AN ABSTRACT OF THE DISSERTATION OF Marion L. Brodhagen for the degree of Doctor of Philosophy in Molecular and Cellular Biology presented on June 30, 2003. Title: Pyoluteorin as a Signaling Molecule Regulating Secondary Metabolite Production and Transport Genes in Pseudomonas fluorescens Pf-5. Redacted for privacy Abstract Approved: Joyce E. Loper A major factor in the ability of Pseudomonasfluorescens Pf-5 to act as a biological control agent is its production of antibiotics, including pyoluteorin (PLT), 2,4-diacetylphloroglucinol (2,4-DAPG) and pyrrolnitrin (PRN). The data provided in this thesis demonstrate that the presence of any of these antibiotics in the extracellular milieu affects production of that same antibiotic, as well as others, by Pf-5. Amending the growth medium with antibiotics had multiple effects on secondary metabolism in Pf-5. i) PLT positively regulated its own production. ii) 2,4-DAPG positively regulated its own production. iii) PLT suppressed 2,4DAPG production. iv) 2,4DAPG inhibited PLT production. v) PLT suppressed transcription of a heterologous ferric-pyoverdine uptake gene. vi) PRN exerted a slight inhibitory effect on PLT gene transcription and production. PLT autoinduction by Pf-5 was extensively characterized, and was shown to require concentrations of exogenous PLT in the nanomolar range. These low concentrations are comparable to those of many molecules proposed to function in signaling roles. PLT served as a signal between distinct populations of cells within the rhizosphere, where it prompted autoinduction by those cells. Aside from effects of Pf5 antibiotics on one another, I also described the positive effect of exogenous PLT on expression of a set of transport genes flanking the PLT biosynthetic gene cluster. Sequence data and experimental evidence suggests that these genes encode a transport apparatus for PLT. The deduced amino acid sequences for four adjacent open reading frames together resemble Type I secretion apparatuses, which typically function in transport of proteins rather than secondary metabolites. The intact transporter genes are necessary for optimal PLT production. Taken together, the data from the studies described herein demonstrate that i) the production of PLT by Pf-5 can affect the production of PLT by neighboring cells, and ii) PLT and other exogenous secondary metabolites have both autoregulatory and cross-regulatory effects in culture. Because Pf-5 derivatives engaged in PLT crossfeeding in the rhizosphere, it is likely that cross-feeding occurs for other secondary metabolites as well. Thus, production of an antibiotic by one cell can profoundly affect secondary metabolism in neighboring cells occupying natural habitats. ©Copyright by Marion L. Brodhagen June 30, 2003 All Rights Reserved Pyoluteorin as a Signaling Molecule Regulating Secondary Metabolite Production and Transport Genes in Pseudomonasfluorescens Pf-5. by Marion L. Brodhagen A DISSERTATION submitted to Oregon State University in partial fulfillment of the requirements for the degree of Doctor of Philosophy Presented June 30, 2003 Commencement June, 2004 Doctor of Philosophy dissertation of Marion L. Brodhagen presented on June 30, 2003. APPROVED: Redacted for privacy Major Ptofessor, representing Molecular and Cellular Biology Redacted for privacy Director Molecular and Cellular Biology Program Redacted for privacy Dean of the graduate School I understand that my dissertation will become part of the permanent collection of Oregon State University libraries. My signature below authorizes release of my dissertation to any reader upon request. Redacted for privacy Marion L. Bro'hagen, Author ACKNOWLEDGEMENTS I would like to express my gratitude to members of the Loper Lab past and present who have made my tenure as a PhD student a truly wonderful time. Thank you for the guidance, the ideas, and most of all, the friendship that you shared. Thanks especially to Nancy Chaney, Marcella Henkels, Dr. Caroline Press, Brenda Shaffer, and Dr. Virginia Stockwell for technical suggestions at crucial times, for conversations that helped me understand the larger context of my research, for mentoring me at the bench and in my development as a scientist, and for sharing your broad perspectives - the knowledge you shared helped to make my time productive and rewarding, and the warmth and humor you shared made it fun. Thanks to Nathan Corbell, Cheryl Whistler, and Todd Temple for collegial interactions that have been so helpful to me, but also for camaraderie on late nights in the lab and elsewhere! I also am grateful to Meadow Anderson, Chelsea Meltzer, and Kevin Hockett for all the things they did to make life easy for me in the lab - and for the sparkle they bring to our lab! Special thanks go to Cara Taormina, Brie Sawtelle, and Amy Davis who have made research contributions to this thesis and who were each very much fun to collaborate with! Finally, the USDA Horticultural Crops Research Laboratory has been a very special place to spend my time as a PhD student, and I can't begin to list all the people that have befriended me here. I cannot imagine a more genial group! The same is true for my friends, colleagues, and mentors within the Botany and Plant Pathology department, the Molecular and Cellular Biology Program, and all over the OSU campus. From backyard bonfires to weekend hikes to much-needed coffee breaks, I could never have made it though this degree without the friendship and support of countless friends. Dr. Linda Thomashow generously provided me the opportunity to do research in her lab; I am very grateful for the wonderful month that I spent at Washington State University. Dr. Dmitri Mavrodi went out of his way again and again to make the time spent there a success, as did everyone else in the lab. I have many fond memories of my time there. Dr. Dan Arp also welcomed me into his lab, and Luis Sayavedra-Soto and Alisa Vangnai spent many days mentoring me there. I am grateful for the chance to have worked with so many warm people, doing exciting science and having fun! My committee, Dr. Dan Arp, Dr. Lynda Ciuffetti, Dr. Walt Ream, and Dr. Oscar Riera-Lizarazu, have guided me through this process and provided advice and encouragement at every step. My sincere thanks for all they have done to help me develop as a scientist. Finally, Dr. Joyce Loper has been a true mentor in every sense of the word, and a scientific and personal role model of the highest caliber. She has given me tremendous freedom, along with constant encouragement and support in all my endeavors. I am very grateful for all the opportunities that she created for me both inside and outside of her laboratory. Lastly, I am grateful for financial support provided by a fellowship from the Environmental Protection Agency Science to Achieve Results program. CONTRIBUTION OF AUTHORS Dr. Joyce Loper advised all of the work reported herein. Marcella Henkels was involved in the design and implementation of experiments to demonstrate pyoluteorin cross-feeding in the rhizosphere. This work is discussed in Chapter Two. Dr. Ian Paulsen provided privileged access to information about sequence quality from the region flanking pkJ in the unfinished Pf-5 genome being sequenced at The Institute for Genomic Research. This work is discussed in Chapter Four. TABLE OF CONTENTS Chapter 1. 1.1. Introduction. The Role of Secondary Metabolites in Biological Control and in the Physiology of R1iizosphere Pseudomonads ................. 1 Biological control of plant pathogens by microorganisms........................ 1 1.1.1. 1.1.2. 1.2. Antibiosis and its importance in biological control by Pseudomonasspp ...................................................................................... 3 1.2.1. 1.2.2. 1.2.3. 1.2.4. 1.2.5. 1.3. Limitations to biological control........................................................ 1 Mechanisms of biological control...................................................... 3 Evidence that antibiosis contributes significantly to biocontrol in the soil environment ....................................................................... 5 Regulation of antibiotic production in Pseudomonas spp .................. 9 Transport of secondary metabolites from cells................................ 14 Regulation of pyoluteorin production by P. fluorescens Pf-5, a model biocontrol organism ............................................................ 14 Biosynthesis of pyoluteorin.............................................................. 18 Secondary metabolites as signal molecules ............................................. 21 1.3.1. 1.3.2. Role of secondary metabolites in Pseudomonas spp ....................... 22 Existing support for the hypothesis that secondary metabolites function as signal molecules in Pseudomonas spp ........................... 23 1.3.3. Signaling properties of antibiotics in non-Pseudomonas species....24 1.3.4. Regulation of antibiotic production by other secondary metabolites ............................................................ 25 1.4. Statement of research objectives....................................................................25 1.5. References...................................................................................................... 28 TABLE OF CONTENTS (Continued) Page Chapter 2. Positive Autoregulation of the Antibiotic Pyoluteorin in the Biological Control Organism Pseudomonasfluorescens Pf-5 .................................. 43 2.1. Abstract....................................................................................................44 2.2 Introduction.............................................................................................45 2.3. Methods ................................................................................................... 47 2.3.1. 2.3.2. Bacterial strains, plasmids, and culture conditions.......................... 47 Influence of exogenous antibiotics on antibiotic production and PLT gene expression by Pf5 ............................................................. 50 2.3.3. Antibiotic quantification ..................................................................51 2.3.4. Transcriptional activity from PLT biosynthetic genes..................... 51 2.3.5. Assessment of PLT cross-feeding on roots ....................................... 52 2.3.6. Data analysis ....................................................................................53 2.4. Results ..................................................................................................... 53 2.4.1. 2.4.2. 2.4.3. 2.4.4. 2.4.5. PLT production is autoregulated...................................................... 53 PLT and 2,4-DAPG exhibit mutual inhibition................................. 58 Effects of PLT and 2,4-DAPG on PLT production occur at the transcriptional level .......................................................................... 62 PLT autoinduction is superseded by genetically and environmentally repressive backgrounds ......................................... 64 PLT autoinduction occurs in the rhizosphere .................................. 64 2.5. Discussion................................................................................................ 67 2.6. Acknowledgements ................................................................................. 71 2.7. References ............................................................................................... 72 TABLE OF CONTENTS (Continued) Chapter 3. Preliminary Evidence for Effects of Pyrrolnitrin, 2,4-Diacetyiphioroglucinol, and Pyoluteorin on Secondary Metabolism in Pseudomonasfluorescens Pf-5 ............................................................ 79 3.1. Introduction ............................................................................................. 79 3.2. Methods ................................................................................................... 80 3.3. Results and Discussion ............................................................................. 83 3.3.1. 3.3.2. 3.33. 3.3.4. Lack of influence by exogenous PLT on PRN................................. 83 Minimal influence by exogenous PRN on PLT production ............. 85 Autoinduction by 2,4-DAPG in P. fluorescens Pf-5 ........................ 87 Exogenous PLT represses expression of pyoverdine biosynthesis genes ................................................................................................. 89 3.4. Conclusions ............................................................................................. 92 3.5. References...............................................................................................94 Chapter 4. Membrane Transport Genes Play a Role in Pyoluteorin Production by Pseudomonasfluorescens Pf-5 ........................................................... 97 4.1. Introduction............................................................................................. 97 4.2. Methods .................................................................................................101 4.2.1. 4.2.2. 4.2.3. 4.2.4. 4.2.5. 4.2.6. 4.2.7. 4.2.8. 4.2.9. Bacterial strains, plasmids, and culture conditions........................ 101 DNA manipulations....................................................................... 105 RT-PCR.......................................................................................... 196 Nucleotide sequencing of pltH,pltI, and plt.J ................................. 107 Derivation of chromosomal p/tI: :inaZ and plt.J::inaZ transcriptional fusions by marker exchange mutagenesis .............. 108 Derivation of chromosomalpltF:: aacCl mutant by marker exchange mutagenesis.................................................................... 113 Antibiotic quantification................................................................ 114 Transcriptional activity from PLT biosynthetic genes................... 114 Statistical analysis .......................................................................... 