foliar blight and fruit rot in processing pumpkin fields in Illinois. Plant Disease 84:1345-1345.

Babadoost, M., Tian, D., Islam, S.Z., and Pavon, C. 2008. Challenges and options in managing Phytophthora blight (Phytophthora capsici) of cucurbits. Cucurbitaceae 2008: Proceedings of the IXth EUCARPIA Meeting on Genetics and Breeding of Cucurbitaceae:399-406.

Babalola, O.O. 2010. Beneficial bacteria of agricultural importance. Biotechnology Letters 32:1559-1570.

Bais, H., Fall, R., and Vivanco, J. 2004. Biocontrol of Bacillus subtilis against infection of Arabidopsis roots by Pseudomonas syringae is facilitated by biofilm formation and surfactin production. Plant Physiology 134(1):307-319.

Bais, H.P., Weir, T.L., Perry, L.G., Gilroy, S., and Vivanco, J.M. 2006. The role of root exudates in rhizosphere interations with plants and other organisms. Annual Review of Plant Biology 57.

Bakker, P., Ran, L.X., Pieterse, C.M.J., and van Loon, L.C. 2003. Understanding the involvement of rhizobacteria-mediated induction of systemic resistance in biocontrol of plant diseases. Canadian Journal of Plant Pathology-Revue Canadienne De Phytopathologie 25.

Bargabus, R.L., Zidack, N.K., Sherwood, J.E., and Jacobsen, B.J. 2003. Oxidative burst elicited by Bacillus mycoides isolate Bac J, a biological control agent, occurs independently of hypersensitive cell death in sugar beet. Molecular Plant-Microbe Interactions 16:1145-1153.

Barilli, E., Sillero, J.C., and Rubiales, D. 2010. Induction of systemic acquired resistance in pea against rust (Uromyces pisi) by exogenous application of biotic and abiotic inducers. Journal of Phytopathology 158:30-34.

Beneduzi, A., Ambrosini, A., and Passaglia, L.M.P. 2012. Plant growth-promoting rhizobacteria (PGPR): Their potential as antagonists and biocontrol agents. Genetics and Molecular Biology 35:1044-1051.

Beneduzi, A., Peres, D., Vargas, L.K., Bodanese-Zanettini, M.H., and Passaglia, L.M.P. 2008. Evaluation of genetic diversity and plant growth promoting activities of nitrogen-fixing bacilli isolated from rice fields in South Brazil. Applied Soil Ecology 39.

Page 165: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

165

Benhamou, N., Kloepper, J.W., QuadtHallman, A., and Tuzun, S. 1996. Induction of defense-related ultrastructural modifications in pea root tissues inoculated with endophytic bacteria. Plant Physiology 112:919-929.

Bishop-Hurley, S.L., Mounter, S.A., Laskey, J., Morris, R.O., Elder, J., Roop, P., Rouse, C., Schmidt, F.J., and English, J.T. 2002. Phage-displayed peptides as developmental agonists for Phytophthora capsici zoospores. Applied and Environmental Microbiology 68:3315-3320.

Bittel, P., and Robatzek, S. 2007. Microbe-associated molecular patterns (MAMPs) probe plant immunity. Current Opinion in Plant Biology 10.

Boller, T., and He, S.Y. 2009. Innate immunity in plants: an arms race between pattern recognition receptors in plants and effectors in microbial pathogens. Science 324.

Bolstridge, N., Card, S., Stewart, A., and Jones, E.E. 2009. Use of rifampicin-resistant bacterial biocontrol strains for monitoring survival in soil and colonisation of pea seedling roots. New Zealand Plant Protection 62:34-40.

Bradley, C.A. 2008. Effect of fungicide seed treatments on stand establishment, seedling disease, and yield of soybean in North Dakota. Plant Disease 92:120-125.

Branda, S.S., Gonzalez-Pastor, J.E., Ben-Yehuda, S., Losick, R., and Kolter, R. 2001. Fruiting body formation by Bacillus subtilis. Proceedings of the National Academy of Sciences of the United States of America 98:11621-11626.

CafeFilho, A.C., and Duniway, J.M. 1996. Effect of location of drip irrigation emitters and position of Phytophthora capsici infections in roots on phytophthora root rot of pepper. Phytopathology 86:1364-1369.

Chen, C.Q., Belanger, R.R., Benhamou, N., and Paulitz, T.C. 2000. Defense enzymes induced in cucumber roots by treatment with plant growth-promoting rhizobacteria (PGPR) and Pythium aphanidermatum. Physiological and Molecular Plant Pathology 56:13-23.

Chen, X.H., Koumoutsi, A., Scholz, R., Eisenreich, A., Schneider, K., Heinemeyer, I., Morgenstern, B., Voss, B., Hess, W.R., Reva, O., Junge, H., Voigt, B., Jungblut, P.R., Vater, J., Suessmuth, R., Liesegang, H., Strittmatter, A., Gottschalk, G., and Borriss, R. 2007. Comparative analysis of the complete genome sequence of the plant growth-promoting bacterium Bacillus amyloliquefaciens FZB42. Nature Biotechnology 25.

Chen, Y., Yan, F., Chai, Y., Liu, H., Kolter, R., Losick, R., and Guo, J.-h. 2013. Biocontrol of tomato wilt disease by Bacillus subtilis isolates from natural environments depends on conserved genes mediating biofilm formation. Environmental Microbiology 15:848-864.

Page 166: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

166

Chiou, A.L., and Wu, W.S. 2001. Isolation, identification and evaluation of bacterial antagonists against Botrytis elliptica on lily. Journal of Phytopathology-Phytopathologische Zeitschrift 149:319-324.

Chithrashree, Udayashankar, A.C., Nayaka, S.C., Reddy, M.S., and Srinivas, C. 2011. Plant growth-promoting rhizobacteria mediate induced systemic resistance in rice against bacterial leaf blight caused by Xanthomonas oryzae pv. oryzae. Biological Control 59:114-122.

Choudhary, D., and Johri, B. 2009. Interactions of Bacillus spp. and plants - With special reference to induced systemic resistance (ISR). Microbiological Research:493-513.

Choudhary, D.K., Prakash, A., and Johri, B.N. 2007. Induced systemic resistance (ISR) in plants: mechanism of action. Indian Journal of Microbiology 47:289-297.

Compant, S., Clement, 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 & Biochemistry 42:669-678.

Compant, S., Duffy, B., Nowak, J., Clement, C., and Barka, E. 2005. Use of plant growth-promoting bacteria for biocontrol of plant diseases: Principles, mechanisms of action, and future prospects. Applied and Environmental Microbiology:4951-4959.

Compant, S., Brader, G., Muzammil, S., Sessitsch, A., Lebrihi, A., and Mathieu, F. 2013. Use of beneficial bacteria and their secondary metabolites to control grapevine pathogen diseases. BioControl 58:435-455.

Conrath, U., Pieterse, C., and Mauch-Mani, B. 2002. Priming in plant-pathogen interactions. Trends in Plant Science 7:210-216.

Conrath, U., Beckers, G.J.M., Flors, V., Garcia-Agustin, P., Jakab, G., Mauch, F., Newman, M.-A., Pieterse, C.M.J., Poinssot, B., Pozo, M.J., Pugin, A., Schaffrath, U., Ton, J., Wendehenne, D., Zimmerli, L., Mauch-Mani, B., and Prime, A.P.G. 2006. Priming: Getting ready for battle. Molecular Plant-Microbe Interactions 19:1062-1071.

Cook, R.J. 1993. Making greater use of introduced microorganaisms for biological control of plant-pathogens. Annual Review of Phytopathology 31:53-80.

Cooper, J.E., and Rao, J.R. 2006. Molecular approaches to soil, rhizosphere and plant microorganism analysis. Molecular approaches to soil, rhizosphere and plant microorganism analysis.

Costerton, J.W. 1995. Overview of microbial biofilms. Journal of Industrial Microbiology 15:137-140.

Page 167: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

167

Cunniffe, N.J., and Gilligan, C.A. 2011. A theoretical framework for biological control of soil-borne plant pathogens: Identifying effective strategies. Journal of Theoretical Biology 278.

Curtis, H., Noll, U., Stormann, J., and Slusarenko, A.J. 2004. Broad-spectrum activity of the volatile phytoanticipin allicin in extracts of garlic (Allium sativum L.) against plant pathogenic bacteria, fungi and Oomycetes. Physiological and Molecular Plant Pathology 65:79-89.

