11.03.2014 Views

School of Engineering and Science - Jacobs University

School of Engineering and Science - Jacobs University

School of Engineering and Science - Jacobs University

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Siderophore production <strong>of</strong> Pseudomonas syringae <strong>and</strong> its implications on<br />

the biological control <strong>of</strong> bacterial blight <strong>of</strong> soybean<br />

by<br />

Annette Wensing<br />

A thesis submitted in partial fulfilment<br />

<strong>of</strong> the requirements for the degree <strong>of</strong><br />

Doctor <strong>of</strong> Philosophy<br />

Approved, Thesis Committee<br />

Pr<strong>of</strong>. Dr. Matthias Ullrich<br />

International <strong>University</strong> Bremen<br />

Pr<strong>of</strong>. Dr. Ulrich Schwaneberg<br />

International <strong>University</strong> Bremen<br />

Pr<strong>of</strong>. Dr. Georgi Muskhelishvili<br />

International <strong>University</strong> Bremen<br />

Dr. Beate Völksch<br />

Friedrich-Schiller-<strong>University</strong> Jena<br />

June 2006<br />

<strong>School</strong> <strong>of</strong> <strong>Engineering</strong> <strong>and</strong> <strong>Science</strong>


Carpe noctem.


Acknowledgements<br />

This project was supported by grants from Deutsche Forschungsgemeinschaft<br />

<strong>and</strong> by IUB.<br />

Pr<strong>of</strong>. Dr. Matthias Ullrich gave me the possibility to work on this project <strong>and</strong><br />

supported my studies with guidance <strong>and</strong> advice.<br />

Dr. Beate Völksch, Pr<strong>of</strong>. Ulrich Schwaneberg <strong>and</strong> Pr<strong>of</strong>. Georgi Muskhelishvili<br />

agreed to take part in my thesis committee <strong>and</strong> were always available for<br />

discussion <strong>of</strong> the project.<br />

I would like to thank our external cooperation partners:<br />

Dr. Beate Völksch (Jena, Germany) started the investigations on the antagonist<br />

Pss22d <strong>and</strong> agreed on a cooperation project on the investigation <strong>of</strong> biocontrol<br />

mechanisms. Discussions with her <strong>and</strong> her group members Petra Büttner <strong>and</strong><br />

Sascha D. Braun significantely stimulated this project.<br />

Pr<strong>of</strong>. Dr. Dominique Expert (Paris, France) supplied me with the<br />

Dikeya chrysanthemi strain for achromobactin production <strong>and</strong> siderophore<br />

cross-feeding experiments were performed in her lab. Pr<strong>of</strong>. Dr. Jean-Marie<br />

Meyer (Strasbourg, France) allowed me to visit his laboratory <strong>and</strong> taught me<br />

how to purify <strong>and</strong> analyse siderophores. Pr<strong>of</strong>. Dr. Herbert Budzikiewicz<br />

(Cologne, Germany) agreed to perform NMR-analysis <strong>of</strong> purified Pss22d<br />

siderophore.<br />

I would not have been able to accomplish the presented thesis without the<br />

assistance <strong>of</strong> Dr. Helge Weingart. His most valuable recommendations <strong>and</strong> the<br />

many discussions with him formed the basis for the presented results.<br />

All present <strong>and</strong> past members <strong>of</strong> the AG Ullrich, I have to thank for their help as<br />

well as for the great working atmosphere. A special thanks to Yvonne Braun,<br />

for a “hurricane proven” friendship.<br />

I would like to thank Marcus Lindemann for his patience while explaining<br />

fluorescence measurements to a biologist.<br />

I am deeply grateful to my family. Without your support, patience <strong>and</strong><br />

motivation I would not have been able to finish my studies.


Manuscripts in preparation:<br />

Wensing, A., P. Büttner, J. M. Meyer, B. Völksch, M. Ullrich <strong>and</strong> H. Weingart.<br />

Siderophore production <strong>of</strong> the soybean pathogen Pseudomonas syringae pv.<br />

syringae 1a/96 <strong>and</strong> its antagonist Pseudomonas syringae pv. syringae 22d/93.<br />

Wensing, A., H. Weingart, D. Expert, H. Budzikiewicz, B. Völksch <strong>and</strong> M.<br />

Ullrich.<br />

Structure determination <strong>of</strong> an achromobactin-like siderophore produced by<br />

plant-associated P. syringae.


Abbreviations<br />

Ach<br />

achromobactin-like siderophore<br />

AHL<br />

acetyl homoserine lactone<br />

CAS<br />

chrom azurol S<br />

DAPG<br />

2,4-diacetylphloroglucinol<br />

DFOM<br />

deferoxaminmesylate<br />

HDTMA<br />

hexadecyltrimethylammonium bromide<br />

HPLC<br />

high performance liquid chromatography<br />

HR<br />

hypersensitive response<br />

ISR<br />

induced systemic resistance<br />

PGPR<br />

plant growth promoting rhizobacteria<br />

Psg<br />

Pseudomonas syringae pv. glycinea<br />

Pss<br />

Pseudomonas syringae pv. syringae<br />

Pvd<br />

pyoverdin<br />

ROS<br />

reactive oxigene species<br />

SAR<br />

systemic aquired resistance<br />

Sid<br />

siderophore<br />

TLC<br />

thin layer chromatographie<br />

Tris<br />

2-amino-2-(hydroxymethyl)-1,3-prop<strong>and</strong>iole<br />

TTSS<br />

type three secretion system<br />

QS quorum sensing


Contents<br />

1 Summary ....................................................................................................... 1<br />

2 Introduction................................................................................................... 2<br />

2.1 Biological control <strong>of</strong> plant disease ....................................................... 3<br />

2.1.1 Advantages <strong>and</strong> limitations ............................................................ 3<br />

2.1.2 Host, pathogen <strong>and</strong> antagonist – more than a “ménage à trois” 6<br />

2.2 Iron <strong>and</strong> its implications for disease <strong>and</strong> disease control................ 11<br />

2.2.1 The biology <strong>of</strong> iron......................................................................... 11<br />

2.2.2 Iron in biological control <strong>of</strong> bacteria............................................ 12<br />

2.3 Biological control <strong>of</strong> bacterial blight <strong>of</strong> soybean .............................. 13<br />

2.4 Aim <strong>of</strong> this study .................................................................................. 18<br />

3 Material ........................................................................................................ 19<br />

3.1 Equipment............................................................................................. 19<br />

3.2 Chemicals, antibiotics <strong>and</strong> enzymes .................................................. 19<br />

3.3 Kits ........................................................................................................ 20<br />

3.4 Media ..................................................................................................... 20<br />

3.4.1 Complex medium for Escherichia coli......................................... 21<br />

3.4.2 Complex medium for Pseudomonas syringae ............................ 21<br />

3.4.3 Minimal media for Pseudomonas syringae ................................. 21<br />

3.4.4 Siderophore detection medium .................................................... 23<br />

3.5 Microorganisms.................................................................................... 24<br />

3.6 Plasmids................................................................................................ 26<br />

3.7 Oligonucleotides .................................................................................. 27<br />

4 Methods ....................................................................................................... 28<br />

4.1 Bacterial growth conditions ................................................................ 28<br />

4.1.1 Escherichia coli ............................................................................. 28


Contents<br />

4.1.2 Pseudomonas syringae................................................................. 28<br />

4.1.3 Other bacteria ................................................................................ 29<br />

4.1.4 Determination <strong>of</strong> cell density........................................................ 29<br />

4.1.5 Storage <strong>of</strong> bacterial strains........................................................... 29<br />

4.2 Molecular biology methods................................................................. 29<br />

4.2.1 Plasmid DNA isolation................................................................... 29<br />

4.2.1.1 Isolation <strong>of</strong> plasmid DNA by alkaline lysis ............................... 29<br />

4.2.1.2 Plasmid DNA isolation by NucleoSpin plasmid kit .................. 30<br />

4.2.2 Preparation <strong>of</strong> genomic DNA ........................................................ 30<br />

4.2.3 DNA separation by gel electrophoresis ....................................... 31<br />

4.2.4 Polymerase chain reaction (PCR) ................................................ 32<br />

4.2.5 Cloning techniques........................................................................ 33<br />

4.2.5.1 Digestion <strong>of</strong> plasmid DNA with endonucleases....................... 33<br />

4.2.5.2 De-phosphorylation <strong>of</strong> digested DNA ....................................... 33<br />

4.2.5.3 Converting overhanging sticky ends to blunt ends................. 34<br />

4.2.5.4 Precipitation <strong>of</strong> DNA by ethanol ................................................ 34<br />

4.2.5.5 Extraction <strong>of</strong> DNA fragments out <strong>of</strong> agarose gels ................... 34<br />

4.2.5.6 QIAquick PCR <strong>and</strong> reaction purification kit.............................. 35<br />

4.2.5.7 Estimation <strong>of</strong> DNA concentration.............................................. 35<br />

4.2.5.8 Ligation <strong>of</strong> DNA........................................................................... 35<br />

4.2.5.9 Ligation <strong>of</strong> PCR fragments by T/A cloning ............................... 36<br />

4.2.5.10 Preparation <strong>of</strong> electrocompetent E. coli cells ........................ 36<br />

4.2.5.11 Preparation <strong>of</strong> electrocompetent P. syringae cells................ 37<br />

4.2.5.12 Electroporation <strong>of</strong> DNA ............................................................ 37<br />

4.2.5.13 Conjugation by triparental mating........................................... 37<br />

4.2.6 DNA-analysis by Southern blot technique .................................. 38<br />

4.2.6.1 Generation <strong>of</strong> digoxigenin labelled probes .............................. 38


Contents<br />

4.2.6.2 Alkaline transfer <strong>of</strong> DNA............................................................. 39<br />

4.2.6.3 Hybridisation <strong>and</strong> detection....................................................... 39<br />

4.2.7 DNA-sequencing............................................................................ 40<br />

4.3 Analysis <strong>of</strong> bacterial growth in vitro................................................... 40<br />

4.4 Analysis <strong>of</strong> bacterial growth in planta................................................ 40<br />

4.4.1 Wound inoculation......................................................................... 40<br />

4.4.2 Spray inoculation........................................................................... 41<br />

4.5 Analysis <strong>of</strong> siderophore-production................................................... 41<br />

4.5.1 Uptake <strong>of</strong> 59 Fe................................................................................. 41<br />

4.5.2 Determination <strong>of</strong> siderophore concentration by quantitative<br />

CAS-assay ...................................................................................... 42<br />

4.5.3 Quantification <strong>of</strong> pyoverdin by fluorescence measurement...... 43<br />

4.5.4 Purification <strong>of</strong> siderophores by XAD4 ......................................... 43<br />

4.5.5 Isoelectric focussing <strong>of</strong> siderophores ......................................... 44<br />

4.5.6 Gel-filtration ................................................................................... 44<br />

4.5.7 Thin-layer chromatography .......................................................... 45<br />

5 Results......................................................................................................... 46<br />

5.1 Analysis <strong>of</strong> siderophores produced by Pss22d................................. 46<br />

5.1.1 Construction <strong>of</strong> a pyoverdin-deficient mutant <strong>of</strong> Pss22d .......... 47<br />

5.1.2 Southern blot analysis <strong>of</strong> Psg1a <strong>and</strong> Pss22d for the presence <strong>of</strong><br />

yersiniabactin biosynthetic genes ............................................... 51<br />

5.1.3 Screening for an additional siderophore <strong>of</strong> Pss22d by r<strong>and</strong>om<br />

mutagenesis................................................................................... 53<br />

5.1.4 Construction <strong>of</strong> a marker exchange mutant <strong>of</strong> the<br />

achromobactin-like siderophore in Pss22d................................. 57<br />

5.1.5 In vitro comparison <strong>of</strong> Pss22d <strong>and</strong> its siderophore mutants..... 59<br />

5.1.6 Purification <strong>of</strong> the achromobactin-like siderophore for structure<br />

determination ................................................................................. 61<br />

5.2. Comparison <strong>of</strong> siderophore production in Psg1a <strong>and</strong> Pss22d ....... 65


Contents<br />

5.3 Influence <strong>of</strong> siderophore production on biocontrol activity <strong>and</strong><br />

epiphytic fitness <strong>of</strong> Pss22d ................................................................ 69<br />

5.4 Distribution <strong>of</strong> the achromobactin–like siderophore synthesis<br />

among plant associated P. syringae ................................................. 71<br />

6 Discussion................................................................................................... 75<br />

6.1 Identification <strong>of</strong> an achromobactin-like siderophore in P. syringae 75<br />

6.2 Possible regulation <strong>of</strong> the achromobactin-like siderophore ............ 78<br />

6.3 Influence <strong>of</strong> siderophore production on epiphytic fitness <strong>of</strong> Pss22d<br />

.............................................................................................................. 80<br />

6.4 Distribution <strong>of</strong> the achromobactin-like siderophore among<br />

P. syringae pathovars......................................................................... 81<br />

6.5 Outlook.................................................................................................. 82<br />

Literature ........................................................................................................ 83


Summary<br />

1 Summary<br />

The use <strong>of</strong> naturally occurring microbial antagonists to suppress plant diseases<br />

<strong>of</strong>fers a favorable alternative to classical methods <strong>of</strong> plant protection. The<br />

soybean epiphyte, Pseudomonas syringae pv. syringae strain 22d/93 (Pss22d),<br />

exhibits a strong potential to control P. syringae pv. glycinea 1a/96 (Psg1a), the<br />

causal agent <strong>of</strong> bacterial blight <strong>of</strong> soybean. Antagonism <strong>of</strong> Pss22d against<br />

Psg1a has been proven under greenhouse <strong>and</strong> field conditions but the<br />

underlying mechanism(s) remained unknown. In frame <strong>of</strong> this work it was<br />

attempted to elucidate the influence <strong>of</strong> siderophore production on biocontrol. It<br />

was verified that Pss22d <strong>and</strong> Psg1a both produce the typical pyoverdin-type<br />

siderophore <strong>of</strong> P. syringae pathovars. In addition, the presence <strong>of</strong> a second<br />

siderophore in Pss22d was demonstrated. A first characterization <strong>of</strong> this novel<br />

achromobactin-like iron uptake system is presented. The distribution <strong>of</strong> this<br />

second siderophore among different pathovars <strong>of</strong> P. syringae was investigated.<br />

It could be shown that many but not all do produce it. Interestingly, Psg1a also<br />

belongs to the strains tested positive for this siderophore. In consequence,<br />

pathogen Psg <strong>and</strong> antagonist Pss22d produced the same set <strong>of</strong> siderophores.<br />

Thus, a direct role <strong>of</strong> competition for iron in the examined biocontrol system<br />

became unlikely. Comparison <strong>of</strong> the siderophore production in vitro showed<br />

some interesting differences in the regulation <strong>of</strong> siderophore biosynthesis<br />

between Pss22d <strong>and</strong> Psg1a. Finally, the impact <strong>of</strong> the two siderophores <strong>of</strong><br />

Pss22d on in planta performance <strong>of</strong> respective siderophore mutants was<br />

investigated. As expected, the Pss22d siderophore mutants suppressed the<br />

pathogen in a similar manner as the wild type. Different application techniques<br />

were compared for an assessment <strong>of</strong> the impact <strong>of</strong> siderophores on epiphytical<br />

fitness <strong>of</strong> Pss22d. While no differences in survival <strong>and</strong> growth <strong>of</strong> Pss22d wt <strong>and</strong><br />

its respective mutants were observed when wound inoculation technique was<br />

used, the siderophore mutants performed slightly poorer than the wild type<br />

when spray inoculation was applied. The phenotypes were quite variable <strong>and</strong> a<br />

potential interference by alternative iron-uptake systems need to be considered.<br />

This study presents novel information on siderophore production <strong>of</strong> plantassociated<br />

bacteria as well as on the role <strong>of</strong> iron supply for epiphytic fitness<br />

<strong>and</strong> biological control in the phyllosphere. Therefore, it can provide important<br />

information for the development <strong>of</strong> new biocontrol strategies targeting foliar<br />

plant diseases. Our results confirm the complexity <strong>of</strong> biological control systems<br />

<strong>and</strong> the need for more general information on the interaction between epiphytic<br />

microbes <strong>and</strong> plant-pathogen systems.<br />

1


Introduction<br />

2 Introduction<br />

Like humans <strong>and</strong> animals, plants are affected by a broad range <strong>of</strong> pathogenic<br />

microorganisms. Despite expensive pest control measure, infectious plant<br />

diseases significantly interfere with crop production. It is estimated that they<br />

diminish the attainable overall production by approximately 14% while another<br />

6-12% <strong>of</strong> the actual harvest are lost due to post-harvest diseases (Agrios,<br />

2005). These losses differ between crop varieties <strong>and</strong> climatic regions, but in<br />

general their percentage is higher in developing countries as compared to the<br />

industrialized world (Oerke et al., 1994). Worldwide, approximately 35 billion<br />

US Dollars are spent on chemicals for plant protection per year (Agrios, 2005).<br />

The high losses (equaling a value <strong>of</strong> more than 220 billion US Dollar) in spite <strong>of</strong><br />

this expenses have various reasons. Increased worldwide trading <strong>and</strong> traveling<br />

brings about a fast distribution <strong>of</strong> foreign pathogens to new regions. For<br />

example, the fire blight pathogen E. amylovora was supposed to be indigenous<br />

to North America, but it has been spreading all over Europe by now (Eastgate,<br />

2000). Monocultures grown over large areas favor epidemical spread <strong>of</strong><br />

pathogens, while the continued use <strong>of</strong> pesticides <strong>and</strong> fungicides to suppress<br />

such outbreaks promotes the occurrence <strong>of</strong> resistant pathogens (Baker et al.,<br />

1997; McManus et al., 2002). Resistance traits from plant pathogenic bacteria<br />

can be transmitted to human pathogens, so the use <strong>of</strong> antibiotics in agriculture<br />

is largely restricted (McManus et al., 2002; Teuber, 1999). Many other agrichemicals<br />

have been proven to cause environmental damage, <strong>and</strong><br />

consequently have been prohibited. Growing public concern on side effects <strong>of</strong><br />

pesticides led to a revision <strong>of</strong> European commission directive 91/414/CEE<br />

regulating safety requirements <strong>and</strong> accreditation <strong>of</strong> plant protection products.<br />

The alteration <strong>of</strong> this directive in 2002 resulted in the direct withdrawal <strong>of</strong> 320<br />

substances from the European market while at the same time the legal status<br />

<strong>of</strong> another 150 products became questionable (European Commission press<br />

release, 2002). As these 470 products equal 60% <strong>of</strong> all plant protection<br />

products previously available, the need for alternatives becomes most obvious.<br />

2


Introduction<br />

2.1 Biological control <strong>of</strong> plant disease<br />

A promising alternative approach to minimize infectious plant disease is the<br />

application <strong>of</strong> antagonizing organisms with the ability to suppress pathogen<br />

development. This concept is referred to as “biological control” <strong>and</strong> the<br />

operative organism is called biocontrol agent (Wilson, 1997a).<br />

2.1.1 Advantages <strong>and</strong> limitations<br />

The targeted use <strong>of</strong> microbial antagonists <strong>of</strong>fers several benefits. Mass<br />

production <strong>of</strong> the active substance (i.e. fermentation <strong>of</strong> the control organism) is<br />

low-priced compared to chemical synthesis <strong>of</strong> pesticides (Shoda, 2000). Since<br />

biocontrol products can by applied by conventional techniques like spraying or<br />

drenching, their ability to proliferate <strong>and</strong> establish stable populations reduces<br />

application cost to a minimum (Montesinos, 2003). The impact <strong>of</strong> a biocontrol<br />

agent on the on indigenous microbial communities or other organisms in the<br />

ecosystem is less severe as compared to broad-spectrum pesticides (Emmert<br />

& H<strong>and</strong>elsman, 1999). Another significant advantage <strong>of</strong> biological control<br />

results from the flexibility <strong>of</strong> the control agent as living organism. The control<br />

organism can interact with a pathogen in many different ways <strong>and</strong> it is able to<br />

co-evolve with target <strong>and</strong> environment. This minimizes the risk <strong>of</strong> resistance<br />

formation <strong>and</strong> contributes to the potency <strong>of</strong> disease control (Emmert &<br />

H<strong>and</strong>elsman, 1999; H<strong>and</strong>elsman & Stabb, 1996). Biological control systems<br />

can be classified into three major groups: Rhizospherical biocontrol, biological<br />

control in the phyllosphere, <strong>and</strong> post harvest disease control. Post-harvest<br />

disease control <strong>of</strong>fers the unique possibility to keep environmental conditions<br />

constant <strong>and</strong> adjust them in favor <strong>of</strong> the control organism (Janisiewicz &<br />

Korsten, 2002). Disease control in the rhizosphere underlies more variable<br />

conditions, yet root exudates provided by the plant stabilize nutrient supply <strong>and</strong><br />

increase microbial density by orders <strong>of</strong> magnitude. Biocontrol agents have to<br />

cope with high microbial competition, but they can be assisted through direct<br />

support by the host plant (Weller et al., 2002).<br />

3


Introduction<br />

In contrast to the rhizosphere, the phyllosphere represents a rather hostile<br />

environment with low nutrient content, sudden changes in temperature <strong>and</strong><br />

humidity, <strong>and</strong> exposure to UV-radiation. These obstacles resulted in rather<br />

pessimistic predictions for biological control in the phyllosphere, but several<br />

control agents able to sustain these harsh conditions have been identified<br />

(Braun-Kiewnick et al., 2000; Ji & Wilson, 2003; Stromberg et al., 2004;<br />

Völksch & May, 2001b; Wilson, 1997b). The versatility <strong>of</strong> plant-associated<br />

microorganisms provides a large pool <strong>of</strong> potential biocontrol agents. Many<br />

screenings for disease suppressive strains have been conducted, yet the<br />

number <strong>of</strong> registered patents is rather low (Fig.1) <strong>and</strong> even less products have<br />

become commerciallized (Tab. 1; Agrios, 2005; Montesinos, 2003).<br />

Despite the advantages listed above, the practical significance <strong>of</strong> biological<br />

control in agriculture is negligible at present time. What is the reason for this?<br />

The biggest advantages <strong>of</strong> biological control, the flexibility <strong>of</strong> the control agent<br />

<strong>and</strong> its complex interactions with the environment, represent the biggest<br />

drawbacks <strong>of</strong> the system as well. First <strong>of</strong> all, in many cases there is no<br />

correlation between in vitro inhibition <strong>and</strong> in planta efficiency against the<br />

pathogen. This complicates screening for a suitable control agent. Even thriving<br />

100<br />

93<br />

Number <strong>of</strong> patents<br />

80<br />

60<br />

40<br />

20<br />

0<br />

64<br />

11 9<br />

5<br />

Bacteria Fungi Yeast Nematode Virus<br />

Biocontrol organism<br />

Fig. 1. Patents on biocontrol agents.<br />

Number <strong>of</strong> patents on biocontrol agents according to the<br />

type <strong>of</strong> active organism. Source: Montesinos, 2003<br />

4


Introduction<br />

performance <strong>of</strong> a control agent during in planta assays under laboratory<br />

conditions does not guarantee a successful application in greenhouse or in the<br />

field (Montesinos et al., 2002; Shoda, 2000). The performance <strong>of</strong> a biocontrol<br />

agent under changing conditions (climate, vicinity etc.) is <strong>of</strong>ten unpredictable<br />

<strong>and</strong> for field applications <strong>of</strong> biocontrol products varying results are reported<br />

(Thomashow, 1996). The outcome <strong>of</strong> a biological control depends on the<br />

complex interaction <strong>of</strong> host plant, pathogen, control agent, <strong>and</strong> environment<br />

<strong>and</strong> most <strong>of</strong> these relations are poorly understood. An increased knowledge on<br />

these interactions is key towards reliable application <strong>of</strong> biological control (Duffy<br />

et al., 2003; H<strong>and</strong>elsman & Stabb, 1996; Montesinos, 2003).<br />

Tab. 1. Bacterial biological control products (USA).<br />

Commercially available biocontrol products based on bacteria in 2003 (USA only).<br />

Source <strong>of</strong> data: Agrios, 2005<br />

Product name Control agent Crop Disease/Tagert Organism<br />

Galltrol<br />

Agrobacterium radiobacter Fruit/ornamental nurserey stock, Crown gall/ Agrobacterium tumafaciens<br />

strain 84<br />

grape, brambles<br />

Nogall<br />

Companion<br />

Agrobacterium radiobacter<br />

strain K1026<br />

Bacillus subtilis str. GB03 <strong>and</strong><br />

other<br />

Fruit/ornamental nurserey stock,<br />

nut<br />

Crown gall/ Agrobacterium tumafaciens<br />

Many in greenhouse <strong>and</strong> nursery Pytium , Phytophthora , Fusarium ,<br />

Rhizoctonia<br />

HiStick N/T Bacillus subtilis str. MB1600 Legumes Fusarium , Rhizoctonia , Aspergillus<br />

Kodiak Bacillus subtilis GB03 Cotton <strong>and</strong> Legumes Fusarium , Rhizoctonia , Alternaria ,<br />

Aspergillus<br />

Deny Burkholderia cepacia Cotton, Legumes, grain crops Pythium , Rhizoctonia , Fusarium , <strong>and</strong><br />

nematodes<br />

Intercept Burkholderia cepacia Maize, vegetables, cotton Pythium , Rhizoctonia , Fusarium<br />

BioJect Spot-Less Pseudomonas aure<strong>of</strong>aciens turf <strong>and</strong> other Dollar spot, anthracnose, Pythium , pink<br />

snow mold<br />

Bio-save 10LP, 110 Pseudomonas syringae Pome fruit, citrus, cherries,<br />

potatoes<br />

BlightBan A506<br />

Pseudomonas fluorescens<br />

A506<br />

Pome <strong>and</strong> stone fruit, potatoes,<br />

tomatoes, strawberries<br />

Postharvest Botrytis , Mucor , Penicillium ,<br />

Geotrichum<br />

Frost damage, Erwinia amylovora ,<br />

russeting bacteria<br />

Dagger G Pseudomonas fluorescens Field crops, vegetables Rhizoctonia , Pythium<br />

Cedemon Pseudomonas chloraphis Grain cereals Barley, oat leaf spots, Fusarium<br />

5


Introduction<br />

2.1.2 Host, pathogen <strong>and</strong> antagonist – more than a “ménage à trois”<br />

The first step to underst<strong>and</strong> complexity <strong>of</strong> biological control is the determination<br />

<strong>of</strong> crucial features <strong>of</strong> the individual relationships; the virulence determinants <strong>of</strong><br />

the pathogen, defense mechanisms <strong>of</strong> the host plant, <strong>and</strong> mode <strong>of</strong> action <strong>of</strong> the<br />

control agent. If these basics are known, the impact <strong>of</strong> environmental conditions<br />

<strong>and</strong> competing organisms can be assessed.<br />

The best characterized part <strong>of</strong> the biological control complex is the interaction<br />

between plant <strong>and</strong> pathogen. Since the discovery <strong>of</strong> microbes as causal agents<br />

<strong>of</strong> infectious plant disease in 1878, numerous plant diseases have been studied<br />

for virulence determinants <strong>of</strong> the pathogens <strong>and</strong> defense mechanisms <strong>of</strong> host<br />

plants (Burrill, 1878; Montesinos, 2000). The following principles have been<br />

determined as common themes in plant disease: Prior to disease development,<br />

most pathogens undergo an epiphytic growth phase. The pathogen attaches to<br />

the plant surface <strong>and</strong> proliferates in this environment until it finds an entry site,<br />

for example wounds or stomata where the pathogen can overcome structural<br />

defense barriers provided by cuticula, cell wall etc. (Agrios, 2005; Montesinos<br />

et al., 2002). The next step during disease formation would be the induction <strong>of</strong><br />

virulence determinants. This genetic switch to the pathogenic life style is tightly<br />

regulated. Virulence factors <strong>of</strong> a pathogen could be perceived by the host <strong>and</strong><br />

lead to defense induction, thus they are only expressed under disease-favoring<br />

conditions after the pathogen has reached a certain threshold population size.<br />

Quorum sensing is considered an important regulation mechanism for many<br />

virulence factors (Lugtenberg et al., 2002; Montesinos, 2000; Staskawicz et al.,<br />

2001). Most pathogen attacks do not result in disease formation. In case the<br />

host plant recognizes the pathogen, it starts an immediate defense reaction, the<br />

so-called “hypersensitive response” (HR). HR is a localized reaction <strong>of</strong> the<br />

infected tissue. The affected areas undergo programmed cell death consisting<br />