115 TABLE OF CONTENTS (Continued) 4.3. Results ................................................................................................... 115 4.3.1. 4.3.2. Nucleotide sequence analysis ........................................................ 115 Phenotypic characterization of piti, p/tJ mutants ........................... 123 4.4. Discussion .............................................................................................. 138 4.4.1. 4.4.2. 4.4.3. 4.4.4. 4.4.5. PLT production requires functionality of transport genes flanking the PLT biosynthetic gene cluster .................................................. 138 Exogenous PLT induces expression of p/tI and p/tJ...................... 139 pitH ................................................................................................. 141 p/tI .................................................................................................. 143 pltJ .................................................................................................. 145 4.5. References ............................................................................................. 147 Chapter 5. Concluding Remarks............................................................................. 156 5.1. Repercussions of PLT production for the rhizosphere community....... 156 5.2. Functionality of bioactive compounds in a soil environment............... 158 5.3. Secondary metabolite production by Pf-5 is affected by the profile of metabolites already present ................................................................... 158 5.4. Site of perception of the PLT signal may be cytoplasmic or extracytoplasmic .................................................................................... 159 5.5 PLT accumulation requires a set of flanking transporter genes, which are in turn regulated by PLT ....................................................... 160 5.6. Significance ........................................................................................... 160 5.7 References..................................................................................................... 162 Bibliography ................................................................................................................ 164 Appendices .................................................................................................................. 187 LIST OF FIGURES Figure ig 1.1. Antibiotics produced by P. fluorescens Pf-5 that have been implicated in biological control......................................................................... 15 1.2. Factors known to influence pyoluteorin biosynthetic gene transcription in P. fluorescens Pf-5 ........................................................................................ 17 1.3. Proposed pyoluteorin biosynthetic pathway..................................................... 20 2.1. Structures of pyoluteorin (a) and 2,4-diacetylphloroglucinol (b) ..................... 46 2.2. Influence of exogenous PLT on PLT and 2,4-DAPG production byderivatives of Pf-5 ........................................................................................ 55 2.3. Minimum effective concentration of PLT for autoinduction in Pf-5 ................ 57 2.4. Influence of exogenous 2,4-DAPG on PLT production by derivatives ofPf-5 ............................................................................................................... 2.5. Minimum effective concentration of 2,4-DAPG for repression of PLT production in Pf-5 ..................................................................................... 61 2.6. Positive and negative influences by PLT and 2,4-DAPG, respectively, on transcriptional activity within the PLT biosynthetic gene cluster................ 63 2.7. Autoinduction of PLT as a result of cross-feeding in the rhizosphere ..............66 3.1. Influence of exogenous PLT and PRN on PRN produced by Pf-5 iceC ......... 84 3.2. Positive and negative influences by PLT and PRN, respectively, on transcriptional activity of the PLT biosynthetic gene, pltB .......................... 86 3.3. Influence of exogenous 2,4-DAPG on production of 2,4-DAPG by derivativesof Pf-5 ............................................................................................. 88 3.4. Negative influence by PLT on transcriptional activity of a ferric-pyoverdine uptake gene ........................................................................... 91 4.1. Map of PLT biosynthetic gene cluster and relationship to subclones used as templates for sequencing templates ...................................................... 99 4.2. Cloning strategy used to construct pJEL6300, which was used in marker-exchange mutagenesis to derive Pf-5 pltI::inaZ ................................. 110 4.3. Map and features of open reading frames flanking one end of the PLT biosynthetic gene region ......................................................................... 116 LIST OF FIGURES (Continued) Figure 4.4. Alignment of predicted amino acid sequence of PItH with that of the Ph1H sequence from P.fluorescens CHA0 ..................................................... 118 4.5. Alignment of predicted amino acid sequence of PitT with HIyD consensus domain ..................................................................................... 120 4.6. Alignment of predicted amino acid sequence of PI1J with the consensus ABC-binding cassette of ABC transporter proteins........................................ 122 4.7. PLT production by Pf-5 and bypltl::inaZ,pltJ::inaZ, and pltB::inaZ derivatives of Pf-5 ........................................................................ 124 4.9. Influence of exogenous PLT on PLT production by Pf-5, and by Pf-5 pltJ::inaZ and Pf-5 pltB::inaZ...........................................................128 4.10. Possible P1tR binding sites from the pltR-pltL and pitH-p/tI intergenic regions ............................................................................................................. 129 4.11. INA activity from pltB::inaZ (a), pitI::inaZ (b), and pltJ::inaZ (c) transcriptional fusions in wild-type (circles), pltF.:aacCl (triangles) orpitR::aacCl (squares) backgrounds ..................... 131 4.12. Positive influence by PLT on transcriptional activity from pltB::inaZ in a wild type Pf-5 background (a); pltJ::inaZ in a wild type Pf-5 background (b);pltJ::inaZin apltF background (c), and pit.J::inaZ in apltR background (d) .................................................................................. 134 4.13. Positive influence by PLT on transcriptional activity from pltB::inaZ in a wild type Pf-5 background (a); p/tI: :inaZ in a wild type Pf-5 background (b); p/tI: :inaZ in a pltF background (c), and pltI::inaZ in apltR background (d) .................................................................................. 136 LIST OF TABLES Table Page 1.1. Antibiotics produced by Pseudomonas spp. and implicated in biological control................................................................................................................. 4 2.1: Bacterial strains and plasmids used in this study.............................................. 49 3.1. Bacterial strains and plasmids used in this study .............................................. 82 3.2. Influence of exogenous PLT on PRN production by derivatives of Pf-5 ......... 83 4.1. Bacterial strains and plasmids used in this study............................................ 102 4.2. PCR primers used in this study....................................................................... 106 LIST OF APPENDICES Page Appendix 1. Overexpression of PItR in Escherichia coli........_......................... 188 Al.l. Introduction ................................................................................... 188 Al .2. Characterization of P1tR ................................................................. 190 Al.2.l. A transcriptional reporter for pltR........................................... 191 A1.2.2. Overexpression of P1tR in E. coli ........................................... 193 A1.3. Future recommendations................................................................ 200 Al .4. Acknowledgements ........................................................................ 201 Al.5. References ...................................................................................... 202 Appendix 2. Nucleotide and Amino Acid Sequence for pitH, piti, pltJ, orJK, orJP, orfQ, orfU, orJX, and orf)'..................................................... 204 A2. 1. Predicted amino acid sequences and positions of open reading frames in nucleotide sequence ........................................................ 204 A2.2. Relevant nucleotide sequence ........................................................ 207 LIST OF APPENDIX FIGURES Figure Al .1. Proposal for PItR/PLT interaction at PLT promoters ..................................... 189 Al .2. Proposal for dual binding by PItR of PLT (co-activator) and 2,4-DAPG (co-repressor).................................................................................................. 190 LIST OF APPENDIX TABLES Table Page Al .1. Expression of P1tR from pTYB 1 in E. coli B ER2566 cells, as visualized byimmunoblotting .......................................................................................... 198 Al.2. Expression of PItR in various hostivector combinations ................................ 199 1642) - (1564 Galilei Galileo them. discover to is point the discovered; are they once understand to easy are truths All Pyoluteorin as a Signaling Molecule Regulating Secondary Metabolite Production and Transport Genes in Pseudomonasfluorescens Pf-5. Chapter 1: The Role of Secondary Metabolites in Biological Control and in the Physiology of Rliizosphere Pseudomonads The feeding of our burgeoning human population places ever-greater demands on agriculture for increased crop yields. Although the past century has certainly seen revolutionary advances in the efficiency of agricultural crop production, some advances have come at the expense of ecological sustainability or environmental health. For example, affordable, organic pesticides introduced in the 1940's were powerfully effective at stemming crop yield losses due to plant diseases. However, the drawbacks of widespread use of such toxins were soon realized, as demonstrated by the prompt enactment of the Federal Insecticide, Fungicide, and Rodenticide Act (FIFRA) in the United States in 1947. The struggle for effective plant disease control without riskato environmental or human health continues: the FIFRA has been recently updated by the Food Quality and Protection Act (FQPA) of 1996, which requires the US Environmental Protection Agency to reevaluate the toxicity of pesticide residues to human health based on cumulative exposure from multiple sources and to groups of pesticides with similar mechanisms for toxicity (http://www.epa.gov/opppspsl /fqpa/). Concerns about environmental health also develop along with our understanding of the impact of chemical usage: the Montreal Protocol of 1991 requires that usage of methyl bromide, a broadly effective and widely used pesticide, be completely discontinued in developed nations by 2005 because of its adverse effects on the Earth's ozone layer. As a result, alternate control measures must be found to replace methyl bromide use in over 100 crops, for control of plant pathogens, nematodes, weeds, and insects (http://w-ww.ars.usda.gov/is/mb/mebrweb.htm). The status of synthetic pesticides in agriculture is suffering from such increasingly restrictive legislation governing pesticide use (10, 36, 54,121), and from concerns of growers and consumers alike about effects of agrochemicals on the environment and on human health (25, 54). Tn addition, the use of synthetic pesticides can lead to unwanted development of resistant populations (114,121) and can have adverse effects on non-target organisms (e.g. 38, 40). Moreover, despite modem agricultural practices and technologies, Agrios estimated in 1988 (3) that approximately 12% of the potential yield for a given agricultural crop is lost to plant disease before harvest, and up to another 20% may be lost after harvest. Thus, there is room for improvement in both the efficacy (see also reference 9) and safety of plant disease control. 1.1. Biological control of plant pathogens by microorganisms. For the past two or three decades, biological control organisms have been a focus of academic study and have increasingly gained commercial favor as viable alternatives to synthetic pesticides in combatting plant disease (24, 43, 53, 83, 89,102, 135). Biological control, as defined by Cook (23), is the control of one organism by another and implies the reduction in either pest populations or the severity of plant disease. Biological control organisms have the following advantages over synthetic pesticides, in the control of plant pathogens: 1) greater safety to human health; 2) minimal potential for known harmful environmental effect; 3) low likelihood of inducing resistance by target organisms; and 4) possibility for combination with conventional (chemical) controls, if necessary for greater efficacy (9, 23, 54). 1.1.1. Limitations to biolo2ical control. As with chemical control, biological control has its share of shortcomings. In a market sense, these include a brief shelf life (122), the necessity for a lag time for establishment (and thus a less immediate effect than that of chemical controls), incomplete control of pathogen populations, and low profit margins for producers and users alike (9). There is also a persistent problem of variable performance in the field (9, 101,130,134) which may stem from any one, or a combination of, a number of biological factors. These factors can include poor or patchy colonization, loss of ecological competence of the biocontrol strain, susceptibility to plant and soil influences, failure to express biocontrol traits, phytotoxicity of secreted metabolites, tolerance of target organisms to secreted metabolites, and influx of non-target pathogens due to alteration of the environment by the biocontrol agent (9, 50). Understanding the mechanisms by which microorganisms exert biological control, and the environmental constraints on the success of those mechanisms, can lead to adoption of strains and application strategies with improved rates of biocontrol success. 