Danhorn, T., and Fuqua, C. 2007. Biofilm formation by plant-associated bacteria. Annual Review of Microbiology 61.

Demidchik, V. 2012. Reactive Oxygen Species and Oxidative Stress in Plants. Plant Stress Physiology:24-58.

Doornbos, R.F., van Loon, L.C., and Bakker, P.A.H.M. 2012. Impact of root exudates and plant defense signaling on bacterial communities in the rhizosphere. A review. Agronomy for Sustainable Development 32.

El-Hadrami, A., El-Hadrami, I., and Daayf, F. 2009. Suppression of induced plant defence responses by fungal and oomycete pathogens. Molecular plant-microbe interactions:231-268.

El-Hassan, S.A., and Gowen, S.R. 2006. Formulation and delivery of the bacterial antagonist Bacillus subtilis for management of lentil vascular wilt caused by Fusarium oxysporum f. sp lentis. Journal of Phytopathology 154:148-155.

Enebak, S.A., and Carey, W.A. 2000. Evidence for induced systemic protection to fusiform rust in loblolly pine by plant growth-promoting rhizobacteria. Plant Disease 84:306-308.

Ertek, A., Sensoy, S., Kucukyumuk, C., and Gedik, I. 2004. Irrigation frequency and amount affect yield components of summer squash (Cucurbita pepo L.). Agricultural Water Management 67:63-76.

Erwin, D.C., and Ribeiro, O.K. 1996. Phytophthora diseases worldwide. Phytophthora diseases worldwide: 562.

Fan, B., Chen, X.H., Budiharjo, A., Bleiss, W., Vater, J., and Borriss, R. 2011. Efficient colonization of plant roots by the plant growth promoting bacterium Bacillus amyloliquefaciens FZB42, engineered to express green fluorescent protein. Journal of Biotechnology 151:303-311.

Flemming, H.-C., and Wingender, J. 2010. The biofilm matrix. Nature Reviews Microbiology 8:623-633.

Page 168: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

168

French-Monar, R., Jones, J., Ozores-Hampton, M., and Roberts, P. 2007. Survival of inoculum of Phytophthora capsici in soil through time under different soil treatments. Plant Disease 91:593-598.

Gachango, E., Kirk, W., Schafer, R., and Wharton, P. 2012. Evaluation and comparison of biocontrol and conventional fungicides for control of postharvest potato tuber diseases. Biological Control 63:115-120.

Garbeva, P., van Veen, J.A., and van Elsas, J.D. 2004. Microbial diversity in soil: Selection of microbial populations by plant and soil type and implications for disease suppressiveness. Annual Review of Phytopathology 42:243-270.

Garcia-Ruiz, H., and Murphy, J.F. 2001. Age-related resistance in Bell Pepper to Cucumber mosaic virus. Annals of Applied Biology 139:307-317.

Gellert, M., Mizuuchi, K., Odea, M.H., and Nash, H.A. 1976. DNA gyrase- enyzme that introduces superhelical turns ino DNA. Proceedings of the National Academy of Sciences of the United States of America 73:3872-3876.

Gevens, A.J., Roberts, R.D., McGovern, R.J., and Kucharek, T.A. (2008). Vegetable diseases caused by Phytophthora capsici in Florida. In Plant Pathology Department document SP159.

Ghorbani, R., Wilcockson, S., Koocheki, A., and Leifert, C. 2008. Soil management for sustainable crop disease control: a review. Environmental Chemistry Letters 6:149-162.

Gilardi, G., Gullino, M.L., and Garibaldi, A. 2010. Evaluation of spray programmes for the management of leaf spot incited by Pseudomonas syringae pv. syringae on tomato cv. Cuore di bue. Crop Protection 29:330-335.

Gong, M., Wang, J.D., Zhang, J., Yang, H., Lu, X.F., Pei, Y., and Cheng, J.Q. 2006. Study of the antifungal ability of Bacillus subtilis strain PY-1 in vitro and identification of its antifungal substance (iturin A). Acta Biochimica Et Biophysica Sinica 38:233-240.

Goretty Caamal-Chan, M., Souza-Perera, R., and Juan Zuniga-Aguilar, J. 2011. Systemic induction of a Capsicum chinense nitrate reductase by the infection with Phytophthora capsici and defence phytohormones. Plant Physiology and Biochemistry 49:1238-1243.

Granke, L.L., and Hausbeck, M.K. 2010. Effects of temperature, concentration, age, and algaecides on Phytophthora capsici zoospore infectivity. Plant Disease 94:54-60.

Granke, L.L., Quesada-Ocampo, L., Lamour, K., and Hausbeck, M.K. 2012. Advances in Research on Phytophthora capsici on Vegetable Crops in The United States. Plant Disease 96:1588-1600.

Page 169: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

169

Granke, L.L., Windstam, S.T., Hoch, H.C., Smart, C.D., and Hausbeck, M.K. 2009. Dispersal and Movement Mechanisms of Phytophthora capsici Sporangia. Phytopathology 99:1258-1264.

Hammami, I., Jaouadi, B., Ben Bacha, A., Rebai, A., Bejar, S., Nesme, X., and Rhouma, A. 2012. Bacillus subtilis bacteriocin Bac 14B with a broad inhibitory spectrum: Purification, amino acid sequence analysis, and physicochemical characterization. Biotechnology and Bioprocess Engineering 17:41-49.

Hammerschmidt, R. 1999. Induced disease resistance: how do induced plants stop pathogens? Physiological and Molecular Plant Pathology 55:77-84.

Hanley, M.E., Fenner, M., and Edwards, P.J. 1995. The effect of seedling age on the likelihood of herbivory by the slug Deroceras reticulatum. Functional Ecology 9:754-759.

Hartmann, A., Rothballer, M., and Schmid, M. 2008. Lorenz Hiltner, a pioneer in rhizosphere microbial ecology and soil bacteriology research. Plant and Soil 312:7-14.

Hausbeck, M.K., and Lamour, K.H. 2004. Phytophthora capsici on vegetable crops: Research progress and management challenges. Plant Disease 88:1292-1303.

Hayat, R., Ali, S., Amara, U., Khalid, R., and Ahmed, I. 2010. Soil beneficial bacteria and their role in plant growth promotion: a review. Annals of Microbiology 60:579-598.

Heerklotz, H., and Seelig, J. 2007. Leakage and lysis of lipid membranes induced by the lipopeptide surfactin. European Biophysics Journal with Biophysics Letters 36:305-314.

Henry, G., Thonart, P., and Ongena, M. 2012. PAMPs, MAMPs, DAMPs and others: an update on the diversity of plant immunity elicitors. Biotechnologie Agronomie Societe Et Environnement 16.

Henry, G., Deleu, M., Jourdan, E., Thonart, P., and Ongena, M. 2011. The bacterial lipopeptide surfactin targets the lipid fraction of the plant plasma membrane to trigger immune-related defence responses. Cellular Microbiology 13:1824-1837.

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.

Hochmuth, G., and Hanlon, E. 2010.A Summary of N, P, and K Research with Squash in Florida.

Hsieh, F., Li, M., Lin, T., and Kao, S. 2004. Rapid detection and characterization of surfactin-producing Bacillus subtilis and closely related species based on PCR. Current Microbiology 49:186-191.

Page 170: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

170

Ishida, A., Souza, R., Resende, M., Cavalcanti, F., Oliveira, D., and Pozza, E. 2008. Rhizobacterium and acibenzolar-S-methyl (ASM) in resistance induction against bacterial blight and expression of defense responses in cotton. Tropical Plant Pathology 33:27-34.

Ishii, H. 2006. Impact of fungicide resistance in plant pathogens on crop disease control and agricultural environment. Jarq-Japan Agricultural Research Quarterly 40.

Janisiewicz, W.J., Tworkoski, T.J., and Sharer, C. 2000. Characterizing the mechanism of biological control of postharvest diseases on fruits with a simple method to study competition for nutrients. Phytopathology 90:1196-1200.

Jetiyanon, K. 2007. Defensive-related enzyme response in plants treated with a mixture of Bacillus strains (IN937a and IN937b) against different pathogens. Biological Control 42:178-185.