<strong>of</strong> cellular electrolyte leakage <strong>and</strong> “oxidative burst” i.e. the rapid formation <strong>of</strong><br />

reactive oxygen species (ROS). Apart from the direct damage to the pathogen,<br />

6


Introduction<br />

the following necrosis <strong>of</strong> the adjacent tissue cuts the pathogen <strong>of</strong>f from its<br />

nutrients. It thus is unable to spread <strong>and</strong> multiply <strong>and</strong> will finally die.<br />

(Hutcheson, 2001; Klement, 1963). The localized response is followed by a<br />

systemic defense reaction, which is mediated by the plant messenger<br />

molecules salicylic acid (Baker et al., 1997). This reaction is referred to as<br />

systemic acquired resistance (SAR; (Ross, 1962; van Loon et al., 1998a).<br />

A pathogen that is recognized by the plant is called “incompatible” to the<br />

respective host plant. The plant is non-susceptible. In case <strong>of</strong> a so-called<br />

“compatible” pathogen host interaction, the pathogen suppresses or modifies<br />

plant defense in such a way, that it is not attacked (Staskawicz et al., 2001).<br />

It proliferates causing visible disease symptoms as it nourishes on plant<br />

material. New inoculum is spread from the infested tissue <strong>and</strong> the disease cycle<br />

continues (Fig.2).<br />

Fig. 2. Life cycle <strong>of</strong> plant pathogens.<br />

The simplified life cycle <strong>of</strong> plant pathogenic<br />

microorganisms divides into two parts.<br />

During the epiphytic phase (left half <strong>of</strong> the<br />

cycle) the pathogen is distributed to new<br />

locations were it colonizes plant surfaces.<br />

Under favourable environmental conditions<br />

<strong>and</strong> in the presence <strong>of</strong> suitable entry sites,<br />

the pathogen invades the plant tissue. In<br />

an incompatible system the pathogen is<br />

recognized by the host <strong>and</strong> plant defence<br />

is activated. During hypersensitive<br />

response (HR) the pathogen is killed <strong>and</strong><br />

the respective plant tissue becomes<br />

necrotic. The surviving plant enters a state<br />

<strong>of</strong> enhanced disease resistance (SAR).<br />

In a compatible system the pathogen<br />

remains undetected or is able to<br />

manipulate plant defence reactions. After<br />

achieving a threshold density it enters into<br />

the pathogenic phase (right half <strong>of</strong> the<br />

cycle). By expression <strong>of</strong> virulence factors it<br />

attacks surrounding tissue <strong>and</strong> further<br />

proliferates on the nutrients gained from<br />

the plant. The pathogen is distributed from<br />

the infected tissue <strong>and</strong> again enters into<br />

the epiphytic phase.<br />

7


Introduction<br />

The relationship between plants <strong>and</strong> epiphytic microorganism is understood to<br />

a much lesser extent. One well-characterized feature is the correlation between<br />

root exudates <strong>and</strong> the high number <strong>of</strong> microorganisms associated to the<br />

rhizosphere. By this mechanism plants are able to recruit biocontrol agents,<br />

according to the type <strong>of</strong> pathogen that threatens the plant (Weller et al., 2002).<br />

An example for this is the frequently observed spontaneous decline <strong>of</strong> take-all<br />

disease. Take-all disease <strong>of</strong> wheat is caused by the fungus<br />

Gaeumannomyces graminis var. tritici which results (as the name implies) in<br />

disastrous losses. The take-all decline is defined as the spontaneous decrease<br />

in the incidence <strong>and</strong> severity <strong>of</strong> take-all that occurs with monoculture <strong>of</strong> wheat<br />

or other susceptible host crops after one or more severe outbreaks <strong>of</strong> the<br />

disease (Mazzola, 2002). It was observed many times in different locations all<br />

over the world <strong>and</strong> has been correlated to 2,4-diacetylphloroglucinol (DAPG)-<br />

producing bacteria (Keel et al., 1996). The relative abundance <strong>of</strong> DAPG<br />

producers in the wheat rhizosphere increases significantly during take-all<br />

disease, while it does not increase if the pathogen is absent. Furthermore, a<br />

temporal break in monocultures <strong>of</strong> wheat by a non-susceptible crop leads to a<br />

decrease in DAPG producer populations <strong>and</strong> a loss <strong>of</strong> take-all suppression by<br />

the respective soil. This stresses the active part <strong>of</strong> the host plant in supporting<br />

the biocontrol agent (Mazzola, 2002; Weller et al., 2002).<br />

Rhizobacteria can actively support the growth <strong>of</strong> their host plants. The<br />

respective group <strong>of</strong> bacteria is called plant growth-promoting rhizobacteria<br />

(PGPR). They produce plant hormones that influence root morphogenesis <strong>and</strong><br />

result in the overproduction <strong>of</strong> root hairs <strong>and</strong> lateral roots. The subsequent<br />

increase <strong>of</strong> ion uptake accompanied by enhanced solubilisation <strong>of</strong> this ions due<br />

to microbial activity results in enhanced plant nourishment <strong>and</strong> subsequently in<br />

better plant health <strong>and</strong> growth (Persello-Cartieaux et al., 2003; Schippers et al.,<br />

1987). Screening <strong>of</strong> PGNRs for biological control abilities revealed an<br />

interesting new mechanism <strong>of</strong> induced plant resistance. Some PGNRs induce a<br />

plant defense reaction similar to SAR. This so-called induced systemic<br />

8


Introduction<br />

resistance (ISR) can be triggered by several determinants including<br />

lipopolysaccharides, siderophores, or salicylic acid (van Loon et al., 1998a).<br />

The signaling pathways <strong>of</strong> SAR <strong>and</strong> ISR are supposed to differ, as<br />

simultaneous induction <strong>of</strong> both shows additive effects (Pieterse et al., 2001). All<br />

these interactions take place in the rhizosphere. Phyllospheric interactions<br />

between plants <strong>and</strong> epiphytes are investigated to a much lesser extent. One<br />

such relationship that has been focused on is the so called “cross protection”.<br />

Here, the biocontrol agent subsidizes for an incompatible pathogen. As it tries<br />

to invade the plant as described above, it activates plant defense. The initiated<br />

SAR-pathway leads to elimination <strong>of</strong> the target pathogen (Shoda, 2000). In the<br />

direct interaction, biocontrol organisms rarely eradicate a pathogen completely,<br />

but they reduce pathogen population density below the threshold necessary for<br />

disease formation. Control organisms interfere with different steps <strong>of</strong> the<br />

pathogen life cycle <strong>and</strong> different active principles have been described (Fig.3).<br />

The most direct approach is toxin production by the control agent, like DAPG<br />

(Keel et al., 1996). Production <strong>of</strong> antibiotics or inhibitory metabolites that are<br />

active against competitors is a common way <strong>of</strong> microorganisms to enhance<br />

their own ecological fitness. Indeed, the prototypes <strong>of</strong> most commercially<br />

available antibiotics are <strong>of</strong> microbial origin. Many biocontrol agents have been<br />

shown to act by antibiosis (Raaijmakers et al., 2002). For example, the<br />

production <strong>of</strong> phenazine is a major determinant in the control <strong>of</strong> Fusarium wilt <strong>of</strong><br />

chickpea by Pseudomonas aeruginosa (Anjaiah et al., 1998). But what is the<br />

advantage <strong>of</strong> the use <strong>of</strong> toxin-producing biocontrol organisms as compared to<br />

direct toxin application? First <strong>of</strong> all, there is a big difference in the used dosage.<br />

A biocontrol organism produces only minute amounts <strong>of</strong> the active compound<br />

compared to direct application <strong>of</strong> toxins. Second, the active compound is<br />

released only in a small microsite on the plant surface while direct application <strong>of</strong><br />

such substances leads to their distribution within the entire ecosystem. This<br />

mode <strong>of</strong> action is believed to decrease the selection pressure for the<br />

9


Introduction<br />

developement <strong>of</strong> resistances in the pathogen since the pathogen is exposed to<br />

the antibiotic only during a short period <strong>of</strong> its life cycle (Duffy et al., 2003).<br />

Competition for limiting nutrient resources is a fundamental ecological principle.<br />

In biocontrol systems, the pathogen <strong>and</strong> its antagonistic control agent have to<br />

compete for nutrients <strong>and</strong> space. Overlapping ecological niches enable a<br />

control agent to reduce pathogen density by starvation <strong>and</strong> block possible entry<br />

sites by colonizing them (Janisiewicz & Korsten, 2002). Particularly, the<br />

competition for iron has been emphasized on in this respect (H<strong>and</strong>elsman &<br />

Stabb, 1996; Loper & Buyer, 1991; Thomashow, 1996).<br />

Fig. 3. Interference <strong>of</strong> biological control with the pathogens disease cycle.<br />

The scheme summarizes possible means <strong>of</strong> biocontrol agents to interfere with the disease<br />

cycle. Activating measures are indicated by green, arrows while suppressive measures are<br />

represented by red, blocked arrows.<br />

10


Introduction<br />

An impact <strong>of</strong> iron uptake systems on biocontrol was proven for many different<br />

pathosystems including the control <strong>of</strong> Gaeumannomyces graminis var. tritici by<br />

fluorescent strains <strong>of</strong> Pseudomonas or the control <strong>of</strong> Pythium-induced<br />

damping-<strong>of</strong>f <strong>of</strong> tomato by Pseudomonas aeruginosa 7NSK2 (Buysens et al.,<br />

1996; Hamdan et al., 1991).<br />

As a result <strong>of</strong> their effective iron uptake systems many screenings for new<br />

biocontrol organisms are limited to the group <strong>of</strong> fluorescent Pseudomonads<br />

(Haas et al., 2000; Haas & Defago, 2005). Another recently discovered<br />

biocontrol mechanisms is quorum sensing silencing (Dong et al., 2004). Many<br />

virulence determinants are regulated in a cell-density dependent manner. In<br />

Gram-negative bacteria the quorum sensing signals are N-acyl homoserine<br />

lactones (AHLs). The gram-positive bacterium, Bacillus thuringiensis, a<br />

common biological control agent, produces an AHL-lactonase that breaks down<br />

AHL <strong>and</strong> thus silences virulence gene expression <strong>of</strong> gram-negative pathogens<br />

(Dong et al., 2004).<br />

The interaction between pathogen <strong>and</strong> control agent is not unidirectional. Plant<br />

pathogens have evolved their own defense mechanisms against antagonizing<br />

microbes, thus newly identified antagonists should always be tested for their<br />

efficiency against several different strains <strong>of</strong> the pathogen (Duffy et al., 2003).<br />

2.2 Iron <strong>and</strong> its implications for disease <strong>and</strong> disease control<br />

2.2.1 The biology <strong>of</strong> iron<br />

Iron is the fourth most abundant element in the outer crust <strong>of</strong> earth. Despite<br />

this, the low water solubility <strong>of</strong> Fe 3+ at neutral pH limits the availability <strong>of</strong> free<br />

iron in most habitats (Loper & Buyer, 1991). The importance <strong>of</strong> iron as c<strong>of</strong>actor<br />

for enzymes involved in electron-transfer makes it an essential nutrient for<br />

microbial growth. In contrast, high internal iron concentrations are deleterious<br />

for the cell as it catalyses the formation <strong>of</strong> active oxygen species such as<br />

hydrogen peroxide (H 2 O 2 ) <strong>and</strong> the hydroxyl radical (˙OH) by the Fenton<br />

reaction:<br />

11


Introduction<br />

Fe 2+ + H 2 O 2 → Fe 3+ + ˙OH<br />

In consequence, microorganisms did not only evolve high affinity iron-uptake<br />

systems, but also storage proteins for internal iron <strong>and</strong> both features are tightly<br />

regulated by the iron status <strong>of</strong> the cell (Poole & McKay, 2003; Smith, 2004).<br />

2.2.2 Iron in biological control <strong>of</strong> bacteria<br />

In order to increase iron supply, bacteria produce siderophores - molecules <strong>of</strong><br />

low molecular weight that chelate <strong>and</strong> thus solubilize iron. Siderophores are<br />

excreted to the surrounding <strong>and</strong> the lig<strong>and</strong>/iron complex is selectively<br />

transported back into the cells. After intracellular release <strong>of</strong> iron, the<br />

siderophore is recycled (Braun & Braun, 2002; Guerinot, 1994; Neil<strong>and</strong>s, 1995;<br />

Visca et al., 2002). Siderophore high affinity iron uptake systems are important<br />

virulence factors in human <strong>and</strong> plant pathogenic bacteria (Finlay & Falkow,<br />

1997). Different organisms vary in quality <strong>and</strong> quantity <strong>of</strong> the siderophores<br />

produced, <strong>and</strong> an antagonist with a siderophore <strong>of</strong> high affinity might be able to<br />

out-compete a rivaling pathogen (Haas & Defago, 2005). Some siderophores<br />

can form complexes with metal ions other than iron. The siderophore pyochelin,<br />

for example, can form complexes with Zn, Cu or Mo (Visca et al., 1992).<br />

Although the binding constants <strong>of</strong> such complexes are significantly lower as<br />

compared to the iron complex, this could make certain trace elements<br />

unavailable for competitors. Some examples for iron-competition dependent<br />

biocontrol were mentioned under 2.1.2. However, siderophores have been<br />

associated with an additional control mechanism. They seem to interfere with a<br />

number <strong>of</strong> plant defense-mediated biocontrol parameters. First <strong>of</strong> all,<br />

siderophores can directly trigger plant defense response. Pyoverdin <strong>of</strong><br />

Pseudomonas fluorescence CHAO can induce SAR in tobacco <strong>and</strong> radish<br />

reacts towards the pyoverdin <strong>of</strong> P. fluorescence WCS374 by expression <strong>of</strong> an<br />

ISR (van Loon et al., 1998b). Secondly, siderophores such as pyochelin or<br />

pseudomonin are derived from salicylic acid as their precursor (Mercado-<br />

Blanco et al., 2001; Quadri et al., 1999). It has been proposed that this could<br />

12


Introduction<br />

interfere with salicylic acid-depending defense mechanisms. An induction <strong>of</strong><br />

plant defense can not be excluded in general, even though no role for<br />

bacterially produced salicylic acid could be demonstrated in rhizobacterial<br />

induction <strong>of</strong> systemic resistance in Arabidopsis (Ran et al., 2005). Finally, the<br />

type <strong>of</strong> siderophores produced by a bacterium influence the damage caused by<br />

oxidative burst. The ROS formation during oxidative burst favors occurrence <strong>of</strong><br />

the Fenton reaction, yet not only free iron interacts, but also iron-lig<strong>and</strong><br />

complexes. Structural features <strong>of</strong> the lig<strong>and</strong> as well as the lig<strong>and</strong> : iron ratio<br />

varies the reaction characteristics <strong>of</strong> the radical formation (Engelmann et al.,<br />

2003). For example, Dikeya chrysanthemi produces the high-affinity<br />

siderophore chrysobactin. This catechol-type siderophore can form radicals<br />

during oxidative burst (Hirai et al., 2005). In addition D. chrysanthemi produces<br />

a low-affinity siderophore called achromobactin, which does not form radicals<br />

(Expert, 1999). So, siderophores are supposed to be involved in induction as<br />

well as in survival <strong>of</strong> plant defense reactions.<br />

2.3 Biological control <strong>of</strong> bacterial blight <strong>of</strong> soybean<br />

The plant-microbe interaction between soybean (Glycine max) <strong>and</strong> the bacterial<br />

blight pathogen P. syringae pv. glycinea (Psg) has been used extensively as a<br />

model system to investigate plant pathogen relationships. Virulence factors <strong>of</strong><br />

the pathogen have been elucidated <strong>and</strong> detailed studies on their regulation are<br />

available (Bender, 1999; Smirnova & Ullrich, 2004; Weingart et al., 2004).<br />

Furthermore, the defense response <strong>of</strong> soybean toward compatible <strong>and</strong><br />

incompatible strains <strong>of</strong> the pathogen has been studied <strong>and</strong> the impact <strong>of</strong><br />

pathogen signaling was examined (Cui et al., 2005; Zou et al., 2005). All this<br />

luxuriant background information makes this pathosystem an especially<br />

feasible model for investigation <strong>of</strong> biological control. Even the third player is<br />

already known: The antagonist Pseudomonas syringae pv. syringae strain<br />

22d/93 (Pss22d) has been identified as an efficient biocontrol agent against<br />

Psg in vitro, in planta <strong>and</strong> - most importantly - under field conditions (May et al.,<br />

13


Introduction<br />

1997a; May et al., 1997b; Völksch et al., 1996; Völksch & May, 2001a). What is<br />

lacking for our underst<strong>and</strong>ing <strong>of</strong> this biocontrol system is the determination <strong>of</strong><br />

the control principles involved. In the following, a review <strong>of</strong> the available<br />

information is given.<br />

Pseudomonas syringae is a plant associated bacterial species belonging to the<br />

γ -subgroup <strong>of</strong> proteobacteria. It stains Gram-negative. Different strains can by<br />

either saprophytic or plant-pathogenic <strong>and</strong> are divided into more than 50<br />

different pathovars (pv.) according to their host range (Doudor<strong>of</strong>f & Palleroni,<br />

1974). The pathogen Psg has been discovered in 1919 as a leaf spot pathogen<br />

causing necrotic spots surrounded by chlorotic halos (Fig.4) that are developed<br />

Fig. 4. Bacterial blight <strong>of</strong> soybean.<br />

Typical symptoms <strong>of</strong> Pseudomonas syringae pv. glycinea infected soybean leafs: Leaf spots<br />

surrounded by chlorotic halos that develop into lesions later on. The picture is taken from the<br />

first publication describing this disease. The title page shown on the right. Source: Coerper,<br />

1919)<br />

14


Introduction<br />

from initial water-soaked lesions (Coerper, 1919). Psg is a cold-weather<br />

pathogen, i.e. the disease occurs only after periods <strong>of</strong> cold <strong>and</strong> humid weather,<br />

<strong>and</strong> virulence determinates <strong>of</strong> Psg are regulated in a temperature-dependent<br />

manner (Smirnova & Ullrich, 2004; Weingart et al., 2004). The best studied<br />

virulence factor <strong>of</strong> Psg is the phytotoxin coronatine, responsible for chlorosis<br />

formation in infected tissue. Coronatine is a structural analogue <strong>of</strong> the plant<br />

signaling compound, methyl jasmonate. In host plant tissue, it is able to induce<br />

a systemic resistance response that is targeted against insects. By this<br />

meddling with plant signal pathways, Psg misleads defense response <strong>of</strong> the<br />

host <strong>and</strong> suppresses the SAR-pathway that would correctly address the<br />

bacterial pathogen (Bender et al., 1999; Cui et al., 2005).<br />

The antagonist Pss22d belongs to the same species, but to a different<br />

pathovar. It has been isolated as possible biocontrol agent in a screening for<br />

soybean phyllosphere epiphytes (May et al., 1997a). During that study, 82<br />

epiphytes isolated from healthy soybean plants were tested for their ability to<br />

suppress different strains <strong>of</strong> Psg in agar diffusion assay (in vitro) <strong>and</strong> in a<br />

laboratory in planta assay (Tab. 2). Due to its excellent performance Pss22d<br />

was selected for further studies.<br />

Tab. 2. Screening <strong>of</strong> soybean phyllosphere for antagonists against Psg.<br />

Correlation between in vitro <strong>and</strong> in planta inhibition <strong>of</strong> Psg by 82 tested<br />

soybean epiphytes. Source <strong>of</strong> data: May et al., 1997a<br />

Isolates<br />

Number <strong>of</strong> strains with<br />

inhibitory activity<br />

Number <strong>of</strong><br />

tested strains in vitro in planta<br />

in vitro <strong>and</strong><br />

in planta<br />

Pseudomonas spp. 11 6 4 3<br />

Pantoea agglomerans<br />

(formerly Erwinia herbicola )<br />

27 13 9 4<br />

Enterobacter/Erwinia 11 5 2 2<br />

Xanthomonas spp. 8 2 1 0<br />

other Gram-negative bacteria 11 1 2 0<br />

Gram-positive bacteria 14 2 1 0<br />

15


Introduction<br />

Pss22d demonstrated a high epiphytic fitness. After artificial wound inoculation<br />

<strong>of</strong> greenhouse-grown soybean plants it established a stable population that<br />

stayed constant for several weeks whereas other saprophytes could not be reisolated<br />

after more than three days post inoculation (May et al., 1997a; May et<br />

al., 1997b).<br />

In the follow-up field trials Pss22d showed a satisfactory epiphytic fitness. It<br />

formed a population <strong>of</strong> about 10 6 colony forming units (cfu) per gram fresh<br />

weight <strong>of</strong> plant material after spray inoculation. This population size was stable<br />

over the entire vegetation period <strong>and</strong> made up 10-40 % <strong>of</strong> the total epiphytic<br />

population. Simultaneous application <strong>of</strong> Pss22d <strong>and</strong> Psg1a lead to a significant<br />

suppression <strong>of</strong> the pathogen. The population density <strong>of</strong> Psg was reduced by as<br />

much as two orders <strong>of</strong> magnitude. When the antagonist was inoculated four<br />

weeks prior to pathogen application, this inhibitory effect was even enhanced.<br />

At the end <strong>of</strong> the growing season the pathogen population reached only 1/800<br />

(ca. 10 4 cfu/g fresh weight) <strong>of</strong> the population size it reaches in absence <strong>of</strong> the<br />

antagonist, i.e. ca. 10 7 cfu/g fresh weight (Völksch & May, 2001b).<br />

As mentioned above, the molecular basis <strong>of</strong> this interesting control system is<br />

undetermined until now. Nevertheless, there are the following hypotheses.<br />

Pss22d belongs to the group <strong>of</strong> fluorescent pseudomonads. Although the active<br />

principles determined for this group so far have been studied mostly in the<br />

rhizosphere, they comprise a starting point for the assessment <strong>of</strong> similar<br />

mechanisms in the phyllosphere. Consequently, antibiosis <strong>and</strong> siderophores<br />

production come into focus <strong>of</strong> two separate studies. Toxin production by<br />

Pss22d has already been demonstrated in vitro, yet it is necessary to<br />

investigate the importance <strong>of</strong> antibiosis in planta (Völksch et al., 1996). This is<br />

currently investigated by Sascha D. Braun at the <strong>University</strong> <strong>of</strong> Jena. The<br />

general importance <strong>of</strong> siderophore production for biological control by<br />

fluorescent pseudomonads gave rise to a first assessment <strong>of</strong> siderophore<br />

production by Pss22d <strong>and</strong> Psg. CAS-agar is a general indicator medium for the<br />

16


Introduction<br />

production <strong>of</strong> siderophores (Schwyn & Neil<strong>and</strong>s, 1987). Comparison <strong>of</strong> both<br />

Pss22d <strong>and</strong> Psg1a on this medium shows remarkably different phenotype <strong>of</strong><br />

the strains (Fig.5A). The antagonist produces a much larger halo as compared<br />

to the pathogen. The same tendency was observed in several low-iron liquid<br />

media after quantification <strong>of</strong> siderophore activity in the supernatants (Fig.5B;<br />

Büttner, 2003). This difference is especially interesting, as the various<br />

pathovars <strong>of</strong> P. syringae are known to produce the same pyoverdin-type<br />

siderophore <strong>and</strong> so far there has been no experimental prove for the presence<br />

<strong>of</strong> other siderophores aside <strong>of</strong> this pyoverdin (Jülich et al., 2001). Siderophore<br />

production <strong>of</strong> Pss22d <strong>and</strong> its influence on the biocontrol system was addressed<br />

by the current thesis.<br />

I) II)<br />

DFOM-Equiv. [µM]<br />

180<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

Pss22d/93<br />

Psg1a/96<br />

0<br />

CAA KB Pipes 5b SM<br />

Medium<br />

Fig. 5A. Siderophore-production on<br />

CAS-agar. Bacterial syspension <strong>of</strong><br />

I) Psg1a <strong>and</strong> II) Pss22d was applied to<br />

CAS-agar plate <strong>and</strong> incubated at 28°C<br />

for 48 h.<br />

Fig. 5B. Siderophore-production in different<br />

low-iron media after 48h <strong>of</strong> growth.<br />

Siderophore production <strong>of</strong> Pss22d ( ) <strong>and</strong><br />

Psg1a ( ) was determined by CAS-assay <strong>and</strong><br />

normalized to OD 600nm =1. Source <strong>of</strong> data:<br />

Büttner, 2003<br />

17


Introduction<br />

2.4 Aim <strong>of</strong> this study<br />

The antagonism <strong>of</strong> P. syringae pv. syringae 22d/93 (Pss22d) against bacterial<br />

blight <strong>of</strong> soybean caused by P. syringae pv. glycinea (Psg) has been<br />

successfully demonstrated in vitro, in planta, <strong>and</strong> under field conditions (May et<br />

al., 1997a; May et al., 1997b; Völksch et al., 1996; Völksch & May, 2001a). The<br />

outst<strong>and</strong>ing performance <strong>of</strong> Pss22d accompanied by the multiple background<br />

information on the pathosystem sets this biological control apart from the large<br />

number <strong>of</strong> screenings for new control agents. It <strong>of</strong>fers a unique opportunity to<br />

investigate active principle(s) <strong>of</strong> biological control in detail. First results implied<br />

a potential involvement <strong>of</strong> siderophore production (Fig.5). It was the aim <strong>of</strong> this<br />

study to investigate the influence <strong>of</strong> siderophore production on the biological<br />

control system by methods <strong>of</strong> molecular biology. A primary step for this was the<br />

determination <strong>of</strong> the siderophores produced by Pss22d. Mutational approaches<br />

such as marker-exchange mutagenesis <strong>and</strong> transposon mutagenesis were<br />

applied in order to identify the respective gene cluster responsible for<br />

siderophore biosynthesis. Pss22d <strong>and</strong> its siderophore mutants were analysed<br />

for their ability to grow under iron limited conditions in vitro. A detailed<br />

quantification <strong>of</strong> siderophores during in vitro growth was performed for a better<br />

comparison between siderophore production <strong>of</strong> Pss22d <strong>and</strong> Psg1a. The<br />

in planta application <strong>of</strong> Pss22d <strong>and</strong> the generated siderophore mutants was<br />

conducted to directly address the impact <strong>of</strong> siderophore production on<br />

biological control ability <strong>and</strong> epiphytic fitness <strong>of</strong> Pss22d. Finally, the distribution<br />

<strong>of</strong> a newly identified siderophore among different pathovars <strong>of</strong> P. syringae was<br />

analysed.<br />

18


Material <strong>and</strong> Methods<br />

3 Material<br />

3.1 Equipment<br />

Tab. 3. Equipment used in these studies.<br />

Equipment Name Company<br />

Refrigerated Incubator shaker Innova 4230 New Brunswick Scientific<br />

(Nürtingen)<br />

beanch top centrifuges MiniSpin Eppendorf (Engelsdorf)<br />

5415R<br />

Eppendorf (Engelsdorf)<br />

5810R<br />

Eppendorf (Engelsdorf)<br />

Centrifuge with rotors Avanti J20XP Beckman Coulter<br />

with rotor JLA-81000<br />

Speed-vacuum centrifuge Concentrator 5301 Eppendorf (Hamburg)<br />

freeze drying aparatus Micro Modulyo Edwards<br />

rotation evaporator Laborota 4000 Hydolph<br />

2-D electrophoresis apparatus Multiphore II electrophorese unit with Pharmacia Biotech (Freiburg)<br />

Power supply EPS 3500XL<br />

Power supply Power Pack P200 Biometra<br />

DNA electrophoresis chamber MINI SUB DNA CELL BioRad (München)<br />

WIDE MINI SUB CELL<br />

AGS (Heidelberg)<br />

Gel documentation system Gel Jet Imager Intas (Göttingen)<br />

UV cross linker CL-1000 UVP (Cambridge, Engl<strong>and</strong>)<br />

Hybrisization oven Shake'n Stack Hybaid, (Middlesex, Engl<strong>and</strong>)<br />

Phosphoimager FLA-3000 Raytest (Straubenhardt)<br />

Electroporation apparatus GenePulser II BioRad (München)<br />

Thermocycler MyCycler BioRad (München)<br />

Thermomixer Eppendorf Thermomixer 5436 Eppendorf (Hamburg)<br />

Spectrophotometer Ultrospec 2100pro Pharmacia Biotech (Freiburg)<br />

Spectr<strong>of</strong>luorometer F-2500 Digilab (Krefeld)<br />

HPLC system Sykam 2000 Sykam (Fürstenfeldbruck)<br />

HPLC column<br />

250x4 mm, Spherisorb, C18-reversed<br />

phase column Sykam (Fürstenfeldbruck)<br />

3.2 Chemicals, antibiotics <strong>and</strong> enzymes<br />

Chemicals <strong>and</strong> antibiotics were purchased from BioRad (München), Biomol<br />