1.1.2. Mechanisms of biological control. Biological control may be exerted by biological control agents directly or indirectly. Direct effects result from i) competition for space, or for nutrients such as C, N, or Fe; ii) antagonism of the pathogen by parasitism or predation, and/or iii) antibiosis. Biological control may also be the result of an indirect interaction with the pathogen, via the plant. In this case, physiological activities of the biological control agent trigger plant host defense responses, which thwart the progress of the pathogen (e.g. reviewed in 52, 130, and 135). 1.2. Antibiosis and its importance in biological control by Pseudomonas spp. Of the mechanisms described above, the work described in this dissertation is focused solely on antibiosis. Antibiotic production is important to the success of many biological control agents. Examples include the production of zwitteromycin A and kanosamine production by Bacillus cereus (123,124), oligomycin A production by Streptomyces libani (74), xanthobaccins from Stentrophomonas spp. (93), and herbicolins 0 and I from Pantoea agglomerans (69). Among fluorescent Pseudoinonas spp., possibly the largest, best characterized, and most promising group of biocontrol bacteria (101,130), the importance of antibiotic production has been especially well highlighted. Pseudomonas spp. produce an array of secondary 4 metabolites, many of which have been studied in the context of their utility as antibiotics that inhibit microorganisms pathogenic to agricultural crops (Table 1.1). For brevity, I will limit my discussion to four well-characterized Pseudomonas antibiotics - phenazine, 2,4-diacetylphloroglucinol, pyoluteorin, and pyrrolnitrin. Table 1.1. Antibiotics produced by Pseudomonas spp. and implicated in biological control. Antibiotic Producing Target Pathogen Host Plant species Selected Reference Anthranhlate P. aeruginosa Fusarium oxysporum f sp. ciceris chickpea 5 2,4- P. fluorescens Gaeumannomyces grainini var. tritici tobacco 71 P. fluorescens Thielaviopsis basicola tobacco 71 P. fluorescens Pythium ultimum sugar beet 41 P. aurantiaca Fusarium oxysporum wheat 106 Hydrogen cyanide P. fluorescens Thielaviopsis basicola tobacco 133 Oomycin A P. fluorescens Pythium ultimum cotton 68 Phenazine-1-carhoxamide P. chloraphis Fusarium oxysporum f. sp. radicislycopersici tomato 22 Phenazine-1-carboxylic acid P.fiuorescens Gauemannomyces gramini var tritici wheat 128 P. aureofaciens Gauemannomyces gramini var tritici wheat 108 Pyocyanin P. aeruginosa Seploria tritici wheat 42 Pyoluteorin P.fluorescens Pythium ultimum cotton 67 P. fluorescens Pythium ultimum cress 84 P. fluorescens Rhizoctonia solani cotton 67 P. cepacia Aphanoinyces cochliodes sugar beet 64 P. fluorescens Pyrenophora tritici-repentis wheat 105 P.fluorescens Scierotinia homoeocarpa bluegrass 112 P. fluorescens Pythium ultimum sugar beet 131 diacetylphloroglucinol Pyrroluitrin Viscosinainide From Dowling and O'Gara, 1994; Camacho, 2001; Hans and Keel, 2003. 1.2.1. Evidence that antibiosis contributes siviificantly to biocontrol in the soil environment. The production of antibiotics by soilborne bacteria has long been postulated to have evolved as a defense mechanism, similar to allelopathy in plants. That antibiosis truly plays a significant ecological role in the soil environment has been supported by multiple lines of evidence, as described below. 1.2.1.1. Inhibition of a pathogen by a pure antibiotic. In most cases, the establishment of an antibioticts contribution to biological control originates with evidence that a biologically active constituent is produced in vitro, has been isolated from a producing bacterium, and has been chemically characterized. Inhibition by the compound is then established on agar plates or in rhizosphere assays during which the pathogen is presented with the compound in its purified form. In this manner, phenazine derivatives produced by P. fluorescens and P. chioraphis have been shown to suppress Gaeumannomyces graminis var. tritici, which causes take-all disease of wheat (48), and Fusarium oxysporum f. sp. radicislycopersici, the causal agent of root rots for greenhouse plants such as tomato (22). Similarly, pure pyrrolnitrin (a compound produced by P. fluorescens and related bacteria) added to cotton seeds prior to planting inhibited the cotton seedling pathogens Rhizoctonia solani, Thielaviopsis basicola, Alternaria sp., and Verticillium dahliae (66). Agar plate assays were used to demonstrate that 2,4diacetyiphioroglucinol, a compound produced by P. fluorescens and P. aureofaciens, is an important factor in antagonism of T. basicola, which causes black root rot of tobacco, (70, 71), Pythium ultimum, which causes damping-off of sugar beet seedlings (71, 86), and the wheat take-all pathogen, G. g. tritici (55, 71). Pyoluteorin, a compound produced by P. fluorescens, inhibited P. ultimum on agar plates (86) and on cotton and cress seedlings (67, 84). Bacteria are thought to inhibit plant pathogens by a number of mechanisms employed simultaneously, as described above. Thus, while these studies demonstrate that antibiotics can suppress both pathogens and plant disease, they do not prove a requirement for antibiosis during biocontrol in the rhizosphere, when other mechanisms may also contribute in varying degrees to pathogen suppression. 1.2.1.2. Comparison ofbiocontrol efficacy between antibiotic-deficient mutants and wild type strains In a number of systems, genetic studies have extended the evidence described above to support an integral role for antibiosis in biological control. Pathogen inhibition by strains bearing genes necessary for antibiotic biosynthesis has been compared with pathogen inhibition by strains deficient in such genes. For example, a phenazine-1-carboxylate producing P. fluorescens 2-79 suppresses G. g. var. tritici, while a Tn5 mutant derivative deficient in phenazine-1-carbooxylate production does not suppress the pathogen (129). P. chioraphis PL1391, which produces the derivative phenazine- 1 -carboxamide, inhibited F. oxysporum in the tomato rhizosphere, but a non-producing mutant did not (22). Similarly, Hill et al. (61) and Pfender et al. (105) demonstrated that disruption of genes necessary for pyrrolnitrin biosynthesis diminished biological control of R. solani and Pyrenophora tritici-repentis, respectively, by P. fluorescens strains. Likewise, genes for 2,4diacetylphloroglucinol biosynthesis were required for optimal suppression by P. fluorescens strains of the pathogens 7'. basicola, P. ultimum, and G. g. var. tritici, and F. oxysporum (41, 70, 71, 86). Genetic analysis has also proven useful in determining whether in vitro observations about pathogen inhibition could be extrapolated to the rhizosphere environment. P. fluorescens 2-79 inhibited F. oxysporum in agar plate assays, but the in vitro effect was not observed in the tomato rhizosphere, presumably because strain 2-79 does not produce phenazine-1-carboxamide, the phenazine derivative most effective at suppression ofF. oxysporum (22). In the case of pyoluteorin, antibioticproducing strains suppressed P. ultimum in agar plate assays (75) and protected cress 7 against P. ultimum while pyoluteorin-deficient mutants did not (75,84). However, the pyoluteorin-deficient mutants were not compromised in their ability to inhibit P. ultimum infection on cucumber seedlings (75, 84). Later, it was shown that when P. fluorescens colonized cucumber seeds, pyoluteorin biosynthesis genes were expressed late, after Pythium infection was expected to have already occurred(76). Thus, while the presence or absence of antibiotic genes allows correlation of antibiotic production with biological control, tracking the nuances of expression of those genes in the rhizosphere may reveal sources of inconsistencies in the contribution of antibiosis to biological control. 1.2.1.3. Expression of antibiotic genes in the natural environment. Expression of antibiotic genes in the rhizosphere is predicted if they indeed play a role in antagonism of plant pathogens on the root surface. Reporter gene systems have been used to test this hypothesis in various antibiotic-producing biological control agents. For example, genes for phenazine (22, 140), 2,4diacetylphloroglucinol (96, 97), and pyoluteorin (16, 76) biosynthesis are expressed in the rhizosphere. Further, environmental factors such as plant species and fungal exudates have been shown to modify antibiotic gene expression (e.g. 75, 97). 1.2.1.4. Isolation ofantibiotics from soil. Because of the nature of soil and the tendency of hydrophobic secondary metabolites to become immobilized by soil constituents such as clay particles and humic substances, extraction of antibiotics from soil is a daunting feat. Yet, antibiotic extractions from the rhizosphere provide the best direct evidence that antibiotics are produced (and, by extrapolation, function antagonistically) in their natural habitat. 1.2.1.4.1. Phenazine. Phenazine was extracted from soil surrounding wheat roots inoculated with P. fluorescens 2-79 and P. aureofaciens 3 0-84, strains suppressive to take-all decline of wheat. The antibiotic was extracted from natural soil in growth chamber experiments and from field soil. A positive correlation was observed between phenazine recovery and suppression of take-all disease in the presence of G. g. tritici (128). h a separate study, P. putida WCS358r, or derivatives modified to contain the genes for phenazine biosynthesis, were inoculated onto wheat seeds that were planted into a field plot for two consecutive years. Soil surrounding roots of mature plants was extracted and found to contain phenazine, and the authors suggest that varying amounts of phenazine extracted from each strain might have accounted for different influences of each strain on the surrounding fungal microflora (44). 1 .2.1 .4.2. 2,4-diacetylphloroA'lucjnoL The polyketide 2,4-diacetylphloroglucinol has been extracted from gnotobiotic soil (clay, sand, and quartz powder) in which wheat seedlings inoculated with P. fluorescens CHAO were grown. Similar amounts of the compound were recovered in the presence or absence of the take-all pathogen of wheat, G.g. tritici (71). In later experiments, methods were developed to extract 2,4diacetylphloroglucinol from raw, virgin soils in which wheat plants colonized with P. fluorescens Q2-87 were grown (14). Finally, it was demonstrated that 2,4diacetylphloroglucinol was produced by naturally occurring fluorescent pseudomonads colonizing wheat roots in their native, take-all suppressive soils, but the compound was not produced by the same isolates when they colonized wheat roots grown in take-all conducive soils (110). Suppressive soils, which naturally develop resistance to a pathogen after repeated seasons of planting with the same crop, have been a source of inspiration for biological control upon which many research dollars and hours have been expended. By extrapolation, this work of Raaijmakers et al. (110) demonstrated that production of a pathogen-inhibiting antibiotic is one factor that delineates a suppressive from a non-suppressive soil. 1.2. 1. 4. 3. Pyolu(eorin and pyrrolnilrjn. Pyotuteorin and 2,4-diacetyiphioroglucinol together have been extracted from soil surrounding the roots of wheat seedlings inoculated with P. fluorescens CHAO in a gnotobiotic system. The presence of the damping-off pathogen P. ultimum appeared to increase levels of both compounds recovered from the wheat rhizosphere, although the reason for this is unknown. It is hypothesized that pathogen attack of wheat roots might release nutrients conducive to antibiotic production (85). Pyrrolnitrin has also been extracted from soil surrounding Pseudomonas-colonized roots of cotton and sugar beet seedlings, and from barley seeds (130, and references therein). 1.2.2. Regulation of antibiotic production in Pseudomonas spp. A complex circuitry governs production of Pseudomonas antibiotics, and includes factors ranging from sigma factors to noncoding RNAs. Examples are described below. 1.2.2.1. GacS/GacA. Two-component regulatory systems are protein pairs in which a histidine protein kinase located in the cell exterior or the cytoplasm senses signals, and uses a phosphorylation cascade to relay signal information to a response regulator that activates transcription from promoters of genes encoding appropriate physiological responses. The GacS/GacA two-component regulatory system is widespread among gram-negative bacteria and frequently controls the production of exoproducts. These exoproducts are often important for interactions with a eukaryotic host; therefore, 10 gacA or gacS mutations result in phenotypes associated with the host such as reduced biocontrol ability and reduced virulence. Excellent reviews of the GacS/GacA regulatory cascade and molecular mechanisms are available (50, 56). Of the Pseudomonas antibiotics important for biocontrol listed in Table 1.1, 2,4diacetyiphioroglucinol, pyrrolnitrin, pyoluteorin, phenazine, and hydrogen cyanide are known to be under the positive control of GacS/GacA. There is strong evidence that the expression of GacA-regulated exoproduct genes is post-transcriptional. First, promoter elements preceding GacA-regulated exoproduct genes that bind GacA have not been identified (26, 50), demonstrating that control of expression of these genes by the GacA transcriptional regulator is indirect. Second, in the case of genes for biosynthesis of HCN in P. fluorescens CHAO, GacS/GacA regulation of hcnA expression does not require the hcnA promoter. Rather, GacA control requires a region surrounding the hcnA nbosome binding site (11, 13, and 56), implying that the GacA effect is posttranscriptional. Whether GacA posttranscriptionally regulates all the exoproduct genes under its jurisdiction remains to be determined. 