Jetiyanon, K., and Kloepper, J.W. 2002. Mixtures of plant growth-promoting rhizobacteria for induction of systemic resistance against multiple plant diseases. Biological Control 24:285-291.

Ji, P., Wilson, M., Campbell, H.L., and Lindow, S.E. 1997. Determination of the importance of pre-emptive carbon source use by nonpathogenic bacteria in the biocontrol of bacterial speck of tomato. Phytopathology 87:S48-S48.

Jones, J.D.G., and Dangl, J.L. 2006. The plant immune system. Nature 444:323-329.

Joshi, R., and Gardener, B.B.M. 2006. Identification and characterization of novel genetic markers associated with biological control activities in Bacillus subtilis. Phytopathology 96:145-154.

Jourdan, E., Henry, G., Duby, F., Dommes, J., Barthelemy, J.P., Thonart, P., and Ongena, M. 2009. Insights into the defense-related events occurring in plant cells following perception of surfactin-type lipopeptide from Bacillus subtilis. Molecular Plant-Microbe Interactions 22:456-468.

Judelson, H.S., and Blanco, F.A. 2005. The spores of Phytophthora: Weapons of the plant destroyer. Nature Reviews Microbiology 3:47-58.

Katagiri, F., and Tsuda, K. 2010. Understanding the Plant Immune System. Molecular Plant-Microbe Interactions 23:1531-1536.

Keinath, A.P. 2007. Sensitivity of populations of Phytophthora capsici from South Carolina to mefenoxam, dimethomorph, zoxamide, and cymoxanil. Plant Disease 91:743-748.

Keinath, A.P., Everts, K.L., Langston, D.B., Jr., Egel, D.S., and Holmes, G.J. 2007. Multi-state evaluation of reduced-risk fungicides and Melcast against Alternaria leaf blight and gummy stem blight on muskmelon. Crop Protection 26:1251-1258.

Page 171: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

171

Kemble, J.M., Sikora, E.J., Patterson, M.G., Zehnder, G.W., and E., B. 2005.Guide to commercial summer squash production.

Kessler, A. 2011. Plant defense: warding off attack by pathogens, herbivores, and parasitic plants. Quarterly Review of Biology 86:356-357.

Kilani-Feki, O., Khiari, O., Culioli, G., Ortalo-Magne, A., Zouari, N., Blache, Y., and Jaoua, S. 2010. Antifungal activities of an endophytic Pseudomonas fluorescens strain Pf1TZ harbouring genes from pyoluteorin and phenazine clusters. Biotechnology Letters 32:1279-1285.

Kim, S.D., De La Fuente, L., Weller, D.M., and Thomashow, L.S. 2012. Colonizing Ability of Pseudomonas fluorescens 2112, Among Collections of 2,4-Diacetylphloroglucinol-Producing Pseudomonas fluorescens spp. in Pea Rhizosphere. Journal of Microbiology and Biotechnology 22:763-770.

Kinsella, K., Schulthess, C.P., Morris, T.F., and Stuart, J.D. 2009. Rapid quantification of Bacillus subtilis antibiotics in the rhizosphere. Soil Biology & Biochemistry 41:374-379.

Kinsinger, R.F., Shirk, M.C., and Fall, R. 2003. Rapid surface motility in Bacillus subtilis is dependent on extracellular surfactin and potassium ion. Journal of Bacteriology 185:5627-5631.

Kloepper, J., Ryu, C., and Zhang, S. 2004. Induced systemic resistance and promotion of plant growth by Bacillus spp. Phytopathology:1259-1266.

Kloepper, J., Gutierrez-Estrada, A., and McInroy, J. 2007. Photoperiod regulates elicitation of growth promotion but not induced resistance by plant growth-promoting rhizobacteria. Canadian Journal of Microbiology:159-167.

Kloepper, J.W., and Schroth, M.N. 1978. Plant growth-promoting rhizobacteria on radishes. Proc. of the 4th Internat. Conf. on Plant Pathogenic Bacter 2:879-882.

Knee, E.M., Gong, F.C., Gao, M.S., Teplitski, M., Jones, A.R., Foxworthy, A., Mort, A.J., and Bauer, W.D. 2001. Root mucilage from pea and its utilization by rhizosphere bacteria as a sole carbon source. Molecular Plant-Microbe Interactions 14:775-784.

Kone, D., Csinos, A., Jackson, K., and Ji, R. 2009. Evaluation of systemic acquired resistance inducers for control of Phytophthora capsici on squash. Crop Protection:533-538.

Koumoutsi, A., Chen, X.H., Henne, A., Liesegang, H., Hitzeroth, G., Franke, P., Vater, J., and Borriss, R. 2004. Structural and functional characterization of gene clusters directing nonribosomal synthesis of bioactive cyclic lipopeptides in Bacillus amyloliquefaciens strain FZB42. Journal of Bacteriology 186:1084-1096.

Page 172: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

172

Kousik, C.S., Donahoo, R.S., and Hassell, R. 2012. Resistance in watermelon rootstocks to crown rot caused by Phytophthora capsici. Crop Protection 39:18-25.

Kuc, J. 1982. Induced immunity to plant-disesae. Bioscience 32:854-860.

Kumar, A., Prakash, A., and Johri, B.N. 2011a. - Bacteria in Agrobiology: Crop Ecosystems:- 59.

Kumar, A.S., Lakshmanan, V., Caplan, J.L., Powell, D., Czymmek, K.J., Levia, D.F., and Bais, H.P. 2012. Rhizobacteria Bacillus subtilis restricts foliar pathogen entry through stomata. Plant Journal 72:694-706.

Kumar, K.V.K., Yellareddygari, S.K.R., Reddy, M.S., Kloepper, J.W., Sudini, H., Zhou, X.G., Reddy, E.C.S., Yuan, Z., Yang, Z., Wang, Q., Du, B., Zhang, S.A., Reddy, B.R., Reddy, B.V.B., Yuan, Z.L., Yang, Z.L., Wang, Q., and Du, B.H. 2011b. Over-view of commercial potential of plant growth-promoting rhizobacteria (PGPR) in the USA. Plant growth-promoting rhizobacteria (PGPR) for sustainable agriculture. Proceedings of the 2nd Asian PGPR Conference, Beijing, China, 21-24 August, 2011:539-539.

Kumar, V.K., Reddy, M., Kloepper, J.W., Lawrence, K.S., Raju, S.K., Groth, D., and Miller, M. 2010. Field efficacy of Bacillus subtilis MBI 600 (Integral(R)) for managing rice sheath blight caused by Rhizoctonia solani. Phytopathology 100:S65-S65.

Kunst, F., Ogasawara, N., Moszer, I., Albertini, A.M., Alloni, G., Azevedo, V., Bertero, M.G., Bessieres, P., Bolotin, A., Borchert, S., Borriss, R., Boursier, L., Brans, A., Braun, M., Brignell, S.C., Bron, S., Brouillet, S., Bruschi, C.V., Caldwell, B., Capuano, V., Carter, N.M., Choi, S.K., Codani, J.J., Connerton, I.F., Cummings, N.J., Daniel, R.A., Denizot, F., Devine, K.M., Dusterhoft, A., Ehrlich, S.D., Emmerson, P.T., Entian, K.D., Errington, J., Fabret, C., Ferrari, E., Foulger, D., Fritz, C., Fujita, M., Fujita, Y., Fuma, S., Galizzi, A., Galleron, N., Ghim, S.Y., Glaser, P., Goffeau, A., Golightly, E.J., Grandi, G., Guiseppi, G., Guy, B.J., Haga, K., Haiech, J., Harwood, C.R., Henaut, A., Hilbert, H., Holsappel, S., Hosono, S., Hullo, M.F., Itaya, M., Jones, L., Joris, B., Karamata, D., Kasahara, Y., KlaerrBlanchard, M., Klein, C., Kobayashi, Y., Koetter, P., Koningstein, G., Krogh, S., Kumano, M., Kurita, K., Lapidus, A., Lardinois, S., Lauber, J., Lazarevic, V., Lee, S.M., Levine, A., Liu, H., Masuda, S., Mauel, C., Medigue, C., Medina, N., Mellado, R.P., Mizuno, M., Moestl, D., Nakai, S., Noback, M., Noone, D., Oreilly, M., Ogawa, K., Ogiwara, A., Oudega, B., Park, S.H., Parro, V., Pohl, T.M., Portetelle, D., Porwollik, S., Prescott, A.M., Presecan, E., Pujic, P., Purnelle, B., Rapoport, G., Rey, M., Reynolds, S., Rieger, M., Rivolta, C., Rocha, E., Roche, B., Rose, M., Sadaie, Y., Sato, T., Scanlan, E., Schleich, S., Schroeter, R., Scoffone, F., Sekiguchi, J., Sekowska, A., Seror, S.J., Serror, P., Shin, B.S., Soldo, B., Sorokin, A., Tacconi, E., Takagi, T., Takahashi, H., Takemaru, K., Takeuchi, M., Tamakoshi, A., Tanaka, T., Terpstra, P., Tognoni, A., Tosato, V., Uchiyama, S., Vandenbol, M., Vannier, F., Vassarotti, A., Viari, A., Wambutt, R., Wedler, E., Wedler, H., Weitzenegger, T., Winters, P., Wipat, A., Yamamoto, H., Yamane, K.,

Page 173: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

173

Yasumoto, K., Yata, K., Yoshida, K., Yoshikawa, H.F., Zumstein, E., Yoshikawa, H., and Danchin, A. 1997. The complete genome sequence of the Gram-positive bacterium Bacillus subtilis. Nature 390:249-256.