(Hamburg), Roth (Karlsruhe), Serva (Heidelberg), Sigma (Deisenh<strong>of</strong>en),<br />

Qiagen (Hilden), Applichem (Darmstadt), <strong>and</strong> Merck (Darmstadt). Enzymes<br />

used in this study were purchased from Amersham-Pharmacia Biotech<br />

(Freiburg), Roche (Mannheim), New Engl<strong>and</strong> Biolabs (Schwalbach),<br />

Stratagene (Heidelberg) <strong>and</strong> MBI fermentas (St. Leon-Rot).<br />

19


Material <strong>and</strong> Methods<br />

Tab. 4. Antibiotic substances used in these studies.<br />

Antibiotic<br />

Stock solution<br />

concentration<br />

End<br />

concentration in<br />

medium<br />

Ampicillin (Amp r ) 50 mg/ml 50 mg/L<br />

Chloramphenicol (Cm r ) 25 mg/ml 25 mg/L<br />

Kanamycin (Km r ) 25 mg/ml 25 mg/L<br />

Spectinomycin (Sp r ) 25 mg/ml 25 mg/L<br />

3.3 Kits<br />

Tab. 5. Kits used in these studies.<br />

Kit<br />

DIG DNA Labeling <strong>and</strong> Detection Kit<br />

Taq core Kit<br />

pGEM-Teasy Vector system I<br />

NucleoSpin Plasmid<br />

QIAquick Gel extraction kit<br />

QIAquick PCR Purification kit<br />

Manufacturer<br />

Roche (Mannheim)<br />

Qiagen (Hilden)<br />

Promega (Mannheim)<br />

Macherey-Nagel (Düren)<br />

Qiagen (Hilden)<br />

Qiagen (Hilden)<br />

3.4 Media<br />

All media were sterilized in an autoclave (30 min, 121°C, 1.3 bar). Some<br />

components <strong>of</strong> the media were sterilized by filtration through a sterile filter <strong>of</strong><br />

0.2 µm pore size (Whatman). 1.5 % agar was added to a medium used for<br />

bacterial growth on a plate.<br />

20


Material <strong>and</strong> Methods<br />

3.4.1 Complex medium for Escherichia coli<br />

LB-medium<br />

Lurea-Bertani-medium (Sambrook et al., 1989), pH 7.0<br />

10 g Bacto-trypton<br />

10 g NaCl<br />

5 g Yeast extract<br />

Adjust to 1 L with H 2 O<br />

3.4.2 Complex medium for Pseudomonas syringae<br />

KB-medium<br />

King’B medium (King et al., 1954), pH 7.2<br />

20 g Peptone<br />

1.5 g K 2 HPO 4<br />

1.5 g MgSO 4 x 7 H 2 O<br />

10 ml glycerol<br />

Adjust to 1 L with H 2 O<br />

3.4.3 Minimal media for Pseudomonas syringae<br />

MG-medium<br />

Mannitol-Glutamate medium (Keane et al., 1970), pH 7.0<br />

10 g Mannitol<br />

2 g L-Glutamic acid<br />

0.5 g KH 2 PO4<br />

0.2 g NaCl<br />

0.2 g MgSO 4 x 7 H 2 O<br />

Adjust to 1 L with H 2 O<br />

21


Material <strong>and</strong> Methods<br />

Iron-free 5b medium<br />

Iron free 5b-medium (Bütter, 2003)<br />

buffer (ad 500 ml):<br />

2.6 g KH 2 PO 4<br />

5.5 g Na 2 HPO 4<br />

2.5 g NH 4 Cl<br />

1 g Na 2 SO 4<br />

carbon source (ad 500 ml):<br />

10 g Glucose<br />

100 mg MgCl x 6 H 2 O<br />

10 mg MnSO 4 x 4 H 2 O<br />

Buffer solution <strong>and</strong> carbon source are autoclaved separately.<br />

Iron-free Pipes medium<br />

Iron free Pipes medium (Schwyn & Neil<strong>and</strong>s, 1987), pH 7.0<br />

buffer (ad 500 ml):<br />

30.24 g Pipes<br />

0.3 g KH 2 PO4<br />

1 g NH 4 Cl<br />

1 g Na 2 SO 4<br />

carbon source (ad 500 ml):<br />

10 g Glucose<br />

100 mg MgCl x 6 H 2 O<br />

10 mg MnSO 4 x 4 H 2 O<br />

CAA medium<br />

Casamino-acids medium (Cornelis, Anjaiah et al., 1992)<br />

5 g Difco Bacto casaminoacids<br />

0.9 g K 2 HPO 4 x 3 H 2 O<br />

0.25 g MgSO 4 x 7 H 2 O<br />

Adjust to 1 L with H 2 O<br />

22


Material <strong>and</strong> Methods<br />

Succinat medium without nitrogen source<br />

Succinat medium without nitrogen source (Meyer et al., 2002), pH 7.0<br />

4 g succinic acid<br />

6 g K 2 HPO 4<br />

3 g KH 2 PO 4<br />

0.2 g MgSO 4 x 7H 2 O<br />

This medium was used as incubation medium for the 59 Fe-uptake experiment.<br />

As it does not contain a nitrogen source, the growth <strong>of</strong> bacteria during the<br />

assay is inhibited.<br />

3.4.4 Siderophore detection medium<br />

CAS agar for siderophore detection<br />

The chrome azurol S (CAS) agar medium used in this study was prepared<br />

according to an enhanced protocol (Alex<strong>and</strong>er & Zuberer, 1991).<br />

Solution 1, CAS indicator:<br />

Mix 10 ml <strong>of</strong> 1 mM FeCl 3 x 6H 2 O (in 10 mM HCl) with 50 ml <strong>of</strong> an aqueous<br />

solution <strong>of</strong> CAS (1.21 mg/ml). The resulting dark blue mixture is slovely added<br />

to 40 ml <strong>of</strong> an aqueous solution <strong>of</strong> hexadecyltrimethylammonium bromide<br />

(HDTMA; concentration 1.82 mg/ml)<br />

Solution 2, Pipes buffer:<br />

30.24 g Pipes were dissolved in 750 ml <strong>of</strong> salt solution containing 0.3 g<br />

KH 2 PO 4 , 0.5 g NaCl <strong>and</strong> 1 g NH 4 Cl. The pH was adjusted to 6.8 with<br />

50 % KOH, 15 g Agar <strong>and</strong> the volume was filled to 800 ml with water.<br />

23


Material <strong>and</strong> Methods<br />

Solution3, carbon source:<br />

2 g glucose, 2 g mannitol, 493 g MgSO 4 x 7 H 2 O, 11 mg CaCl 2 , 1.17 mg<br />

MnSO 4 x 7 H 2 O, 1.4 mg H 3 BO 3 , 0.04 CuSO 4 x 5 H 2 O, 1.2 mg ZnSO 4 x 7 H 2 O<br />

<strong>and</strong> 1 mg Na 2 MoO 4 x 2 H 2 O were dissolved in 70 ml <strong>of</strong> water.<br />

Solution 4, casamino acids :<br />

30 ml <strong>of</strong> 10 % (w/v) <strong>of</strong> casamino acids<br />

Solutions 1-3 were autoclaved separately <strong>and</strong> solution 4 was filter sterilized. All<br />

components were mixed after cooling down to 50°C. Subsequently the agar<br />

plates were poored. The plates were kept in the dark prior to usage, as the<br />

CAS-indicator is light sensitive.<br />

The trace metal composition <strong>of</strong> solution 3 was used for complementation test <strong>of</strong><br />

Pipes-medium<br />

3.5 Microorganisms<br />

Tab. 6. Bacterial strains used in these studies.<br />

Bacterial strain Relevant characteristics Reference or source<br />

Escherichia coli :<br />

DH5α<br />

S17-1 λ-pir<br />

supE 44 ∆lacU169 (φ80 lacZ<br />

∆M15) hsdR 17 recA 1 endA 1<br />

gyrA 96 thi -1 relA 1 Hanahan, 1983<br />

λ-pir lysogen <strong>of</strong> E.coli S17-1 (thi<br />

pro hsdR - hsdM + recA RP4 2 -<br />

Tc::MU-Km::Tn7 (Tet R /Sm R ), host<br />

strain for plasmid pCAM-Not Wilson, et al . 1995<br />

24


Material <strong>and</strong> Methods<br />

Tab. 6 continued. Bacterial strains used in these studies.<br />

Bacterial strain Relevant characteristics Reference or source<br />

Pantoea agglomerans 48b epiphyt, DFFE producer Völksch et al. , 1996<br />

Dickeya chrysanthemi cbsE-1 tonB60<br />

Xanthomonas campestris DSM 1050<br />

cbsE-1 , tonB60 , Cbs – , TonB – ,<br />

Spec R , Km R Enard <strong>and</strong> Expert 2000<br />

Ach - , but contains homologue<br />

siderophore synthethase<br />

German Culture Collection (DSM)<br />

Sinorhizobium meliloti DSM 6047 Produces rhizobactin German Culture Collection (DSM)<br />

Pseudomonas syringae pv.<br />

syringae 22d/93 (Pss22d) antagonist wild type; Pvd + , Ach + Völksch et al. , 1996<br />

Pss22d∆Pvd<br />

Pss22d∆Ach<br />

Pss22d∆Sid<br />

derivative <strong>of</strong> Pss22d; Pvd - , Ach + ,<br />

Km r<br />

derivative <strong>of</strong> Pss22d; Pvd + , Ach - ,<br />

Km r<br />

derivative <strong>of</strong> Pss22d∆Pvd; Pvd - ,<br />

Ach - , Km r , Sp r<br />

this study<br />

this study<br />

this study<br />

glycinea 1a/96 pathogen wild type; Pvd+, Ach+ Völksch <strong>and</strong> May, 2002<br />

glycinea PG4180 Bender et al. , 1993<br />

aceris CFBP 2339<br />

actinidiae MAFF 302091<br />

aptata GSPB 1080<br />

atr<strong>of</strong>aciens GSPB 1440<br />

atropurporea MAFF 301309<br />

berberis CFBP 1727<br />

broussonetiae MAFF 810036<br />

J-M Meyer, Strasbourg, France<br />

David Guttman, Toronto, Canada<br />

Helge Weingart / K. Rudolph<br />

German Culture Collection (DSM)<br />

B. Völksch , Jena<br />

J-M Meyer, Strasbourg, France<br />

David Guttman, Toronto, Canada<br />

castaneae CFBP 4217<br />

J-M Meyer, Strasbourg, France<br />

coronafaciens DSM 50261<br />

strains<br />

tested for the<br />

Helge Weingart, Bremen<br />

daphniphylli CFBP 4219<br />

achromobactin-like J-M Meyer, Strasbourg, France<br />

siderophore<br />

delfinii ICMP 529 Kawalleck et al. (1995)<br />

mori MAFF 301020<br />

David Guttman, Toronto, Canada<br />

mosprunorum GSPB 1013 K. Rudolph, Göttingen<br />

maculicola DC3000<br />

papulans ICMP 4048<br />

phaseolicola 1448A<br />

Elina Roine, Helsinki, Finnl<strong>and</strong><br />

John Lydon, Beltsville, USA<br />

John W. Mansfield<br />

pisi GSPB 1477 K. Rudolph, Göttingen<br />

savastanoi GSBP 2259 K. Rudolph, Göttingen<br />

tabaci GSPB 117 K. Rudolph, Göttingen<br />

tomato 487<br />

B. Völksch , Jena<br />

25


Material <strong>and</strong> Methods<br />

3.6 Plasmids<br />

Tab. 7. Plasmids used in these studies.<br />

Plasmid Relevant characteristics Reference/Source<br />

pGEM-TEasy cloning vector for T/A cloning, Amp r Promega<br />

pPVSA1<br />

pPVSA2<br />

1,4 kb PCR-product <strong>of</strong> pvsA (fragment1) in pGEM-T Easy,<br />

primer combination pvsA_fwd1_Spe <strong>and</strong> pvsA_rev6_Kpn ,<br />

template Pss22d, Amp r<br />

1,6 kb PCR-product <strong>of</strong> pvsA (fragment2) pGEM-T Easy, primer<br />

combination pvsA_fwd7_Kpn und pvsA_rev4_Bam , template<br />

Pss22d, Amp r<br />

This study<br />

This study<br />

pPVSA3 Spe I-Kpn I fragment <strong>of</strong> pPVSA1 in pPVSA2 This study<br />

pPVSA4<br />

pPVSA3 containing Km r -cassette (Kpn I-Fragment), suizidvector<br />

for pvsA-mutagenesis in Pss22d<br />

This study<br />

pACS1<br />

1.3 kb PCR fragment amplified by primer combination<br />

Achr1_fwd <strong>and</strong> Achr2_Kpn I_rev. in pGEM-T Easy, Amp r<br />

This study<br />

pACS2<br />

2 kb PCR fragment amplified by primer combination<br />

Achr5_EcoRV_fwd <strong>and</strong> Achr6_rev in pGEM-T Easy, Amp r<br />

This study<br />

pACS3 pACS1 with Km r resistance cassette, Amp r , Km r This study<br />

pACS4<br />

pACS3 insert in pACS2, suizide vector for mutagenesis <strong>of</strong> acsgenes<br />

in Pss22d, Amp r , Km r<br />

This study<br />

pBBR-MCS1 cloning vector, Cm r Kovach et al., 1994<br />

pBBR-11-Sid 6 kb Ps tI-fragment <strong>of</strong> Pss22d∆Sid in pBBR-MCS1, Cm r , Sp r This study<br />

pBBR-23-Sid 6 kb Ps tI-fragment <strong>of</strong> Pss22d∆Sid-23 in pBBR-MCS1, Cm r , Sp r This study<br />

pMKm vector containing kanamycin cassette, Km r Murillo et al., 1994<br />

pCAM-Not<br />

Sp r Amp r , carries the mTn5 SS40 transposon<br />

Wilson et al., 1995, H.<br />

Weingart<br />

26


Material <strong>and</strong> Methods<br />

3.7 Oligonucleotides<br />

The following oligonucleotides were used in this study:<br />

Pyoverdin biosynthesis-mutation:<br />

PvsA_rev6_Kpn:<br />

TAAGGTACCACGTCGAGGCTGAGCGGATC<br />

PvsA_fwd7_Kpn:<br />

TTAGGTACCTCGAACTTGGCCTCGCGGCTG<br />

pvsA_rev4_BamHI:<br />

TTTGGATCCGGCAGACCGTGGCTGAG<br />

pvsA_fwd1_Spe<br />

AATACTAGTGGATCCTGATGCGACTGGCCTTCGATC<br />

Achromobactin biosynthesis mutation:<br />

Achr1_fwd :<br />

AGCGAGGACTCACAGATGTTG<br />

Achr2_KpnI_rev :<br />

GGTACCCAATGCTGCTGAATGGCAAC<br />

Achr5_EcoRV_fwd:<br />

GATATCAACTATGTGCGTCTTGCGTC<br />

Achr6_rev:<br />

ACGAATGCCACCAGACAGG<br />

Yersiniabactin-Southern probe<br />

South-irp1-fwd:<br />

AATACCGCTGATCTCGAC<br />

South-irp1-rev:<br />

ATTGTTGCTGCCAGTGTG<br />

27


Material <strong>and</strong> Methods<br />

PCR-Screening for achromobactin-like siderophore production:<br />

iuc1-fwd:<br />

TGCTGGACTGGTTCGATG<br />

iuc1-rev:<br />

GATCAGCAGGCAATAGG<br />

iuc2-fwd:<br />

AAATCGAGCAGACCCAGC<br />

iuc2-rev:<br />

TGTTGGTCATGCTCATCG<br />

4 Methods<br />

4.1 Bacterial growth conditions<br />

4.1.1 Escherichia coli<br />

E. coli cells were grown overnight at 37°C on LB agar plates containing the<br />

appropriate antibiotics. Subsequently, E. coli cells were cultured overnight at<br />

37°C by shaking at 250 rpm in LB medium containing the appropriate<br />

antibiotics. 2 ml plastic reaction tubes <strong>and</strong> test tubes were used for growth <strong>of</strong><br />

small volumes <strong>of</strong> E. coli cultures (1-5 ml), <strong>and</strong> Erlenmeyer flasks were used for<br />

growth <strong>of</strong> big volumes <strong>of</strong> E. coli cultures (50-500 ml).<br />

4.1.2 Pseudomonas syringae<br />

P. syringae cells were grown for 2-5 days at 28°C on MG agar plates<br />

containing appropriate antibiotics. Subsequently, P. syringae cells were<br />

inoculated in liquid media (KB, 5b or Pipes) <strong>and</strong> grown at 28°C by shaking at<br />

280 rpm. Test tubes were used for growth <strong>of</strong> small volumes <strong>of</strong> P. syringae<br />

cultures (5 ml) <strong>and</strong> Erlenmeyer flasks were used for growth <strong>of</strong> big volumes <strong>of</strong><br />

P. syringae cultures (20-250 ml).<br />

28


Material <strong>and</strong> Methods<br />

4.1.3 Other bacteria<br />

Cells <strong>of</strong> D. chrysanthemi, X. campestris <strong>and</strong> S. meliloti were grown for 1-3 days<br />

at 28°C on LB or KB agar plates. Liquid cultures (20-250 ml) <strong>of</strong><br />

D. chrysanthemi were grown in Erlenmeyer flask at 28°C by shaking at 280rpm.<br />

4.1.4 Determination <strong>of</strong> cell density<br />

Cell density <strong>of</strong> bacterial suspensions or liquid culture was determined by<br />

photometric measurement. The optical density (OD) <strong>of</strong> the sample was<br />

determined at λ = 600nm. Samples with an OD above 0.6 were diluted in the<br />

appropriate medium. It is assumed that an OD 600nm <strong>of</strong> one correlates to a cell<br />

density <strong>of</strong> 10 9 (Miller, 1992).<br />

4.1.5 Storage <strong>of</strong> bacterial strains<br />

Fresh grow bacterial colonies or lawns from agar plates were resuspended in<br />

15% sterile glycerol solution <strong>and</strong> stored at -80°C.<br />

4.2 Molecular biology methods<br />

4.2.1 Plasmid DNA isolation<br />

4.2.1.1 Isolation <strong>of</strong> plasmid DNA by alkaline lysis<br />

Small amounts <strong>of</strong> plasmid DNA were isolated by the 1-2-3 protocol based on<br />

alkaline lysis (Sambrook et al., 1989).1-1.5 ml <strong>of</strong> E. coli overnight culture was<br />

harvested by centrifugation at 13,000 rpm for 2-5 min. The pellet was<br />

resuspended in 150 µl <strong>of</strong> pre-cooled buffer P1 (100 µg/ml RNase A, 50 mM<br />

Tris/HCl, 10 mM EDTA-Na at a pH <strong>of</strong> 8.0). Alternative, cells from an overnight<br />

LB-plate were resuspended in 150 µl <strong>of</strong> pre-cooled buffer P1. Subsequently,<br />

the cells were lysed upon addition <strong>of</strong> 150 µl <strong>of</strong> buffer P2 (200 mM NaOH,<br />

1% SDS) for 5 min, <strong>and</strong> the lysate was neutralized by addition <strong>of</strong> 150 µl <strong>of</strong> precooled<br />

3 M potassium acetate solution (pH 5.5). This fast neutralisation led to<br />

the renaturation <strong>of</strong> small plasmid DNA. Larger molecules as proteins <strong>and</strong><br />

chromosomal DNA, <strong>and</strong> cellular debris stay denaturated. Following incubation<br />

29


Material <strong>and</strong> Methods<br />

on ice for 10 min <strong>and</strong> subsequent centrifugation at 13,000 rpm for 15 min, the<br />

supernatant containing plasmid DNA was transferred into another plastic<br />

reaction tube. DNA was precipitated by addition <strong>of</strong> 0.7 volume (300 µl) <strong>of</strong> 2-<br />

propanol <strong>and</strong> subsequent centrifugation for 30 min. Then the DNA pellet was<br />

washed by 400-500 µl <strong>of</strong> 70% ethanol solution. After centrifugation for 15 min,<br />

the DNA pellet was dried in a speedvacuum centrifuge for 5 min, <strong>and</strong> dissolved<br />

in 20 µl <strong>of</strong> deionised water.<br />

4.2.1.2 Plasmid DNA isolation by NucleoSpin plasmid kit<br />

For preparation <strong>of</strong> larger quantities <strong>of</strong> plasmid DNA, the NucleoSpin kit<br />

provided by Macherey-Nagel (Düren) was applied according to the protocol<br />

provided by the manufacturer. The NucleoSpin preparation is based on alkaline<br />

lysis, too, but instead <strong>of</strong> precipitation the plasmid DNA is recovered by binding<br />

to a silica-membrane. The application <strong>of</strong> chaotrophic salts in the buffer system<br />

<strong>of</strong> the kit leads to a high purity <strong>of</strong> the obtained DNA, thus this method was<br />

applied for DNA preparations used for subsequent cloning or sequencing.<br />

4.2.2 Preparation <strong>of</strong> genomic DNA<br />

Genomic DNA <strong>of</strong> P. syringae was isolated by chlor<strong>of</strong>orm phenol extraction. As<br />

the pseudomonads produce large quantities <strong>of</strong> exopolysaccharides that<br />

interfere with DNA extraction, a cethylhexatrimethylammoniumbromide (CTAB)<br />

extraction <strong>of</strong> exopolysaccharides was applied (Ausubel et al., 1987).<br />

1.5 ml <strong>of</strong> a over night culture <strong>of</strong> P. syringae were harvested by centrifugation.<br />

Supernatant was discarded <strong>and</strong> the cell pellet was resuspended in 567 µl <strong>of</strong><br />

buffer P1 (4.2.1.1). 30 µl <strong>of</strong> a 10% SDS solution a 3µl <strong>of</strong> 20 mg/ml proteinase K<br />

were added. The sample was mixed <strong>and</strong> incubated at 37°C for one hour. 100 µl<br />

<strong>of</strong> 5M NaCl was added to the viscous lysate <strong>and</strong> mixed thoroughly. Then, 80µl<br />

<strong>of</strong> a 65°C preheated CTAB/NaCl solution (10% CTAB in 0.7 M NaCl) was<br />

added. The sample was mixed <strong>and</strong> incubated at 65°C for 10 minutes.<br />

Afterwards, the CTAB/polysaccharide complexes were removed by chlor<strong>of</strong>orm-<br />

30


Material <strong>and</strong> Methods<br />

extraction. For this, 1 volume <strong>of</strong> chlor<strong>of</strong>orm/isoamyl alcohol (24:1; v/v) was<br />

added to the sample, mixed <strong>and</strong> centrifuged for 10 min. in a microcentrifuge.<br />

Subsequently, the upper phase was transferred to a fresh plastic reaction tube<br />

<strong>and</strong> mixed with 1 volume <strong>of</strong> phenol/chor<strong>of</strong>orm/isoamyl alcohol (25:24:1; v/v).<br />

After centrifugation, the upper phase was transferred to a fresh plastic reaction<br />

tube, <strong>and</strong> mixed with 1 volume <strong>of</strong> chlor<strong>of</strong>orm/isoamyl alcohol (24:1; v/v) to<br />

remove residues <strong>of</strong> phenol. After centrifugation, the upper phase was<br />

transferred to a fresh plastic reaction tube <strong>and</strong> the genomic DNA was<br />

precipitated by addition <strong>of</strong> 0.6 volumes <strong>of</strong> isopropanol. The precipitate was<br />

harvested by 30 min centrifugation in a microcentrifuge <strong>and</strong> washed once with<br />

70 % ethanol. The resulting DNA-pellet was dried <strong>and</strong> resuspended in 50-<br />

300 µl <strong>of</strong> sterile water. DNA was stored at –20°C.<br />

4.2.3 DNA separation by gel electrophoresis<br />

DNA-fragments can be separated in an agarose gel matrix upon application <strong>of</strong><br />

an electric field. Migration <strong>of</strong> the DNA fragments depends on the electrical<br />

charge <strong>of</strong> the fragment <strong>and</strong> its structural interactions with the polymer mesh <strong>of</strong><br />

the agarose gel. Therefore for linear DNA fragments the migration distance<br />

correlates to the fragment size. Upon addition <strong>of</strong> ethidium bromide which is<br />

intercalated into GC-pairs, DNA becomes fluorescent <strong>and</strong> can be visualized<br />

under UV-light (Sambrook et al., 1989). Agarose gels were prepared in TAE<br />

buffer (1x: 40 mM Tris-acetate, 1.3 mM EDTA-Na, 0.47 mM acetic acid). For<br />

agarose gels that were subsequently blotted to a membrane an agarose<br />

concentration <strong>of</strong> 0.8 % was used, while gels for st<strong>and</strong>ard applications<br />

contained 1 % agarose. DNA samples were mixed with 1/6 volume <strong>of</strong> loading<br />

buffer (6x: 0.25 % bromphenol blue, 0.25 % xylene cyanol, 40 % sucrose) <strong>and</strong><br />

separated at 80 V in TAE buffer. The size <strong>of</strong> DNA fragments was determined by<br />

parallel loading to the gel <strong>of</strong> the 1-kb DNA ladder (GibcoBRL) or the gene ruler<br />

(MBI fermentas). Both markers are suitable to estimate the size <strong>of</strong> DNA<br />

fragments in a range from 500 bp to 12 kb. After electrophoresis, the gel was<br />

31


Material <strong>and</strong> Methods<br />

stained in ethidium bromide solution (10 µg/ml in TAE buffer) for 5-15 min.<br />

Then the gel was analysed <strong>and</strong> documented under UV-light using the gel jet<br />

imager system (INTAS).<br />

4.2.4 Polymerase chain reaction (PCR)<br />

Polymerase chain reaction (PCR) is a method used to amplify a specific DNA<br />

sequence in vitro by repeated cycles <strong>of</strong> synthesis with specific primers <strong>and</strong><br />

thermostable Taq DNA polymerase (Saiki et al., 1988). Specific primers are<br />

complementary to sequences that lie on opposite str<strong>and</strong>s <strong>of</strong> the template DNA<br />

<strong>and</strong> flank the segment <strong>of</strong> DNA that is to be amplified. The template DNA was<br />

first denatured by heating in the presence <strong>of</strong> a large molar excess <strong>of</strong> each <strong>of</strong><br />

the two primers <strong>and</strong> dNTPs. The reaction mixture was then cooled to a<br />

temperature that allows the primers to anneal to their target sequences, <strong>and</strong><br />

subsequently the annealed primers were extended with DNA polymerase.<br />

Cycles <strong>of</strong> denaturation, annealing, <strong>and</strong> DNA synthesis were repeated several<br />

times. Reaction mixture was prepared according to the supplier’s manual.<br />

As the primers used in this study were designed from heterologues sequence<br />

information, touch-down PCR was applied to enhance the yield <strong>of</strong> product.<br />

A typical PCR thermopr<strong>of</strong>ile was programmed as follows:<br />

Initial denaturation 95°C 5 min 1x<br />

Denaturation 95°C 1 min<br />

Hybridisation 65-50° 30 sec. 10x<br />

Polymerisation 65°C 1 min/kb<br />

Denaturation 95°C 1 min<br />

Hybridisation 55°C-45°C (-0.5°C/cycle) 30 sec. 25x<br />

Polymerisation 65°C 1 min/kb<br />

Final extension 65°C 5 min<br />

32


Material <strong>and</strong> Methods<br />

4.2.5 Cloning techniques<br />

4.2.5.1 Digestion <strong>of</strong> plasmid DNA with endonucleases<br />

Type II restriction endonucleases recognize a DNA region with a specific<br />

palindromic sequence <strong>of</strong> 4-8 bp <strong>and</strong> cut it. Cleavage <strong>of</strong> DNA linearizes it,<br />

generating either blunt or 3’- or 5’-overhanging sticky ends. For DNA restriction,<br />

sample DNA was incubated in an appropriate restriction buffer with enzyme(s)<br />

for 2-16 hours either at 30°C or 37°C according to the protocol provided by the<br />

supplier. The amount <strong>of</strong> enzyme added to a reaction mixture <strong>and</strong> the duration<br />

<strong>of</strong> digestion were determined according to the assumption that 1 U <strong>of</strong> enzyme<br />

cleaves 1 µg <strong>of</strong> DNA for 1 h in 50 µl reaction volume. For subsequent cloning<br />

procedures, the restriction enzyme was either inactivated by heating or by<br />

purification <strong>of</strong> the reaction with the Quiaquick PCR-purification system<br />