1.2.2.2. Regulatory RNAs. Untranslated RNAs have important functions in both eukaryotes (e.g. 28) and bacteria (e.g. see reference 95). The functionality of these small RNAs has been gaining so much recognition that databases of protein families have been joined by databases for non-coding RNA families (e.g. http://rfam.wustl.edu; http://www.imb- jena.deIRNA.html). In bacteria, untranslated RNAs can perform various functions by binding to mRNAs (e.g. 4, 78, 79), to phospholipid bilayers (73), and to proteins (e.g. 141). For example, the binding of untranslated RNA to a translational repressor, RsmA, is part of a regulatory cascade governing exoproduct formation (e.g. 20, 30, 31, 80, and 92) among the Enterobacteriaceae (29) and pseudomonads (12, 104). The RsmAIRsmB system has been well characterized in Erwinia carotovora subsp. carotovora. In. E. carotovora, the small protein RsinA inhibits the production of various extracellular products by stimulating transcript degradation, presumably 11 because RsmA binds exoproduct mRNAs in a destabilizing manner (29). However, RsmA inhibition is opposed by a small untranslated RNA, RsmB, which entangles RsmA in an inactive nucleoprotein complex (81). Homologs of RsmA and RsmB have been found in P.fluorescens; while RsmA sequences are highly conserved among strains, the nucleotide sequences of the RsmB homologs are not. However, the function of RsmA binding appears to remain constant. In P. fluorescens, the RsmB homolog has been called either RsmZ (57) or PrrB (1). Like their counterparts in Erwinia spp., RsmA and RsmZ control exoproduct formation. The RsmA homolog is a translational repressor in P. fluorescens (12) and RsmZ appears to function by sequestering RsmA or a homologous protein away from mRNA. This is accomplished by RNA secondary structures containing multiple stem-loops bearing ribosome-binding site-like AGG(G)A motifs, which presumably occupy RNA-binding sites on RsmA homologs. RsmZ expression is under the control of GacA, and although GacA-controlled exoproduct genes do not require GacA for transcription, the rsmZ gene from P. fluorescens CHAQ, which encodes an untranslated regulatory RNA, is dependent on GacA for transcription. Expression of RsmZ requires an element upstream of the rsmZ transcriptional start, implying that GacA must exert direct or indirect transcriptional control over RsmZ expression (57). Another untranslated RNA, RsmY, has recently been discovered in P. fluorescens CHAO; RsmY displays a similar secondary structure to RsmZ and is under GacA control (50). 1.2.2.3. Sigma factors. Sigma factors constitute a pooi of subunits for RNA polymerase that, when bound to the core polymerase, complete the holoenzyme. The identity of the bound sigma factor confers a primary level of transcriptional regulation, by directing the polymerase to bind to a specific subset of promoters in the cell. In P. fluorescens, the stationary phase and stress response sigma factor RpoS (c?) represses the production 12 of pyoluteorin, 2,4-diacetyiphioroglucinol and HCN, but is required for pyrrolnitrin production (116, 137). In contrast, the housekeeping sigma factor RpoD (a70) positively affects pyoluteorin and 2,4-diacetylphloroglucinol production (118). c? is best described in E. co/i, where it confers transcriptional specificity for a subset of genes expressed predominantly during stationary phase or during nutrient, osmotic, oxidative, or other stresses (37, 58, 59, 94, and 126). In Pf-5, & is similarly implicated in tolerance to osmotic and oxidative stresses (116), and is partially under the control of the histidine protein kinase/response regulator pair GacS/GacA (137). 1.2.2.4. Pathway-linked regulatory genes. Often, genes for biosynthesis of secondary metabolites are clustered, and associated with these clusters are genes for regulation of expression of the same genes. Among the compounds listed in Table 1.1, pathway-linked regulators have been especially well characterized for the phenazine and 2,4-diacetylphioroglucinol gene clusters. The phenazine biosynthetic genes are flanked by phzl and phzR, which encode LuxIILuxR homologs involved in the synthesis and perception, respectively, of an AHL signal (108). Although the phenazine biosynthetic pathway is responsive to this quorum-sensing system, PhzI/PhzR do not appear to influence production of other secondary metabolites (50) in phenazine-producing strains. The 2,4diacetyiphioroglucinol gene cluster contains two TetR-type transcriptional regulators, ph/F and phlH. A function for the latter has not been reported, but ph/F has been characterized. In P. fluorescens, Ph1F binds repressively to the ph/A CBD promoter in the absence of 2,4-diacetyiphioroglucinol, but dissociates from its target sequence in the presence of 2,4-diacetylphloroglucinol (2). The pyoluteorin gene cluster contains the linked gene pltR, encoding a LysR-type transcriptional regulator. The pltR gene is necessary for transcription of genes within the cluster (100). A DNA hybridization array showed that effects of PItR on gene expression are not limited to the pyoluteorin biosynthetic genes (Caroline Press, unpublished data), although it is unclear whether these pleiotropic effects are direct or indirect. Another gene within the pyoluteorin 13 biosynthetic gene cluster, pitH, encodes a member of the TetR transcriptional regulator family (Chapter 4). No pathway-linked regulator has been found for the pyrrolnitrin gene cluster, but the sequence of the genome of P. fluorescens Pf-5 may reveal candidate regulatory genes for this and other previously less-characterized pathways. 1.2.2.5. Quorum sensing. Extracellular concentrations of AHLs appear to govern production of various antibiotics byP. aeruginosa, P.fiuorescens, and P. aureofaciens (e.g. 39, 107, 139, and 140). Although AHLs have not been implicated in production of the majority of antibiotics listed in Table 1.1, phenazine production is under the control of quorum sensing in P. aureofaciens and P. chiororaphis (21, 140). Additionally, AHLs can mediate cross-talk among bacterial populations in the rhizosphere environment, orchestrating the production of phenazines (107). 1.2.2.6. Autoregulation of antibiotic and secondary metabolite production. Another layer in the regulation of Pseudomonas secondary metabolites is autoregulation. For example, 2,4-diacetylphloroglucinol, which is produced by many pseudomonads (e.g. 72) including Pf-5 and CHAO (71), is positively autoregulated in CHAO (120; discussed further in Chapter 2). Oomycin A is produced byP. fluorescens Hv37aR2 and has been shown to suppress Pythium on cotton and other crops. A lacZ transcriptional fusion to afuE, a gene required for oomycin A biosynthesis, was highly expressed in a wild type background while little or no expression could be detected when the fusion was introduced into an oomycin A- deficient mutant (49, 68). The phenazine operon experiences positive feedback through PhzJlR autoregulation (51). Other examples of positive autoregulation by extracellular metabolites in Pseudomonas include those of the siderophores pyochelin (111) and pyoverdine (77) in P. aeruginosa. In the latter case, binding of extracellular 14 pyoverdine to the ferripyoverdine receptor protein FpvA in the outer membrane initiates a regulatory cascade resulting in increased production of pyoverdine itself, and of two other secreted virulence factors, exotoxin A and an endoprotease (PrpL). 1.2.3. Transport of secondary metabolites from cells. Although many of the compounds described in Table 1.1 are known to be extracellular, the mode of escape from the cell is known for few. The 2,4diacetyiphioroglucinol gene cluster in P. fluorescens Q2-87 contains phiE, a gene encoding a putative integral membrane permease (7). The phiE gene resembles sugar transporters; its role in 2,4-diacetyiphioroglucinol transport has not been reported, but intact phlE is required for optimal 2,4-diacetylphloroglucinol production (6). 1.2.4. Regulation of pyoluteorin production by P. biocontrol organism. P. fluorescens Pf-5 is a fluorescens Pf5g a model model biocontrol organism. It has been a focus of study since its isolation from the rhizosphere of cotton seedlings over two decades ago (66). It produces a suite of secondary metabolites, including HCN, pyrrolnitrin, 2,4diacetyiphioroglucinol, and pyoluteorin, that contribute to biocontrol of phytopathogens. The associations of pyrrolnitrin and pyoluteorin with disease suppression represent two of the earliest assignments of biochemical mechanisms to biocontrol (66, 67). The sequencing of the genome of Pf-5 is currently nearing completion at The Institute for Genomic Research (http://www.tigr.org/tdb/ufmgl). Because Pf-5 shares biochemical traits with other biocontrol pseudomonads, valuable comparisons can be made and insights gained about their genetic underpinnings. One such trait is the production of antibiotics, which is especially well characterized in Pf5 and the similar strain P. fluorescens CHAO. The studies described in this dissertation further our understanding of factors regulating antibiotic production in Pf5, with a specific focus on pyoluteorin. 15 Hydrogen cyanide HCEN OH Pyoluteorin EEIII;L11JiII>ct OH o H OH 2,4-diacetyiphloroglucinol CI Pyrrolnitrin Fig. 1.1. Antibiotics produced by P. fluorescens Pf-5 that have been implicated in biological control. In Pf-5, pyoluteorin production is affected by external and intracellular conditions. Susceptibility of pyoluteorin production to environmental influences contributes to variability in the overall efficacy of biocontrol. Pyoluteorin production in Pf-5 is affected by nutrient source (76, 99; Brodhagen and Loper, unpublished data). Pyoluteorin production is also affected by osmotic potential (115), temperature (98), and cell density (32). The links between extracellular conditions and intracellular regulatory pathways are not yet well understood. However, a complex picture is emerging of the intracellular regulatory circuit in Pf-5 that governs production of pyoluteorin and other secreted products, and links the production of these exoproducts to the physiological status and stress response of the cell (Fig. 1.1). For example, the sigma factors RpoS (&) and RpoD (a70) influence pyoluteorin production as described above. The repression of pyoluteorin production when the stationary phase/stress response/starvation sigma factor as is present implies that the antibiotic is not made under conditions of limited nutrient supply. Indeed, pyoluteorin production is low in minimal media when compared to production in complex and rich 16 media (unpublished data). The serine protease Lon, which accumulates during heat shock and degrades unstable, nonfunctional, or denatured proteins (46), represses pyoluteorin production. Heat shock induces Lon accumulation in Pf-5. Moreover, the ion gene is preceded by heat shock sigma factor (ar') recognition sequences, and like & and a70, H plays a role, albeit indirect, in pyoluteorin biosynthetic gene expression (138). PtsP, a paralog of Enzyme I of the sugar phosphotransferase system, also represses pyoluteorin production (136). The GacS/GacA pair positively regulates production of antibiotics in Pf-5, including pyoluteorin, 2,4- diacetyiphioroglucinol, pyrrolnitrin, hydrogen cyanide (27, 75). In addition to the global regulatory proteins that affect transcription of pyoluteorin biosynthetic genes, more localized regulatory elements also exist. A cluster of ten genes is necessary for pyoluteorin biosynthesis, and transcriptional activity from all promoters within the cluster requires the linked gene pltR, as described above. In the biochemically similar strain P. fluorescens CHAO, pyoluteorin is also produced and is subject to regulation by additional intracellular and extracellular factors. These factors include negative regulation by pyrrolquinoline quinone (119), a cofactor necessary for activity of glucose and alcohol dehydrogenases, and both positive and negative influences on pyoluteorin production by a number of mineral salts (34). These regulatory mechanisms are thought to function similarly in Pf-5, although that assumption has not been verified. The regulatory mechanisms described above could arguably provide mechanisms to Pf-5 for "considering" the physiological status of the cell for regulation of pyoluteorin production. Thus, in a broad sense, pyoluteorin, and perhaps other antibiotics, may be thought of as carrying information about the physiological status of a cell to the outside environment, and perhaps ultimately, to neighboring cells. 17 Starvation Unknown signal jr Heat shock I Carbon Source Temperature Fig. 1.2. Factors known to influence pyoluteorin biosynthetic gene transcription in P. fluorescens Pf-5. 18 1.2.5. Biosynthesis of pyoluteorin. A cluster of genes required for pyoluteorin biosynthesis has been identified, and gene functions have been predicted by sequence analysis, although these predictions remain to be verified biochemically (100). 1.2.5.1. Organization ofgenes within the pyoluteorin biosynthetic gene cluster. The pyoluteorin biosynthetic genes (Fig. 1.3) are contained on at least three separate operons. The divergent transcription of pltL and pltR implies the existence of separate promoters between these genes. There is evidence thatpltR andp/tMdo not form a single operon, as transcription of pltM is not correlated with that of pltR (Chapter 4). At least two promoters drive transcription of the p1tLABCDEFG gene cluster. One of these has been shown experimentally to reside upstream of pltB (76), presumably in the pltR-pltL intergenic region; another has been shown, via activation in trans of a promoterless reporter gene, to precede eitherpltD orpltE (76). 