Kus, J.V., Zaton, K., Sarkar, R., and Cameron, R.K. 2002. Age-related resistance in Arabidopsis is a developmentally regulated defense response to Pseudomonas syringae. Plant Cell 14:479-490.

Lahlali, R., Peng, G., Gossen, B.D., McGregor, L., Yu, F.Q., Hynes, R.K., Hwang, S.F., McDonald, M.R., and Boyetchko, S.M. 2013. Evidence that the biofungicide Serenade (Bacillus subtilis) suppresses clubroot on canola via antibiosis and induced host resistance. Phytopathology 103:245-254.

Lamour, K.H., and Hausbeck, M.K. 2000. Mefenoxam insensitivity and the sexual stage of Phytophthora capsici in Michigan cucurbit fields. Phytopathology 90:396-400.

Latha, P., Anand, T., Rappathi, N., Prakasam, V., and Samiyappan, R. 2009. Antimicrobial activity of plant extracts and induction of systemic resistance in tomato plants by mixtures of PGPR strains and Zimmu leaf extract against Alternaria solani. Biological Control 50:85-93.

Laura Tonelli, M., Furlan, A., Taurian, T., Castro, S., and Fabra, A. 2011. Peanut priming induced by biocontrol agents. Physiological and Molecular Plant Pathology 75:100-105.

Leclere, V., Marti, R., Bechet, M., Fickers, P., and Jacques, P. 2006. The lipopeptides mycosubtilin and surfactin enhance spreading of Bacillus subtilis strains by their surface-active properties. Archives of Microbiology 186:475-483.

Leclere, V., Bechet, M., Adam, A., Guez, J.S., Wathelet, B., Ongena, M., Thonart, P., Gancel, F., Chollet-Imbert, M., and Jacques, P. 2005. Mycosubtilin overproduction by Bacillus subtilis BBG100 enhances the organism's antagonistic and biocontrol activities. Applied and Environmental Microbiology 71:4577-4584.

Leelasuphakul, W., Hemmanee, P., and Chuenchitt, S. 2008. Growth inhibitory properties of Bacillus subtilis strains and their metabolites against the green mold pathogen (Penicillium digitatum Sacc.) of citrus fruit. Postharvest Biology and Technology 48:113-121.

Leonian, L.H. 1922. Stem and fruit blight of peppers caused by Phytophthora capsici sp nov. Phytopathology 12:401-408.

Levenfors, J.P., Eberhard, T.H., Levenfors, J.J., Gerhardson, B., and Hokeberg, M. 2008. Biological control of snow mould (Microdochium nivale) in winter cereals by Pseudomonas brassicacearum, MA250. Biocontrol 53:651-665.

Li, Y.C., Klassen, W., Lamberts, M., and Olczyk, T. 2006.Summer squash production in miami-dade county, Florida.

Page 174: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

174

Liang, J.-g., Tao, R.-x., Hao, Z.-n., Wang, L.-p., and Zhang, X. 2011. Induction of resistance in cucumber against seedling damping-off by plant growth-promoting rhizobacteria (PGPR) Bacillus megaterium strain L8. African Journal of Biotechnology 10:6920-6927.

Lim, J., and Kim, S. 2010. Biocontrol of Phytophthora Blight of Red Pepper Caused by Phytophthora capsici Using Bacillus subtilis AH18 and B. licheniformis K11 Formulations. Journal of the Korean Society For Applied Biological Chemistry 53:766-773.

Lin, S., and Jiang, H. 1997. Recovery and purification of the lipopeptide biosurfactant of Bacillus subtilis by ultrafiltration. Biotechnology Techniques 11:413-416.

Lindsay, D., Brozel, V.S., and Von Holy, A. 2005. Spore formation in Bacillus subtilis Biofilms. Journal of Food Protection 68:860-865.

Liu, L., Kloepper, J.W., and Tuzun, S. 1995. Induction of systemic resistance in cucumber against Fusarium wilt by plant growth-promoting rhizobacteria. Phytopathology 85:695-698.

Liu, Y., Wang, H., Sun, X., Yang, H., Wang, Y., and Song, W. 2011. Study on mechanisms of colonization of nitrogen-fixing PGPB, Klebsiella pneumoniae NG14 on the root surface of rice and the formation of biofilm. Current Microbiology 62:1113-1122.

Locke, E., Coronado, G.D., Thompson, B., and Kuniyuki, A. 2009. Seasonal Variation in Fruit and Vegetable Consumption in a Rural Agricultural Community. Journal of the American Dietetic Association 109:45-51.

Lopez, D., Fischbach, M.A., Chu, F., Losick, R., and Kolter, R. 2009. Structurally diverse natural products that cause potassium leakage trigger multicellularity in Bacillus subtilis. Proceedings of the National Academy of Sciences of the United States of America 106:280-285.

Lopez-Bucio, J., Cruz-Ramirez, A., and Herrera-Estrella, L. 2003. The role of nutrient availability in regulating root architecture. Current Opinion in Plant Biology 6:280-287.

Lopez-Raez, J.A., Kohlen, W., Charnikhova, T., Mulder, P., Undas, A.K., Sergeant, M.J., Verstappen, F., Bugg, T.D.H., Thompson, A.J., Ruyter-Spira, C., and Bouwmeester, H. 2010. Does abscisic acid affect strigolactone biosynthesis? New Phytologist 187.

Louws, F.J., Wilson, M., Campbell, H.L., Cuppels, D.A., Jones, J.B., Shoemaker, P.B., Sahin, F., and Miller, S.A. 2001. Field control of bacterial spot and bacterial speck of tomato using a plant activator. Plant Disease 85:481-488.

Lugtenberg, B., and Kamilova, F. 2009. Plant-Growth-Promoting Rhizobacteria. Annual Review of Microbiology:541-556.

Page 175: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

175

Lugtenberg, B., Kravchenko, L., and Simons, M. 1999. Tomato seed and root exudate sugars: composition, utilization by Pseudomonas biocontrol strains and role in rhizosphere colonization. Environmental Microbiology 1:439-446.

Lugtenberg, B.J.J., Dekkers, L., and Bloemberg, G.V. 2001. Molecular determinants of rhizosphere colonization by Pseudomonas. Annual Review of Phytopathology 39.

Madhaiyan, M., Reddy, B.V.S., Anandham, R., Senthilkumar, M., Poonguzhali, S., Sundaram, S.P., and Sa, T.M. 2006. Plant growth-promoting Methylobacterium induces defense responses in groundnut (Arachis hypogaea L.) compared with rot pathogens. Current Microbiology 53:270-276.

Maheshwari, D. 2011. Bacteria in Agrobiology: Crop Ecosystems. Bacteria in Agrobiology: Crop Ecosystems.

Mannanov, R.N., and Sattarova, R.K. 2001. Antibiotics produced by Bacillus bacteria. Chemistry of Natural Compounds 37.

Mariutto, M., Duby, F., Adam, A., Bureau, C., Fauconnier, M.-L., Ongena, M., Thonart, P., and Dommes, J. 2011. The elicitation of a systemic resistance by Pseudomonas putida BTP1 in tomato involves the stimulation of two lipoxygenase isoforms. Bmc Plant Biology 11.