(Quiagen).<br />

4.2.5.2 De-phosphorylation <strong>of</strong> digested DNA<br />

In order to prevent self-ligation <strong>of</strong> vector DNA during cloning procedures, a<br />

treatment with shrimp alkaline phosphatase (SAP) was applied. This removes<br />

the phosphate groups from 5’-ends <strong>of</strong> linearized DNA. SAP can be added<br />

directly to a digestion mix because the SAP buffer is usually compatible with<br />

the buffers for restriction endonucleases. 1/10 volume <strong>of</strong> the SAP buffer were<br />

added to a sample. The amount <strong>of</strong> SAP, which was added to the reaction,<br />

equalled the amount <strong>of</strong> endonuclease that had been used to linearize the<br />

vector DNA. The reaction mixture was incubated for 30 min at 37°C to<br />

accomplish de-phosphorylation. Afterwards, SAP <strong>and</strong> restriction enzymes were<br />

inactivated by heating at 65°C for 15 min. This step is important in order to<br />

avoid de-phosphorylation putative insert DNA fragment in subsequent ligation<br />

reactions. To remove salts, the reaction was purified by Quiaquick PCRpurification<br />

system.<br />

33


Material <strong>and</strong> Methods<br />

4.2.5.3 Converting overhanging sticky ends to blunt ends<br />

In order to remove overhanging ends from endonuclease treated DNAfragments<br />

the samples were incubated with Klenow polymerase. After<br />

inactivation <strong>of</strong> endonuclease, 1/20 volume <strong>of</strong> 0.5 mM dNTPs, 1/10 volume <strong>of</strong><br />

the appropriate 10x-buffer <strong>and</strong> 1-5 U <strong>of</strong> Klenow fragment were added to the<br />

reaction mixture <strong>and</strong> the sample was incubated at 30°C for 20 minutes. The<br />

reaction was stopped by heat inactivation at 75°C for 10 min. The reaction was<br />

purified by Quiaquick PCR-purification system.<br />

4.2.5.4 Precipitation <strong>of</strong> DNA by ethanol<br />

One possible method to concentrate or purify DNA is a precipitation by salt <strong>and</strong><br />

cold ethanol. 1/9 Volume <strong>of</strong> 3 M sodium acetate (pH 5.8) <strong>and</strong> 2 volumes <strong>of</strong> precooled<br />

absolute ethanol were added to the DNA solution. The mixture was<br />

incubated for at least an hour in an ethanol/ice bath at -20°C or for half an hour<br />

at –80°C. Then the sample was centrifuged at 13,000 rpm for 30 min at 4°C.<br />

The supernatant was discarded, <strong>and</strong> the precipitated DNA was washed in 70%<br />

ethanol solution followed by another centrifugation at 13,000 rpm for 15 min at<br />

4°C. The resulting DNA pellet was dried in a speed-vacuum centrifuge <strong>and</strong><br />

dissolved in 10-20 µl <strong>of</strong> sterile water. DNA samples were stored at –20°C.<br />

4.2.5.5 Extraction <strong>of</strong> DNA fragments out <strong>of</strong> agarose gels<br />

For re-isolation <strong>of</strong> DNA fragments after separation by agarose gel<br />

electrophoresis the QIAquick extraction Kit provided by Qiagen was used<br />

according to the supplier’s protocol. This kit applies a silica matrix in<br />

combination with a buffer system containing chaotrophic salts to selectively<br />

bind DNA fragments out <strong>of</strong> the agarose – DNA Mixture. The DNA fragment <strong>of</strong><br />

appropriate size was cut from the ethidium bromide strained gel under<br />

illumination with UV-light. To avoid formation <strong>of</strong> thymine dimmers, the exposure<br />

to the UV light was reduced to a minimum. The resulting agarose slice was<br />

mixed with 300 µl QG-buffer per 100 mg <strong>of</strong> gel. Then the gel piece was<br />

34


Material <strong>and</strong> Methods<br />

dissolved in the buffer by 10 min incubation at 50°C in a thermomixer. During<br />

this incubation the sample was mixed every two minutes. After complete<br />

disolvance <strong>of</strong> the agarose piece, column purification was performed in<br />

accordance to the manual. The DNA was eluted in water <strong>and</strong> stored at –20°C.<br />

4.2.5.6 QIAquick PCR <strong>and</strong> reaction purification kit<br />

The QIAquick-kit (Qiagen, Hilden) was used to re-isolate DNA out <strong>of</strong> reaction<br />

mixtures. 5 volumes <strong>of</strong> buffer PB were added to the reaction. After that the<br />

mixture was loaded to a QIAquick spin column in a 2 ml collection tube by short<br />

centrifugation. Then the column was washed with 0.75 ml <strong>of</strong> buffer PE <strong>and</strong><br />

placed into a 1.5 ml plastic reaction tube. To elute, 30 µl <strong>of</strong> sterile water was<br />

added to the center <strong>of</strong> the column <strong>and</strong> incubated at room temperature for<br />

1-3 minutes. The sample was shortly centrifuged (1 min, 13,000 rpm). DNA<br />

samples were stored at –20°C.<br />

4.2.5.7 Estimation <strong>of</strong> DNA concentration<br />

Concentration <strong>of</strong> DNA samples was determined photometrically, as an<br />

absorption <strong>of</strong> 1 at λ = 260nm correlates to approx. 50 µg DNA per ml solution<br />

(Sambrook et al., 1989).<br />

4.2.5.8 Ligation <strong>of</strong> DNA<br />

T4 DNA ligase catalyses the formation <strong>of</strong> phosphodiester bonds at juxtaposed<br />

5’-phosphate <strong>and</strong> 3’-hydroxyl ends in duplex DNA. The appropriate insertion <strong>of</strong><br />

a piece <strong>of</strong> insert into an opened vector by this reaction is most efficient, if a<br />

relatively small amount <strong>of</strong> vector (20-100 ng) is mixed with an 3-8 fold excess <strong>of</strong><br />

insert DNA. To high concentration <strong>of</strong> vector facilitates self-ligation <strong>of</strong> vector<br />

DNA while to low concentration <strong>of</strong> insert DNA minimizes ligation efficiency. The<br />

ligation reaction contained 1 µl <strong>of</strong> 10x buffer <strong>and</strong> 1 µl <strong>of</strong> T4 DNA ligase. It was<br />

incubated at 18°C for about 20 hours. Subsequently the ligated DNA was<br />

transformed into E. coli DH5α. The ligation reaction was stored at –20°C.<br />

35


Material <strong>and</strong> Methods<br />

4.2.5.9 Ligation <strong>of</strong> PCR fragments by T/A cloning<br />

One attribute <strong>of</strong> the Taq polymerase is the creation <strong>of</strong> a so-called A-overhang.<br />

After finishing the polymerisation <strong>of</strong> the according PCR fragment, a terminal<br />

adenine is attached to the fragment. The pGEM-T Easy cloning kit provided by<br />

Promega uses this feature for an enhanced cloning <strong>of</strong> PCR fragments. The kit<br />

provides linearized DNA <strong>of</strong> the pGEM-T Easy vector extended by a terminal<br />

thymine. The combination <strong>of</strong> T- <strong>and</strong> A-overhang provide a sticky end that highly<br />

increases ligation efficacy.<br />

For pGEM-T Easy cloning the purified PCR fragment was mixed with 0.5 µl <strong>of</strong><br />

vector solution, 2µl <strong>of</strong> 5x rapid ligation buffer <strong>and</strong> 5 U <strong>of</strong> high concentrated T4<br />

ligase in a 10 µl reaction mixture. The ligation was incubated at 18°C over night<br />

<strong>and</strong> subsequently transformed into E. coli DH5α. Ligation reaction was stored<br />

at –20°C.<br />

4.2.5.10 Preparation <strong>of</strong> electrocompetent E. coli cells<br />

Electrocompetent cells <strong>of</strong> E. coli DH5α were prepare according to the st<strong>and</strong>ard<br />

protocol (Sambrook et al., 1989). A fresh over night culture <strong>of</strong> DH5α was<br />

inoculates at a dilution <strong>of</strong> 1:100 <strong>and</strong> incubated at 37°C until exponential growth<br />

phase was reached (OD 600nm = 0.5 - 0.8). The cells were incubated on ice for<br />

30 min. <strong>and</strong> then harvested by centrifugation (600rpm; 4°C). Next, medium<br />

residues <strong>and</strong> salts were removed from the cells by three subsequent washing<br />

steps (two times ice cold water, one ice cold 10% glycerol). The resuspension<br />

volume was reduced in each step <strong>and</strong> after the final washing step the cells<br />

were resuspended in 1/500 volume <strong>of</strong> the initial culture volume. 40 µl aliquots<br />

<strong>of</strong> electrocompetent cells were filled into 0.5 ml plastic reaction tubes <strong>and</strong><br />

directly shock frozen in liquid nitrogen. The cells were stored at –80°C.<br />

36


Material <strong>and</strong> Methods<br />

4.2.5.11 Preparation <strong>of</strong> electrocompetent P. syringae cells<br />

One inoculation loop <strong>of</strong> P. syringae cells was scratched from an over night KB<br />

agar plate. The cells were resuspended in 1ml <strong>of</strong> ice-cold water <strong>and</strong> kept on ice<br />

for 15 to 30 minutes. Then the cells were pelleted by centrifugation <strong>and</strong> washed<br />

with 1 ml <strong>of</strong> ice-cold water two times. Thereafter the cells were washed another<br />

two times in 0.5 ml <strong>of</strong> ice cold 10% glycerol. After the final washing step, the<br />

cell pellet was resuspended in 200 µl <strong>of</strong> 10% glycerol. The cells were not stored<br />

but directly used for electroporation.<br />

4.2.5.12 Electroporation <strong>of</strong> DNA<br />

40 µl <strong>of</strong> electrocompetent cells (either E.coli or P. syringae) were mixed with<br />

0.1-1 µg <strong>of</strong> plasmid DNA or 1-2 µl <strong>of</strong> ligation mixture, <strong>and</strong> exposed to electric<br />

shock in a pre-chilled electroporation cuvette. The GenePulser-Apparatus<br />

(BioRad) used for this experiment was set to 2.5 kV <strong>and</strong> 200 Ω. Directly after<br />

the electric pulse, cells were resuspended in 1 ml <strong>of</strong> the appropriate medium<br />

(LB medium for E. coli <strong>and</strong> KB medium for P. syringae) The cells were then<br />

incubated for 60 min. by shaking at 280 rpm at the respective growth<br />

temperature, <strong>and</strong> plated onto agar plates (LB medium or KB medium)<br />

containing the appropriate antibiotic(s). The plates were incubated until single<br />

colonies became visible.<br />

4.2.5.13 Conjugation by triparental mating<br />

The native mechanism <strong>of</strong> horizontal gene transfer by plasmid conjugation via<br />

‘triparental mating’ was used to transfer broad-host range plasmids into P.<br />

syringae. Donor plasmids with the mob function are mobilized <strong>and</strong> transferred<br />

into the recipient P. syringae strain by the use <strong>of</strong> the helper strain E. coli HB101<br />

(pRK2013) carrying the transfer function. Recipient P. syringae strains were<br />

grown for 2 days on MG agar plates prior to the conjugation. Approximately one<br />

inoculation loop <strong>of</strong> recipient bacteria was resuspended in 1 ml <strong>of</strong> sterile water.<br />

Subsequently, 1/3 <strong>of</strong> an inoculation loop each <strong>of</strong> the donor <strong>and</strong> helper E. coli<br />

37


Material <strong>and</strong> Methods<br />

strains were scratched from an over night grown plate <strong>and</strong> resuspended in the<br />

same volume. Then 2 µl <strong>of</strong> each bacterial suspension was mixed together. The<br />

resulting 60µl <strong>of</strong> bacterial suspension was divided into three aliquots <strong>of</strong> 20 µl<br />

<strong>and</strong> spotted onto an LB plate. The suspension was allowed to dry <strong>and</strong> the plate<br />

was subsequently incubated over night at 28°C. Then the mating spots were<br />

scraped <strong>of</strong>f the plate <strong>and</strong> resuspended in 1 ml sterile water, each. This cell<br />

suspension was diluted from 10 -1 to 10 -5 . 200-300 µl <strong>of</strong> each dilution step were<br />

plated onto MG plates with the appropriate antibiotic(s) to select for<br />

transconjugants. Subsequently, plates were incubated at 28° for 3-5 days until<br />

single colonies <strong>of</strong> P. syringae became visible. The selected transconjugants <strong>of</strong><br />

P. syringae were restreaked twice on MG agar plates in order to eliminate<br />

residual E. coli cells.<br />

4.2.6 DNA-analysis by Southern blot technique<br />

4.2.6.1 Generation <strong>of</strong> digoxigenin labelled probes<br />

Digoxigenin labelling <strong>of</strong> DNA-probes for DNA-DNA hybridisation was performed<br />

according to the manual <strong>of</strong> the “DIG DNA labelling <strong>and</strong> detection kit”. 15 µl <strong>of</strong><br />

sample DNA was incubated at 95°C for 10 min. <strong>and</strong> subsequently shock-cooled<br />

in an ethanol/ice-bath. The denaturated sample DNA was the mixed with 2 µl<br />

10x hexanucleotidemix, 2 µl <strong>of</strong> dNTP mixture (1 mM dATP/dCTP/dGTP;<br />

0.65 mM dTTP <strong>and</strong> 0.35 mM DIG-dUTP) <strong>and</strong> 1 µl Klenow fragment. The<br />

reaction mixture was incubated at 37°C over night. In order to remove residual<br />

nucleotides, the labelled DNA probe was precipitated by addition <strong>of</strong> 1 µl 0.5 M<br />

EDTA, 2.5 µl <strong>of</strong> 4M LiCl <strong>and</strong> 75 µl <strong>of</strong> pre-cooled EtOH (96%). The mixture was<br />

incubated for 30min. at –80°C <strong>and</strong> subsequently centrifuged. The pelleted DNA<br />

was washed with 1 ml <strong>of</strong> cold 70% EtOH <strong>and</strong> dried completely. The labelled<br />

DNA was resuspended in 50 µl <strong>of</strong> water <strong>and</strong> stored at –20°C.<br />

38


Material <strong>and</strong> Methods<br />

4.2.6.2 Alkaline transfer <strong>of</strong> DNA<br />

The DNA for Southern blot analysis was separated by gelelectrophoresis<br />

(4.2.3) on a 0.8% agarose gel. After documentation <strong>of</strong> the gel, the DNA was<br />

transferred to a positively charged nylon membrane (Hybond N + ) by capillary<br />

transfer using 5x SSC/10 mM NaOH buffer. The transfer was performed over<br />

night. After transfer the DNA was fixed to the membrane using a UVcrosslinker.<br />

The dried membrane could be stored at room temperature until<br />

hybridisation.<br />

Composition <strong>of</strong> 20x SSC is 3 M NaCl <strong>and</strong> 0.3 M tri-sodium citrate.<br />

4.2.6.3 Hybridisation <strong>and</strong> detection<br />

Prior to hybridisation with the probe, the membrane was incubated with<br />

hybridisation buffer (0.1% n-laurylsarcosin, 0.02% SDS <strong>and</strong> 1% blocking<br />

reagent in 5x SSC) for 1 hour. In the meantime, the DNA-probe was<br />

denaturated for 10 min. at 95°C <strong>and</strong> then shock-cooled in an ethanol/ice bath.<br />

Then the probe was resuspended in 10 ml <strong>of</strong> pre-warmed hybridisation buffer.<br />

Then the probe/buffer mixture was added to the membrane <strong>and</strong> hybridised for<br />

12-16 hours at the respective temperature. After hybridisation the membrane<br />

was washed two times under non-stringent conditions (5 min. room<br />

temperature in 2x SSC, 0.1% SDS) <strong>and</strong> two times under stringent conditions<br />

(15 min at hybridisation temperature in 0.1x SSC, 0.1% SDS). Thereafter the<br />

membrane was equilibrated at room temperature for 1 min. in buffer 1 (100 mM<br />

Tris/HCl at pH 7.5 with 150 mM NaCl). Then the membrane was incubated at<br />

room temperature for 30 min. in the blocking buffer 2 (0.5% blocking reagent in<br />

buffer 1). After a washing the membrane in buffer 1, the antibody solution (2 µl<br />

<strong>of</strong> anti-digoxygenin-ap, fab-fragment by Roche, Mannheim, in 20 ml <strong>of</strong> buffer 1)<br />

was added to the membrane <strong>and</strong> incubated at room temperature for 30 min..<br />

Any excess <strong>of</strong> antibody was removed by two washing steps in buffer 1 (15 min.,<br />

room temperature). Finally, the membrane was equilibrated for 1 min. in the<br />

detection buffer (0.1 M Tris/HCL, 100 mM NaCl, <strong>and</strong> 50 mM MgCl 2 at an pH <strong>of</strong><br />

39


Material <strong>and</strong> Methods<br />

9.5). For signal generation, the membrane was incubated with the ECFsubstrate<br />

(Amersham Pharmacia) on a glass tray. The signal was detected on<br />

a phosphoimager (FLA-3000, Raytest). The membrane was scanned at<br />

excitation wavelength 475 nm <strong>and</strong> emission signals above 520 nm were<br />

detected.<br />

4.2.7 DNA-sequencing<br />

DNA sequencing was performed at MWG Biotech (Ebersberg).<br />

4.3 Analysis <strong>of</strong> bacterial growth in vitro<br />

For growth curve analysis <strong>of</strong> P. syringae bacterial cultures were inoculated to<br />

an OD 600nm <strong>of</strong> 0.05 to 0.1 with an exponentially growing pre-culture. The<br />

cultures were incubated at 28°C by shaking at 280 rpm. For each strain 3<br />

separate cultures were analysed in parallel.<br />

4.4 Analysis <strong>of</strong> bacterial growth in planta<br />

For in planta analysis, soybean plants <strong>of</strong> the cultivar “Maple Arrow” were grown<br />

in light shelves (Conviron) at approx. 23°C with a light period <strong>of</strong> 16 hours.<br />

Relative humidity was approx. 40%.<br />

4.4.1 Wound inoculation<br />

Bacterial inoculum was prepared by resuspending cells from a fresh KB-agar<br />

plate in sterile water. OD 600nm <strong>of</strong> the suspension was adjusted to 0.01. For<br />

inoculation <strong>of</strong> the antagonist Pss22d, this suspension was mixed 2:1 with sterile<br />

water. For inoculation <strong>of</strong> the pathogen Psg1a, the suspension was diluted 1:2<br />

with sterile water. For co-inoculation, 1 volume <strong>of</strong> pathogen suspension was<br />

mixed with 2 volumes <strong>of</strong> antagonist suspension. Subsequently, leafs <strong>of</strong> 4-weekold<br />

soybean plants were punched with a sterile needle (10 wounds per leaflet).<br />

Each wound was inoculated with 5 µl <strong>of</strong> bacterial suspension. For<br />

40


Material <strong>and</strong> Methods<br />

determination <strong>of</strong> bacterial population 20 such wounds were analysed per timepoint.<br />

Disks <strong>of</strong> 7 mm diameter surrounding the wound-site were punched from<br />

the leaflet using a cork-borer. The disks were grinded down in 20 ml <strong>of</strong> sterile<br />

NaCl (0.9%). A dilution serious <strong>of</strong> the resuspension was plated on KB-agar <strong>and</strong><br />

cell-forming units were counted after 48 hours <strong>of</strong> incubation.<br />

4.4.2 Spray inoculation<br />

For spray inoculation experiments, bacterial inoculum was prepared by<br />

resuspending cells from a fresh KB-agar plate in sterile water. OD 600nm <strong>of</strong> the<br />

suspension was adjusted to 0.01. 100 ml <strong>of</strong> this suspension were sprayed on<br />

three 4 weak old soybean plants. For determination <strong>of</strong> bacterial population<br />

fresh soybean leaflets were cut <strong>and</strong> grinded down in sterile 0.9% NaCl (10ml <strong>of</strong><br />

NaCl per gram fresh weight). A dilution serious <strong>of</strong> the resuspension was plated<br />

on KB-agar <strong>and</strong> cell-forming units were counted after 48 hours <strong>of</strong> incubation.<br />

4.5 Analysis <strong>of</strong> siderophore-production<br />

4.5.1 Uptake <strong>of</strong> 59 Fe<br />

59 Fe can be utilized to monitor the bacterial uptake <strong>of</strong> iron/siderophore<br />

complexes (Meyer et al., 2002). In this experiment, bacterial cells from 40 h <strong>of</strong><br />

culture in CAA medium were harvested by centrifugation, washed once with<br />

distilled water, <strong>and</strong> resuspended at an OD 600nm <strong>of</strong> 0.33 in an incubation medium<br />

made <strong>of</strong> succinate medium with the nitrogen source omitted. Label mix<br />

containing 59 Fe/siderophore complex consisted <strong>of</strong> 5 µl <strong>of</strong> the commercial 59 Fe 3+<br />

solution (iron chloride in 0.1 M HCl, specific activity 110 to 925 MBq/mg <strong>of</strong> iron;<br />

Amersham) was diluted first with 100 µl <strong>of</strong> water <strong>and</strong> then mixed with 10 µl <strong>of</strong> a<br />

6.5-mg/ml siderophore solution. Alternatively, 5 µl <strong>of</strong> the 59 Fe 3+ solution were<br />

mixed with 100 µl <strong>of</strong> siderophore-containing culture supernatants. The final<br />

volume <strong>of</strong> the label mix was adjusted after 30 min <strong>of</strong> incubation at room<br />

temperature to 1 ml with incubation medium. Bacterial suspension (1.8 ml) was<br />

mixed at time zero with 0.2 ml <strong>of</strong> label mix. After 20 min <strong>of</strong> incubation under<br />

41


Material <strong>and</strong> Methods<br />

gentle shaking in a waterbath at 25°C, 1 ml <strong>of</strong> the bacterial suspension was<br />

rapidly filtered through a Whatman nitrocellulose filter (0.45-µm pore size), <strong>and</strong><br />

the filter was washed twice with 2 ml <strong>of</strong> fresh incubation medium. Each filter<br />

was then wrapped in aluminum foil, <strong>and</strong> counts were determined in a Gamma<br />

4000 counter (Beckman). The radioactivity in the remaining 1 ml <strong>of</strong> bacterial<br />

suspension was counted directly to determine the total amount <strong>of</strong> radioactivity<br />

present in the assay. Control assays without bacteria were performed to verify<br />

the complete solubility <strong>of</strong> labelled iron through siderophore complexation.<br />

4.5.2 Determination <strong>of</strong> siderophore concentration by quantitative CASassay<br />

Quantification <strong>of</strong> siderophore activity was performed according to the CASassay<br />

(Schwyn & Neil<strong>and</strong>s, 1987). This photometric assay quantifies a colour<br />

change <strong>of</strong> a blue coloured iron complex to red/orange after removal <strong>of</strong> the iron<br />

by a siderophore. The indicator solution was prepared as follows:<br />

1.5 ml <strong>of</strong> 1 mM FeCl 3+ in 10 mM HCl was mixed with 7.5 ml <strong>of</strong> a 2 mM<br />

chrome azurol S (CAS) solution <strong>and</strong> slowely added to 6 ml <strong>of</strong> a 10 mM<br />

hexadecyltrimethylammonium bromide (HDTMA) solution. The resulting mixture<br />

was subsequently added to 80 ml <strong>of</strong> a buffer solution (4.307 g piperazine in<br />

80 ml <strong>of</strong> water, neutralized by addition <strong>of</strong> 6.25ml <strong>of</strong> conz. HCl). This indicator<br />

mixture was stored in the dark. Directly before application, 4 mM sulfosalicylic<br />

acid, a substance that enhances siderophore/iron complexe formation. For the<br />

quantification assay, the indicator solution was diluted ½ by addition <strong>of</strong> the<br />

sample or a dilution <strong>of</strong> the sample. The mixture was incubated for 1 hour <strong>and</strong><br />

the siderophore-dependent colour change was determined at λ=630nm using a<br />

spectrophotometer (Ultrospec 2100pro, Pharmacia Biotech, Freiburg).<br />

Deferoxaminmesylate (DFOM) was applied as a quantification st<strong>and</strong>ard. The<br />

CAS reaction was linear to a concentration <strong>of</strong> 0-20 µM DFOM.<br />

42


Material <strong>and</strong> Methods<br />

4.5.3 Quantification <strong>of</strong> pyoverdin by fluorescence measurement<br />

The presence <strong>of</strong> pyoverdin was estimated by fluorescence measurements<br />

using the F-2500 spectr<strong>of</strong>luorometer (Digilab, Krefeld). Purified pyoverdin <strong>of</strong><br />

Pss (courtesy <strong>of</strong> Pr<strong>of</strong>. J.M. Meyer, Strasbourg, France) was used as st<strong>and</strong>ard.<br />

For comparison between st<strong>and</strong>ard <strong>and</strong> sample, the pyoverdin was dissolved in<br />

Pipes-medium (1 mg/ml). The excitation <strong>and</strong> emission spectrum was analysed<br />

<strong>and</strong> an excitation maximum at λ= 397 nm as well as a maximum <strong>of</strong> emission at<br />

λ= 458 nm was determined. Thus these parameters were applied for the<br />

quantification measurements. A dilution series <strong>of</strong> pyoverdin in Pipes medium<br />

served as equilibration curve for quantification. The fluorescence <strong>of</strong> the sample<br />

was linear to pyoverdin concentration in a range <strong>of</strong> 0.5-7 µg pyoverdin/ml.<br />

Culture supernatants were diluted until the fluorescence <strong>of</strong> the sample was in<br />

the linear range.<br />

4.5.4 Purification <strong>of</strong> siderophores by XAD4<br />

Siderophore-containing supernatant <strong>of</strong> 5b cultures were harvested after 48 h <strong>of</strong><br />

incubation. The pH <strong>of</strong> the supernatant was subsequently adjusted to pH = 6 for<br />

pyoverdin purification <strong>and</strong> to pH = 3 for purification <strong>of</strong> achromobactin or the<br />

achromobactin-like siderophore. 50 ml to 150 ml <strong>of</strong> XAD4-resin (Sigma) were<br />

loaded to a column. The column was pre-washed with water (1 L) <strong>and</strong> then the<br />

supernatant was applied by gravity flow. Subsequently, the column was<br />

washed with water (2 times the supernatant volume). Finally, the siderophore<br />

was eluted in 100% methanol, until no more siderophore activity could be<br />

measured in the eluate. The eluate was brought to dryness in a rotation<br />

evaporator <strong>and</strong> eluted in 5-10 ml volume <strong>of</strong> water. This sample was brought to<br />

dryness using a freeze drying apparatus (Edwards).<br />

For fast screening <strong>of</strong> siderophores, this purification method was downscaled.<br />

2 ml <strong>of</strong> culture supernatant were harvested <strong>and</strong> applied on a Mobicon minicolumn<br />

(Mobitech, Göttingen) containing 2 ml <strong>of</strong> XAD4-resin. Supernatant was<br />

applied to the column twice in order to improve siderophore binding. The<br />

43


Material <strong>and</strong> Methods<br />

columns were washed twice with 5 ml <strong>of</strong> water <strong>and</strong> eluted with 500 µl <strong>of</strong> 100%<br />

methanol. The eluate was brought to dryness using an Eppendorf<br />

Concentrator.<br />

4.5.5 Isoelectric focussing <strong>of</strong> siderophores<br />

The method <strong>of</strong> Koedam et al. was adapted to perform isoelectric focussing <strong>of</strong><br />

siderophores (Koedam et al., 1994). The Multiphore II electrophorese unit<br />

(Amersham) was used for IEF. Polyacrylamid gels consisted <strong>of</strong> 4 ml Readysolve<br />

IEF mixture (Amersham), 4 ml <strong>of</strong> 25% glycerol, <strong>and</strong> 1 ml <strong>of</strong> ampholyte 3-<br />

10 (Biorad). Water was added to a volume <strong>of</strong> 20 ml <strong>and</strong> 25 µl TEMED as well<br />

as 50 µl <strong>of</strong> 10% ammoniompersulfate were added to start polymerisation. After<br />

polymerisation, the gel was placed on the cooling plate <strong>of</strong> the Multiphore II<br />

elecrophoresis unit. Per lane, 1µl <strong>of</strong> sample was applied. A three-step<br />

electrophoresis with 15 min at 100 V, 15 min at 200 V, <strong>and</strong> 1 h at 450 V was<br />

performed at 4°C. Immediately after electrophoresis, the siderophore b<strong>and</strong>s<br />

were visualized by CAS overlay. For this the CAS indicator solution (4.5.2) was<br />

supplemented with 1% agarose <strong>and</strong> a 0.1-2 mm gel was cast. The indicator gel<br />

was overlayed over the polyacrylamid gel <strong>and</strong> bubbles were carefully removed.<br />