1.2.5.2. Predicted functions ofgenes within the pyoluteorin biosynthetic gene cluster. The predicted functions of eight structural genes in the pyoluteonn biosynthetic gene cluster are sufficient to account for pyoluteorin biosynthesis (100). Briefly, pltB and pltC encode type I polyketide synthases (PKSs) and are predicted to be involved in synthesis of the resorcinol (phenolic) moiety of pyoluteorin. The precursor for the growing polyketide chain is likely to be proline, and the acyl-CoA synthetase P1tF is predicted to activate this molecule. Neither PltB nor PItC contain a thioesterase domain, which is normally necessary for cleavage of the growing polyketide chain from the PKS. However, pltG is predicted to encode this function. The pyrrole (proline-derived) moiety of pyoluteorin is activated by PltF as described above, and modified by PItE. The pilE gene encodes an acyl-CoA dehydrogenase 19 thought to perform oxidation of the pyrrole ring prior to its incorporation as a precursor for the polyketide ring. By analogy with homologous gene products, PltE activity probably results in double bond formation adjacent to the ketone linkage of the final pyoluteorin product. The pyrrole ring in pyoluteorin is aromatic; it is thought that the second oxidation may occur spontaneously. Finally, pitA, pltD, and pltM encode halogenase enzymes. Because pyoluteorin exhibits only two chlorination sites, the existence of three chlorinating enzymes is enigmatic. However, PltD lacks a predicted NADH cofactor-binding site, and may therefore lack catalytic activity. Nevertheless, a mutation inpltD resulted in loss of pyoluteorin production. A role for pitL has not been predicted. 20 a jItB/G H HQJ7 !!!* FItB/C/ PItE . OH OH çç_ç HO 0 PItAITj H a (&X0 HO 0 H Pyoiuteorb çc OH HO 0 H p1 tM pltL 1 p1W 1kb Gene pliM pliR pilL p/IA puB p/IC p/ID pilE p/IF puG Putative protein function halogenase transcriptional regulator unknown halogenase polyketide synthase polyketide synthase halogenase acyl-00A dehydrogenase adenylating enzyme thioesterase Fig. 1.3. Proposed pyoluteorin biosynthetic pathway. From Nowak-Thompson et al. (100). 21 1.3. Secondary metabolites as signal molecules. Primary metabolites are, by definition, necessary for survival of the producing cells and therefore understood to play vital roles in the cell's physiology. In many cases, these roles are quite well studied; it is also becoming clear that many metabolites are not limited to a single role within the cell but play parts in various reactions. On the other hand, secondary metabolites do not appear to be necessary for survival to bacterial cells, and the impetus behind the production of these energetically expensive and frequently toxic molecules remains, in many instances, enigmatic. As described above, the role of antibiotic has been assigned to certain secondary metabolites based on their demonstrated ability to suppress the growth of organisms that might potentially co-inhabit their environment, and compete for niches or nutrients. Indeed, phenazine antibiotics have been shown to contribute to ecological fitness of Pf-5 in natural soils, but not in pasteurized soils in which resident microfiora have already been destroyed (87). However, not all Pseudomonas antibiotics have been shown to contribute to rhizosphere competence. A short term (12 day) experiment suggested a lack of a role for oomycin A in competence of P. fluorescens in cotton rhizospheres (68), but a longer colonization period may be necessary to see differences, as demonstrated for phenazine (87). Rhizosphere populations of wild type and 2,4-diacetylphloroglucinol-deficient strains of P. fluorescens Fl 13 remained similar in natural soils containing sugarbeets over a period of 270 days (18). Thus, a contribution to ecological fitness does not explain the selection pressure to retain 2,4- diacetylphioroglucinol biosynthesis genes. Still, these genes are highly conserved among Pseudomonas spp. collected from locations all around the world (72; L. de la Fuente, personal communication). Are there additional roles for antibiotics beside that of competition with rhizosphere neighbors? Many other roles for secondary metabolites have been suggested. Secondary metabolites have been postulated to serve as metal transporters, 22 metabolic waste or detoxification products, byproducts of primary metabolism, ancient or extant effectors of catalytic RNA reactions, reactants with cellular macromolecules (e.g. DNA and cell wall components), and providers of a pool of "biochemical diversity" for the cell (to name a few). Secondary metabolites have also been suggested to serve as signal molecules. Signaling roles are expected to include those essential to eukaryote-bacterium interactions, as well as those involved in communication between neighboring bacterial cells (19). 1.3.1. Role of secondary metabolites in Pseudomonas spp. Based on the discussion above, it is possible that metabolites such as those listed in Table 1.1 may have roles additional to that of extracellular toxins that contribute to the overall physiology of the bacteria. It is apparent that production of these compounds is highly regulated (e.g. see discussion above) and dependent on specific growth phases, stress response, and metabolic and physiological status of the cell. Moreover, at least two antibiotics, 2,4-diacetyiphioroglucinol and pyoluteorin, affect the metabolism of the cell when applied exogenously (16, 120). Thus, I propose a second role for secondary metabolites of pseudomonads. I hypothesize that some, or possibly all, secondary metabolites of Pseudomonas spp. may function as signal molecules. In this way, production of secondary metabolites would represent not "dead ends" of the branches of regulatory cascades that govern their production, but rather, intermediate points in cellular regulatory processes at which a new signal molecule (the metabolite) is generated. According to this hypothesis, the generation of a secondary metabolite would represent a node in the cell's regulatory circuitry, and, much like the phosphorylation of a regulatory protein, could be expected to generate multiple phenotypic changes. Hypothesis: Secondary metabolites of Pseudomonas spp. that function as extracellular toxins and are important for biological control may additionally function as signal molecules. 23 1.3.2. ExiStjn2 support for the hypothesis that secondary inetabolites function as siEnal molecules in Pseudomonas spp. One well-documented example of secondary metabolites that act as signal molecules in pseudomonads as well as other gram-negative bacteria is that of N-acyl homoserine lactones (AHLs). These diffusible molecules mediate cell-densitydependent regulation (quorum sensing). Their re-uptake across cell membranes presumably serves as an indicator of either cell density or confinement, and initiates upregulation of their own biosynthetic genes as well as others in regulatory cascades, resulting in profound changes in the cellular phenotype. For example, two AHL systems, las and rhi, are together estimated to govern expression of 1-4% of all genes in P. aeruginosa (139). However, some strains, like P. fluorescens Pf-5 and P. fluorescens CHAO, produce no detectable AHLs (57; J. E. Loper, unpublished data). It is unlikely that these rhizosphere bacteria lack extracellular signaling molecules. Rather, it is more likely that AHLs are only one class of signal molecule of which microbiologists have become acutely aware in the past two decades; the possibility that others exist is quite logical. Indeed, secondary metabolites of other structural classes have been shown to function as signal molecules in Pseudomonas. For instance, a quinolone antibiotic from P. aeruginosa, PQS, was found to be an integral part of the las/rhi regulatory network, which governs the production of virulence factors and secondary metabolites (88, 103). Pyoverdine, a siderophore that chelates ferric iron and returns it to the cell, was shown to have signaling properties with repercussions for the synthesis of three secreted virulence factors, including itself, in P. aeruginosa (8, 77). N-acyl homoserine lactones typically bind intracellular receptors from the LuxR family; recently, it was found that a family of cyclic dipeptides called diketopiperazines, or DKPs, from P. aeruginosa can induce LuxR- regulated functions. This observation led to the hypothesis that diketopiperazines function as signal molecules, and serve as modulators of quorum sensing pathways (63). Similarly active diketopiperazines were isolated from other gram-negative bacteria as well (63), and at least one of the active DKPs is also produced by the fungal pathogen Alternaria alternata, and confers phytotoxicity to spotted knapweed 24 (Centaurea maculosa) (125). This latter observation invites the tempting speculation that diketopiperazines could play a role in Pseudomonas-plant interactions, although this proposal has not been experimentally tested. 1.3.3. Si2nalin properties of antibiotics in non-Pseudomonas species. In non-Pseudomonas systems, other examples of secondary metabolites with signaling properties also exist. An exhaustive listing of these is beyond the scope of this chapter. However, it is worth noting that subinhibitory concentrations of antibiotics commonly used in medicinal practice can induce non-lethal physiological responses in their targets. Recently, it was shown that exposure to subinhibitory concentrations of erythromycin and rifampicin resulted in altered transcription of as many as 5% of genes in Salmonella typhimurium (45). Subinhibitory concentrations of antibiotics have been observed to have non-toxic effects on other bacterial species as well: clindamycin affects exoproduct expression in Staphylococcus aureus (60), and quinolone antibiotics affect flmbrial expression and adherence in Escherichia coli (15). Low concentrations of quinolone antibiotics also block intercellular signaling and the production of virulence factors in P. aeruginosa (47, 127, and 132). These observations lend credence to the ofi-touted view that antibiotics and other microbial toxins may have intracellular significance in addition to their extracellular roles as toxins. It seems intuitive to question the effect of bacterially produced antibiotics on the cells that produce them. For example, erythromycin, which affected Salmonella gene expression, is produced by Streptomyces erythreus. Tetracycline, streptomycin, cycloheximide, neomycin, and kanamycin are other familiar antibiotics produced by Streptomyces spp., while polymyxin B and bacitracin are produced by Bacillus spp. However, I am unaware of studies that explore the effect of these antibiotics on their producing organisms. 1.3.4. Re2ulation of antibiotic production by other secondary metabolites. Production of Pseuc/omonas antibiotics is itself modulated by secondary metabolites from both producing cells, and from other organisms. AHLs are a wellstudied example of such metabolite - metabolite regulation, and are discussed in a separate section. However, other classes of secondary metabolites regulate antibiotic production as well. 2,4-diacetylphioroglucinol production in P. fluorescens CHAO is repressed by at least three secondary metabolites: salicylic acid, fusaric acid, and pyoluteorin (35, 97, and 120). Salicylic acid and pyoluteorin are produced by CHAO, but fusaric acid is secreted by the pathogenic fungus F. oxysporum, a potential rhizosphere coinhabitant. Moreover, salicylic acid is produced by many organisms, including plants, where it serves as a signaling molecule and is integral in mounting induced systemic resistance against pathogens. "Foreign" secondary metabolite influences do not appear to be unique. For example, production of the secondary metabolite syringomycin, a phytotoxin produced by P. syringae pv. syringae, is induced by various phenolic compounds produced by the host plant (90, 91, and 109). 1.4. Statement of research objectives. As described above, the genes responsible for biosynthesis of pyoluteorin have been sequenced, and several intra- and extracellular factors affecting its production levels have been elucidated. This knowledge distinguishes pyoluteorin production in Pf-5 as an enviable model system for the study of both intracellular and environmental effects on regulation of antibiotic biosynthesis. I employed Pf-5 to test the hypothesis, stated above, of signaling roles for Pseudomonas antibiotics, with specific regard to pyoluteorin. Pyoluteorin, pyrrolnitrin, and 2,4diacetylphioroglucinol share facets of global regulation in Pf-5. Therefore, possible shared or interrelated signaling roles for pyoluteorin, pyrrolnitrin and 2,4- diacetylphlorglucinol were also explored. A recurring theme in antibiotic regulation by biocontrol pseudomonads is autoregulation, as described above (51). Thus, the first 26 test for a signaling role for pyoluteorin was one to determine the influence of exogenous pyoluteorin on its own production. Early results indicated that pyoluteorin, like 2,4-diacetylphloroglucinol, was positively autoregulated, and my larger goal became the characterization of the pyoluteorin autoinduction phenomenon. I also dissected the involvement of pyrrolnitrin and 2,4-diacetyiphioroglucinol in pyoluteorin production, and vice versa. The phenomenon of pyoluteorin autoinduction implies that exogenous pyoluteorin is either i) perceived at the cell surface, or ii) transported from the extracellular milieu back into the cytoplasm, where it is perceived by an intracellular pyoluteorin receptor. Pathway-linked LysR- and TetR-type transcriptional regulators are often activated by cofactor binding, and cofactors frequently are, in turn, products of the biosynthetic pathway under regulation (e.g. 62, 115). By regulating transcription from their own promoters, thus enhancing or repressing cellular levels of the cofactor receptor, LysR and TetR-type regulators can in some instances facilitate positive or negative autoregulation by products of the pathways that they regulate. Two such linked regulatory genes, pltR and pitH, occur in the pyoluteorin pathway and encode a LysR- and TetR-type regulator, respectively. In the related stain CHAO, autoinduction of 2,4-diacetylphloroglucinol is mediated by the pathway-linked TetR- type transcriptional regulator, PhlF. These observations added appeal to the hypothesis of an intracellular pyoluteorin receptor. Because an intracellular pyoluteorin receptor would necessitate the existence of a pyoluteorin uptake mechanism, a search was initiated for potential pyoluteorin transporter genes. Genes encoding enzymatic functions for biosynthesis of secondary metabolites, like pyoluteorin, are often flanked by genes for immunity to, or transport of, the metabolite (reviewed by 82). Examples of such clustering are given in Chapter 4. Therefore, the pyoluteorin biosynthetic gene cluster was further delineated. The early discovery of both regulatory and transport genes on one end of the antibiotic gene cluster prompted a shift in focus from sequencing to characterization of those genes. In light of its potential as a pyoluteorin receptor, attempts were also made to biochemically characterize the pathway-linked regulator PltR. 27 The results presented herein do not completely delineate the pyoluteorin biosynthetic pathway, nor do they provide a comprehensive understanding of the intracellular roles for Pf-5 antibiotics. However, they lay solid groundwork for fuzther studies which are already ongoing to better understand the function of pyoluteorinregulated transport genes that flank the pyoluteorin biosynthetic gene cluster, and to elucidate on a more global scale the physiological responses of Pf-5 to its antibiotic, pyoluteorin. 