Marx, R., Stein, T., Entian, K.D., and Glaser, S.J. 2001. Structure of the Bacillus subtilis peptide antibiotic subtilosin a determined by H-1-NMR and matrix assisted laser desorption/ionization time-of-flight mass spectrometry. Journal of Protein Chemistry 20:501-506.

Mayer, A.M. 2006. Polyphenol oxidases in plants and fungi: Going places? A review. Phytochemistry 67.

McManus, P.S., Stockwell, V.O., Sundin, G.W., and Jones, A.L. 2002. Antibiotic use in plant agriculture. Annual Review of Phytopathology 40:443.

Meyer, M.D., and Hausbeck, M.K. 2013. Using soil-applied fungicides to manage Phytophthora crown and root rot on summer squash. Plant Disease 97:107-112.

Mitchell, D.J., and Kannwischer-Mitchell, M.E. 1992. PHYTOPHTHORA. Singleton, L. L., J. D. Mihail and C. M. Rush (Ed.). Methods for research on soilborne phytopathogenic fungi. American Phytopathological Society: St. Paul, Minnesota, USA. Illus. Paper:31-38.

Moll, G.N., Konings, W.N., and Driessen, A.J.M. 1999. Bacteriocins: mechanism of membrane insertion and pore formation. Antonie Van Leeuwenhoek International Journal of General and Molecular Microbiology 76:185-198.

Montesinos, E. 2007. Antimicrobial peptides and plant disease control. Fems Microbiology Letters 270:1-11.

Page 176: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

176

Morgan, J.A.W., Bending, G.D., and White, P.J. 2005. Biological costs and benefits to plant-microbe interactions in the rhizosphere. Journal of Experimental Botany 56.

Morikawa, M. 2006. Beneficial biofilm formation by industrial bacteria Bacillus subtilis and related species. Journal of Bioscience and Bioengineering 101:1-8.

Morin, L., Gianotti, A.F., Barker, R., and Johnston, P.R. 1998. Favourable conditions for the bioherbicide candidate Fusarium tumidum to infect and cause severe disease on gorse (Ulex europaeus) in a controlled environment. Biocontrol Science and Technology 8:301-311.

Mossler, M.A., and Nesheim, O.N. 2001.Florida Crop/Pest Management Profile: Squash.

Murphy, J.F., Zehnder, G.W., Schuster, D.J., Sikora, E.J., Polston, J.E., and Kloepper, J.W. 2000. Plant growth-promoting rhizobacterial mediated protection in tomato against Tomato mottle virus. Plant Disease 84:779-784.

Myresiotis, C.K., Karaoglanidis, G.S., Vryzas, Z., and Papadopoulou-Mourkidou, E. 2012. Evaluation of plant-growth-promoting rhizobacteria, acibenzolar-S-methyl and hymexazol for integrated control of Fusarium crown and root rot on tomato. Pest Management Science 68:404-411.

Nakkeeran, S., Kavitha, K., Chandrasekar, G., Renukadevi, P., and Fernando, W.G.D. 2006. Induction of plant defence compounds by Pseudomonas chlororaphis PA23 and Bacillus subtilis BSCBE4 in controlling damping-off of hot pepper caused by Pythium aphanidermatum. Biocontrol Science and Technology 16:403-416.

NASS. 2011. Vegetables 2010 summary, United State Department of Agriculture.

NASS. 2012. Vegetables 2011 summary,United State Department of Agriculture.

NASS. 2013. Vegetables 2012 summary,United State Department of Agriculture.

Nicaise, V., Roux, M., and Zipfel, C. 2009. Recent Advances in PAMP-Triggered Immunity against Bacteria: Pattern Recognition Receptors Watch over and Raise the Alarm. Plant Physiology 150:1638-1647.

Nihorimbere, V., Ongena, M., Smargiassi, M., and Thonart, P. 2011. Beneficial effect of the rhizosphere microbial community for plant growth and health. Biotechnologie Agronomie Societe Et Environnement 15:327-337.

Nihorimbere, V., Cawoy, H., Seyer, A., Brunelle, A., Thonart, P., and Ongena, M. 2012. Impact of rhizosphere factors on cyclic lipopeptide signature from the plant beneficial strain Bacillus amyloliquefaciens S499. Fems Microbiology Ecology 79:176-191.

Nihorimbere, V., Ongena, M., Cawoy, H., Brostaux, Y., Kakana, P., Jourdan, E., and Thonart, P. 2010. Beneficial effects of Bacillus subtilis on field-grown tomato in

Page 177: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

177

Burundi: Reduction of local Fusarium disease and growth promotion. African Journal of Microbiology Research 4:1135-1142.

Niu, D.-D., Liu, H.-X., Jiang, C.-H., Wang, Y.-P., Wang, Q.-Y., Jin, H.-L., and Guo, J.-H. 2011. The Plant Growth-Promoting Rhizobacterium Bacillus cereus AR156 Induces Systemic Resistance in Arabidopsis thaliana by Simultaneously Activating Salicylate- and Jasmonate/Ethylene-Dependent Signaling Pathways. Molecular Plant-Microbe Interactions 24:533-542.

Noling, J.W., and Becker, J.O. 1994. The challenge of research and extension to define and implement alternatives to methyl-bromide. Journal of Nematology 26:573-586.

O'Toole, G., Kaplan, H.B., and Kolter, R. 2000. Biofilm formation as microbial development. Annual Review of Microbiology 54:49-79.

Oelke, L.M., Bosland, P.W., and Steiner, R. 2003. Differentiation of race specific resistance to phytophthora root rot and foliar blight in Capsicum annuum. Journal of the American Society for Horticultural Science 128:213-218.

Ongena, M., and Jacques, P. 2008. Bacillus lipopeptides: versatile weapons for plant disease biocontrol. Trends in Microbiology 16:115-125.

Ongena, M., Duby, F., Jourdan, E., Beaudry, T., Jadin, V., Dommes, J., and Thonart, P. 2005. Bacillus subtilis M4 decreases plant susceptibility towards fungal pathogens by increasing host resistance associated with differential gene expression. Applied Microbiology and Biotechnology 67:692-698.

Ongena, M., Jourdan, E., Adam, A., Paquot, M., Brans, A., Joris, B., Arpigny, J.-L., and Thonart, P. 2007. Surfactin and fengycin lipopeptides of Bacillus subtilis as elicitors of induced systemic resistance in plants. Environmental Microbiology 9.

Ozoreshampton, M., Schaffer, B., Bryan, H.H., and Hanlon, E.A. 1994. Nutrient concentrations, growth, and yield of tomato and squash in municipal solid-waste-ameneded soil. Hortscience 29:785-788.

Padley, L.D., Kabelka, E.A., Roberts, P.D., and French, R. 2008. Evaluation of Cucurbita pepo Accessions for Crown Rot Resistance to Isolates of Phytophthora capsici. Hortscience 43:1996-1999.

Pandey, A., and Palni, L.M.S. 1997. Bacillus species: The dominant bacteria of the rhizosphere of established tea bushes. Microbiological Research 152:359-365.

Paris, H.S. 1996. Summer squash: history, diversity, and distribution. HortTechnology 6:6-13.

Paris, H.S. 2008. Summer squash, Vegetable I, handbook of Plant Breeding.

Page 178: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

178

Passardi, F., Penel, C., and Dunand, C. 2004. Performing the paradoxical: how plant peroxidases modify the cell wall. Trends in Plant Science 9:534-540.

Pavon, C.F., Babadoost, M., and Lambert, K.N. 2008. Quantification of Phytophthora capsici oospores in soil by sieving-centrifugation and real-time polymerase chain reaction. Plant Disease 92:143-149.

Pedersen, K. 1990. Biofilm development on stainless-steel and PVC surfaces in dringking-water. Water Research 24:239-243.

Persello-Cartieaux, F., Nussaume, L., and Robaglia, C. 2003. Tales from the underground: molecular plant-rhizobacteria interactions. Plant Cell and Environment 26.

Pieterse, C.M.J., Leon-Reyes, A., Van der Ent, S., and Van Wees, S.C.M. 2009. Networking by small-molecule hormones in plant immunity. Nature Chemical Biology 5.