After short incubation (5-10 min) siderophore active b<strong>and</strong>s became visible due<br />

to the colour change <strong>of</strong> the CAS-indicator.<br />

4.5.6 Gel-filtration<br />

The Biorad Econo-system was used for gel-filtration. The column size used was<br />

0.5 cm x 20 cm. Three different column-materials were tested (Fractogel TSK<br />

HW 50s, Merck; Bio-Gel P-30, Biorad; <strong>and</strong> G25 Sephadex, Amersham). All<br />

materials were equilibrated in water. XAD4-purified siderophore was<br />

resuspended in 0.5-1 ml <strong>of</strong> water <strong>and</strong> applied to the columns. Mobile phase for<br />

the separation was water at a flow speed <strong>of</strong> 1 ml/min. Fractions <strong>of</strong> 2 ml volume<br />

were collected <strong>and</strong> subsequently tested for siderophore activity by CAS assay<br />

(4.4.2)<br />

44


Material <strong>and</strong> Methods<br />

4.5.7 Thin-layer chromatography<br />

Thin layer chromatography was performed on silica-based TLC plates<br />

(Macherey-Nagel, Düren). Samples were applied as methanol solutions <strong>and</strong><br />

brought to dryness completely prior to separation. Separation was performed in<br />

glass tanks with equilibrated gas phase. The respective mobile phase was<br />

prepared freshly prior to TLC separation. After separation, substances were<br />

detected by aut<strong>of</strong>luorescence at an excitation wavelength <strong>of</strong> λ=365nm or<br />

quenching <strong>of</strong> a fluorescence indicator within the silica matrix (excitation<br />

λ=254nm). Alternatively, the TLC plate was scanned on the phosphoimager<br />

(Fla-3000, Raytest) as described under 4.2.6.3. Detection <strong>of</strong> siderophore<br />

activity was achieved by CAS overlay as described under 4.5.5.<br />

45


Results<br />

5 Results<br />

5.1 Analysis <strong>of</strong> siderophores produced by Pss22d<br />

Under iron-limiting conditions many pseudomonads produce a yellow-greenish,<br />

fluorescent pigment, thus they are classified as “fluorescent pseudomonads”.<br />

These pigments are peptide-type siderophores called pyoverdins. All<br />

pyoverdins share the same dihydroxyquinoline chromophore responsible for<br />

their fluorescence, with an amidically connected small dicarboxylic acid (or its<br />

monoamide). The peptide chain bound to the chromophore carboxyl group<br />

varies. It consists <strong>of</strong> 6-12 amino acids <strong>of</strong>ten including modified amino acids or<br />

D-forms (Fuchs et al., 2001). Approximately 50 different pyoverdins have been<br />

identified so far <strong>and</strong> the type <strong>of</strong> pyoverdin produced can vary even among<br />

strains <strong>of</strong> the same species, for example 3 different pyoverdins have been<br />

characterized for P. aeruginosa (de Chial et al., 2003). In contrast to their<br />

otherwise prominent diversity, all pathovars <strong>of</strong> P. syringae tested so far produce<br />

the same pyoverdin (Fig.6) (Bultreys et al., 2001; Bultreys et al., 2003; Bultreys<br />

et al., 2004; Jülich et al., 2001). In order to validate that pathogen Psg1a <strong>and</strong><br />

antagonist Pss22d do produce this pyoverdin (Pvd Ps ), a<br />

59 Fe-uptake<br />

experiment was performed with the help <strong>of</strong> our collaboration partner<br />

HO<br />

L-Thr 1<br />

HO<br />

H<br />

NH CO N<br />

CO<br />

OC<br />

HN<br />

HO COOH<br />

D-OHAsp 1<br />

OC L-Thr 2<br />

OH<br />

HN<br />

COOH OH<br />

L-Ser 2<br />

L-Ser 1<br />

D-OHAsp 2<br />

A<br />

OC<br />

HN<br />

HO<br />

CO<br />

NH<br />

COOH<br />

HO<br />

HO<br />

NH 2<br />

CH<br />

CH2<br />

CH 2 L-Lys<br />

CH 2<br />

CH2<br />

NH<br />

H<br />

N +<br />

O<br />

NH<br />

NHCOCH 2 CH 2 CONH 2<br />

1, L-Ser 1<br />

2, Gly<br />

B<br />

HO<br />

OC<br />

HN<br />

OC<br />

HN<br />

HO<br />

OC<br />

HN<br />

O<br />

HO<br />

L-Thr 1<br />

L-Thr 2<br />

CO<br />

CO<br />

NH<br />

NH<br />

COOH<br />

CO<br />

HO<br />

D-OHAsp<br />

OH<br />

H<br />

N<br />

COOH<br />

D-OHAsp 1<br />

L-Ser 2<br />

CO<br />

HO<br />

HO<br />

NH 2<br />

CH<br />

CH2<br />

CH 2 L-Lys<br />

CH2<br />

CH 2<br />

NH<br />

H<br />

N<br />

O<br />

+<br />

NH<br />

NHCOCH2CH2CONH2<br />

Fig. 6. The siderophore pyoverdin <strong>of</strong> P. syringae.<br />

The pyoverdin typical for P. syringae has a seven amino-acid side chain that can be linear (A) or<br />

can form a ring between the second threonine <strong>and</strong> the terminal serine (B).<br />

Modified after Bultreys et al., (2004)<br />

46


Results<br />

Pr<strong>of</strong>. J. M. Meyer (Strasbourg, France). Pantoea agglomerans strain 48b<br />

served as control. The cells <strong>of</strong> Psg1a <strong>and</strong> Pss22d could both incorporate<br />

labeled iron in the presence <strong>of</strong> reference pyoverdin <strong>of</strong> the P. syringae type,<br />

while none could utilize a 59 Fe – desferrioxamine E complex. In contrast, cells<br />

<strong>of</strong> P. agglomerans 48b incorporated only the labeled desferrioxamine E <strong>and</strong><br />

not pyoverdin (Tab. 8). The uptake experiment indicated that both strains,<br />

Psg1a <strong>and</strong> Pss22d, produce a pyoverdin typical for P. syringae.<br />

5.1.1 Construction <strong>of</strong> a pyoverdin-deficient mutant <strong>of</strong> Pss22d<br />

The phenotypic difference between Pss22d <strong>and</strong> Psg1a on CAS-agar could<br />

either originate from quantitative differences in pyoverdin production or the<br />

production <strong>of</strong> other siderophores aside <strong>of</strong> pyoverdin. To address this question,<br />

a pyoverdin-negative mutant <strong>of</strong> Pss22d was constructed by marker exchange<br />

mutagenesis. The biosynthesis <strong>of</strong> pyoverdines is not completely understood,<br />

but several essential enzymes have been determined (Ravel & Cornelis, 2003).<br />

As the pyoverdin chromophore is essential for a functional pyoverdin <strong>and</strong> as its<br />

structure is highly conserved, the enzymes involved in chromophore formation<br />

represent good targets to generate a biosynthesis mutant. In P. fluorescense,<br />

the non-ribosomal peptide synthetase (NRPS) PvsA has been determined to<br />

mediate the formation <strong>of</strong> the chromophore precursor. A homologue <strong>of</strong> the<br />

respective pvsA gene in P. syringae pv. maculicola DC3000 (PsmDC3000) had<br />

already been described (Mossialos et al., 2002).<br />

Tab. 8. 59 Fe uptake experiment.<br />

Plant associated bacteria where tested for their ability to incorporate different iron<br />

siderophore complexes labelled by 59 Fe.<br />

59<br />

Fe incorporation mediated by<br />

strain Desferrioxamine E P.syringae pyoverdin<br />

Psg1a - +<br />

Pss22d - +<br />

P. agglomerans 48b + -<br />

47


Results<br />

The 13-kb open reading frame (ORF) <strong>of</strong> pvsA encodes for a multi-domain<br />

enzyme consisting <strong>of</strong> four different modules (Fig.7).<br />

A<br />

module 1 module 2<br />

(L-Glu)<br />

module 3<br />

(D-Tyr → L-Tyr)<br />

module 4<br />

(L-Dab)<br />

B<br />

C<br />

4 3 2 1<br />

PVSA<br />

2 kb PCR<br />

fragment 2<br />

PCR<br />

fragment 1<br />

Fig. 7. Construction <strong>of</strong> a pvsA deletion mutant in Pss22d.<br />

The NRPS PvsA is involved in the biosynthesis <strong>of</strong> the precursor <strong>of</strong> the pyoverdin chromophore<br />

(A). It consists <strong>of</strong> four different modules <strong>and</strong> comprises the domains for activation <strong>of</strong> three<br />

amino acids, L-glutamate, D-tyrosine <strong>and</strong> L-2, 4-diaminobutyric acid, while the function <strong>of</strong><br />

module 1 is undetermined up to now (B). The domains are abbreviated as follows: AL = acyl<br />

CoA ligase, A = adenylation, T = 4´phosphopantheteine binding, E = epimerisation, <strong>and</strong> C =<br />

condensation<br />

For construction <strong>of</strong> the deletion mutant, fragments <strong>of</strong> the 13kb ORF pvsA were amplified by<br />

PCR (C). The respective primer-binding sites are numbered 1-4. The first fragment is the<br />

1,4kb amplificate <strong>of</strong> primer pvsA_fwd1_Spe (1) <strong>and</strong> pvsA_rev6_Kpn (2) <strong>and</strong> the second<br />

fragment is the 1,6kb region amplified by the primers pvsA_fwd7_Kpn (3) <strong>and</strong><br />

pvsA_rev4_Bam (4). The fragments were subsequently fused to a Km r -cassette <strong>and</strong> brought<br />

back into pvsA by double homologues recombination. In the recombinant mutant, a 1,7kb<br />

fragment containing the adenylation <strong>and</strong> part <strong>of</strong> the termination domain <strong>of</strong> module 2 is<br />

exchanged by the Km r -cassette. Modified after Mossialos (2003)<br />

48


Results<br />

While the function <strong>of</strong> module 1 is unclear, modules 2-4 are responsible for<br />

activation <strong>and</strong> condensation <strong>of</strong> glutamate, tyrosine, <strong>and</strong> diaminobutyric acid,<br />

respectively (Mossialos et al., 2002). It is well-known that the modular structure<br />

<strong>of</strong> NRPS allows some genetical modifications. If one <strong>of</strong> the modules is deleted<br />

completely the rest <strong>of</strong> the enzyme remains functional, producing a similar<br />

peptide just missing the respective amino acid residue (Finking & Marahiel,<br />

2004). Taking this into account, deletion <strong>of</strong> the complete adenylation domain<br />

<strong>and</strong> part <strong>of</strong> the condensation domain <strong>of</strong> module 2 in the marker exchange<br />

construct should block chromophore synthesis. We could identify a homologue<br />

<strong>of</strong> pvsA in the genome sequence <strong>of</strong> PssB728a.<br />

This strain is closely related to Pss22d <strong>and</strong> primers derived from the nucleotide<br />

sequence <strong>of</strong> PssB728a were used to amplify a PCR product <strong>of</strong> Pss22d. A first<br />

PCR fragment <strong>of</strong> 1.4 kb was amplified by primer pair pvsA_fwd1_Spe <strong>and</strong><br />

pvsA_rev6_Kpn. The amplificate was purified <strong>and</strong> subcloned in the T/A cloning<br />

vector pGEM-T Easy (Promega) resulting in plasmid pPVSA1. The second<br />

PCR fragment, 1.6 kb in size, was amplified by primers pvsA_fwd7_Kpn <strong>and</strong><br />

pvsA_rev4_Bam. It was purified <strong>and</strong> ligated into pGEM-T Easy, resulting in<br />

plasmid pPVSA2. Subsequently, the SpeI/KpnI fragment <strong>of</strong> pPVSA 1 was<br />

cleaved, purified <strong>and</strong> inserted into pPVSA2, resulting in plasmid pPVSA3.<br />

Finally, a 1.7 kb KpnI fragment <strong>of</strong> plasmid pMKm, carrying the kanamycin<br />

resistance cassette, was inserted into the unique KpnI site <strong>of</strong> pPVSA3. The<br />

resulting plasmid, pPVSA4, carries the entire marker exchange cassette<br />

consisting <strong>of</strong> the kanamycin cassette flanked by the two PCR fragments.<br />

pPVSA4 was transferred into Pss22d by electroporation. As the pGEM-T Easy<br />

vector backbone is not able to replicate in Pss22d, the plasmid acts as a<br />

suicide plasmid. Consequently, the resistance against kanamycin is transferred<br />

to Pss22d only if the plasmid recombines into the chromosome. If integration <strong>of</strong><br />

the resistance cassette occurs due to a single cross-over, the complete plasmid<br />

integrates <strong>and</strong> the transformants carry the resistance <strong>of</strong> the plasmid backbone<br />

(i.e. ampicillin), too.<br />

49


Results<br />

If integration is due to a homologues double cross over event, the plasmid<br />

backbone is lost. Thus the Pss22d transformants were screened for kanamycin<br />

resistance <strong>and</strong> the resulting c<strong>and</strong>idates were than screened for a lack <strong>of</strong><br />

ampicillin resistance. Finally, integration <strong>of</strong> the marker exchange cassette by<br />

double crossover was verified by Southern blot analysis (Fig.8). Plasmid<br />

pPVSA4 (control) <strong>and</strong> genomic DNA <strong>of</strong> wild type (wt) Pss22d <strong>and</strong> <strong>of</strong> 5 possible<br />

transformants were digested by ApaLI, separated by agarose gelelectrophoresis,<br />

transferred to a membrane <strong>and</strong> then hybridized at 60°C. Three<br />

separate blots were analyzed by hybridization with probes against the<br />

pGEM-T Easy vector backbone, the kanamycin cassette, <strong>and</strong> pvsA PCR<br />

fragment1. Transformant no. 5 showed the expected hybridization pattern for<br />

double cross-over insertion: There is no signal for the vector backbone, but<br />

there is a signal for the kanamycin cassette.<br />

probe:<br />

vector backbone kanamycin cassette<br />

pvsA fragment 1<br />

K M wt 1 2 3 4 5 K M wt 1 2 3 4 5 K M wt 1 2 3 4 5<br />

8<br />

5<br />

4<br />

3<br />

2<br />

1,6<br />

size (kb)<br />

Fig. 8. Verification <strong>of</strong> the Pss22d∆Pvd genotype by Southern blot analysis.<br />

Five different transconjugants were tested for correct insertion <strong>of</strong> the marker-exchange cassette by<br />

double-homologues recombination. Sample DNA was digested with ApaL1 <strong>and</strong> loaded as follows:<br />

K= control plasmid pPVSA4; M= 1kb molecular weight st<strong>and</strong>ard; wt= Pss22d; 1-5 = Transformants<br />

<strong>of</strong> Pss22d electroporated with pPVSA4 <strong>and</strong> screened for kanamycin resistance. Three separate<br />

blots were analysed by hybridisation with probes against the pGEM-T Easy®-vector backbone, the<br />

kanamycin cassette, <strong>and</strong> pvsA PCR fragment 1, respectively. The red boxes indicate the<br />

hybridisation pattern <strong>of</strong> Pss22d wild type <strong>and</strong> the double cross-over mutant no. 5.<br />

50


Results<br />

The signal for PCR fragment 1 is shifted as compared to Pss22d wt. This shift<br />

is due to the insertion <strong>of</strong> the resistance cassette <strong>and</strong> the loss <strong>of</strong> a chromosomal<br />

ApaLI site within the deleted sequence. Transformant no. 5 was designated<br />

Pss22d∆Pvd <strong>and</strong> further analyzed for its siderophore phenotype. As expected,<br />

Pss22d∆pvsA was no longer fluorescent when grown on King’s B medium<br />

indicating the deficiency <strong>of</strong> the strain to produce pyoverdin (Fig.9A). But<br />

surprisingly, it still produced a halo on CAS-agar almost as large as the wild<br />

type strain (Fig.9B). This halo clearly indicated that mutant Pss22d∆pvsA<br />

produces a functional siderophore although the pyoverdin biosynthesis <strong>of</strong> this<br />

strain disrupted successfully.<br />

5.1.2 Southern blot analysis <strong>of</strong> Psg1a <strong>and</strong> Pss22d for the presence <strong>of</strong><br />

yersiniabactin biosynthetic genes<br />

At the time <strong>of</strong> construction <strong>of</strong> Pss22d∆Pvd, the first genome sequence <strong>of</strong> a<br />

P. syringae was completed, namely the sequence <strong>of</strong> PsmDC3000. With the<br />

finished annotation <strong>of</strong> the genome project, the presence <strong>of</strong> a yersiniabactin<br />

biosynthesis cluster in PsmDC3000 was reported (Buell et al., 2003).<br />

Pss22d∆Pvd<br />

Pss22d<br />

A<br />

B<br />

Fig. 9. Siderophore production <strong>of</strong> Pss22d <strong>and</strong> Pss22d∆Pvd.<br />

A: Fluorescence <strong>of</strong> pyoverdin on Kings B medium; B: Siderophore production<br />

on CAS agar. Bacterial suspension was applied to the respective medium <strong>and</strong><br />

incubated at 28°C for 48h.<br />

51


Results<br />

Yersiniabactin is a siderophore first identified in Yersinia pestis but also<br />

reported for a number <strong>of</strong> other bacteria like for example uropathogenic E. coli<br />

(Schubert et al., 1998). BLAST analysis <strong>of</strong> the available sequence for<br />

PssB728a did not show any possible c<strong>and</strong>idates <strong>of</strong> yersiniabactin biosynthetic<br />

genes. Since it was unclear whether Pss22d produces yersiniabactin, a<br />

Southern blot analysis was performed. The NRPS encoded by the irp1 gene is<br />

necessary for yersiniabactin synthesis, thus a probe was derived against the<br />

irp1 homologue <strong>of</strong> DC3000 via PCR <strong>and</strong> used for hybridization <strong>of</strong> KpnI-digested<br />

genomic DNA <strong>of</strong> Psg1a, Pss22d, PssB728a <strong>and</strong> PsmDC3000. Hybridization<br />

temperature was 55°C. The only signal detected was that <strong>of</strong> the control,<br />

PsmDC3000 (Fig.10). Although we can not completely exclude the presence <strong>of</strong><br />

a yersiniabactin biosynthetic gene cluster in Psg <strong>and</strong> Pss, for the tested strains<br />

the presence <strong>of</strong> yersiniabactin biosynthetic genes is unlikely.<br />

1<br />

2 3 4<br />

10 kb<br />

Fig. 10. Screening for the yersiniabactin<br />

biosynthesis gene irp1 by southern blot<br />

analysis.<br />

KpnI digested genomic DNA <strong>of</strong> Psg1a (1),<br />

Pss22d (2), PssB728a (3) <strong>and</strong><br />

PsmDC3000 (4) was hybridised with a<br />

probe against the yersiniabactin NRPS<br />

irp1.<br />

52


Results<br />

5.1.3 Screening for an additional siderophore <strong>of</strong> Pss22d by r<strong>and</strong>om<br />

mutagenesis<br />

In order to identify biosynthetic genes responsible <strong>of</strong> the potential second<br />

siderophore <strong>of</strong> Pss22d, a r<strong>and</strong>om mutagenesis was conducted. For this we<br />

used a modified mini-Tn5 construct derived from pCAM140 (Wilson et al.,<br />

1995). The plasmid pCAM-Not was constructed by Dr. Helge Weingart via<br />

deletion <strong>of</strong> the gusA gene by NotI digest <strong>of</strong> pCAM140. R<strong>and</strong>om mutagenesis<br />

was performed by conjugating the plasmid pCAM-Not in both, Pss22d∆pvsA<br />

<strong>and</strong> Pss22d wt. The r<strong>and</strong>om character <strong>of</strong> transposon insertion-loci was verified<br />

by analysis <strong>of</strong> size distribution among the transposon-tagged fragments in<br />

Pss22d∆pvsA tansconjugants, as follows. Genomic DNA <strong>of</strong> 17 Tn5 mutants <strong>of</strong><br />

Pss22d∆Pvd was digested with PstI (Fig.11A) <strong>and</strong> ClaI (Fig.11B) <strong>and</strong><br />

hybridized with a probe against Tn5 at 60°C. The resulting distribution pattern<br />

did not show any redundancy <strong>of</strong> fragment sizes, thus it could be assumed that<br />

the mutagenesis had resulted in r<strong>and</strong>om Tn5 insertion.<br />

Size (kb)<br />

A<br />

B<br />

Fig. 11. Distribution <strong>of</strong> mini-Tn5 insertions in Pss22d∆Pvd.<br />

17 Tn5 mutants <strong>of</strong> Pss22d∆Pvd were analysed for their loci <strong>of</strong> transposon<br />

insertion. Genomic DNA was digested with PstI (A) <strong>and</strong> ClaI (B) <strong>and</strong> examined by<br />

Southern blot analysis for the size <strong>of</strong> the Tn5 insertion fragment. The probe was<br />

derived against the BamHI fragment <strong>of</strong> the transposon.<br />

53


Results<br />

Next, a large number <strong>of</strong> transconjugants was screened on CAS agar for<br />

siderophore-negative (Sid - ) phenotype (Fig.12). The colony size <strong>of</strong> the resulting<br />

mutants on the indicator plates had a remarkable influence on the size <strong>of</strong> the<br />

siderophore-derived halos. Therefore, all clones that did not produce a halo in<br />

the first screening where re-streaked on a CAS agar plate. Out <strong>of</strong> 500 Tn5-<br />

mutants <strong>of</strong> Pss22d∆Pvd, three siderophore-negative clones were obtained. In<br />

contrast <strong>and</strong> as expected, screening <strong>of</strong> 1500 Tn5-mutants derived from Pss22d<br />

wild type did not result in any siderophore-negative mutants.<br />

The loci <strong>of</strong> transposon insertions in the siderophore-negative mutants were<br />

subsequently identified by cloning. Genomic DNA <strong>of</strong> mutants Pss22d∆Sid,<br />

Pss22d∆Sid-23 <strong>and</strong> Pss22d∆Sid-25 was treated with the following restrictases:<br />

PstI, ClaI, XbaI, KpnI, SacI, SacII <strong>and</strong> SpeI. These digests were ligated into<br />

plasmid pBBR-Mcs1 cleaved with the respective restrictase. In order to select<br />

for ligation products containing the transposon, the ligation was transformed<br />

into E. coli DH5α <strong>and</strong> transformants were screened for spectinomycin<br />

resistance mediated by the mini-Tn5. Two plasmides were derived from this<br />

experiment: pBBR-11-Sid <strong>and</strong> pBBR-23-Sid, containing a PstI insert <strong>of</strong><br />

Pss22d∆Sid <strong>and</strong> Pss22d∆Sid-23, respectively. Insert sizes <strong>of</strong> both plasmids<br />

were approx. 6 kb. The plasmids were sent to MWG biotech for sequencing.<br />

Fig. 12. Screening <strong>of</strong> Tn5-mutants for<br />

siderophore-negative phenotype on CAS-agar.<br />

Transposon mutants <strong>of</strong> Pss22d∆Pvd were picked on<br />

CAS-agar plate <strong>and</strong> incubated for 48h.The white<br />

arrow indicates a sid - clone.<br />

54


Results<br />

For both plasmids a 500 bp sequence starting from the pBBR T7 primer binding<br />

site was obtained. The sequences were then used for BLAST analysis. Figure<br />

13 summarizes the result obtained for pBBR-11-Sid. For pBBR-23-Sid, the<br />

same sequence was obtained, indicating that the same PstI Fragment had<br />

been subcloned.<br />

i- mini-Tn5 -o<br />

2,8 kb 1,3 kb<br />

A<br />

acr<br />

acsF<br />

acsD<br />

acsC<br />

acsA<br />

acsE yhcA<br />

acsB<br />

cbrA-D<br />

B<br />

acr<br />

acsF<br />

acsD<br />

acsC acsA<br />

acsE yhcA acsB<br />

cbrA-D<br />

C<br />

acsF<br />

acr<br />

acsD<br />

acsC acsA<br />

acsE yhcA acsB<br />

cbrA-D<br />

2 kb<br />

Fig. 13. Location <strong>of</strong> mini-Tn5 insertion 11-Sid - in Pss22d∆Pvd.<br />

BLAST analysis <strong>of</strong> the insert sequence obtained for pBBR-11-Sid - showed significant<br />

similarities to the sequence <strong>of</strong> the achromobactin biosynthesis cluster in Dikeya chrysanthemi<br />

(C) <strong>and</strong> the two genome sequences <strong>of</strong> PssB728a (A) <strong>and</strong> Psp1448A (B). The part <strong>of</strong> the insert<br />

fragment that could be sequenced was located approx. 2.8 kb upstream <strong>of</strong> the transposon i-<br />

end <strong>and</strong> it aligned with the iucAC-like siderophore synthetase acsD. This indicated, that the<br />

transposon itself is inserted in the yhcA-permease. Comparison <strong>of</strong> the three clusters shows a<br />

high similarity in sequence <strong>and</strong> organisation <strong>of</strong> the genes, yet the siderophore receptor acr is<br />

arranged differently (indicated in red). ORF’s indicated in green could be part <strong>of</strong> the<br />

achromobactin cluster as well.<br />

55


Results<br />

Nevertheless the site <strong>of</strong> transposon insertion varies a little, as restriction<br />

analysis could demonstrate (data not shown). Blast analysis <strong>of</strong> the pBBR-11-<br />

Sid 500bp sequence showed significant sequence similarity to genes <strong>of</strong> the<br />

Dickeya chrysanthemi (formerly Erwinia chrysanthemi) achromobactin<br />

biosynthesis cluster (GenBank accession number: F416740 <strong>and</strong> AF416739;<br />

Franza, et. al. 2005). Best similarity was obtained for the acsD siderophore<br />

synthetase from within this gene cluster. Calculation <strong>of</strong> the respective fragment<br />

sizes indicated, that the localization <strong>of</strong> transposon insertion should be within the<br />

gene yhcA encoding for a permease.<br />

The insertion into yhcA was next confirmed by a PCR subcloning technique<br />

(Daria Zhurina, personal communication). By the time this sequence<br />

information was obtained, draft annotation <strong>of</strong> the genome sequence <strong>of</strong><br />

PssB728a became available at GenBank (www.ncbi.nlm.nih.gov/Taxonomy/<br />

taxonomyhome.html, TxID: 205918).<br />

In line with our results for Pss22d, the draft annotation indicated the presence<br />

<strong>of</strong> a second siderophore biosynthetic gene cluster in PssB728a. This cluster<br />

showed similar organization <strong>and</strong> high sequence identities to the achromobactin<br />

biosynthetic gene cluster from D. chrysanthemi. As the genome sequence <strong>of</strong><br />

P. syringae pv. phaseolicola 1448A became available, we could identify a<br />

similar siderophore biosynthesis cluster in this pathovar, too (Fig.13). There is<br />

no homologue <strong>of</strong> this cluster in strain PsmDC3000.<br />

Comparison <strong>of</strong> the biosynthesis cluster in the three positive strains allowed the<br />

following observations:<br />

1) All genes that have been assigned to be involved in achromobactin<br />

biosynthesis are present in every strain (Fig.13; grey arrows, genes<br />

labeled according to Franza, et. al. 2005).<br />

2) The aminotransferase acsF is located differently in D. chrysanthemi as<br />

compared to P. syringae (Fig.13). While it is located upstream <strong>of</strong> the<br />

56


Results<br />

siderophore receptor gene acr in D. chrysanthemi, it is downstream <strong>of</strong><br />

this ORF in the pseudomonads.<br />

3) Certain additional genes could potentially be part <strong>of</strong> the siderophore<br />

cluster (Fig.13 indicated in green). Following cbrD, there is a putative<br />

methyltransferase gene, conserved in all three strains. In both<br />

pseudomonads a putative regulator containing a sensory domain with<br />

week similarity to chemotaxis/aerotaxis <strong>and</strong> a regulator containing a<br />