1.5. References I. Aarons, S., A. Abbas, C. Adams, A. Fenton, and F. O'Gara. 2000. A regulatory RNA (PrrB RNA) modulates expression of secondary metabolite genes in Pseudomonasfluorescens Fl 13. J. Bacteriol. 182:3913-3919. 2. Abbas, A., J. P. Morrissey, P. Carnicero Marquez, M. M. Sheehan, I. R. Delany, and F. O'Gara. 2002. Characterization of interactions between the transcriptional repressor PhIF and its binding site at the phiA promoter in P. fluorescens F113. J. Bacteriol. 184:3008-3016. 3. 4. Agrios, G. N. 1988. Plant Pathology. Academic Press, San Diego, CA. Altuvia, S., A. Zhang, L. Argaman, A. Tiwari, and G. Storz. 1998. The Escherichia coli OxyS regulatory RNA represses jhlA translation by blocking ribosome binding. EMBO J. 17:6069-6075. 5. Anjaiah, V., N. Koedam, B. Nowak-Thompson, J. E. Loper, M. Höfte, J. T. Tambong, and P. Cornelis. 1998. Involvement of phenazines and anthranilate in the antagonism of Pseudomonas aeruginosa PNA1 and Tn5 derivatives against Fusarium spp. and Pythium spp. Mo!. Plant-Microbe Interact. 11:847-854. 6. Bangera, G. M., and L. S. Thomashow. 1996. Characterization of a genomic locus required for synthesis of the antibiotic 2,4diacetyiphioroglucinol by the biological control agent Pseudomonas fluorescens Q2-87. Mo!. Plant-Microbe Interact. 9:83-90. 7. Bangera, G. M., and L. S. Thomashow. 1999. Identification and characterization of a gene cluster for synthesis of the polyketide antibiotic 2,4-diacetyiphioroglucinol from Pseudomonasfluorescens Q2-87. J. Bacteriol. 181 :3155-3163. 8. Beare, P. A., R. J. For, L. W. Martin, and I. L. Lamont. 2003. Siderophore-mediated cell signalling in Pseudomonas aeruginosa: divergent pathways regulate virulence factor production and siderophore receptor synthesis. Mol. Microbiol. 47:195-207. 9. Becker, J. 0., and F. J. Schwinn. 1993. Control of soil-borne pathogens with living bacteria and fungi: status and outlook. Pestic. Sci. 37:355-363. 10. Beubrook, C. M., E. Groth, J. M. Ilalloran, M. K. Hansen, and S. Marquardt. 1996. Pest management at the crossroads. 272 pp. Consumers Union, Yonkers, NY. 11. Blumer, C., and D. Haas. 2000. Iron regulation of the hcnABC genes encoding hydrogen cyanide synthase depends on the anaerobic regulator ANR rather than on the global activator GacA in Pseudomonasfluorescens CHAO. Microbiology 146:2417-2424. 12. Blumer, C., and D. Haas. 2000. Multicopy suppression ofagacA mutation by the infC operon in Pseudomonasfluorescens CHAO: competition with the global translational regulator RsmA. FEMS Microbiol. Lett. 187:53-58. 13. Biumer, C., S. Heeb, G. Pessi, and D. Haas. 1999. Global GacA-steered control of cyanide and exoprotease production in Pseudomonasfluorescens involves specific ribosome binding sites. Proc. Natl. Acad. Sci. USA 96:14073-14078. 14. Bonsai!, R. F., D. M. Weller, and L. S. Thomashow. 1997. Quantification of 2,4-diacetylphloroglucinol produced by fluorescent Pseudomonas spp. in vitro and in the rhizosphere of wheat. App!. Environ. Microbiol. 63:951-955. 15. Breines, D., and J. Burnham. 1994. Modulation of Escherichia coli type 1 fimbrial expression and adherence to uroepithelial cells following exposure of logarithmic phase cells to quinolones at subinhibitoiy concentrations. J. Antimicrob. Chemother. 34:205-221. 16. Brodhagen, M., M. D. Henkels, and J. E. Loper. 17. Camacho, M. M. 2001. Molecular characterization of the role of Type IV pili, NDH-I and PyrR in rhizosphere colonization of Pseudomonas fluorescens WCS36S. PhD Thesis. Universiteit Leiden. 18. Positive autoregulation of the antibiotic pyoluteorin in the biological control organism Pseudomonasfluorescens Pf-5. Submitted. 2003. Carroll, H., Y. Moënne-Loccoz, D. N. Dowling, and F. O'Gara. 1995. Mutational disruption of the biosynthesis genes coding for the antifungal metabolite 2,4-diacetylphloroglucinol does not influence the ecological fitness of Pseudomonasfluorescens Fl 13 in the rhizosphere of sugarbeets. Appl. Environ. Microbiol. 61:3002-3007. 19. Chadwick, D. J., and J. Whelan (eds.). 1992. Secondary metabolites: their function and evolution. John Wiley and Sons, Chichester, England. 318 pp. 30 20. Chatterjee, A., Y. Cui, Y. Liu, C. K. Dumenyo, and A. K. Chatterjee. 1995. Inactivation of rsmA leads to overproduction of extracellular pectinases, cellulases, and proteases in Erwinia carotovora subsp. carotovora in the absence of starvation/cell density-sensing signal, N-(3oxohexanoyl)-L-homoserine lactone. App!. Environ. Microbiol. 61:19591967. 21. Chin-A-Woeng, T. F. C., D. van den Broek, G. de Voer, D. M. G. M. van der Drift, S. Tuinman, 3. E. Thomas-Oates, B. 3.3. Lugtenberg, and G. V. Bloemberg. 2001. Phenazine-1-carboxamide production in the biocontrol strain Pseudomonas chiororaphis PCL1391 is regulated by multiple factors secreted into the growth medium. Mo!. Plant-Microbe Interact. 14:969-979. 22. Chin-A-Woeng, T. F. C., G. V. Bloemberg, A. J. van der Bij, K. M. G. M. van der Drift, 3. Schripsema, B. Kroon, IL J. Scheffer, C. Keel, P. A. H. M. Bakker, H. V. Tichy, F. J. de Bruijn, 3. E. Thomas-Oates, and B. J. J. Lugtenberg. 1998. IBiocontrol by phenazine-1-carboxamideproducing Pseudomonas chioraphis PCL1391 of tomato root rot caused by Fusarium oxysporum f. sp. radicis-lycopersici. Mol. Plant-Microbe Interact. 11:1069-1077. 23. Cook, J. R. 1988. Biological control and holistic plant-health care in agriculture. American Journal of Alternative Agriculture 3:51-62. 24. Cook, R. J. 1993. Making greater use of introduced microorganisms for biological control of plant pathogens. Annu. Rev. Phytopath. 31:53-80. 25. Cook, IL J., W. L. Bruckart, J. R. Coulson, M. S. Goettel, IL A. Humber, R. D. Lumsden, J.V. Maddox, M. L. McManus, L. Moore, S.F. Meyer, P.C. Quimby, Jr., J.P. Stack, and 3. L. Vaughn. 1996. Safety of microorganisms intended for pest and plant disease control: a framework for scientific evaluation. Biol. Control 7:333-351. 26. Corbell, N. 1999. A two-component regulatory system controlling antibiotic production by Pseudomonasfluorescens Pf-5. PhD Thesis. Oregon State University. 27. Corbell, N., and J.E. Loper. 1995. A global regulator of secondary metabolite production in Pseudomonasfluorescens Pf-5. J. Bacteriol. 177(21): 6230-6236. 28. Couzin, J. 2002. Breakthrough of the year #1, the winner: Small RNAs make big splash. Science 298:2296-2297. 31 29. Cui, Y., A. Chatterjee, Y. Liu, C. K. Dumenyo, and A. K. Chatterjee. 1995. Identification of a global repressor gene, rsmA, of Erwinia carotovora subsp. carotovora that controls extracellular enzymes, N-(3oxohexanoyl)-L-homoserine lactone, and pathogenicity in sort-rotting Erwinia spp. J. Bacteriol. 177:5108-5115. 30. Cui, Y., A. Mukherjee, C. K. Dumenyo, Y. Liu, and A. K. Chatterjee. 1999. rsmC of the soft-rotting bacterium Erwinia carotovora subsp. carotovora negatively controls extracellular enzyme and harpinE production and virulence by modulating levels of regulatory RNA (rsmB) and RNA-binding protein (RsmA). J. Bacteriol. 181:6042-6052. 31. Cui, Y., Chatterjee, A., and A. K. Chatterjee. 2001. Effects of the twocomponent system comprising GacA and GacS of Erwinia carotovora subsp. carotovora on the production of global regulatory rsmB RNA, extracellular enzymes, and harpin&. Mol. Plant-Microbe Interact. 14:5 16526. 32. Cuppels, D.A., C.R. Howell, R.D. Stipanovic, A. Stoessl, and J.B. Stothers. 1986. Biosynthesis of pyoluteorin: a mixed polyketidetricarboxylic acid cycle origin demonstrated by [1 ,21 3C2 acetate incorporation. Z. Naturforsch. 41c: 532-536. 33. Dowling, D. N., and F. O'Gara. 1994. Metabolites of Pseudomonas involved in the biocontrol of plant disease. Trends Biotech. 12:133-144. 34. Duffy, B. K. and G. Défago. 1999. Environmental factors modulating antibiotic and siderophore biosynthesis by Pseudomonasfluorescens biocontrol strains. Appl. Environ. Microbiol. 65(6): 2429-2438. 35. Duffy, B. K., and G. Défago. 1997. Zinc improves biocontrol of Fusarium crown and root rot of tomato by Pseudomonasfluorescens and represses the production of pathogen metabolites inhibitory to bacterial antibiotic biosynthesis. Phytopathology 87:1250-1267. 36. Duke, S. 0., Menn, J. J., and J. R. Plimmer. 1993. Challenges of pest control with enhanced toxicological and environmental safety. p. 1-13 In: Duke, S. 0., J. J. Menn, and J. R. Plimmer (eds.). Pest control with enhanced environmental safety. American Chemical Society, Washington, D.C. 37. Eisenstark, A., Calcutt, M.J., Becker-Hapak, M., and A. Ivanova. 1999. Role of Escherichia coli rpoS and associated genes in defense against oxidative damage. Free Rad. Biol. Med. 21: 975-993. 32 38. Elmholt, S. 1991. Side effects of propiconazole (Tilt 250 EC) on nontarget soil fungi in a field trial compared with natural stress effects. Microb. Ecol. 22:99-108. 39. El-Sayed, A. K., J. Hothersall, and C. M. Thomas. 2001. Quorumsensing-dependent regulation of biosynthesis of the polyketide antibiotic mupirocin in Pseudomonasfluorescens NCIMB 10586. Microbiol. 147:2127-2139. 40. Felton, G. W., and D. L. Dablman. 1984. Nontarget effects of a fungicide: toxicity of Maneb to the parasitoid Microplitis croceipes (Hymenoptera: Braconidae) [Heliothis virescens]. J. Econ. Entomol. 77:842-850. 41. Fenton, A. M., P. M. Stephens, J. Crowley, M. O'Callaghan, and F. O'Gara. 1992. Exploitation of gene(s) involved in 2,4diacetylphloroglucinol biosynthesis to confer a new biocontrol capability to a Pseudomonas strain. App!. Environ. Microbiol. 58:3873-3878. 42. Flaishman, M., Z. EyaI., C. Voisard, and D. Haas. 1990. Suppression of Septoria tritici by phenazine or siderophore-deficient mutants of Pseudomonas. Curr. Microbiol. 20:121-124. 43. Froyd, J. D. 1997. Can synthetic pesticides be replaced with biologicallybased alternatives? - an industry perspective. Journal of Industrial Microbiology and Biotechnology 19:192-195. 44. Glandorf, D. C. M., P. Verheggen, T. Jansen, J.-W. Jorritsma, E. Smit, P. Leeflang, K. Wernars, L. S. Thomashow, E. Laureijs, J. E. ThomasOates, P. A. H. M. Bakker, and L. C. van Loon. 2001. Effect of genetically modified Pseudomonas putida WCS358r on the füngal rhizosphere microfiora of field-grown wheat. App!. Environ. Microbio!. 67:3371-3378. 45. Goh, E. -B., G. Yim, W. Tsui, J. McClure, M. G. Surette, and i. Davies. 2002. Transcriptional modulation of bacterial gene expression by subinhibitory concentrations of antibiotics. Proc. Nati. Acad. Sci. USA. 99:17025-17030. 46. Gottesman, S. 1996. Proteases and their targets in Escherichia co/i. Annu. Rev. Genet. 30: 465-506. 47. Grimwood, K., M. To, H. R. Rabin, and D. E. Woods. 1989. Inhibition of Pseudomonas aeruginosa exoenzyme expression by subinhibitory antibiotic concentrations. Antimicrob. Agents Chemother. 33:41-47. 33 48. Gurusiddaiah, S., D. M. Weller, A. Sarkar, and R. J. Cook. 1986. Characterization of an antibiotic produced by a strain of Pseudomonas fluorescens inhibitory to Gaeumannomyces graminis var. tritici and Pythium spp. Antimicrob. Agents Chemother. 29 :488-495. 49. Gutterson, N., J.S. Ziegle, G.J. Warren, and T.J. Layton. 1988. Genetic determinants for catabolite induction of antibiotic biosynthesis in Pseudomonasfluorescens HV37a. J. Bactenol. 170(1): 380-385. 50. Haas, D., and C. Keel. 2003. Regulation of antibiotic production in rootcolonizing Pseudomonas spp. and relevance for biological control of plant disease. Annu. Rev. Phytopathol. Online. doi: 10.1 146/annurev.phyto.41 .052002.095656 51. Haas, D., C. Blumer, and C. Keel. 2000. Biocontrol ability of fluorescent pseudomonads genetically dissected: importance of positive feedback regulation. Curr. Opin. Biotech. 11:290-297. 52. Handelsman, J., and E. V. Stabb. 1996. Biocontrol of soilbome plant pathogens. Plant Cell 8:1855-1869. 53. Hardy, R. W. F. 1993. Biologically based pest management in managed ecosystenis: Increasing its acceptance. p. 2-9, In: Lumsden, R. D. and J. L. Vaughn (eds.). Pest management: Biologically based technologies. American Chemical Society, Washington, D.C. 54. Hardy, R. W. F., Beachy, R. N., Browning, H., Caulder, J. D., Charudattan, R., Faulkner, P., Gould, F. L., Hinkle, M. K., Jaffee, B. A., Knudson, M. K., Lewis, W. J., Loper, J. E., Mahr, D. L., and van Alfen, N. K. 1996. NRC Report. Ecologically based pest management: New solutions for a new century. 144 pp. NatI. Acad. Sci. Press. 