Pieterse, C.M.J., Van Wees, S.C.M., Ton, J., Van Pelt, J.A., and Van Loon, L.C. 2002. Signalling in rhizobacteria-induced systemic resistance in Arabidopsis thaliana. Plant Biology 4:535-544.

Pimentel, D., Acquay, H., Biltonen, M., Rice, P., Silva, M., Nelson, J., Lipner, V., Giordano, S., Horowitz, A., and D'Amore, M. 1992. Assessment of environmental and economic impacts of pesticide use. The pesticide question: environment, economics and ethics.:47-84.

Pinton, R., Varanini, Z., and Nannipieri, P. 2001. The rhizosphere. Biochemistry and organic substances at the soil-plant interface. The rhizosphere: biochemistry and organic substances at the soil-plant interface: 424.

Ploetz, R., Heine, G., Haynes, J., and Watson, M. 2002. An investigation of biological attributes that may contribute to the importance of Phytophthora capsici as a vegetable pathogen in Florida. Annals of Applied Biology 140:61-67.

Ploetz, R.C., and Haynes, J.L. 2000. How does Phytophthora capsici survive in squash fields in Southeastern Florida during the off-season? Proceedings of the Florida State Horticultural Society 113:211-215.

Podile, A.R., and Prakash, A.P. 1996. Lysis and biological control of Aspergillus niger by Bacillus subtilis AF 1. Canadian Journal of Microbiology 42:533-538.

Poleatewich, A.M., Ngugi, H.K., and Backman, P.A. 2012. Assessment of Application Timing of Bacillus spp. to Suppress Pre- and Postharvest Diseases of Apple. Plant Disease 96:211-220.

Page 179: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

179

Portmann, R.W., Daniel, J.S., and Ravishankara, A.R. 2012. Stratospheric ozone depletion due to nitrous oxide: influences of other gases. Philosophical Transactions of the Royal Society B-Biological Sciences 367:1256-1264.

Price, N.P.J., Rooney, A.P., Swezey, J.L., Perry, E., and Cohan, F.M. 2007. Mass spectrometric analysis of lipopeptides from Bacillus strains isolated from diverse geographical locations. Fems Microbiology Letters 271:83-89.

Raaijmakers, J., Paulitz, T., Steinberg, C., Alabouvette, C., and Moenne-Loccoz, Y. 2009a. The rhizosphere: a playground and battlefield for soilborne pathogens and beneficial microorganisms. Plant and Soil 321:341-361.

Raaijmakers, J.M., Vlami, M., and de Souza, J.T. 2002. Antibiotic production by bacterial biocontrol agents. Antonie Van Leeuwenhoek International Journal of General and Molecular Microbiology 81:537-547.

Raaijmakers, J.M., de Bruijn, I., and de Kock, M.J.D. 2006. Cyclic lipopeptide production by plant-associated Pseudomonas spp.: Diversity, activity, biosynthesis, and regulation. Molecular Plant-Microbe Interactions 19:699-710.

Raaijmakers, J.M., Paulitz, T.C., Steinberg, C., Alabouvette, C., and Moenne-Loccoz, Y. 2009b. The rhizosphere: a playground and battlefield for soilborne pathogens and beneficial microorganisms. Plant and Soil 321.

Raaijmakers, J.M., Leeman, M., Vanoorschot, M.M.P., Vandersluis, I., Schippers, B., and Bakker, P. 1995a. Dose-response Relationships in Biological-control of Fusarium-wilt of Radish by Pseudomonas spp. Phytopathology 85:1075-1081.

Raftoyannis, Y., and Dick, M.W. 2006. Zoospore encystment and pathogenicity of Phytophthora and Pythium species on plant roots. Microbiological Research 161:1-8.

Ramaekers, L., Remans, R., Rao, I.M., Blair, M.W., and Vanderleyden, J. 2010. Strategies for improving phosphorus acquisition efficiency of crop plants. Field Crops Research 117:169-176.

Ramamoorthy, V., Viswanathan, R., Raguchander, T., Prakasam, V., and Samiyappan, R. 2001. Induction of systemic resistance by plant growth promoting rhizobacteria in crop plants against pests and diseases. Crop Protection 20:1-11.

Ramarathnam, R., Bo, S., Chen, Y., Fernando, W.G.D., Gao, X., and de Kievit, T. 2007. Molecular and biochemical detection of fengycin- and bacillomycin D-producing Bacillus spp., antagonistic to fungal pathogens of canola and wheat. Canadian Journal of Microbiology 53:901-911.

Ramey, B.E., Koutsoudis, M., von Bodman, S.B., and Fuqua, C. 2004. Biofilm formation in plant-microbe associations. Current Opinion in Microbiology 7:602-609.

Page 180: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

180

Raupach, G., and Kloepper, J. 1998a. Mixtures of plant growth-promoting rhizobacteria enhance biological control of multiple cucumber pathogens. Phytopathology 88:1158-1164.

Raupach, G.S., and Kloepper, J.W. 1998b. Mixtures of plant growth-promoting rhizobacteria enhance biological control of multiple cucumber pathogens. Phytopathology 88.

Ristaino, J.B. 1991. Influnce of Rainfull, Drip Irrigation, and Inoculum Density on the Development of Phytophthora Root and Crown Rot Epridemics and Yield in Bell Pepper. Phytopathology 81:922-929.

Riviere, M.-P., Marais, A., Ponchet, M., Willats, W., and Galiana, E. 2008. Silencing of acidic pathogenesis-related PR-1 genes increases extracellular beta-(1 -> 3)-glucanase activity at the onset of tobacco defence reactions. Journal of Experimental Botany 59.

Roberts, P.D., Urs, R.R., French-Monar, R.D., Hoffine, M.S., Seijo, T.E., and McGovern, R.J. 2005. Survival and recovery of Phytophthora capsici and oomycetes in tailwater and soil from vegetable fields in Florida. Annals of Applied Biology 146:351-359.

Roesti, D., Gaur, R., Johri, B.N., Imfeld, G., Sharma, S., Kawaljeet, K., and Aragno, M. 2006. Plant growth stage, fertiliser management and bio-inoculation of arbuscular mycorrhizal fungi and plant growth promoting rhizobacteria affect the rhizobacterial community structure in rain-fed wheat fields. Soil Biology & Biochemistry 38:1111-1120.

Romero, D., de Vicente, A., Rakotoaly, R.H., Dufour, S.E., Veening, J.-W., Arrebola, E., Cazorla, F.M., Kuipers, O.P., Paquot, M., and Perez-Garcia, A. 2007. The iturin and fengycin families of lipopeptides are key factors in antagonism of Bacillus subtilis toward Podosphaera fusca. Molecular Plant-Microbe Interactions 20:430-440.

Ron, E.Z., and Rosenberg, E. 2001. Natural roles of biosurfactants. Environmental Microbiology 3:229-236.

Rouatt, J.W., Katznelson, H., and Payne, T.M.B. 1960. Statistical evaluation of the rhizosphere effect. Soil Sci Soc Amer Proc 24.

Rudrappa, T., Czymmek, K., Pare, P., and Bais, H. 2008. Root-Secreted Malic Acid Recruits Beneficial Soil Bacteria. Plant Physiology:1547-1556.

Ryu, C., Hu, C., Reddy, M., and Kloepper, J. 2003. Different signaling pathways of induced resistance by rhizobacteria in Arabidopsis thaliana against two pathovars of Pseudomonas syringae. New Phytologist:413-420.

Page 181: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

181

Salman, M., and Abuamsha, R. 2012. Potential for integrated biological and chemical control of damping-off disease caused by Pythium ultimum in tomato. Biocontrol 57:711-718.

Sande, D., Mullen, J., Wetzstein, M., and Houston, J. 2011. Environmental Impacts from Pesticide Use: A Case Study of Soil Fumigation in Florida Tomato Production. International Journal of Environmental Research and Public Health 8:4649-4661.

Sanogo, S., and Ji, P. 2012. Integrated management of Phytophthora capsici on solanaceous and cucurbitaceous crops: current status, gaps in knowledge and research needs. Canadian Journal of Plant Pathology-Revue Canadienne De Phytopathologie 34:479-492.

Sarwar, N., Zahid, M.H., Ashfaq, S., and Jamil, F.F. 2011. Induced Systemic Resistance in Chickpea against Ascochyta Blight by Safe Chemicals. Pakistan Journal of Botany 43.