GGDEF domain involved in cyclic-diGMP signaling are present directly<br />

after the siderophore cluster. Additionally, in both P. syringae strains two<br />

putative regulatory genes are located in front <strong>of</strong> the siderophore<br />

receptor. The first is a homologue <strong>of</strong> the alternative sigma factor FecI,<br />

involved in ferric citrate uptake. The second shows similarities to the<br />

regulator FecR that interacts with FecI (Visca et al., 2002).<br />

As Pss22d∆Sid is the best characterized <strong>of</strong> the three transposon mutants, this<br />

derivation was used for further analysis.<br />

5.1.4 Construction <strong>of</strong> a marker exchange mutant <strong>of</strong> the achromobactinlike<br />

siderophore in Pss22d<br />

In order to address the impact <strong>of</strong> the achromobactin-like siderophore on the<br />

phenotype <strong>of</strong> Pss22d, a non-polar marker exchange mutant, deficient for the<br />

biosynthesis <strong>of</strong> this siderophore was constructed. The achromobactin<br />

biosynthesis cluster encodes for three copies <strong>of</strong> a siderophore synthetase,<br />

namely AcsA, AcsC, <strong>and</strong> AscD. These synthetases are supposed to be<br />

responsible for the amide bond formation connecting the achromobactin<br />

precursor molecules (Fig.14). The first <strong>of</strong> the three corresponding genes is<br />

acsD. In mutant Pss22d∆Ach, this ORF has been exchanged against a<br />

kanamycin cassette (Fig.15). For the construction <strong>of</strong> this mutant, a 1.3-kb PCR<br />

fragment was amplified by PCR using the primer combination Achr1_fwd <strong>and</strong><br />

Achr2_KpnI_rev. The PCR fragment was purified <strong>and</strong> ligated into the vector<br />

pGemT-easy. The resulting plasmid pACS1 was selected for orientation <strong>of</strong> the<br />

57


Results<br />

Achr2_KpnI_rev binding site next to the SpeI recognition site <strong>of</strong> the vector. A<br />

second PCR fragment <strong>of</strong> 2 kb was amplified by the primer pair<br />

Achr5_EcoRV_fwd <strong>and</strong> Achr6_rev. The PCR product was purified <strong>and</strong> brought<br />

into pGemT-easy, too. This plasmid was designated pACS2. The 1.7-kb PstI<br />

fragment <strong>of</strong> pMKm was purified <strong>and</strong> blunt-ended by application <strong>of</strong> Klenow<br />

fragment. Plasmid pACS1 was linearized by SpeI, blunt-ended <strong>and</strong> fused with<br />

the kanamycin resistance cassette by ligation.<br />

Fig. 14. Structure <strong>of</strong> the siderophore achromobactin.<br />

The oxo-glutaric acid side chains <strong>of</strong> achromobactin can cyclize. The structure shown on the right<br />

prevails under neutral conditions. The siderophore synthetases AcsA, AcsC <strong>and</strong> AcsD are<br />

supposed to form the amide bonds present in achromobactin. Modified after Franza, et al.,<br />

(2005).<br />

The resulting plasmid carrying a 3 kb insert was designated pACS3. This<br />

plasmid was linearized by the restrictase ScaI <strong>and</strong> afterwards treated with<br />

EcoRI, resulting in a 3-kb fragment <strong>of</strong> the insert <strong>and</strong> parts <strong>of</strong> the vector<br />

backbone. The 3-kb fragment was purified, blunt ended <strong>and</strong> ligated into the<br />

EcoRV-opened plasmid pACS2. Correct orientation <strong>of</strong> the fragments to each<br />

other, resulting in the kanamycin cassette being flanked by both PCR<br />

fragmentes was verified by restriction digest with the restrictase PstI. As PCR<br />

fragment 1 carries an internal recognition site for this enzyme in addition to a<br />

single recognition site in the vector sequence, the fragment pattern <strong>of</strong> a plasmid<br />

with insert in correct orientation could be unambiguously identified.<br />

58


Results<br />

Fig. 15. Construction <strong>of</strong> the marker exchange mutant Pss22d∆Ach.<br />

acsD is the first <strong>of</strong> three genes encoding for a siderophore synthetase. In mutant Pss22d∆Ach,<br />

this ORF has been exchanged by a kanamycin cassette. For construction <strong>of</strong> this mutant, a 1,3 kb<br />

PCR fragment was amplified by the primers Achr1_fwd (1) <strong>and</strong> Achr2_KpnI_rev. A second PCR<br />

fragment, 2 kb in size, was amplified by the primer pair Achr5_EcoRV_fwd (4) <strong>and</strong> Achr6_rev (4).<br />

The PCR fragments were subsequently fused to a 1.7-kb fragment containing a kanamycin<br />

resistance cassette <strong>and</strong> incorporated into the chromosome by double homologous recombination.<br />

Dashed arrow indicate the flanking genes acr <strong>and</strong> acsA.<br />

The resulting plasmid, carrying the complete marker exchange construct was<br />

designated pACS4. This plasmid was transferred into Pss22d by<br />

electroporation <strong>and</strong> the resulting tranformants were screened for kanamycin<br />

resistance <strong>and</strong> absence <strong>of</strong> ampicillin resistance. The correct genotype <strong>of</strong> the<br />

c<strong>and</strong>idate transformantes was verified by PCR (data not show). The resulting<br />

mutant was designated Pss22d∆Ach.<br />

5.1.5 In vitro comparison <strong>of</strong> Pss22d <strong>and</strong> its siderophore mutants<br />

The phenotype <strong>of</strong> the Pss22d <strong>and</strong> the single siderophore mutants was<br />

compared on CAS agar (Fig.16).<br />

A<br />

B<br />

C<br />

D<br />

Figure 16. Siderophore production on CAS-agar.<br />

Bacterial suspensions <strong>of</strong> A: Pss22d; B: Pss22d∆Ach;<br />

C: Pss22d∆Sid; <strong>and</strong> D: Pss22d∆Pvd were applied on<br />

a CAS-agar plate <strong>and</strong> incubated at 28°C for 48h.<br />

59


Results<br />

Interestingly the mutant Pss22d∆Pvd shows an as large siderophore halo as<br />

Pss22d wt, while the halo <strong>of</strong> Pss22d∆Ach is reduced.The siderophore negative<br />

mutant Pss22d∆Sid, produced no halo as indicated previously. In order to<br />

investigate this phenotypic difference between Pss22d∆Pvd <strong>and</strong> Pss22d∆Ach<br />

in more detail, siderophore production <strong>of</strong> Pss22d <strong>and</strong> its siderophore mutants<br />

was analysed during in vitro growth in iron free Pipes medium (Fig.17; Tab. 9).<br />

The two single mutants Pss22d∆Pvd <strong>and</strong> Pss22d∆Ach grew similar. Compared<br />

to the wild type, they grew a little bit slower, but the difference was only minor.<br />

The double mutant <strong>and</strong> Pss22d∆Sid grew considerably slower than the wild<br />

type, yet it reached the same OD 600 . The siderophore production during growth<br />

in liquid was analyzed as well (Tab. 9).In contrast to their similar growth,<br />

analysis <strong>of</strong> siderophore production showed a clear difference between both<br />

single mutants. While Pss22d∆Ach did produce as much siderophore activity<br />

<strong>and</strong> fluorescence as the wild type, mutant Pss22d∆Pvd produced a more than<br />

5 times higher siderophore activity.<br />

10<br />

OD600<br />

1<br />

0,100<br />

0,010<br />

0 5 10 15 20 25 30 35 40 45<br />

time (h)<br />

1<br />

2<br />

3<br />

4<br />

Fig. 17. In vitro growth <strong>of</strong><br />

Pss22d <strong>and</strong> its respective<br />

siderophore mutants.<br />

Growth in Pipes medium<br />

under iron-limiting conditions<br />

was compared between<br />

Pss22d wt (1), Pss22d∆Pvd<br />

(2), Pss22d∆Ach (3), <strong>and</strong><br />

Pss22d∆Sid (4). Both single<br />

mutants grew similar, only<br />

slightly slower than the wild<br />

type. The double mutant<br />

Pss22d∆Sid, instead, shows<br />

a significantelly slower<br />

growth.<br />

60


Results<br />

Tab. 9. Siderophore Production <strong>of</strong> Pss22d <strong>and</strong> its mutants.<br />

Siderophore production <strong>and</strong> fluorescence <strong>of</strong> Pss22d <strong>and</strong> the derived<br />

siderophore mutants after 48h <strong>of</strong> growth in Pipes medium. Values are<br />

normalized to OD 600 =1.<br />

µM DFOM<br />

equiv.<br />

fluorescence<br />

[Pvd ps ] in<br />

µg/ml<br />

Pss22d 16 +/- 0 30 +/- 6<br />

P ss22d∆Pvd 107 +/- 0 1 +/- 0<br />

P ss22d∆Ach 20 +/- 0 33 +/- 0<br />

P ss22d∆Sid 0 1 +/- 0<br />

5.1.6 Purification <strong>of</strong> the achromobactin-like siderophore for structure<br />

determination<br />

Analysis <strong>of</strong> the Pss22d∆Sid transposon insertion allowed the conclusion, that<br />

Pss22d produces siderophore similar to achromobactin. Whether this<br />

siderophore is identical to achromobactin (Fig.14) or just similar to it remained<br />

to be determined by structural analysis. NMR analysis is currently been<br />

performed in cooperation with the laboratory <strong>of</strong> Pr<strong>of</strong>. Budzikiewicz, <strong>University</strong> <strong>of</strong><br />

Köln. This laboratory already determined the structure <strong>of</strong> authentic<br />

achromobactin from D. chrysanthemi (Münzinger et al., 2000).<br />

To obtain an NMR-suitable sample <strong>of</strong> the achromobactin-like siderophore <strong>of</strong><br />

Pss22d, this siderophore had to be extracted from the culture supernatant <strong>and</strong><br />

next had to be separated from other secondary metabolites, cell debris <strong>and</strong><br />

residual medium components. For this, column chromatographic methods like<br />

gel-filtration or HPLC are applied frequently. For both methods a concentration<br />

<strong>of</strong> the sample volume prior to applications was necessary. To do so, affinity<br />

chromatography <strong>of</strong>fered the advantage that a large sample volume could be<br />

dealt with at one time without preceding concentration steps. XAD4-resin<br />

(Sigma) had been successfully applied for affinity chromatography <strong>of</strong> pyoverdin<br />

(Jülich et al., 2001), <strong>and</strong> we could successfully modify this method for a first<br />

concentration <strong>of</strong> the achromobactin-like siderophore <strong>of</strong> strain Pss22d∆Pvd.<br />

61


Results<br />

XAD4-purification yielded in a yellow-brownish mixture <strong>of</strong> substances; thus it<br />

had to be further cleaned. The following three gel-filtrations using different<br />

materials (Fractogel TSK HW 50s, Merck; Bio-Gel P-30, Biorad; <strong>and</strong> G25<br />

Sephadex, Amersham) were conducted, resulting in an almost colorless<br />

sample. Interestingly, upon addition <strong>of</strong> iron the color <strong>of</strong> the sample changed to<br />

yellow. This is remarkable, as the D. chrysanthemi achromobactin was named<br />

so because the iron-siderophore complex did stay colorless (Münzinger et al.,<br />

2000). Unfortunately, the sample obtained after gel-filtration still contained<br />

interfering substances <strong>and</strong> the resulting siderophore concentration was rather<br />

low. Thus it was not suitable for NMR (Pr<strong>of</strong>. Budzikiewicz, cologne, personal<br />

communication). Therefore other methods were tested. HPLC separation on a<br />

C18-RP sherisorp column did result in separation <strong>of</strong> several substances, but no<br />

siderophore activity could be determined after the HPLC run (data not shown).<br />

Next, separation <strong>of</strong> the sample by thin layer chromatography (TLC) was<br />

conducted. Compared to column chromatography, TLC allows a rather fast<br />

optimization <strong>of</strong> separation conditions, as it circumvents collection <strong>and</strong> testing <strong>of</strong><br />

numerous fractions per separation step. Theoretical models on the separation<br />

characteristics <strong>of</strong> mobile phase mixtures allow a targeted modification <strong>of</strong><br />

solvent mixtures (Nyiredy, 2004; Snyder, 1978). After successful optimization <strong>of</strong><br />

the separation strategy, preparative TLC can be utilized or the running<br />

conditions can be transferred back to column chromatography. TLC separation<br />

<strong>of</strong> XAD4-purified supernatant <strong>of</strong> Pss22d∆Pvd was performed on silica plates<br />

(Macherey-Nagel) in combination with various mobile phases. Separation <strong>of</strong> the<br />

sample was investigated by aut<strong>of</strong>luorescence <strong>of</strong> the substances at an excitation<br />

wavelength <strong>of</strong> λ=360nm or quenching <strong>of</strong> a fluorescence indicator within the<br />

silica matrix (excitation λ=254nm). The position <strong>of</strong> the achromobactin-like<br />

siderophore was determined by CAS-overlay. A first TLC-separation <strong>of</strong><br />

Pss22d∆Pvd XAD4-purification using an equivolurimetric mixture <strong>of</strong><br />

diethylether, n-butanole, acetic acid, <strong>and</strong> water resulted in the separation <strong>of</strong><br />

several fluorescent substances (Fig.18).<br />

62


Results<br />

Fig. 18. Seperation <strong>of</strong> the achromobactin-containing XAD4<br />

purificationS by TLC.<br />

XAD4 purified supernatant <strong>of</strong> Pss22d∆Pvd separated by TLC on silica<br />

plate (mobile phase: Diethylether : n-butanole: acetic acid: water; ratio<br />

1:1:1:1; v/v). Fluorescence <strong>of</strong> the sample was visualizes by UV 365nm .<br />

Drawn line at the bottom indicates the start <strong>and</strong> the uppermost fluorescent<br />

line represents the running front.<br />

Separations <strong>of</strong> the Pss22d∆Pvd XAD4-purification were compared to that <strong>of</strong> the<br />

Pss22d∆Sid <strong>and</strong> a control achromobactin sample purified from<br />

D. chrysanthemi cbsE-1 tonB60 (Fig.19). In all three samples the majority <strong>of</strong><br />

fluorescent components migrated to more than half <strong>of</strong> the running front. The<br />

CAS-overlay <strong>of</strong> the TLC plates showed a positive reaction for authentic<br />

achromobactin <strong>and</strong> the Pss22d∆Pvd XAD4-purification (Fig19A).<br />

In both samples the CAS-active substance had moved only slightly (retention<br />

factor <strong>of</strong> 0.25). As expected this b<strong>and</strong> was absent in the siderophore-deficient<br />

strain. No fluorescent b<strong>and</strong> could be associated to the CAS-active region on the<br />

plate, indicating that neither the achromobactin-like siderophore nor authentic<br />

achromobactin were visible at the wavelength tested.<br />

A<br />

B<br />

←<br />

1 2 3 1 2 3<br />

Fig. 19. CAS overlay <strong>of</strong> TLC separated XAD4 purification.<br />

XAD4 purified supernatant separated by TLC on silica plate with different<br />

mobile phases: A) Diethylether : n-butanole: acetic acid: water; ratio 1:1:1:1;<br />

(v/v). B) MTBE : Methanol: acetic acid : water; ratio 3:1:1:1 (v:v).<br />

Samples: 1) Pss22d∆Pvd 2) Pss22d∆Sid 3) Achromobactin purified from<br />

D. chrysanthemi cbsE-1 tonB60. The arrow indicates CAS-active b<strong>and</strong> <strong>of</strong><br />

Pss22d∆Pvd, the less intensive signal below this b<strong>and</strong> is due to long<br />

incubation time <strong>of</strong> the overlay.<br />

63


Results<br />

After changing the mobile phase to methyl t-buthyl ether (MTBE), methanol,<br />

acetic acid <strong>and</strong> water (3:1:1:1, v/v) the retention factor <strong>of</strong> the CAS-active<br />

substance was increased to 0.45 (Fig.19B). Interestingly, the retention factor for<br />

authentic achromobactin differs from that (0.35). The smearing <strong>of</strong> the CASactive<br />

b<strong>and</strong> <strong>of</strong> Pss22d∆Pvd (Fig.19B) was likely due to a relatively long<br />

incubation <strong>of</strong> the overlay in order to visualize a second signal above the major<br />

CAS-active b<strong>and</strong>. This additional signal could be observed in both siderophorepositive<br />

samples <strong>and</strong> may correspond to a second conformation <strong>of</strong> the<br />

siderophore.<br />

The original CAS-active b<strong>and</strong> <strong>of</strong> Pss22d∆Pvd corresponds to the more<br />

intensive spot indicated by the arrow (Fig.19B). Approximately the same<br />

amount <strong>of</strong> XAD4-purification was applied for both strains <strong>and</strong> the intensity <strong>of</strong><br />

the different fluorescent b<strong>and</strong>s did not differ significantly between both lanes. In<br />

contrast, the siderophore activity detected for D. chrysanthemi cbsE-1 tonB60<br />

was much lower as compared to Pss22d∆Pvd.<br />

The modified separation conditions are currently applied on the next large-scale<br />

purification. After this step, another TLC-separation using a different stationary<br />

phase will be used for further purification. Macherey-Nagel <strong>of</strong>fers a range <strong>of</strong><br />

differently modified silica plates with varying hydrophobicities <strong>of</strong> the matrices.<br />

The affinity <strong>of</strong> the CAS-active substance to silica modified by addition <strong>of</strong> a diolor<br />

an amino group was tested (Fig.20). The affinity <strong>of</strong> the CAS-active substance<br />

towards the stationary phase was lower for the diol modification <strong>and</strong> higher for<br />

the amino modified plate. Thus the diol modified plate could be optimized for<br />

better focusing <strong>of</strong> the CAS-active b<strong>and</strong>, while the amino-modified phase could<br />

be used to separate other disturbing substances.<br />

64


Results<br />

silica diol am ino silica diol amino<br />

Fig. 20. Comparison <strong>of</strong> different silica matrixes.<br />

XAD4-purified supernatant separated by TLC with different stationary phases: Silica=normal Silica<br />

60; diol= silica with an added diol group; amino= silica with an added amino group; left side shows<br />

fluorescence scane with the phosphoimager while the right side shows the respective CAS-overlay.<br />

Mobile phase was diethylether : n-butanole: acetic acid: water; ratio 1:1:1:1; v/v.<br />

Samples from left to right: Pss22d∆Sid; Pss22d∆Pvd; D. chrysanthemi cbsE-1 tonB60<br />

5.2. Comparison <strong>of</strong> siderophore production in Psg1a <strong>and</strong> Pss22d<br />

BLAST analysis had demonstrated the presence <strong>of</strong> the achromobactin-like<br />

gene cluster in two <strong>of</strong> the three sequenced P. syringae strains <strong>and</strong> in Pss22d,<br />

but what about the pathogen Psg1a? To address this question, XAD4-<br />

concentrated culture supernatants from Psg1a, Pss22d, Pss22d∆Pvd, <strong>and</strong><br />

Pss22d∆Sid were compared by isoelectric focussing (IEF). IEF is generally<br />

applied to separate proteins or peptides on a polyacrylamid gel according to<br />

their isoelectric point, but it has also been successfully applied to investigate<br />

siderophore patterns among different bacteria (Fuchs et al., 2001). For this,<br />

siderophore-containing samples were separated on a polyacryamidgel <strong>and</strong><br />

subsequently visualized by CAS-overlay. Figure 21 shows results <strong>of</strong> such an<br />

experiment. Pvd Ps (courtesy <strong>of</strong> Pr<strong>of</strong>. J.M. Meyer, Strasbourg, France) served as<br />

control. Pss22d <strong>and</strong> Psg1a produced the same b<strong>and</strong> patterns consisting <strong>of</strong> two<br />

pyoverdin b<strong>and</strong>s <strong>and</strong> an additional b<strong>and</strong> corresponding to the achromobactinlike<br />

siderophore. So, despite their phenotypic difference on CAS agar plate,<br />

both strains showed the same siderophore pattern.<br />

65


Results<br />

P<br />

P<br />

A<br />

Fig. 21. Isoelectic focusing <strong>and</strong> CAS-overlay detection <strong>of</strong> purified siderophores.<br />

Lanes: A: reference pyoverdin <strong>of</strong> P. syringae; B: Psg1a; C: Pss22d; D: Pss22d∆Pvd;<br />

E: Pss22d∆Sid; P= pyoverdin; A= achromobactin-like siderophore<br />

Siderophore production <strong>of</strong> antagonist <strong>and</strong> pathogen were analysed in more detail<br />

during growth in iron-free pipes medium (Fig.22). One aim <strong>of</strong> this analysis was to<br />

distinguish between pyoverdin production <strong>and</strong> production <strong>of</strong> the achromobactinlike<br />

siderophore. As demonstrated before, there is no fluorescence signal in<br />

pyoverdin-negative Pss22d strains <strong>and</strong> the siderophore-negative mutant does<br />

not show a reaction in the CAS-assay (Tab. 9). Thus comparison between CASmeasurement<br />

<strong>and</strong> fluorescence determination was used to differentiate between<br />

total siderophore production (= achromobactin-like siderophore plus pyoverdin)<br />

<strong>and</strong> pyoverdin production.<br />

In initial experiments, 250 ml <strong>of</strong> culture were incubated in a 1 L Erlenmeyer flask<br />

(Fig.22A). The antagonist Pss22d reached stationary phase within 30 h at an<br />

OD 600 <strong>of</strong> 2. Psg1a showed a slower growth <strong>and</strong> reached the stationary phase not<br />

before 60 h, yet it reached the same OD 600 . In contrast to our expectations, the<br />

siderophore activity produced by the antagonist Pss22d was significantly (about<br />

10 times) lower than that <strong>of</strong> the pathogen Psg1a. While the activity <strong>of</strong> Psg1a was<br />

in line with previous results (Fig.5B), siderophore production in Pss22d was<br />

66


Results<br />

A<br />

B<br />

Fig. 22. Comparison <strong>of</strong> growth, siderophore production <strong>and</strong> fluorescence <strong>of</strong> Pss22d <strong>and</strong> Psg1a.<br />

Siderophore production <strong>and</strong> fluorescence <strong>of</strong> antagonist Pss22d <strong>and</strong> pathogen Psg1a were compared<br />

during in vitro growth in iron-depleted Pipes medium. The strains were incubated either in 250 ml culture<br />

volume in 1 L Erlenmeyer flasks (A) or in 250 ml culture volume in 2 L Erlenmeyer flasks (B).<br />

67


Results<br />

reduced in comparison to these prior experiments. As the production <strong>of</strong><br />

siderophores is known to be tightly regulated, we next analyzed potential<br />

regulatory parameters in order to reconstitute the original phenotype. Indeed,<br />

siderophore production <strong>of</strong> Pss22d could be increased by addition <strong>of</strong> the trace<br />

elements Zn 2+ , Cu 2+ , <strong>and</strong> Ca 2+ . This supplementation did not influence<br />

siderophore production in Psg1a, but the CAS activity in Pss22d supernatants<br />

was increased to approx. 150 µM DFOM equiv. at OD 600 =1. As the first<br />

measurements were performed in the lab <strong>of</strong> our cooperation partner Dr. Beate<br />

Völksch in Jena, chemicals from other commercial suppliers had been used to<br />

prepare the medium. Thus a slight variation in medium composition due to<br />

variable trace element content has an impact on the quantity <strong>of</strong> siderophores<br />

produced. To investigate this hypothesis, Pipes medium was prepared with<br />

chemicals <strong>and</strong> water supplied by the laboratory in Jena. Doing so, the<br />

siderophore activity produced by Pss22d could be slightly increased; under the<br />

conditions approximately 200 µM DFOM equiv. an OD 600 =1 were produced by<br />

Pss22d.<br />

Finally, the influence <strong>of</strong> culture volume on siderophore production was<br />

investigated. Again, the total siderophore activity produced by Psg1a did not<br />

change, but the response <strong>of</strong> Pss22d to different ratios <strong>of</strong> culture vessel to culture<br />

volume became obvious. 250 ml <strong>of</strong> culture volume were incubated, but this time<br />

a 2 L Erlenmeyer flask as used instead <strong>of</strong> a 1 L-flask (Fig.22B). Under these<br />

conditions, the original phenotype could be reconstituted. At the same time,<br />

growth <strong>of</strong> Pss22d was delayed, indicating a possible stress response. The two<br />

growth curves gave an interesting insight into regulation <strong>of</strong> siderophore<br />

biosynthesis in Psg1a <strong>and</strong> Pss22d. Under conditions <strong>of</strong> lower oxygen supply<br />

(250 ml culture in a 1 L-Erlenmeyer flask), Pss22d produced lower amounts <strong>of</strong><br />

siderophores but relatively high fluorescence. Under high oxygen conditions<br />

(250 ml culture in 2 L-Erlenmeyer flask), Pss22d produced approximately the<br />

same amount <strong>of</strong> fluorescence but more than 10 times the siderophore activity.<br />

68


Results<br />

This result indicated an up-regulation <strong>of</strong> the achromobactin-like siderophore<br />

under high oxygen conditions. The situation in Psg1a differs from this<br />

considerably. In Psg1a, the total amount <strong>of</strong> siderophore production was not<br />

influenced by oxygen supply. However the amount <strong>of</strong> fluorescence differed:<br />

Under low oxygen conditions, less fluorescence was produced than under high<br />

oxygen conditions.<br />

These results indicated a differential regulation <strong>of</strong><br />

siderophore production in Psg1a <strong>and</strong> Pss22.<br />

5.3 Influence <strong>of</strong> siderophore production on biocontrol activity <strong>and</strong> epiphytic<br />

fitness <strong>of</strong> Pss22d<br />

In order to study the influence <strong>of</strong> siderophore production on the antagonism<br />

between Pss22d <strong>and</strong> Psg1a, the siderophore mutants <strong>of</strong> Pss22d were tested for<br />

their ability to suppress growth (<strong>and</strong> disease development) <strong>of</strong> Psg1a in planta.<br />

Bacteria were applied by wound inoculation technique. After punching soybean<br />

leaves with a sterile needle, 5 µl <strong>of</strong> bacterial suspension were applied to the fresh<br />

wounds <strong>and</strong> allowed to dry in. The inoculations were assigned as follows<br />

(Fig.23):<br />

A B C D E F G H I<br />

Fig. 23. Symptom development 14 days after wound inoculation.<br />

After punching soybean leaves with a sterile needle, 5 µl <strong>of</strong> bacterial suspension were applied to the<br />

fresh wounds <strong>and</strong> allowed to dry in. 14 days later the symptom development was assessed. Inoculations<br />

were as follows: A-E single inoculations with A: Psg1a; B: Pss22d; C: Pss22d∆Pvd; D: Pss22d∆Ach;<br />

E: Pss22d∆Sid; F-I Co-inoculation <strong>of</strong> Psg1a in combination with F: Pss22d; G: Pss22d∆Pvd;<br />

H: Pss22d∆Ach; I: Pss22d∆Sid<br />

69


Results<br />

A-E represents single inoculations with A: Psg1a; B: Pss22d; C: Pss22d∆Pvd; D:<br />

Pss22d∆Ach, <strong>and</strong> E: Pss22d∆Sid. F-I showed the results for co-inoculation<br />

experiments <strong>of</strong> Psg1a in combination with F: Pss22d; G: Pss22d∆Pvd; H:<br />

Pss22d∆Ach; I: Pss22d∆Sid. In the co-inoculation experiments, antagonist <strong>and</strong><br />

pathogen were mixed at a ratio <strong>of</strong> 2:1. Symptom development was investigated<br />

14 days post inoculation (Fig.23). Only when a single inoculation <strong>of</strong> Psg1a was<br />

applied, a clear formation <strong>of</strong> chlorosis could be observed (Fig.23A). Wounds<br />

inoculated with the antagonistic strains alone symptom-free (Fig.23B-E). Wounds<br />

co-inoculated with antagonist <strong>and</strong> pathogen showed mostly no symptoms, while<br />

on some single wounds a strongly reduced symptom development could be<br />

observed (for example Fig.23G, uppermost wound).<br />

Determination <strong>of</strong> CFU per wound did not give any evidence for a different<br />

performance <strong>of</strong> Pss22d wild type <strong>and</strong> its siderophore mutants. Therefore data<br />

were grouped in Figure 24. In the co-inoculation experiments, pathogen<br />

population size was reduced by one order <strong>of</strong> magnitude as compared to the<br />

single inoculation <strong>of</strong> Psg1a. Since wound inoculation technique could potentially<br />

interfere with the impact <strong>of</strong> siderophores due to a possible iron leakage from<br />

wound exudates, the in planta epiphytical fitness <strong>of</strong> Pss22d <strong>and</strong> its siderophore<br />

mutants was investigated in a spray inoculation experiment. Preliminary results<br />

indicated a reduction <strong>of</strong> population size for the siderophore mutants as compared<br />

to the wild type (Tab. 10). While both single mutants reached a cell density one<br />

order <strong>of</strong> magnitude below the wild type, the double mutant was reduced by two<br />

orders <strong>of</strong> magnitude. A repetition <strong>of</strong> this experiment showed a very similar<br />

outcome for Pss22d∆Pvd; however, growth <strong>of</strong> Pss22d∆Sid was reduced only by<br />

one order <strong>of</strong> magnitude <strong>and</strong> that <strong>of</strong> Pss22d∆Ach was not affected at all.<br />