55. Harrison, L. A., L. Letendre, P. Kovacevich, E. Pierson, and D. Weller. 1993. Purification of an antibiotic effective against Gaeumannomyces graminis var. tritici produced by a biocontrol agent, Pseudomonas aureofaciens. Soil Biol. Biochem. 25:215-221. 56. Heeb, S., and D. Haas. 2001. Regulatory roles of the GacS/GacA twocomponent system in plant-associated and other gram-negative bacteria. Mol. Plant-Microbe Interact. 14:1351-1363. 57. Heeb, S., C. Blumer, and D. Haas. 2002. Regulatory RNA as mediator in GacAIRsmA-dependent global control of exoproduct formation in Pseudomonasfluorescens CHAO. J. Bacteriol. 183:1046-1056. 34 58. Hengge-Aronis, R. 2000. The general stress response in Escherichia coil. pp. 16 1-178. In: Bacterial Stress Responses. Storz, G. and R. HenggeAronis (Eds.). ASM Press, Washington, D.C. 59. Hengge.-Aronis, IL, Lange, R., Henneberg, N., and D. Fischer. 1993. Osmotic regulation of rpoS-dependent genes in Escherichia coil. J. Bacteriol. 175: 259-265. 60. Herbert, S., P. Barry, and R. P. Novick. 2001. Subinhibitory clindamycin differentially inhibits transcription of exoprotein genes in Staphylococcus aureus. Infect. Immun. 69:2996-3003. 61. Hill, D. S., J. I. Stein, M. R. Torkewitz, A. M. Morse, C. R. Howell, J. P. Pachlatko, J. 0. Becker, and J. M. Ligon. 1994. Cloning of genes involved in the synthesis of pyrrolnitrin from Pseudomonasfluorescens and role of pyrrolnitrin synthesis in biological control of plant disease. Appl. Environ. Microbiol. 60:78-85. 62. Hillen, W., and C. Berens. 1994. Mechanisms underlying expression of TnlO encoded tetracycline resistance. Annu. Rev. Microbiol. 48:345-369. 63. Holden, M. T. G., S. R. Chhabra, R. de Nys, P. Stead, J. J. Bainton, P. J. Hill, M. Manefield, N. Kumar, M. Labatte, D. England, S. Rice, M. Givskov, G. P. C. Salmond, G. S. A. B. Stewart, B. W. Bycroft, S. Kjelleberg, and P. Williams. 1999. Quorum-sensing cross talk: isolation and chemical characterization of cyclic dipeptides from Pseudomonas aeruginosa and other Gram-negative bacteria. Mo!. Microbio!. 33:12541266. 64. Homma, Y. 1994. Mechanism of biological control focussed on the antibiotic pyrrolnitrin. p. 100-103 In: M. H. Ryder, P. M. Stephens, and G. D. Bowen (eds.), Improving plant productivity with rhizobacteria. CS]IRO Division of Soils, Adelaide, Australia. 65. Howell, C. R. (Agricultural Research Station, U.S. Department of Agriculture, College Station, Tex.). 1993. Personal communication. 66. Howell, C. R., and R. D. Stipanovic. 1979. Control of Rhizoctonia solani on cotton seedlings with Pseudomonasfluorescens and with an antibiotic produced by the bacterium. Phytopathology 69:480-482. 67. Howell, C. IL, and R. D. Stipanovic. 1980. Suppression of Pythium ultimum-induced damping-off of cotton seedlings by Pseudomonas fluorescens and its antibiotic, pyoluteorin. Phytopathogy 70:712-715. 35 68. Howie, W. J. and T. V. Susiow. 1991. Role of antibiotic biosynthesis in the inhibition of Pythium ultimum in the cotton spermosphere and rhizosphere by Pseudomonasfiuorescens. Mo!. Plant-Microbe Interact. 4:393-399. 69. Ishimaru, C. A., E. J. Kios, and R. R. Brubaker. 1988. Multiple antibiotic production by Erwinia herbicola. Phytopathology 78:746-750. 70. Keel, C., P. Wirthner, T. Oberhänsli, C. Voisard, Burger, D. Hass, and G. Défago. 1990. Pseudomonads as antagonists of plant pathogens in the rhizosphere: role of the antibiotic 2,4-diacetylphloroglucinol in the suppression of black root rot of tobacco. Symbiosis 9:327-341. 71. Keel, C.., U. Schnider, M. Maurhofer, C. Voisard, J. Laville, U. Burger, P. Wirthner, D. Haas, and G. Défago. 1992. Suppression of root diseases by Pseudomonasfluorescens CHAO: importance of the bacterial secondary metabolite 2,4-diacetyiphioroglucinol. Mo!. Plant-Microbe Interact. 5:413. 72. Keel, C., Weller, D., Natsch, A., Defago, G., Cook, R.J., and L.S. Thomashow. 1996. Conservation of the 2,4-diacetylphloroglucinol biosynthesis locus among fluorescent Pseudomonas strains from diverse geographic locations. App!. Environ. Microbiol. 62: 552-563. 73. Khvorova, A., Y.-G. Kwak, M. Tamkun, I. Majerfield, and M. Yarus. 1999. RNAs that bind and change permeability of phospholipid membranes. Proc. Nat!. Acad. Sci. USA 96:10649-10654. 74. Kim, B. S., S. S. Moon, and 13. K. Hwang. 1999. Isolation, identification, and antifungal activity of a macrolide antibiotic, oligomycin A, produced by Streptomyces libani. Can. J. Bot. 77:850-858. 75. Kraus, J., and J.E. Loper. 1992. Lack of evidence for a role of antifungal metabolite production by Pseudomonasfluorescens Pf-5 in the biological control of Pythium damping-off of cucumber. Phytopathology 82: 264-27 1. 76. Kraus, J., and J.E. Loper. 1995. Characterization of a genomic region required for the production of the antibiotic pyoluteorin by the biological control agent Pseudomonasfluorescens Pf-5. App!. Environ. Microbiol. 61(3): 849-854. 77. Lamont, I. L., P. A. Beare, U. Ochsner, A. I. Vasil, and M. L. Vasil. 2002. Siderophore-mediated signaling regulates virulence factor production in Pseudomonas aeruginosa. Proc. Nat!. Acad. Sci. USA 99:7072-7077. 36 78. Lease, R. A., and M. Belfort. 2000. A trans-acting RNA as a control switch in Escherichia co/i: DsrA modulates function by forming alternative structures. Proc. Nat!. Acad. Sci. USA 97:9919-9924. 79. Lease, R. A., M. E. Cusick, and M. Belfort. 1998. Riboregulation in Escherichia co/i: DsrA RNA acts by RNA:RNA interactions at multiple loci. Proc. Nati. Acad. Sci USA 95:12456-12461. 80. Liii, Y., G. Jiang, Y. Ciii, A. Mukherjee, W. 0. Ma, and A. K. Chatterjee. 1999. kdgRECC negatively regulates genes for pectinases, cellulase, protease, harpinj, and a global RNA regulator in Erwinia carotovora subsp. carotovora. J. Bacteriol. 181:2411-2422. 81. Liu, Y., Y. Ciii, A. Mukherjee, and A. K. Chatterjee. 1998. Characterization of a novel RNA regulator of Erwinia carotovora ssp. carotovora that controls production of extracellular enzymes and secondary metabolites. Mo!. Microbiol. 29:219-234. 82. Martin, J. F., and P. Liras. 1989. Organization and expression genes involved in the biosynthesis of antibiotics and other secondary metabolites. Annu. Rev. Microbiol. 43:173-206. 83. Mathre, D. E., R. J. Cook, and N. W. Callan. 1999. From discovery to use: Traversing the world of commercializing biocontrol agents for plant disease control. Plant Dis. 83:972-983. 84. Maurhofer, M., C. Keel, D. Haas, and G. Defago. 1994. Pyoluteorin production by Pseudomonasfluorescens strain CHAO is involved in the suppression of Pythium damping-off of cress but not of cucumber. Eur. J. Plant Pathol. 100:221-232. 85. Maurhofer, M., C. Keel, D. Haas, and G. Defago. 1995. Influence of plant species on disease suppression by Pseudomonasfluorescens strain CHAO with enhanced antibiotic production. Plant Pathol. 44:40-50. 86. Maurhofer, M., C. Keel, U. Schnider, C. Voisard, D. Haas, and G. Défago. 1992. Influence of enhanced antibiotic production in Pseudomonasfluorescens strain CHAO on its disease suppressive capacity. Phytopathology 82:190-195. 87. Mazzola, M., R. J. Cook, L. S. Thomashow, D. M. Weller, and L. S. Pierson III. 1992. Contribution of phenazine antibiotic biosynthesis to the ecological competence of fluorescent pseudomonads in soil habitats. App!. Environ. Microbiol. 48:2616-2624. 37 88. McKnight, S. L., B. H. Igiewski, and E. C. Pesci. 2000. The Pseudomonas quinolone signal regulates rhi quorum sensing in Pseudomonas aeruginosa. J. Bacteriol. 182:2702-2708. 89. MeSpadden Gardener, B. B., and D. R. Fravel. 2002. Biological control of plant pathogens: Research, commercialization, and application in the USA. Online. Plant Health Progress doi: 10.1094/PHP-2002-0510-01-RV. 90. Mo, Y. Y., and D. C. Gross. 1991. Plant signal molecules activate the syrB gene, which is required for syringomycin production by Pseudomonas syringae pv. syringae. J. Bacteriol. 173:5784-5792. 91. Mo, Y. Y., M. Geibel, R. F. Bonsall, and D. C. Gross. 1995. Analysis of sweet cherry (Prunus avium L.) leaves for plant signal molecules that activate the syrB gene required for synthesis of the phytotoxin, syringomycin, by Pseudomonas syringae pv. syringae. Plant Physiol. 107:603-612. 92. Mukherjee, A., Y. Cui, W. Ma, Y. Liu, A. Ishihama, A. Eisenstark, and A. K. Chatterjee. 1998. RpoS (Sigma-S) controls expression of rsmA, a global regulator of secondary metabolites, harpin, and extracellular proteins in Erwinia carotovora. J. Bacteriol. 180:3629-3634. 93. Nakayama, T., Y. Homma, Y. Hashidoko, J. Mizutani, and S. Tahara. 1999. Possible role of xanthobaccins produced by Stenotrophomonas sp. strain SB-K88 in suppression of sugar beet damping-off disease. Appl. Environ. Microbiol. 65:4334-4339. 94. Nguyen, L.H., Jensen, D.B., Thompson, N.E., Gentry, D.R., and R.R. Burgess. 1993. In vitro functional characterization of overproduced Escherichia coli katF/rpoS gene product. Biochemistry 32: 11112-11117. 95. Nogueira, T., and M. Springer. 2000. Post-transcriptional control by global regulators of gene expression in bacteria. Curr. Opin. Microbiol. 3:154-158. 96. Notz, R., M. Maurhofer, U. Schnider-Keel, B. Duffy, D. Haas, and G. Défago. 2001. Biotic factors affecting expression of the 2,4diacetylphloroglucinol biosynthesis gene phiA in Pseudomonasfluorescens biocontrol strain CHAO in the rhizosphere. Phytopathology 91:873-881. 97. Notz, R., Maurhofer, M., Dubach, III., Haas, D., and G. Defago. 2002. Fusaric acid-producing strains of Fusarium oxysporum alter 2,4diacetylphloroglucinol biosynthetic gene expression in Pseudomonas fluorescens CHAO in vitro and in the rhizosphere of wheat. App. Environ. Microbiol. 68(5): 2229-2235. 98, Nowak-Thompson, B. 1997. An integrative approach to understanding pyoluteorin biosynthesis in the biological control bacterium Pseudomonas fluorescens Pf-5. PhD Thesis. Oregon State University. 99. Nowak-Thompson, B., Gould, S.J., Kraus, J., and J.E. Loper. 1994. Production of 2,4-diacetyiphioroglucinol by the biocontrol agent Pseudomonasfluorescens Pf-5. Can. J. Microbiol.40: 1064-1066. 100. Nowak-Thompson, B., N. Chaney, J. S. Wing, S. J. Gould, and J. E. Loper. 1999. Characterization of the pyoluteorin biosynthetic gene cluster ofPseudomonasfluorescens Pf-5. J. Bacteriol. 181:2166-2174. 101. O'Sullivan, D.J. and F. O'Gara. 1992. Traits of fluorescent Pseudomonas spp. involved in suppression of plant root pathogens. Microbiological Reviews 56:662-676. 102. Paulitz, T. C., and R. 11. Bélanger. 2001. Biological control in greenhouse systems. Annu. Rev. Phytopathol. 39:103-133. 103. Pesci, E. C., J. B. J. Milbank, J. P. Pearson, S. McKnight, A. S. Kende, E. P. Greenberg, and B. H. Iglewski. 1999. Quinolone signaling in the cell-to-cell communication system of Pseudomonas aeruginosa. Proc. Natl. Acad. Sci. USA. 96:11229-11234. 104. Pessi, G., F. Williams, Z. Hindle, K. Heurlier, M. T. G. Holden, M. Camara, D. Haas, and P. Williams. 2001. The global posttranscriptional regulator RsmA modulates production of virulence determinants and Nacyihomoserine lactones in Pseudomonas aeruginosa. J. Bacteriol. 183:6676-6683. 105. Pfender, W. F., J. Kraus, and J. E. Loper. 1993. A genomic region from Pseudomonasfluorescens Pf-5 required for pyrrolnitrin production and inhibition of Pyrenophora tritici-repentis in wheat straw. Phytopathology 83:1223-1228. 106. Pidoplichenko, V. N., and A. D. Garagulya. 1974. Effect of antagonistic bacteria on the development of wheat root rot. Microbilogische Z. 36:599602. 107. Pierson, E. A., D. W. Wood, J. A. Cannon, F. M. Blachere, and L. S. Pierson III. 1998. Interpopulation signaling via N-acyl-homoserine lactones among bacteria in the wheat rhizosphere. Mo!. Plant-Microbe Interact. 11:1078-1084. 108. Pierson, L.. S. III, and E. A. Pierson. 1996. Phenazine antibiotic production in Pseudomonas aureofaciens: role in rhizosphere ecology and pathogen suppression. FEMS Microbiol. Left. 136:101-108. 109. Quigley, N. B., and D. C. Gross. 1994. Syringomycin production among strains of Pseudomonas syringae pv. syringae: conservation of the syrB a and syrD genes and activation of phytotoxin production by plant signal molecules. Mo!. Plant-Microbe Interact. 7:78-90. 110. Raaijmakers, J. M., R. F. Bonsall, and D. M. Weller. 1999. Effect of population density of Pseudomonasfluorescens on production of 2,4diacetylphloroglucinol in the rhizosphere of wheat. Phytopathology 89:470-475. 111. Reimmann, C., L Serino, M. Beyeler, and D. Haas. 1998. Dihydroaeruginoic acid synthetase and pyochelin synthetase, products of the pchEF genes, are induced by extracellular pyochelin in Pseudomonas aeruginosa. Microbiol. 144:3135-3148. 112. Rodriguez, F., and W. F. Pfender. 1997. Antibiosis and antagonism of Scierotinia homoeocarpa and Drechslera poae by Pseudomonas fluorescens Pf-5 in vitro and in planta. Phytopathology 87:614-62 1. 113. Romeo, T. 1998. Global regulation by the small RNA-binding protein CsrA and the non-coding RNA molecule CsrB. Mol. Microbiol. 29:13211330. 114. Russell, P. E. 1995. Fungicide resistance: occurrence and management. J. Agric. Sci. 124:317-323. 115. Sarniguet, A. (JNIRA, Centre de recherches de Rennes, Station de pathologie vegetale, Le Rheu cedex, France). 1997. Personal communication. 116. Sarniguet, A., J. Kraus, M.D. Henkels, A.M. Muehichen, and J.E. Loper. 1995. The sigma factor ó affects antibiotic production and biological control activity of Pseudomonasfluorescens Pf-5. Proc. Nat!. Acad. Sci. USA 92:1225-12259. 117. SchelI, M. 1993. Molecular biology of the LysR family of transcriptional regulators. Annu. Rev. Microbiol. 47:597-626. 40 118. Schnider, U., Keel, C, Blumer, C., Troxler, J., Defago, G., and D. Haas. 1995. Amplification of the housekeeping sigma factor in Pseudomonas fluorescens CHAO enhances antibiotic production and improves biocontrol abilities. J. Bacteriol. 177(18): 5387-5392. 119. Schnider, U., Keel, C., Voisard, C.., Defago, G. and D. Haas. 1995. Tn5directed cloning ofpqq genes from Pseudomonasfluorescens CHAO: mutational inactivation of the genes results in overproduction of the antibiotic pyoluteorin. Appi. Environ. Microbiol. 