Schisler, D.A., Slininger, R.J., Behle, R.W., and Jackson, M.A. 2004. Formulation of Bacillus spp. for biological control of plant diseases. Phytopathology 94:1267-1271.

Schmidt, C.S., Agostini, F., Leifert, C., Killham, K., and Mullins, C.E. 2004. Influence of inoculum density of the antagonistic bacteria Pseudomonas fluorescens and Pseudomonas corrugata on sugar beet seedling colonisation and suppression of Pythium damping off. Plant and Soil 265.

Schmitt, A., Koch, E., Stephan, D., Kromphardt, C., Jahn, M., Krauthausen, H.J., Forsberg, G., Werner, S., Amein, T., Wright, S.A.I., Tinivella, F., van der Wolf, J., and Groot, S.P.C. 2009. Evaluation of non-chemical seed treatment methods for the control of Phoma valerianellae on lamb's lettuce seeds. Journal of Plant Diseases and Protection 116:200-207.

Seneviratne, G., Weerasekara, M.L.M.A.W., Seneviratne, K.A.C.N., Zavahir, J.S., Kecskes, M.L., and Kennedy, I.R. 2010. Importance of Biofilm Formation in Plant Growth Promoting Rhizobacterial Action. Plant Growth and Health Promoting Bacteria 18:81-95.

Shang, H.Z., Chen, J.J., Handelsman, J., and Goodman, R.M. 1999. Behavior of Pythium torulosum zoospores during their interaction with tobacco roots and Bacillus cereus. Current Microbiology 38:199-204.

Shibata, Y., Kawakita, K., and Takemoto, D. 2010. Age-Related Resistance of Nicotiana benthamiana Against Hemibiotrophic Pathogen Phytophthora infestans Requires Both Ethylene- and Salicylic Acid-Mediated Signaling Pathways. Molecular Plant-Microbe Interactions 23:1130-1142.

Page 182: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

182

Shoresh, M., Harman, G.E., and Mastouri, F. 2010. Induced Systemic Resistance and Plant Responses to Fungal Biocontrol Agents. Annual Review of Phytopathology, Vol 48 48:21-43.

Silva, H., Romeiro, R., Carrer, R., Pereira, J., Mizubuti, E., and Mounteer, A. 2004a. Induction of systemic resistance by Bacillus cereus against tomato foliar diseases under field conditions. Journal of Phytopathology 152:371-375.

Silva, H., Romeiro, R., Macagnan, D., Halfeld-Vieira, B., Pereira, M., and Mounteer, A. 2004b. Rhizobacterial induction of systemic resistance in tomato plants: non-specific protection and increase in enzyme activities. Biological Control 29:288-295.

Simunek, J., and Hopmans, J.W. 2009. Modeling compensated root water and nutrient uptake. Ecological Modelling 220:505-521.

Smith, B.D. 1997. The initial domestication of Cucurbita pepo in the Americas 10,000 years ago. Science 276:932-934.

Somers, E., Vanderleyden, J., and Srinivasan, M. 2004. Rhizosphere bacterial signalling: A love parade beneath our feet. Critical Reviews in Microbiology 30.

Spadaro, D., and Gullino, M.L. 2005. Improving the efficacy of biocontrol agents against soilborne pathogens. Crop Protection 24:601-613.

Spaepen, S., Vanderleyden, J., and Remans, R. 2007. Indole-3-acetic acid in microbial and microorganism-plant signaling. Fems Microbiology Reviews 31:425-448.

Stein, T. 2005. Bacillus subtilis antibiotics: structures, syntheses and specific functions. Molecular Microbiology 56:845-857.

Stein, T., Borchert, S., Kiesau, P., Heinzmann, S., Kloss, S., Klein, C., Helfrich, M., and Entian, K.D. 2002. Dual control of subtilin biosynthesis and immunity in Bacillus subtilis. Molecular Microbiology 44:403-416.

Strigul, N.S., and Kravchenko, L.V. 2006. Mathematical modeling of PGPR inoculation into the rhizosphere. Environmental Modelling & Software 21:1158-1171.

Sturz, A.V., Peters, R.D., Carter, M.R., Sanderson, J.B., Matheson, B.G., and Christie, B.R. 2005. Variation in antibiosis ability, against potato pathogens, of bacterial communities recovered from the endo- and exoroots of potato crops produced under conventional versus minimum tillage systems. Canadian Journal of Microbiology 51:643-654.

Sunwoo, J.Y., Lee, Y.K., and Hwang, B.K. 1996. Induced resistance against Phytophthora capsici in pepper plants in response to DL-beta-amino-n-butyric acid. European Journal of Plant Pathology 102:663-670.

Page 183: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

183

Takemoto, D., Jones, D.A., and Hardham, A.R. 2003. GFP-tagging of cell components reveals the dynamics of subcellular re-organization in response to infection of Arabidopsis by oomycete pathogens. Plant Journal 33:775-792.

Tang, E., Hill, C.B., and Hartman, G.L. 2010. Carbon utilization profiles of Fusarium virguliforme isolates. Canadian Journal of Microbiology 56:979-986.

Thennarasu, S., Lee, D.K., Poon, A., Kawulka, K.E., Vederas, J.C., and Ramamoorthy, A. 2005. Membrane permeabilization, orientation, and antimicrobial mechanism of subtilosin A. Chemistry and Physics of Lipids 137:38-51.

Thomashow, L.S., Weller, D.M., Mavrodi, O.V., and Mavrodi, D.V. 2007. Selecting, monitoring, and enhancing the performance of bacterial biocontrol agents: Principles, pitfalls, and progress. Novel Biotechnologies for Biocontrol Agent Enhancement and Management.

Timmusk, S., Grantcharova, N., and Wagner, E.G.H. 2005. Paenibacillus polymyxa invades plant roots and forms biofilms. Applied and Environmental Microbiology 71:7292-7300.

Tjamos, S.E., Flemetakis, E., Paplomatas, E.J., and Katinakis, P. 2005. Induction of resistance to Verticillium dahliae in Arabidopsis thaliana by the biocontrol agent K-165 and pathogenesis-related proteins gene expression. Molecular Plant-Microbe Interactions 18:555-561.

Toro, M., Azcon, R., and Barea, J.M. 1998. The use of isotopic dilution techniques to evaluate the interactive effects of Rhizobium genotype, mycorrhizal fungi, phosphate-solubilizing rhizobacteria and rock phosphate on nitrogen and phosphorus acquisition by Medicago sativa. New Phytologist 138:265-273.

Tran, H., Ficke, A., Asiimwe, T., Hoefte, M., and Raaijmakers, J.M. 2007. Role of the cyclic lipopeptide massetolide A in biological control of Phytophthora infestans and in colonization of tomato plants by Pseudomonas fluorescens. New Phytologist 175.

Tsuge, K., Inoue, S., Ano, T., Itaya, M., and Shoda, M. 2005. Horizontal transfer of iturin A operon, itu, to Bacillus subtilis 168 and conversion into an iturin A producer. Antimicrobial Agents and Chemotherapy 49:4641-4648.

Van Breusegem, F., Vranova, E., Dat, J.F., and Inze, D. 2001. The role of active oxygen species in plant signal transduction. Plant Science 161:405-414.

Van der Ent, S., Van Wees, S.C.M., and Pieterse, C.M.J. 2009. Jasmonate signaling in plant interactions with resistance-inducing beneficial microbes. Phytochemistry 70:1581-1588.

van Loon, L., Rep, M., and Pieterse, C. 2006. Significance of inducible defense-related proteins in infected plants. Annual Review of Phytopathology:135-162.

Page 184: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

184

van Loon, L.C. 2007. Plant responses to plant growth-promoting rhizobacteria. European Journal of Plant Pathology 119:243-254.

vanVeen, J.A., vanOverbeek, L.S., and vanElsas, J.D. 1997. Fate and activity of microorganisms introduced into soil. Microbiology and Molecular Biology Reviews 61:121-&.

Vater, J., Kablitz, B., Wilde, C., Franke, P., Mehta, N., and Cameotra, S.S. 2002. Matrix-assisted laser desorption ionization-time of flight mass spectrometry of lipopeptide biosurfactants in whole cells and culture filtrates of Bacillus subtilis C-1 isolated from petroleum sludge. Applied and Environmental Microbiology 68:6210-6219.