70


Results<br />

1x10 8<br />

1x10 7<br />

1x10 6<br />

1x10 5<br />

1x10 4<br />

1x10 3<br />

1x10 2<br />

Fig. 24. Development <strong>of</strong> bacterial population in planta after wound inoculation.<br />

The formation <strong>of</strong> bacterial populations following wound inoculation was examined. The<br />

inoculation that is shown in Fig. 23: A-E single inoculations A: Psg1a; B: Pss22d; C:<br />

Pss22d∆Pvd; D: Pss22d∆Ach; E: Pss22d∆Sid<br />

F-I Co-inoculation <strong>of</strong> Psg1a in combination with F: Pss22d; G: Pss22d∆Pvd; H:<br />

Pss22d∆Ach; I: Pss22d∆Sid<br />

Series “Psg1a in co-inoculation” represents the mean value <strong>of</strong> the Psg1a counts for<br />

inoculation F-I. Likewise, the series “Pss22d <strong>and</strong> mutants in co-inoculation” represents the<br />

values determined for the antagonists in samples F-I. The series “Pss22d <strong>and</strong> mutants”<br />

comprises the mean value <strong>of</strong> antagonist cfu in single inoculations B-E.<br />

As the st<strong>and</strong>ard derivation <strong>of</strong> these series indicates, there was no difference for in planta<br />

growth <strong>of</strong> Pss22d wt <strong>and</strong> the respective siderophore mutants neither in single nor in coinoculation.<br />

5.4 Distribution <strong>of</strong> the achromobactin–like siderophore synthesis among<br />

plant associated P. syringae<br />

It was demonstrated that the achromobactin-like siderophore is produced by<br />

Pss22d, Psg1a, Psp1448A <strong>and</strong> PssB728a, but not by PsmDC3000. In order to<br />

elucidate the importance <strong>of</strong> this siderophore for the bacteria – plant interaction<br />

a screening among a diverse group <strong>of</strong> plant associated P. syringae was<br />

performed. The method <strong>of</strong> colony-PCR was used for fast detection <strong>of</strong> the<br />

respective biosynthesis genes. Two sets <strong>of</strong> primers were derived from the<br />

nucleotide sequence <strong>of</strong> the siderophore synthetase genes acsD <strong>and</strong> acsC.<br />

71


Results<br />

Tab. 10. Development <strong>of</strong> population sizes after spray inoculation.<br />

Pss22d wild type <strong>and</strong> its siderophore mutants were analysed for their<br />

growth on soybean leafs after application by a single spray inoculation.<br />

cfu/g fresh material<br />

day 1 day 10<br />

Pss22d 5 x 10 4 6.7 x 10 4<br />

Pss22d∆Pvd 1.3 x 10 4 5.4 x 10 3<br />

Pss22d∆Ach 8.9 x 10 4 4.6 x 10 3<br />

Pss22d∆Sid 9.2 x 10 4 4 x 10 2<br />

Primer binding sites were highly conserved in PssB728a, Psp1448A, <strong>and</strong><br />

D. chysanthemi. A 550 bp fragment <strong>of</strong> the ascC gene was amplified by the<br />

primer pair iuc2_fwd <strong>and</strong> iuc2_rev (Fig.25). In case <strong>of</strong> a negative PCR result,<br />

colony PCR was repeated two more times. In case <strong>of</strong> a negative result, a<br />

second PCR using the Primer combination iuc1_fwd <strong>and</strong> iuc1_rev was<br />

performed. The specificity <strong>of</strong> the applied PCR program was controlled by<br />

X S D Pss 1 2 3 4 5 6 DC<br />

Fig. 25. PCR screening for achromobactin biosynthesis genes.<br />

Colony PCRs with the primer combination iuc2_fwd <strong>and</strong> iuc2_rev was performed on<br />

different strains <strong>of</strong> P. syringae <strong>and</strong> control strains. Abbreviations are as follows:<br />

X= X. campestris; S= S. meliloti; D= D. chrysanthemi; Pss= Pss22d; 1-6 different<br />

pathovars <strong>of</strong> P.syringae. DC=PsmDC3000. While sample 1 <strong>and</strong> 4 show a clear<br />

positive reaction, the PCR for reactions like in sample 2, 3, <strong>and</strong> 5 were repeated.<br />

72


Results<br />

colony PCR with cells <strong>of</strong> Sinorhizobium meliloti <strong>and</strong> Xanthomonas campestris.<br />

Both strains are known to contain similar siderophore synthetases but not the<br />

achromobactin biosynthetic gene cluster. While no PCR product was obtained<br />

for S. meliloti, a PCR fragment significantly smaller than the expected product<br />

was detected in X. campestris. A colony PCR <strong>of</strong> PsmDC3000 served as a<br />

negative control for each PCR mix, while cells <strong>of</strong> Pss22d <strong>and</strong> D. chrysanthemi<br />

serve as positive control. Figure 25 gives an example for a screening gel. In<br />

addition, the production <strong>of</strong> the achromobactin-like siderophore was investigated<br />

by IEF analysis, as already shown in Figure 21. Table 11 summarizes the<br />

preliminary results <strong>of</strong> both screening methods. So far, the presence <strong>of</strong> the<br />

achromobactin-like siderophore could be demonstrated for 19 different<br />

pathovars <strong>of</strong> P. syringae, indicating that the biosynthesis <strong>of</strong> the achromobactinlike<br />

siderophore is widespread among different pathovars <strong>of</strong> P. syringae.<br />

73


Results<br />

Tab. 11. Distribution <strong>of</strong> the achromobactin-like siderophore among different pathovars <strong>of</strong><br />

P. syringae.<br />

P. syringae strains were analyzed for their ability to produce the acromobactin-like siderophore<br />

(IEF analysis). Additionally the presence <strong>of</strong> the siderophore biosynthesis gene acsB was<br />

determined by PCR with the primer pair iuc2_fwd <strong>and</strong> iuc2_rev.<br />

pathovar<br />

strain<br />

Achromobactin-like PCR product PCR for<br />

IEF-b<strong>and</strong> product iuc2<br />

syringae Pss22d + +<br />

Pss22d∆Pvd + +<br />

Pss22d∆Sid - +<br />

Pss22d∆Ach - -<br />

B728a + +<br />

B301D + +<br />

B241 + +<br />

C72 + +<br />

aceris CFBP 2339 + +<br />

actinidiae MAFF 302091 + nd<br />

aptata GSPB 1080 + +<br />

atr<strong>of</strong>aciens GSPB 1440 + +<br />

atropurporea MAFF 301309 + (-)<br />

berberis CFBP 1727 + nd<br />

broussonetiae MAFF810036 + nd<br />

castaneae CFBP 4217 + nd<br />

coronafaciens DSM 50261 - -<br />

daphniphylli CFBP 4219 + nd<br />

delfinii ICMP 529 + nd<br />

glycinea Psg1a + +<br />

PG4180 +<br />

mori MAFF 301020 + +<br />

mosprunorum GSPB 1013 + (-)<br />

maculicola DC3000 - -<br />

papulans ICMP 4048 + +<br />

phaseolicola 1448A + +<br />

pisi GSPB 1477 + nd<br />

savastanoi GSPB 2259 + (-)<br />

tabaci GSPB 117 + +<br />

tomato 487 + (-)<br />

74


Discussion<br />

6 Discussion<br />

Public awareness <strong>of</strong> problems with food safety is constantly increasing, as is<br />

the customer dem<strong>and</strong> for food that has been produced by an integrative <strong>and</strong><br />

sustainable agriculture. While this situation provides a good chance for<br />

biological control to prove its commercial applicability, the biggest bottleneck <strong>of</strong><br />

the system, the high variability in efficiency, remains to be solved. For this<br />

reason research on biological control can improve the practical application <strong>of</strong><br />

biocontrol by investigation <strong>of</strong> the underling microbial interactions rather than by<br />

screening for other control agents. Lack <strong>of</strong> knowledge on parameters <strong>of</strong><br />

epiphytical life in the phyllosphere is slowing down the progress in<br />

development <strong>of</strong> biocontrol products. Biological control <strong>of</strong> bacterial blight <strong>of</strong><br />

soybean by the antagonist Pss22d is a good model system to gain more<br />

information on these parameters, as it provides a) a highly efficient biological<br />

control system <strong>and</strong> b) significant amount <strong>of</strong> background information on the<br />

pathosystem involved. Thus the active principles <strong>of</strong> biological control could be<br />

directly interconnected to the complex plant-microbe interaction. In the frame <strong>of</strong><br />

this work it was attempted to elucidate the influence <strong>of</strong> siderophore production<br />

on the above mentioned biocontrol system. The results <strong>of</strong> this study clearly<br />

opposed the initial hypothesis, that competition for iron between pathogen <strong>and</strong><br />

control organism is an active principle <strong>of</strong> this biological control. Despite this, we<br />

could identify a siderophore production system that had not been described in<br />

plant-associated P. syringae before <strong>and</strong> address the importance <strong>of</strong> this<br />

siderophore for epiphytical fitness <strong>of</strong> the antagonist could be addressed. The<br />

production <strong>of</strong> this second, achromobactin-like siderophore is widely distributed<br />

among the pathovars <strong>of</strong> P. syringae.<br />

6.1 Identification <strong>of</strong> an achromobactin-like siderophore in P. syringae<br />

Iron uptake is essential for most bacteria, thus it is not surprising that the<br />

increasing number <strong>of</strong> genome sequences reveal more <strong>and</strong> more information on<br />

75


Discussion<br />

bacteria with redund<strong>and</strong>ent siderophore systems. For example,<br />

Pseudomonas aeroginosa produces pyoverdin <strong>and</strong> pyochelin, some strains <strong>of</strong><br />

P. fluorescense produce pyoverdin <strong>and</strong> pseudomonin, <strong>and</strong> D. chrysanthemi<br />

produces chrysobactin <strong>and</strong> achromobactin (Cox & Adams, 1985; Franza &<br />

Expert, 1991; Franza et al., 2005; Mercado-Blanco et al., 2001; Persmark et<br />

al., 1989). While the production <strong>of</strong> pyoverdin by different pathovars <strong>of</strong><br />

P. syringae had been reported previously (Bultreys et al., 2001; Jülich et al.,<br />

2001), we are the first to pro<strong>of</strong> experimentally the production <strong>of</strong> an addtional<br />

siderophore aside <strong>of</strong> pyoverdin. This study demonstrated that the biosynthesis<br />

gene cluster <strong>of</strong> this siderophore showed high similarity towards the<br />

achromobactin siderophore gene cluster <strong>of</strong> D. chrysanthemi (Fig.13). IEF<br />

analysis <strong>of</strong> achromobactin <strong>and</strong> the second siderophore <strong>of</strong> Pss22d shows a<br />

CAS-active b<strong>and</strong> at the same migration (data not shown). TLC separation <strong>of</strong><br />

achromobactin <strong>and</strong> the second siderophore <strong>of</strong> Pss22d on silica plates using a<br />

mobile phase <strong>of</strong> diethylether, n-butanole, acetic acid, water (1:1:1:1; v/v)<br />

exhibited a similar CAS-active b<strong>and</strong> (Fig.19A). More importantely, the<br />

siderophore <strong>of</strong> Pss22d∆Pvd could enhance growth <strong>of</strong> a siderophore-deficient<br />

D. chrysanthemi mutant under iron limiting conditions, while the growth<br />

deficiency <strong>of</strong> a mutant defective in the achromobactin receptor was not<br />

restored by feeding with Pss22d∆Pvd supernatant (D. Expert, personal<br />

communication). All these results imply a close structural similarity between the<br />

novel Pss22d siderophore with achromobactin from D. chrysanthemi. .<br />

The genetic organization <strong>of</strong> the biosynthesis genes differs between P. syringae<br />

<strong>and</strong> D. chrysanthemi. The gene acsF, encoding for a putative<br />

aminotransferase, is located upstream <strong>of</strong> the achromobactin receptor gene in<br />

D. chrysanthemi, while it follows downstream this receptor gene in P. syringae.<br />

This divergent organization could be the result <strong>of</strong> a genomic reorganization,<br />

however it could also indicate a different origin <strong>and</strong> therefore a different<br />

substrate specificity <strong>of</strong> the two respective aminotransferases. In support <strong>of</strong> this,<br />

the observed color change <strong>of</strong> the Pss22d∆Pvd siderophore upon iron addition<br />

76


Discussion<br />

may indicate a difference in the structures <strong>of</strong> the Pss22d siderophore <strong>and</strong><br />

achromobactin. However, we can not rule out the possibility, that the different<br />

behavior <strong>of</strong> both siderophores could be an artifact due to different purification<br />

procedures. The same might hold true for the different retention factors<br />

obtained after TLC separation (Fig.19B). Eventually, only structure<br />

determination <strong>of</strong> the Pss22d∆Pvd siderophore will show whether it is identical<br />

to achromobactin or not.<br />

Several siderophores structurally related to achromobactin are known. They<br />

are referred to as citrate-type siderophores. Aerobactin produced by E. coli,<br />

rhizobactin <strong>of</strong> S. meliloti, or vibri<strong>of</strong>errin synthesized by Vibrio parahaemolyticus<br />

are some examples for this group (Harris et al., 1979; Lynch et al., 2001;<br />

Tanabe et al., 2003). However, similarities between the P. syringae<br />

siderophore gene cluster <strong>and</strong> the achromobactin gene cluster <strong>of</strong><br />

D. chrysanthemi is much higher than to the other citrate-type siderophores.<br />

Recently, another gene cluster with high similarities towards the achromobactin<br />

biosynthesis genes has been reported in Sodalis glossinidiust (Darby et al.,<br />

2005). This obligate symbiont <strong>of</strong> Tse-Tse flies carries the achromobactin<br />

biosynthesis genes on a plasmid. The organization <strong>of</strong> the respective gene<br />

cluster shows a similar organization as that in D. chrysanthemi the acsF<br />

homologue located in front <strong>of</strong> acr (Darby et al., 2005). In S. glossinidiust the<br />

complete biosynthesis region is flanked by several transposons. This implies a<br />

genetic mobility <strong>of</strong> this region. Indeed, a transposon-mediated transfer <strong>of</strong><br />

siderophore biosynthesis gene clusters via so-called pathogenicity isl<strong>and</strong>s had<br />

been observed before, namely for the distribution <strong>of</strong> aerobactin or<br />

yersiniabactin (Carniel, 2001; Vokes et al., 1999). Achromobactin biosynthesis<br />

could be transferred by such an “isl<strong>and</strong>”. However no obvious transposase<br />

gene is located next to the respective biosynthesis genes in P. syringae. The<br />

organization <strong>of</strong> the siderophore biosynthesis genes together with a putative<br />

regulatory system next to the biosynthesis genes rather indicated that this<br />

gene cluster is native to P. syringae.<br />

77


Discussion<br />

6.2 Possible regulation <strong>of</strong> the achromobactin-like siderophore<br />

The close proximity <strong>of</strong> genes encoding for a FecR <strong>and</strong> a FecI homologue to the<br />

siderophore gene cluster in P. syringae implies the possibility <strong>of</strong> regulatory<br />

similarities between the Pss22d∆Pvd siderophore synthesis <strong>and</strong> ferric citrate<br />

uptake. Ferric citrate is internalized by combination <strong>of</strong> an outer membrane<br />

receptor (FecA) <strong>and</strong> an ATP-binding cassette transporter, ABC-transporter<br />

(Ferguson et al., 2002). FecA is also the first component <strong>of</strong> a signalling<br />

cascade, which is initiated at the cell surface (Fig.26). Binding <strong>of</strong> ferric citrate to<br />

FecA induces a conformational change that results in an interaction <strong>of</strong> FecA<br />

with the response regulator, FecR. Subsequently FecR converts FecI into an<br />

active sigma factor that directs RNA polymerase to the P fec promoter <strong>and</strong> thus<br />

initiates transcription <strong>of</strong> the fecABCDE transport genes. When iron supply is<br />

sufficient, the Fe 2+ -loaded ferric uptake regulation repressor, (Fur), inhibits<br />

transcription <strong>of</strong> the fecIR regulatory gene <strong>and</strong> the fecBCDE transport genes<br />

(Braun & Killmann, 1999). P fur represents the promoters to which Fe 2+ -loaded<br />

Fur repressor binds (Fig.26). A similar organisation is likely for the<br />

achromobactin-like siderophore <strong>of</strong> Pss22d, with the expression <strong>of</strong> the<br />

permease genes cbrABCD potentially being part <strong>of</strong> the feedback regulation<br />

loop (Fig.26, indicated in brackets). Although the influence <strong>of</strong> Fur on<br />

achromobactin biosynthesis genes had be demonstrated (Franza et al., 2005),<br />

no evidence for a FecI/FecR homologue system involved in achromobactin<br />

regulation is currently present for D. chrysanthemi (Franza et al., 2005).<br />

The conserved GGDEF domain within a predicted gene produc encoded<br />

downstream <strong>of</strong> the P. syringae biosynthetic gene cluster <strong>of</strong> the achromobactin–<br />

like siderophore is coupled to a putative sensory domain with similarity to<br />

chemotaxis / aerotaxis signaling domains. GGDEF domains are involved in<br />

cyclic di-GMP signaling, a global regulatory system for bacteria that mediates<br />

response to a wide range <strong>of</strong> stress signals such as pH-shift, osmotic shock, or<br />

78


Discussion<br />

(acr)<br />

(cbrA-D)<br />

Fig. 26. Model <strong>of</strong> transmembrane signalling for regulation <strong>of</strong> the ferric citrate uptake .<br />

The FecA protein is the initial component in the signaling cascade, which starts at the cell surface.<br />

Binding <strong>of</strong> ferric citrate to FecA induces a conformation such that FecA interacts with FecR, which in turn<br />

converts FecI into an active sigma factor that directs RNA polymerase to the P fec promoter <strong>and</strong> initiates<br />

transcription <strong>of</strong> the fecABCDE transport genes. When iron supply is sufficient, the Fe 2+ -loaded Fur<br />

repressor inhibits transcription <strong>of</strong> the regulatory genes fecIR <strong>and</strong> <strong>of</strong> the fecABCDE transport genes. P fur<br />

the promoters to which the Fe 2+ -loaded Fur repressor binds. aa, amino acids encoded. The brackets<br />

indicate homologues from the achromobactin-like siderophore uptake system. Modified after Braun <strong>and</strong><br />

Killman, 1999<br />

79


Discussion<br />

stationary phase transition (Galperin, 2006; Romling & Amikam, 2006). It is<br />

possible that this signaling pathway is involved in the differentiated regulation<br />

<strong>of</strong> siderophore biosynthesis in Pss22d.<br />

The obvious differences in siderophore production observed during growth <strong>of</strong><br />

Psg1a <strong>and</strong> Pss22d in vitro can not directly be associated to the wellinvestigated<br />

iron-mediated regulation <strong>of</strong> siderophore biosynthesis like the Fur<br />

signaling. Although iron availability did not differ between the two growth curve<br />

experiments depicted in Fig.22, the oxygen saturation <strong>of</strong> the culture medium<br />

must have varied between the two settings. Possible oxygen stress is the most<br />

likely explanation for the reduced growth <strong>of</strong> Pss22d shown in Fig.22B. It is<br />

interesting to note that Pss22d reacts to this stress by an up-regulation <strong>of</strong><br />

synthesis <strong>of</strong> the achromobactin-like siderophore. Achromobactin is supposed<br />

to protect D. chrysanthemi against oxidative stress (Expert, 1999). It could be<br />

possible that a similar protection mechanism occurs in Pss22d. However, the<br />

pathogenic Psg1a demonstrated an opposite regulation pattern. Pyoverdin<br />

production was up-regulated under high oxygen conditions. However, while the<br />

growth rate <strong>of</strong> Pss22d was reduced under high oxygen conditions, Psg1a<br />

growth was not effected. This result could indicate that in Psg1a other means<br />

are effective to avoid oxygen stress like for example in increased production <strong>of</strong><br />

superoxid dismutase. Up to now there is no experimental prove for this<br />

assumption.<br />

6.3 Influence <strong>of</strong> siderophore production on epiphytic fitness <strong>of</strong> Pss22d<br />

We could clearly rule out the hypothesis, that siderophore production directly<br />

influences biological control <strong>of</strong> Psg1a by Pss22d via competition for iron.<br />

However, it was also essential to investigate the influence <strong>of</strong> siderophore<br />

production on the overall epiphytic fitness <strong>of</strong> the antagonist in planta. One <strong>of</strong><br />

the most basic requirements for a versatile biological control agent is its ability<br />

to survive in the respective habitat. It can only suppress pathogen development<br />

sufficiently if it is able to compete with other epiphytic microorganisms. The<br />

80


Discussion<br />

wound inoculation technique is not suitable to test the epiphytic fitness <strong>of</strong><br />

Pss22d, as the artificial wound gives rise to many plant born nutrients normally<br />

not present at the plant surface. Thus, spray inoculation experiments were<br />

performed (Tab. 10). A practical problem <strong>of</strong> spray inoculation experiments is a<br />

relatively high variation. In addition, the performance <strong>of</strong> the Pss22d<br />

siderophore mutants could be influenced by the composition <strong>of</strong> the epiphytical<br />

community <strong>of</strong> the inoculated plant. This may not only have affected the total<br />

amount <strong>of</strong> available iron but bacteria also possess a large number <strong>of</strong><br />

siderophore receptors for uptake <strong>of</strong> heterogenic, “foreign” siderophores<br />

(Schubert et al., 1999). For example the PsmDC3000 genome contains at least<br />

23 different TonB-dependent siderophore receptors (Cornelis & Matthijs,<br />

2002). Heterologous iron uptake could have complemented the phenotype <strong>of</strong><br />

the Pss22d mutants, thus the experiment needs to be repeated several times<br />

on different sets <strong>of</strong> plants in order to obtain results for different overall epiphytic<br />

communities.<br />

6.4 Distribution <strong>of</strong> the achromobactin-like siderophore among P. syringae<br />

pathovars<br />

Screening <strong>of</strong> different P. syringae strains for the production <strong>of</strong> the<br />

achromobactin-like siderophore revealed that a high number <strong>of</strong> the tested<br />

strains produced this siderophore. Thus it can be concluded, that despite the<br />

high genetic variability <strong>of</strong> P. syringae pathovars, siderophore biosynthesis <strong>of</strong><br />

Pvd PS as well as the achromobactin-like siderophore is highly conserved<br />

(Bultreys et al., 2003; Bultreys et al., 2004; Jülich et al., 2001), this study). This<br />

uniformity may reflect a lack <strong>of</strong> competitiveness for iron uptake among plantassociated<br />

bacteria. This hypothesis is supported by investigations on the<br />

siderophore production <strong>of</strong> PssB728a. Loper <strong>and</strong> coworkers (1987) reported<br />

that knock-out mutation <strong>of</strong> pyoverdin in PssB728a had no impact on growth<br />

survival or pathogenicity <strong>of</strong> PssB728a on snap bean leafs. They concluded,<br />

that siderophore production has no impact on the interaction between plant<br />

81


Discussion<br />

<strong>and</strong> bacterium (Loper et al., 1984; Loper & Lindow, 1987). What they did not<br />

notice at that time was the presence <strong>of</strong> a second siderophore in these mutants.<br />

Since the CAS-assay for siderophore detection was not used at the time <strong>of</strong> that<br />

study, the “siderophore-negative” phenotype <strong>of</strong> PssB728a mutants was just<br />

determined by lack <strong>of</strong> fluorescence. Thus the presence <strong>of</strong> a second<br />

siderophore stayed unnoticed. In follow-up studies, reporter gene fusions to<br />

iron-regulated promoters were constructed <strong>and</strong> applied for an evaluation <strong>of</strong> iron<br />

availability on the plant surface (Loper & Lindow, 1994; Loper & Henkels,<br />

1997). However, the promoter used for this reporter gene fusion was derived<br />

from pyoverdin synthesis genes <strong>and</strong> therefore the fusions responded only to<br />

severe conditions <strong>of</strong> iron limitation. Therefore, the results obtained by these<br />

fusions were not suitable to evaluate the impact <strong>of</strong> a siderophore that is<br />

already expressed at less restrictive conditions as it has been recently<br />

demonstrated for achromobactin expression in D. chrysanthemi (Franza et al.,<br />

2005). In consequence, the conserved distribution <strong>of</strong> Pvd PS <strong>and</strong> the<br />

achromobactin-like siderophore found in the current study rather may be<br />

attributed to an optimal suitability <strong>of</strong> this siderophores for plant-associated<br />

bacteria.<br />

6.5 Outlook<br />

The results represented in this thesis demonstrated the presence <strong>of</strong> an<br />

achromobactin-like siderophore in Pss22d, but so far we could not clarify<br />

whether this siderophore is structurally identical to achromobactin or not.<br />

Further purification methods need to be developed in order to obtain a<br />

siderophore sample that can be analyzed by NMR. As we could exclude a<br />

direct role <strong>of</strong> siderophore production as an active principle <strong>of</strong> biological control<br />

between Psg1a <strong>and</strong> Pss22d, the molecular basis <strong>of</strong> this antagonism remains<br />

unknown. The impact <strong>of</strong> antibiosis is currentely investigated by Sascha Braun,<br />

Jena, but several other mechanisms, like for example a possible influence <strong>of</strong><br />

plant defense activation remain to be examined.<br />

82


Literature<br />

Literature<br />

Agrios, G. N. (2005). Plant pathology, 5th edn. edn. New York: Elseviewer<br />

Academic Press.<br />

Alex<strong>and</strong>er, D. B. & Zuberer, D. A. (1991). Use <strong>of</strong> chrome azurol S reagents to<br />

evaluate siderophore production by rhizosphere bacteria. Biol Fertil Soil 12, 39-<br />

45.<br />

Anjaiah, V., Koedam, N., Nowakthompson, B., Loper, J. E., H<strong>of</strong>te, M.,<br />

Tambong, J. T. & Cornelis, P. (1998). Involvement <strong>of</strong> phenazines <strong>and</strong><br />

anthranilate in the antagonism <strong>of</strong> Pseudomonas aeruginosa PNA1 <strong>and</strong> Tn5<br />

Derivatives toward Fusarium spp. <strong>and</strong> Pythium spp. Molecular Plant-Microbe<br />

Interactions 11, 847-854.<br />

Ausubel, F. M., Brent, R., Kingston, R. E., Moore, D. D., Seidman, J. G.,<br />

Smith, J. A. & Struhl, K. (1987). Current protocols in molecular biology: John<br />

Wiley <strong>and</strong> sons.<br />

Baker, B., Zambryski, P., Staskawicz, B. J. & Dinesh-Kumar, S. P. (1997).<br />

Signalling in plant-microbe interactions. <strong>Science</strong> 276, 726-733.<br />

Bender, C. (1999). Chlorosis-inducing phytotoxins produced by Pseudomonas<br />

syringae. European Journal <strong>of</strong> Plant Pathology 105, 1–12.<br />

Bender, C. L., Alarcon-Chaidez, F. & Gross, D. C. (1999). Pseudomonas<br />

syringae phytotoxins: mode <strong>of</strong> action, regulation, <strong>and</strong> biosynthesis by peptide<br />

<strong>and</strong> polyketide synthetases. Microbiology <strong>and</strong> molecular biology reviews 63,<br />

266-292.<br />

Braun, V. & Killmann, H. (1999). Bacterial solutions to the iron-supply<br />

problem. Trends in Biochemical <strong>Science</strong> 24, 104-109.<br />

Braun, V. & Braun, M. (2002). Iron transport <strong>and</strong> signaling in Escherichia coli.<br />

FEBS Letters 529, 78-85.<br />

Braun-Kiewnick, A., <strong>Jacobs</strong>en, B. J. & S<strong>and</strong>s, D. C. (2000). Biological<br />

control <strong>of</strong> Pseudomonas syringae pv. syringae, the causal agent <strong>of</strong> basal<br />

kernel blight <strong>of</strong> barley, by antagonistic Pantoea agglomerans. Phytopathology<br />

90, 368-375.<br />

Buell, C. R., Joardar, V., Lindeberg, M. & other authors (2003). The<br />

complete genome sequence <strong>of</strong> the Arabidopsis <strong>and</strong> tomato pathogen<br />

Pseudomonas syringae pv. tomato DC3000. Proceedings <strong>of</strong> the National<br />

Academy <strong>of</strong> <strong>Science</strong>s <strong>of</strong> the United States <strong>of</strong> America, 10181-10186.<br />

Bultreys, A., Gheysen, I., Maraite, H. & de H<strong>of</strong>fmann, E. (2001).<br />