Verhagen, B., Glazebrook, J., Zhu, T., Chang, H., van Loon, L., and Pieterse, C. 2004. The transcriptome of rhizobacteria-induced systemic resistance in Arabidopsis. Molecular Plant-Microbe Interactions 17:895-908.

Viswanathan, R., and Samiyappan, R. 2002. Role of oxidative enzymes in the plant growth promoting rhizobacteria (PGPR) mediated induced systemic resistance in sugarcane against Colletotrichum falcatum. Zeitschrift Fur Pflanzenkrankheiten Und Pflanzenschutz-Journal of Plant Diseases and Protection:88-100.

Vitullo, D., Di Pietro, A., Romano, A., Lanzotti, V., and Lima, G. 2012. Role of new bacterial surfactins in the antifungal interaction between Bacillus amyloliquefaciens and Fusarium oxysporum. Plant Pathology 61:689-699.

Vlamakis, H., Chai, Y., Beauregard, P., Losick, R., and Kolter, R. 2013. Sticking together: building a biofilm the Bacillus subtilis way. Nature Reviews Microbiology 11:157-168.

Watt, M., Hugenholtz, P., White, R., and Vinall, K. 2006. Numbers and locations of native bacteria on field-grown wheat roots quantified by fluorescence in situ hybridization (FISH). Environmental Microbiology 8.

Wei, G., Kloepper, J., and Tuzun, S. 1996. Induced systemic resistance to cucumber diseases and increased plant growth by plant growth-promoting rhizobacteria under field conditions. Phytopathology:221-224.

Weller, D.M., Raaijmakers, J.M., Gardener, B.B.M., and Thomashow, L.S. 2002. Microbial populations responsible for specific soil suppressiveness to plant pathogens. Annual Review of Phytopathology 40:309-+.

Weller, D.M., Mavrodi, D.V., van Pelt, J.A., Pieterse, C.M.J., van Loon, L.C., and Bakker, P.A.H.M. 2012. Induced Systemic Resistance in Arabidopsis thaliana Against Pseudomonas syringae pv. tomato by 2,4-Diacetylphloroglucinol-Producing Pseudomonas fluorescens. Phytopathology 102:403-412.

Page 185: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

185

Wen, P.F., Chen, J.Y., Kong, W.F., Pan, Q.H., Wan, S.B., and Huang, W.D. 2005. Salicylic acid induced the expression of phenylalanine ammonia-lyase gene in grape berry. Plant Science 169:928-934.

West, A.W., Burges, H.D., Dixon, T.J., and Wyborn, C.H. 1985. Survival of Bacillus thuringiensis and Bacillus cereus Spore Inocula in Soil - Effects of pH, Moisture, Nutrient Availability and Indigenous Microorganisms. Soil Biology & Biochemistry 17:657-665.

Whipps, J.M. 2001. Microbial interactions and biocontrol in the rhizosphere. Journal of Experimental Botany 52:487-511.

Yakimov, M.M., Timmis, K.N., Wray, V., and Fredrickson, H.L. 1995. Characterization of a New Lipopeptide Surfactant Produced by Thermmotolerant and Halotolerant Subsurface Bacillus licheniformis BAS50. Applied and Environmental Microbiology 61:1706-1713.

Yan, Z., Reddy, M., Ryu, C., McInroy, J., Wilson, M., and Kloepper, J. 2002. Induced systemic protection against tomato late blight elicited by plant growth-promoting rhizobacteria. Phytopathology:1329-1333.

Yang, J.W., Yu, S.H., and Ryu, C.-M. 2009. Priming of Defense-Related Genes Confers Root-Colonizing Bacilli-Elicited Induced Systemic Resistance in Pepper. Plant Pathology Journal 25.

Yaryura, P.M., Leon, M., Correa, O.S., Kerber, N.L., Pucheu, N.L., and Garcia, A.F. 2008. Assessment of the role of chemotaxis and biofilm formation as requirements for colonization of roots and seeds of soybean plants by Bacillus amyloliquefaciens BNM339. Current Microbiology 56:625-632.

Yi, H.-S., Yang, J.W., and Ryu, C.-M. 2013. ISR meets SAR outside: additive action of the endophyte Bacillus pumilus INR7 and the chemical inducer, benzothiadiazole, on induced resistance against bacterial spot in field-grown pepper. Frontiers in plant science 4:122-122.

Yoshida, S., Hiradate, S., Tsukamoto, T., Hatakeda, K., and Shirata, A. 2001. Antimicrobial activity of culture filtrate of Bacillus amyloliquefaciens RC-2 isolated from mulberry leaves. Phytopathology 91:181-187.

Yoshida, T., Xia, Z.Y., Takeda, K., Katsuta, T., Sugimoto, K., and Funaoka, M. 2005. Peroxidase-catalyzed polymerization and copolymerization of lignin-based macromonomer (lignocresol) having high content of p-cresol and thermal properties of the resulting polymers. Polymers for Advanced Technologies 16:783-788.

Zahir, Z.A., Arshad, M., and Frankenberger, W.T. 2004. Plant growth promoting rhizobacteria: Applications and perspectives in agriculture. Advances in Agronomy, Vol 81 81.

Page 186: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

186

Zamioudis, C., and Pieterse, C.M.J. 2012. Modulation of Host Immunity by Beneficial Microbes. Molecular Plant-Microbe Interactions 25.

Zeng, W., Kirk, W., and Hao, J. 2012a. Field management of Sclerotinia stem rot of soybean using biological control agents. Biological Control 60:141-147.

Zeng, W., Wang, D., Kirk, W., and Hao, J. 2012b. Use of Coniothyrium minitans and other microorganisms for reducing Sclerotinia sclerotiorum. Biological Control 60:225-232.

Zhang, J.X., Xue, A.G., and Tambong, J.T. 2009. Evaluation of Seed and Soil Treatments with Novel Bacillus subtilis Strains for Control of Soybean Root Rot Caused by Fusarium oxysporum and F. graminearum. Plant Disease 93:1317-1323.

Zhang, S., Reddy, M., and Kloepper, J. 2002. Development of assays for assessing induced systemic resistance by plant growth-promoting rhizobacteria against blue mold of tobacco. Biological Control:79-86.

Zhang, S., White, T., Martinez, M., McInroy, J., Kloepper, J., and Klassen, W. 2010. Evaluation of plant growth-promoting rhizobacteria for control of Phytophthora blight on squash under greenhouse conditions. Biological Control:129-135.

Zhang, S., Vallad, G., Roberts, P., Muchovej, R., Kucharek, T., Pernezny, K., and Momol, T. (2007). A Series on Diseases in the Florida Vegetable Garden: Squash (EDIS).

Zhang, S.A., Vallad, G.E., White, T.L., and Huang, C.H. 2011. Evaluation of Microbial Products for Management of Powdery Mildew on Summer Squash and Cantaloupe in Florida. Plant Disease 95:461-468.

Zheng, G.L., Yan, L.Z., Vederas, J.C., and Zuber, P. 1999. Genes of the sbo-alb locus of Bacillus subtilis are required for production of the antilisterial bacteriocin subtilosin. Journal of Bacteriology 181:7346-7355.

Zipfel, C. 2008. Pattern-recognition receptors in plant innate immunity. Current Opinion in Immunology 20.

Page 187: INTERACTIONS BETWEEN PLANT GROWTH-PROMOTING RHIZOBACTERIA

187

BIOGRAPHICAL SKETCH

Xiaodan Mo was born in Jinzu, Guangxi, China. She obtained her bachelor’s

degree in the Department of Plant Protection from China Agricultural University in 2006,

and she was admitted to the master of science program in the Department of Plant

Pathology at China Agricultural University the same year. From 2006 to 2009, she was

involved in the research project mainly focused on the function of β-glucanase and its

heterologous expression in the Bacillus PGPR. In August 2009, Xiaodan was awarded a

fellowship from China Scholarship Council and joined the Ph. D program in the

Department of Plant Pathology at the University of Florida. In 2011, she began working

with Dr. Shouan Zhang on biological control of Phytophthora blight by Bacillus PGPR

strains on squash. As a graduate student, Xiaodan has presented her work in 2012 and

2013 American Phytopathological Society Annual Meetings. She received 2012 Warren

Wood, Sr. Memorial Fellowship from Miami Dade County, FL. Xiaodan received her Ph.

D. from the University of Florida in December 2013.