Characterization <strong>of</strong> fluorescent <strong>and</strong> nonfluorescent peptide siderophores<br />

produced by Pseudomonas syringae strains <strong>and</strong> their potential use in strain<br />

identification. Applied & Environmental Microbiology 67, 1718-1727.<br />

83


Literature<br />

Bultreys, A., Gheysen, I., Wathelet, B., Maraite, H. & de H<strong>of</strong>fmann, E.<br />

(2003). High-Performance Liquid Chromatography analyses <strong>of</strong> pyoverdin<br />

siderophores differentiate among phytopathogenic fluorescent Pseudomonas<br />

species. Applied & Environmental Microbiology 69, 1143-1153.<br />

Bultreys, A., Gheysen, I., Wathelet, B., Schafer, M. & Budzikiewicz, H.<br />

(2004). The pyoverdins <strong>of</strong> Pseudomonas syringae <strong>and</strong> Pseudomonas cichorii.<br />

Zeitschrift für Naturforschung [C] 59, 613-618.<br />

Burrill, T. J. (1878). Remarks made in a discussion. Trans Ill State Hortic Soc<br />

11, 79-80.<br />

Büttner, P. (2003). Siderophore des Pathogens der Sojabohne<br />

Pseudomonas syringae pv. glycinea 1a/96 und seiner bakteriellen<br />

Antagonisten Pseudomonas syringae pv. syringae 22d/93 und Pantoea<br />

agglomerans 48b/90. Diploma Thesis, <strong>University</strong> <strong>of</strong> Jena.<br />

Buysens, S., Huengens, K., Poppe, J. & H<strong>of</strong>te, M. (1996). Involvement <strong>of</strong><br />

pyochelin <strong>and</strong> pyoverdin in suppression <strong>of</strong> Pythium-induced damping-<strong>of</strong>f <strong>of</strong><br />

tomato by Pseudomonas aeruginosa 7NSK2. Applied & Environmental<br />

Microbiology 62, 865-871.<br />

Carniel, E. (2001). The Yersinia high-pathogenicity isl<strong>and</strong>: an iron-uptake<br />

isl<strong>and</strong>. Microbes <strong>and</strong> Infections 3, 561-569.<br />

Coerper, F. M. (1919). Bacterial blight <strong>of</strong> soybean. Journal <strong>of</strong> Agricultural<br />

Research 18, 179-193.<br />

Cornelis, P. & Matthijs, S. (2002). Diversity <strong>of</strong> siderophore-mediated iron<br />

uptake systems in fluorescent pseudomonads: not only pyoverdines.<br />

Environmental Microbiology 4, 787-798.<br />

Cox, C. D. & Adams, P. (1985). Siderophore activity <strong>of</strong> pyoverdin for<br />

Pseudomonas aeruginosa. Infection <strong>and</strong> Immunity 48, 130-138.<br />

Cui, J., Bahrami, A. K., Pringle, E. G., Hern<strong>and</strong>ez-Guzman, G., Bender, C.<br />

L., Pierce, N. E. & Ausubel, F. M. (2005). Pseudomonas syringae<br />

manipulates systemic plant defenses against pathogens <strong>and</strong> herbivores.<br />

Proceedings <strong>of</strong> the National Academy <strong>of</strong> <strong>Science</strong>s <strong>of</strong> the United States <strong>of</strong><br />

America 102, 1791-1806.<br />

Darby, A. C., Lagnel, J., Mathew, C. Z., Bourtzis, K., Maudlin, I. & Welburn,<br />

S. C. (2005). Extrachromosomal DNA <strong>of</strong> the symbiont Sodalis glossinidiust.<br />

Journal <strong>of</strong> Bacteriology 187, 5003-5007.<br />

de Chial, M., Ghysels, B., Beatson, S. A. & other authors (2003).<br />

Identification <strong>of</strong> type II <strong>and</strong> type III pyoverdine receptors from Pseudomonas<br />

aeruginosa. Microbiology 149, 821-831.<br />

84


Literature<br />

Dong, Y. H., Zhang, X. F., Xu, J. L. & Zhang, L. H. (2004). Insecticidal<br />

Bacillus thuringiensis silences Erwinia carotovora virulence by a new form <strong>of</strong><br />

microbial antagonism, signal interference. Appl Environ Microbiol 70, 954-960.<br />

Doudor<strong>of</strong>f, M. & Palleroni, N. J. (1974). Genus I. Pseudomonas Migula. In<br />

Bergey's manual <strong>of</strong> determinative bacteriology, pp. 217-243. Edited by R. E.<br />

Buchanan & N. E. Gibbons. Baltimore, Md.: Williams & Wilkins.<br />

Duffy, B., Schouten, A. & Raaijmakers, J. M. (2003). Pathogen self-defense:<br />

Mechanisms to counteract microbial antagonism. Annual Review <strong>of</strong><br />

Phytopathology 41, 501-538.<br />

Eastgate, J. A. (2000). Erwinia amylovora : the molecular basis <strong>of</strong> fireblight<br />

disease. Molecular Plant Pathology 1, 325-329.<br />

Emmert, E. A. & H<strong>and</strong>elsman, J. (1999). Biocontrol <strong>of</strong> plant disease: a (gram-<br />

) positive perspective. FEMS Microbiolgy Letters 171, 1-9.<br />

Engelmann, M. D., Bobier, R. T., Hiatt, T. & Cheng, I. F. (2003). Variability <strong>of</strong><br />

the Fenton reaction characteristics <strong>of</strong> the EDTA, DTPA, <strong>and</strong> citrate complexes<br />

<strong>of</strong> iron. Biometals 16, 519-527.<br />

Expert, D. (1999). Withholding <strong>and</strong> exchanging iron: Interactions between<br />

Erwinia spp. <strong>and</strong> their plant hosts. Annual Review <strong>of</strong> Phytopathology 37, 307-<br />

334.<br />

Ferguson, A. D., Chakraborty, R., Smith, B. S., Esser, L., van der Helm, D.<br />

& Deisenh<strong>of</strong>er, J. (2002). Structural basis <strong>of</strong> gating by the outer membrane<br />

transporter FecA. <strong>Science</strong> 295, 1715-1719.<br />

Finking, R. & Marahiel, M. A. (2004). Biosynthesis <strong>of</strong> nonribosomal peptides.<br />

Annual Review <strong>of</strong> Microbiology, 453-477.<br />

Finlay, B. B. & Falkow, S. (1997). Common themes in microbial pathogenicity<br />

revisited. Microbiology <strong>and</strong> molecular biology reviews 61, 136-169.<br />

Franza, T. & Expert, D. (1991). The virulence-associated chrysobactin iron<br />

uptake system <strong>of</strong> Erwinia chrysanthemi 3937 involves an operon encoding<br />

transport <strong>and</strong> biosynthetic functions. Journal <strong>of</strong> Bacteriology 173, 6874-6881.<br />

Franza, T., Mahe, B. & Expert, D. (2005). Erwinia chrysanthemi requires a<br />

second iron transport route dependent <strong>of</strong> the siderophore achromobactin for<br />

extracellular growth <strong>and</strong> plant infection. Molecular Microbiology 55, 261-275.<br />

Fuchs, R., Schafer, M., Ge<strong>of</strong>froy, V. & Meyer, J. M. (2001). Siderotyping-a<br />

powerful tool for the characterization <strong>of</strong> pyoverdines. Current Medicinal<br />

Chemistry 1, 31-57.<br />

85


Literature<br />

Galperin, M. Y. (2006). Structural classification <strong>of</strong> bacterial response<br />

regulators: diversity <strong>of</strong> output domains <strong>and</strong> domain combinations. Journal <strong>of</strong><br />

Bacteriology 188, 4169-4182.<br />

Guerinot, M. L. (1994). Microbial iron transport. Annual Review <strong>of</strong><br />

Microbiology 48, 743-772.<br />

Haas, D., Blumer, C. & Keel, C. (2000). Biocontrol ability <strong>of</strong> fluorescent<br />

pseudomonads genetically dissected: importance <strong>of</strong> positive feedback<br />

regulation. Current Opinion in Biotechnology 11, 290-297.<br />

Haas, D. & Defago, G. (2005). Biological control <strong>of</strong> soil-borne pathogens by<br />

fluorescent pseudomonads. Nature Reviews Microbiol 3, 307-319.<br />

Hamdan, H., Weller, D. M. & Thomashow, L. S. (1991). Relative importance<br />

<strong>of</strong> fluorescent siderophores <strong>and</strong> other factors in biological control <strong>of</strong><br />

Gaeumannomyces graminis var. tritici by Pseudomonas fluorescens 2-79 <strong>and</strong><br />

M4-80R. Applied & Environmental Microbiology 57, 3270-3277.<br />

H<strong>and</strong>elsman, J. & Stabb, E. V. (1996). Biocontrol <strong>of</strong> Soilborne Plant<br />

Pathogens. Plant Cell 8, 1855-1869.<br />

Harris, W. R., Carrano, C. J. & Raymond, K. N. (1979). Coordination<br />

chemistry <strong>of</strong> microbial iron transport compounds.16. Isolation, characterization,<br />

<strong>and</strong> formation constants <strong>of</strong> ferric aerobactin. Journal <strong>of</strong> the American Chemical<br />

Society 101, 2722-2727.<br />

Hirai, T., Fukushima, K., Kumamoto, K. & Iwahashi, H. (2005). Effects <strong>of</strong><br />

some naturally occurring iron ion chelators on in vitro superoxide radical<br />

formation. Biology <strong>of</strong> trace element research 108, 77-85.<br />

Hutcheson, S. W. (2001). The molecular biology <strong>of</strong> hypersensitivity to plant<br />

pathogenic bacteria. Journal <strong>of</strong> Plant Pathology 83, 151-172.<br />

Janisiewicz, W. J. & Korsten, L. (2002). Biological control <strong>of</strong> postharvest<br />

diseases <strong>of</strong> fruits. Annual Review <strong>of</strong> Phytopathology 40, 411-441.<br />

Ji, P. & Wilson, M. (2003). Enhancement <strong>of</strong> population size <strong>of</strong> a biological<br />

control agent <strong>and</strong> efficacy in control <strong>of</strong> bacterial speck <strong>of</strong> tomato through<br />

salicylate <strong>and</strong> ammonium sulfate amendments. Applied & Environmental<br />

Microbiology 69, 1290-1294.<br />

Jülich, M., Taraz, K., Budzikiewicz, H., Ge<strong>of</strong>froy, V., Meyer, J. M. &<br />

Gardan, L. (2001). The structure <strong>of</strong> the pyoverdin isolated from various<br />

Pseudomonas syringae pathovars. Zeitschrift für Naturforschung [C] 56, 687-<br />

694.<br />

Keel, C., Weller, D. M., Natsch, A., Defago, G., Cook, R. J. & Thomashow,<br />

L. S. (1996). Conservation <strong>of</strong> the 2,4-diacetylphloroglucinol biosynthesis locus<br />

86


Literature<br />

among fluorescent Pseudomonas strains from diverse geographic locations.<br />

Applied & Environmental Microbiology 62, 552-563.<br />

Klement, Z. (1963). Rapid detection <strong>of</strong> pathogenicity <strong>of</strong> phytopathogenic<br />

pseudomonads. Nature 199, 299-300.<br />

Koedam, N., Wittouck, E., Gaballa, A., Gillis, A., H<strong>of</strong>te, M. & Cornelis, P.<br />

(1994). Detection <strong>and</strong> differentiation <strong>of</strong> microbial siderophores by isoelectric<br />

focusing <strong>and</strong> chrome azurol S overlay. Biometals 7, 287-291.<br />

Loper, J. E., Orser, C. S. & Panopoulos, N. J. (1984). Genetic Analysis <strong>of</strong><br />

Fluorescent Pigment Production in Peudomonas syringae pv. syringae. J Gen<br />

Microbiol 130, 1507-1515.<br />

Loper, J. E. & Lindow, S. E. (1987). Lack <strong>of</strong> evidence for insitu fluorescent<br />

pigment production by Pseudomonas syringae pv. syringae on bean leaf<br />

surfaces. Phytopathology 77, 1449-1454.<br />

Loper, J. E. & Buyer, J. S. (1991). Siderophores in microbial interactions on<br />

plant surfaces. Mol Plant Microbe Interact 4, 5-13.<br />

Loper, J. E. & Lindow, S. E. (1994). A Biological Sensor for Iron Available to<br />

Bacteria in Their Habitats on Plant-Surfaces. Appl Environ Microb 60, 1934-<br />

1941.<br />

Loper, J. E. & Henkels, M. D. (1997). Availability <strong>of</strong> iron to Pseudomonas<br />

fluorescens in rhizosphere <strong>and</strong> bulk soil evaluated with an ice nucleation<br />

reporter gene. Appl Environ Microbiol 63, 99-105.<br />

Lugtenberg, B. J. J., Chin-A-Woeng, T. F. C. & Bloemberg, G. V. (2002).<br />

Microbe–plant interactions: principles <strong>and</strong> mechanisms. Antonie Van<br />

Leeuwenhoek 81, 373–383.<br />

Lynch, D., O'Brien, J., Welch, T., Clarke, P., Cuiv, P. O., Crosa, J. H. &<br />

O'Connell, M. (2001). Genetic organization <strong>of</strong> the region encoding regulation,<br />

biosynthesis, <strong>and</strong> transport <strong>of</strong> rhizobactin 1021, a siderophore produced by<br />

Sinorhizobium meliloti. Journal <strong>of</strong> Bacteriology 183, 2576-2585.<br />

May, R., Völksch, B. & Kampmann, G. (1997a). Antagonistic activities <strong>of</strong><br />

epiphytic bacteria from soybean leaves against Pseudomonas syringae pv.<br />

glycinea in vitro <strong>and</strong> in planta. Microbial Ecology 34, 118-124.<br />

May, R., Völksch, B., Kampmann, G. & Nüske, J. (1997b). Isolation <strong>and</strong><br />

characterization <strong>of</strong> Pseudomonas syringae strain with antagonistic activities<br />

against Pseudomonas syringae pv. glycinea in vitro <strong>and</strong> in planta. In<br />

Pseudomonas syringae pathovars <strong>and</strong> related pathogens, pp. 629-634. Edited<br />

by K. Rudolph, T. J. Burr, J. W. MANSFIELD, D. Stead, A. Vivian & J. von<br />

Kietzell: Kluwer Academic Publischer.<br />

87


Literature<br />

Mazzola, M. (2002). Mechanisms <strong>of</strong> natural soil suppressiveness to soilborne<br />

diseases. Antonie Van Leeuwenhoek 81, 557-564.<br />

McManus, P. S., Stockwell, V. O., Sundin, G. W. & Jones, A. L. (2002).<br />

Antibiotic use in plant agriculture. Annual Review <strong>of</strong> Phytopathology 40, 443-<br />

465.<br />

Mercado-Blanco, J., van der Drift, K. M., Olsson, P. E., Thomas-Oates, J.<br />

E., van Loon, L. C. & Bakker, P. A. (2001). Analysis <strong>of</strong> the pmsCEAB gene<br />

cluster involved in biosynthesis <strong>of</strong> salicylic acid <strong>and</strong> the siderophore<br />

pseudomonine in the biocontrol strain Pseudomonas fluorescens WCS374.<br />

Journal <strong>of</strong> Bacteriology 183, 1909-1920.<br />

Meyer, J. M., Ge<strong>of</strong>froy, V. A., Baida, N., Gardan, L., Izard, D., Lemanceau,<br />

P., Achouak, W. & Palleroni, N. J. (2002). Siderophore typing, a powerful tool<br />

for the identification <strong>of</strong> fluorescent <strong>and</strong> nonfluorescent pseudomonads. Applied<br />

<strong>and</strong> Environmental Microbiology 68, 2745-2753.<br />

Miller, J. H. (1992). A short course in bacterial genetics. A laboratory manual<br />

<strong>and</strong> h<strong>and</strong>book for Escherichia coli <strong>and</strong> related bacteria. Cold Spring Harbour,<br />

N. Y.: Cold Spring Harbour Laboratory.<br />

Montesinos, E. (2000). Pathogenic plant–microbe interactions. What we know<br />

<strong>and</strong> how we benefit. International Microbiology 3, 69-70.<br />

Montesinos, E., Bonaterra, A., Badosa, E., France´ s, J., Alemany, J.,<br />

Llorente, I. & Moragrega, C. (2002). Plant-microbe interactions <strong>and</strong> the new<br />

biotechnological methods <strong>of</strong> plant disease control. Int Microbiol 5, 169–175.<br />

Montesinos, E. (2003). Development, registration <strong>and</strong> commercialization <strong>of</strong><br />

microbial pesticides for plant protection. International Microbiology 6, 245-252.<br />

Mossialos, D., Ochsner, U., Baysse, C. & other authors (2002).<br />

Identification <strong>of</strong> new, conserved, non-ribosomal peptide synthetases from<br />

fluorescent pseudomonads involved in the biosynthesis <strong>of</strong> the siderophore<br />

pyoverdine. Molecular Microbiology 45, 1673-1685.<br />

Münzinger, M., Budzikiewicz, H., Expert, D., Enard, C. & Meyer, J. M.<br />

(2000). Achromobactin, a new citrate siderophore <strong>of</strong> Erwinia chrysanthemi.<br />

Zeitschrift für Naturforschung [C] 55, 328-332.<br />

Neil<strong>and</strong>s, J. B. (1995). Siderophores: structure <strong>and</strong> function <strong>of</strong> microbial iron<br />

transport compounds. Journal <strong>of</strong> Biological Chemistry 270, 26723-26726.<br />

Nyiredy (2004). Separation strategies<strong>of</strong> plant constituents - current status.<br />

Journal <strong>of</strong> Chromatography B 812, 35-51.<br />

Oerke, E. C., Dehne, H. W., Schönbeck, F. & Weber, A. (1994). Crop<br />

production <strong>and</strong> crop protection: Estimated losses in major food <strong>and</strong> cash crops.<br />

Amsterdam: Elseviewer.<br />

88


Literature<br />

Persello-Cartieaux, F., Nussaume, L. & Robaglia, C. (2003). Tales from the<br />

underground: molecular plant-rhizobacteria interactions. Plant Cell <strong>and</strong><br />

Environment 26, 189-199.<br />

Persmark, M., Expert, D. & Neil<strong>and</strong>s, J. B. (1989). Isolation, characterization,<br />

<strong>and</strong> synthesis <strong>of</strong> chrysobactin, a compound with siderophore activity from<br />

Erwinia chrysanthemi. Journal <strong>of</strong> Biological Chemistry 264, 3187-3193.<br />

Pieterse, C. M. J., van Pelt, J. A., van Wees, S. C. & other authors (2001).<br />

Rhizobacteria-mediated induced systemic resistance: triggering, signalling <strong>and</strong><br />

expression. European Journal <strong>of</strong> Plant Pathology 107, 51-61.<br />

Poole, K. & McKay, G. A. (2003). Iron acquisition <strong>and</strong> its control in<br />

pseudomonas aeruginosa: many roads lead to rome. Frontiers in Bioscience 8,<br />

661-686.<br />

Quadri, L. E., Keating, T. A., Patel, H. M. & Walsh, C. T. (1999). Assembly <strong>of</strong><br />

the Pseudomonas aeruginosa nonribosomal peptide siderophore pyochelin: In<br />

vitro reconstitution <strong>of</strong> aryl-4, 2-bisthiazoline synthetase activity from PchD,<br />

PchE, <strong>and</strong> PchF. Biochemistry 38, 14941-14954.<br />

Raaijmakers, J. M., Vlami, M. & de Souza, J. T. (2002). Antibiotic production<br />

by bacterial biocontrol agents. Antonie Van Leeuwenhoek 81, 537-547.<br />

Ran, L. X., van Loon, L. C. & Bakker, P. A. H. M. (2005). No role for<br />

bacterially produced salicylic acid in rhizobacterial induction <strong>of</strong> systemic<br />

resistance in Arabidopsis. Phytopathology 95, 1349-1355.<br />

Ravel, J. & Cornelis, P. (2003). Genomics <strong>of</strong> pyoverdine-mediated iron uptake<br />

in pseudomonads. Trends in Microbiology 11, 195-200.<br />

Romling, U. & Amikam, D. (2006). Cyclic di-GMP as a second messenger.<br />

Current Opinion in Microbiology 9, 218-228.<br />

Ross, A. F. (1962). Systemic acquired resistance induced by localized virus<br />

infections in plants. Virology 14, 340-358.<br />

Saiki, R. K., Gelf<strong>and</strong>, D. H., St<strong>of</strong>fel, S., Scharf, S. J., Higuchi, R., Horn, G.<br />

T., Mullis, K. B. & Erlich, H. A. (1988). Primer-directed enzymatic<br />

amplification <strong>of</strong> DNA with thermostable DNA polymerase. <strong>Science</strong> 239, 487-<br />

491.<br />

Sambrook, J., Fritsch, E. F. & Maniatis, T. E. (1989). Molecular cloning. A<br />

laboratory manual. Cold Spring Harbor, N.Y.: Cold Spring Harbor Laboratory.<br />

Schippers, B., Bakker, A. W. & Bakker, P. A. H. M. (1987). Interactions <strong>of</strong><br />

Deleterious <strong>and</strong> Beneficial Rhizosphere Microorganisms <strong>and</strong> the Effect <strong>of</strong><br />

Cropping Practices. Annual Review <strong>of</strong> Phytopathology 25, 339-358.<br />

89


Literature<br />

Schubert, S., Rakin, A., Karch, H., Carniel, E. & Heesemann, J. (1998).<br />

Prevalence <strong>of</strong> the "high-pathogenicity isl<strong>and</strong>" <strong>of</strong> Yersinia species among<br />

Escherichia coli strains that are pathogenic to humans. Infection <strong>and</strong> Immunity<br />

66, 480-485.<br />

Schubert, S., Fischer, D. & Heesemann, J. (1999). Ferric enterochelin<br />

transport in Yersinia enterocolitica: molecular <strong>and</strong> evolutionary aspects. J<br />

Bacteriol 181, 6387-6395.<br />

Schwyn, B. & Neil<strong>and</strong>s, J. B. (1987). Universal chemical assay for the<br />

detection <strong>and</strong> determination <strong>of</strong> siderophores. Analytical Biochemistry 160, 47-<br />

56.<br />

Shoda, M. (2000). Bacterial control <strong>of</strong> plant diseases. Journal <strong>of</strong> Bioscience<br />

<strong>and</strong> Bioengineering 89, 515-521.<br />

Smirnova, A. V. & Ullrich, M. S. (2004). Topological <strong>and</strong> deletion analysis <strong>of</strong><br />

CorS, a Pseudomonas syringae sensor kinase. Microbiology 150, 2715-2726.<br />

Smith, J. L. (2004). The physiological role <strong>of</strong> ferritin-like compounds in<br />

bacteria. Critical Reviews in Microbiology, 30, 173-185.<br />

Snyder, L. R. (1978). Classification <strong>of</strong> the solvent properties <strong>of</strong> common<br />

liquids. Journal <strong>of</strong> chromatographic science 16, 223-234.<br />

Staskawicz, B. J., Mudgett, M. B., Dangl, J. L. & Galan, J. E. (2001).<br />

Common <strong>and</strong> contrasting themes <strong>of</strong> plant <strong>and</strong> animal diseases. <strong>Science</strong> 292,<br />

2285-2289.<br />

Stromberg, K. D., Kinkel, L. L. & Leonard, K. J. (2004). Quantifying the<br />

effect <strong>of</strong> bacterial antagonists on the relationship between phyllosphere<br />

population sizes <strong>of</strong> Xanthomonas translucens pv. translucens <strong>and</strong> subsequent<br />

bacterial leaf streak severity on wheat seedlings. Biological Control 29, 58-65.<br />

Tanabe, T., Funahashi, T., Nakao, H., Miyoshi, S., Shinoda, S. &<br />

Yamamoto, S. (2003). Identification <strong>and</strong> characterization <strong>of</strong> genes required for<br />

biosynthesis <strong>and</strong> transport <strong>of</strong> the siderophore vibri<strong>of</strong>errin in Vibrio<br />

parahaemolyticus. Journal <strong>of</strong> Bacteriology 185, 6938-6949.<br />

Teuber, M. (1999). Spread <strong>of</strong> antibiotic resistance with food-borne pathogens.<br />

Cell Mol Life Sci 56, 755-763.<br />

Thomashow, L. S. (1996). Biological control <strong>of</strong> plant root pathogens. Current<br />

Opinion in Biotechnology 7, 343-347.<br />

van Loon, L. C., Bakker, P. A. H. M. & Pieterse, C. M. J. (1998a). Systemic<br />

resistance induced by rhizosphere bacteria. Annual Review <strong>of</strong> Phytopathology<br />

36, 453-483.<br />

90


Literature<br />

van Loon, L. C., Bakker, P. A. H. M. & Pieterse, C. M. J. (1998b). Systemic<br />

resistance induced by rhizospheric bacteria. Annual Review <strong>of</strong> Phytopathology<br />

36, 453-483.<br />

Visca, P., Colotti, G., Serino, L., Verzili, D., Orsi, N. & Chiancone, E.<br />

(1992). Metal regulation <strong>of</strong> siderophore synthesis in Pseudomonas aeruginosa<br />

<strong>and</strong> functional effects <strong>of</strong> siderophore-metal complexes. Appl Environ Microbiol<br />

58, 2886-2893.<br />

Visca, P., Leoni, L., Wilson, M. J. & Lamont, I. L. (2002). Iron transport <strong>and</strong><br />

regulation, cell signalling <strong>and</strong> genomics: lessons from Escherichia coli <strong>and</strong><br />

Pseudomonas. Molecular Microbiology 45, 1177-1190.<br />

Vokes, S. A., Reeves, S. A., Torres, A. G. & Payne, S. M. (1999). The<br />

aerobactin iron transport system genes in Shigella flexneri are present within a<br />

pathogenicity isl<strong>and</strong>. Molecular Microbiology 33, 63-73.<br />

Völksch, B., Nüske, J. & May, R. (1996). Characterization <strong>of</strong> two epiphytic<br />

bacteria from soybean leaves with antagonistic activities against Pseudomonas<br />

syringae pv. glycinea. Journal <strong>of</strong> Basic Microbiology 36, 453-462.<br />

Völksch, B. & May, R. (2001a). Biological Control <strong>of</strong> Pseudomonas syringae<br />

pv. glycinea by Epiphytic Bacteria under Field Conditions. Microb Ecol 41, 132-<br />

139.<br />

Völksch, B. & May, R. (2001b). Biological control <strong>of</strong> Pseudomonas syringae<br />

pv. glycinea by epiphytic bacteria under field conditions. Microbial Ecology 41,<br />

132-139.<br />

Weingart, H., Stubner, S., Schenk, A. & Ullrich, M. S. (2004). Impact <strong>of</strong><br />

temperature on in planta expression <strong>of</strong> genes involved in synthesis <strong>of</strong> the<br />

Pseudomonas syringae phytotoxin coronatine. Molecular Plant-Microbe<br />

Interactions 17, 1095-1102.<br />

Weller, D. M., Raaijmakers, J. M., Gardener, B. B. M. & Thomashow, L. S.<br />

(2002). Microbial populations responsible for specific soil suppressiveness to<br />

plant pathogens. Annual Review <strong>of</strong> Phytopathology 40, 309-348.<br />

Wilson, C. L. (1997a). Biological control <strong>and</strong> plant diseases - a new paradigm.<br />

Journal <strong>of</strong> Industrial Microbiology & Biotechnology 19, 158-159.<br />

Wilson, K. J., Sessitsch, A., Corbo, J. C., Giller, K. E., Akkermans, A. D. &<br />

Jefferson, R. A. (1995). beta-Glucuronidase (GUS) transposons for ecological<br />

<strong>and</strong> genetic studies <strong>of</strong> rhizobia <strong>and</strong> other gram-negative bacteria. Microbiology<br />

141 ( Pt 7), 1691-1705.<br />

Wilson, M. (1997b). Biocontrol <strong>of</strong> aerial plant diseases in agriculture <strong>and</strong><br />

horticulture: current approaches <strong>and</strong> future prospects. Journal <strong>of</strong> Industrial<br />

Microbiology <strong>and</strong> Biotechnology 19, 188 - 191.<br />

91


Literature<br />

Zou, J., Rodriguez-Zas, S., Aldea, M., Li, M., Zhu, J., Gonzalez, D. O.,<br />

Vodkin, L. O., DeLucia, E. & Clough, S. J. (2005). Expression pr<strong>of</strong>iling<br />

soybean response to Pseudomonas syringae reveals new defense-related<br />

genes <strong>and</strong> rapid HR-specific downregulation <strong>of</strong> photosynthesis. Molecular<br />

Plant Microbe Interaction 18, 1161-1174.<br />

92

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!