61(11): 3856-3864. 120. Schnider-Keel, U., A. Seematter, M. Maurhofer, C. Blumer, B. Duffy, C. Gigot-Gonnefoy, C. Reimmann, It Notz, G. Défago, D. Haas, and C. Keel. 2000. Autoinduction of 2,4- diacetylphloroglucinol biosynthesis in the biocontrol agent Pseudomonasfluorescens CHAO and repression by the bacterial metabolites salicylate and pyoluteorin. J. Bacteriol. 182: 12151225. 121. Schumann, G. L. 1991. Plant diseases: Their biology and social impact. APS Press, St. Paul, MN. 397 pp. 122. Shah-Smith, D. A., and R. G. Burns. 1997. Shelf-life of a biocontrol Pseudomonasputida applied to sugar beet seeds using commercial coatings. Biocontrol. Sci. Technol. 7:65-74. 123. Shang, H., J. Chen, J. Handelsman, and R. M. Goodman. 1999. Behavior of Pythium torulosum zoospores during their interaction with tobacco roots and Bacillus cereus. Curr. Microbiol. 38:199-204. 124. Silo-Suh, L. A., B. J. Lethbridge, S. J. Raffel, H. He, J. Clardy, and J. Handlesman. 1994. Biological activities of two fungistatic antibiotics produced by Bacillus cereus TJW85. Appi. Environ. Microbiol. 60:20232030. 125. Stierle, A. C., J. H. Cardellina, and G. A. Strobe!. 1988. Maculosin, a host-specific phytotoxin for spotted knapweed from Alternaria alternata. Proc. Natl. Acad. Sci. USA 85: 8008-8011. 126. Tanaka, K., Takayanagi, Y., Fugita, N., Ishihama, A., and H. Takahashi. 1993. Heterogeneity of the principal a factor in Escherichia coli: the rpoS gene product, a38 is a second principal a factor of RNA polymerase in stationary-phase Escherichia coli. Proc. Natl. Acad. Sci. USA 90: 3511-3515. 41 127. Tateda, K., R. Comte, J.-C. Pechere, T. Köhler, K. Yamaguchi, and C. van Delden. 2001. Azithromycin inhibits quorum sensing in Pseudomonas aeruginosa. Antimicrob. Agents Chemother. 45:1930-1933. 128. Thomashow, L. A., D. M. Weller, R. F. Bonsall, and L. S. Pierson III. 1990. Production of the antibiotic phenazine-1-carboxylic acid by fluorescent Pseudomonas species in the rhizosphere of wheat. App!. Environ. Microbiol. 56:908-9 12. 129. Thomashow, L. S., and D. M. Weller. 1988. Role of a phenazine antibiotic from Pseudomonasfluorescens in biological control of Gaeumannomyces graminis var. tritici. J. Bacteriol. 170:3499-3508. 130. Thomashow, L. S., and D. M. Weller. 1996. Current concepts in the use of introduced bacteria for biological disease control: mechanisms and antifungal metabolites. p. 236-27 1 In: Plant-Microbe interactions. Stacey, G. and N. T. Keen (eds.). Chapman and Hall, NY. 131. Thrane, C., T. H. Neilsen, M. N. Neilsen, J. Sørensen, and S. Olsson. 2000. Viscosinamide-producing Pseudomonasfluorescens DR54 exerts a biocontrol effect on Pythium ultimum in sugar beet rhizosphere. FEMS Microbiol. Ecol. 33:139-146. 132. Trancassini, M., A. Giordano, A. Magni, and P. Cipriani. 1993. Subinhibitory concentrations of antibacterial drugs and Pseudomonas aeruginosa virulence factors. New Microbiol. 16:275-279. 133. Voisard, C., C. Keel, D. Haas, and G. Défago. 1989. Cyanide production by Pseudomonasfluorescens helps suppress black root rot of tobacco under gnotobiotic conditions. EMBO J. 8:35 1-358. 134. Weller, D.M. 1988. Biological control of soilborne plant pathogens in the rhizosphere with bacteria. Annu. Rev. Phytopathol. 26: 379-407. 135. Whipps, J. M. 2001. Microbial interactions and biocontrol in the rhizosphere. J. Exp. Bot. 52:487-511. 136. Whistler, C. A. 2000. Regulation of antibiotic production and stress response by the biological control organism Pseudomonasfluorescens Pf-5. PhD Thesis. Oregon State University. 137. Whistler, C. A., Corbell, N. A., Sarniguet, A., Ream, W., and J. E. Loper. 1998. The two-component regulators GacS and GacA influence 42 accumulation of the stationary-phase sigma factor S and the stress response in Pseudomonasfluorescens Pf-5. J. Bacterjol. 180(24): 6635-6641. 138. Whistler, C.A., V.0. Stockwell, and J.E. Loper. 2000. Lon protease influences antibiotic production and UV tolerance of Pseudomonas fluorescens Pf-5. App!. Environ. Microbiol. 66(7): 2718-2725. 139. Whiteley, M., K. M. Lee, and E. P. Greenberg. 1999. Identification of genes controlled by quorum sensing in Pseudomonas aeruginosa. Proc. Nat!. Acad. Sci. USA 96:13904-13909. 140. Wood, D. W., F. Gong, M. M. Daykin, P. Williams, and L. S. Pierson III. 1997. N-acyl-homoserine lactone-mediated regulation of phenazine gene expression by Pseudomonas aureofaciens 3 0-84 in the wheat rhizosphere. J. Bacteriol. 179:7663-7670. 141. Zhang, A., S. Altuvia, A. Tiwari, L. Argaman, R. Hengge-Aronis, and G. Storz. 1998. The OxyS regulatory RNA represses rpoS translation and binds the Hfq (HF-I) protein. EMBO J. 17:6061-6068. 43 Chapter 2 Positive Autoregulation of the Antibiotic Pyoluteorin in the Biological Control Organism Pseudomonasfluorescens Pf-5 Marion Brodhagen, Marcella D. Henkels, and Joyce E. Loper Submitted to Applied and Environmental Microbiology American Society for Microbiology 2.1. Abstract Pseudomonasfluorescens Pf-5, a rhizosphere bacterium, produces a suite of secondary metabolites that are toxic to seed- and root-rotting plant patbogens. Among these are the polyketide compounds pyoluteorin (PLT) and 2,4-diacetyiphioroglucinol (2,4-DAPG). We provide evidence that PLT production is influenced by positive autoregulation. Addition of PLT to liquid cultures of Pf-5 enhanced PLT production. In addition, PLT and 2,4-DAPG mutually inhibit one another's production in Pf-5. For PLT, both positive autoregulation and negative influences on production by 2,4DAPG were demonstrated at the transcriptional level, by measuring activity from transcriptional fusions of an ice nucleation reporter gene to three separate PLT biosynthetic genes. The occurrence of PLT autoregulation in the rhizosphere was assessed on cucumber seedlings in pasteurized soil, using cross-feeding experiments. In the rhizosphere, a Plt derivative of Pf-5 carrying a reporter gene fusion to a PLT biosynthetic gene responded positively to PLT produced by coinoculated cells of Pf-5 or a derivative. These data establish for the first time that the polyketide PLT is an autoregulatory compound and functions as a signal molecule influencing the spectrum of secondary metabolites produced by the bacterial cell. 2.2 Introduction Microbial biological control agents such as the rhizosphere bacterium Pseudomonasfluorescens Pf-5 represent alternatives to synthetic chemicals for combating plant disease in agriculture. An important aspect of plant disease suppression by rhizosphere bacteria is the production of low molecular weight metabolites with antibiotic properties against certain plant pathogens (reviewed in 8, 12, 17, 18, 36). P. fluorescens Pf-5 produces an array of secondary metabolites that inhibit plant pathogens, including pyoluteorin (PLT), pyrrolnitrin, 2,4- diacetyiphioroglucinol (2,4-DAPG) and hydrogen cyanide (23,24, 26, 29, 43, 47). PLT and 2,4-DAPG (Fig. 2.1) are synthesized by polyketide synthase complexes (4, 42, 44) and are secreted from cells of Pf-5. The focus of this research is PLT, a chlorinated polyketide compound (3) that suppresses the Oomycete Pythium ultimum, a seed- and root-rotting pathogen (24, 37). PLT production is conferred by a biosynthetic gene cluster encompassing nine structural genes whose predicted functions are sufficient for the biochemical transformations required for PLT biosynthesis from acetate and proline precursors (11, 42, 44). Also within the PLT biosynthetic gene cluster is pltR, which encodes a LysR transcriptional regulator and is required for transcription of PLT biosynthesis genes and resultant PLT production (42). PLT production by Pf-5 is affected by nutrient source (30, 43), temperature (41), and cell density (11). Although the links between these extracellular factors and the intracellular regulatory pathways controlling PLT production are not yet understood, a complex picture is emerging that links regulation of production of PLT and other exoproducts to the physiological status of the cell. 46 0 0 ççrc' OH a Figure 2.1. Structures of pyoluteorin (a) and 2,4-diacetylphioroglucinol (b). Several global regulatory genes control PLT biosynthesis in Pf-5 and in P. fluorescens CHAO, a biochemically similar strain in which regulation of PLT production also has been extensively characterized. The production of many antibiotics and exoproducts by Pseudomonas spp. and other Gram-negative bacteria requires the histidine protein kinase/response regulator pair GacS/GacA (e.g. see 21); the signal(s) that activates GacS is not known. In Pf-5, the GacS/GacA pair positively regulates production of several secreted products, including PLT, 2,4-DAPG, pyrrolnitrin, hydrogen cyanide and an extracellular protease (10, 29). At least two common components of stress response systems, the stationary-phase and stressresponse sigma factor RpoS (aS) and Lon protease, are implicated in regulation of antibiotic production in Pf-5. Relative levels of as and the housekeeping sigma factor RpoD (&) influence PLT, 2,4-DAPG, and pyrrolnitrin production. Abundant as is required for the production of pyrrolnitrin, but is repressive to PLT and 2,4-DAPG production (50, 52). Lon, a serine protease induced by heat shock, also is repressive to PLT production (58). The importance of nutrient status to PLT production is corroborated by the observation that pyrrolquinoline quinone, a cofactor required by glucose and alcohol dehydrogenases, represses PLT production (53). Thus, PLT production in Pf-5 is linked to the physiological status of the cell. In this manuscript, we present evidence that the regulatory circuit governing PLT production includes positive autoregulation. The most familiar example of positive autoregulation is that of the N-acyl homoserine lactones (AHLs). In Gramnegative bacteria, AULs are synthesized and perceived, respectively, by Luxl/LuxR 47 homologs, and are postulated to function as signals of cell density and/or confinement. AHL production, along with that of other cellular products, is upregulated as a consequence of AHL accumulation in the extracellular milieu and re-entry into producer cells. AHLs have been demonstrated to function as intercellular signal molecules in the rhizosphere, where their production by rhizosphere inhabitants can induce phenazine antibiotic production by Pseudomonas auerofaciens (48, 59). NonAHL extracellular metabolites produced by Pseudomonas spp. also undergo positive autoregulation, including the antibiotics 2,4-DAPG (54) and oomycin A (16,25), and the siderophores pyochelin (49) and pyoverdine (31). This report adds PLT to the growing list of non-AHL autoinducers known in Pseudomonas spp. Because Pf-5 produces several antibiotics, we investigated the potential for regulation of PLT production by a heterologous antibiotic. We demonstrate that PLT production and the transcription of PLT biosynthesis genes are influenced by the presence of another antibiotic produced by Pf-5, 2,4-DAPG. Finally, we present evidence that PLT produced by Pf-5 can induce the expression of PLT biosynthesis genes by neighboring bacterial cells in the rhizosphere. 2.3. Methods 2.3.1. Bacterial strains plasmids, and culture conditions. Bacterial strains and plasmids used in this study are listed in Table 2.1. Pf-5 and derivative strains were cultured routinely on King's Medium B agar (KMB; 28) containing streptomycin (100 pg/ml). KMB was supplemented with other antibiotics (kanamycin, 50 jig/mI; gentamycin, 40 j.tg/ml) when appropriate. Inoculum of Pf-5 and derivatives was grown overnight with shaking (200 rpm) at 27°C in either KMB broth or KMBGIu, modified from KMB broth by substituting the glycerol with 1% (w/v) glucose to repress PLT production. For experiments evaluating antibiotic production or PLT biosynthetic gene transcription, cultures were grown at 20°C with shaking (200 rpm) in nutrient broth (pH 6.8; Difco Laboratories, Detroit, MI) without 48 selection and supplemented with either 2% (w/v) glycerol for NBG1y, a medium conducive to PLT production, or with 2% (w/v) glucose for NBGIu, a medium repressive to PLT production by Pf-5 (30, 43). For experiments evaluating antibiotic production or PLT biosynthetic gene transcription in defined medium, bacterial cultures were grown in OS medium (45) amended with 0.5% (w/v) glycerol (OSG1y) at 20°C with shaking (200 rpm). Table 2.1: Bacterial strains and plasmids used in this study Strain Strain Number Relevant JL4474 Pseudomonasfluorescens Pf-5 Reference or Source characteristicsa P. fluorescens Pf-5 23 rhizosphere isolate; Pitt, Ina Pf-5 iceC JL4303 Pf-5 pltE::inaZ JL4365 Pf-5 (pJELI 703); Kmr Plt, Ina 34 pltE::Tn3 -nice derivative of Pf-5, P1t, 30 Ina, Kmr Pf-5pItB::inaZ JL4389 pItB::Tn3-n ice derivative of Pf-5, P1t, 30 Ina, Kmr Pf-5pltF::inaZ JL4390 pltF::Tn3 -nice derivative of Pf-5, PIf, Ina, KThr 30 Pf-5 rpoS::lacZ JL4489 rpoS::lacZ derivative of Pf-5, Plt, Ina 57 Pf-5 gacA:.aphl JL4577 gacA ::aphl derivative of Pf-5; Pif, Ina, 9 Pf-5 pltR::aacCl JL4563 pltR::aacCl derivative of Pf-5; Pif, Ina 42 GmT Plasmids pJELI 703 iceC from P. syringae cloned into pVSP61 (nucleotide sequences of iceC and inaZ are nearly identical) 34 pVSP61 Stable plasmid vector containing pVS1 replicon and mobilization region and p1 5A replicon; Mob, W. Tucker, DNA Plant Technolog KThr yCorp., Oakland, Calif. aplt+ produces pyoluteorin; Plf, does not produce detectable pyoluteorin; P1t, overproduces pyoluteorin relative to the parental strain; 1na, expresses ice nucleation activity at -5°C; In&, does not express ice nucleation activity at -5°C; KmT and GmT, resistant to kanamycin and gentamycin, respectively. 50 2.3.2. Influence of exogenous antibiotics on antibiotic production and PLT gene expression by Pf-5. Cultures of experimental strains were grown overnight in KBGIu, and cells were harvested by centrifugation (2,500 x g, 5 mm), washed twice in sterile KMBGIu to eliminate traces of exported metabolites, and resuspended in 5 ml of NBG1y to achieve an OD6 of 0.08. Purified PLT and synthetic 2,4-DAPG standards dissolved in methanol were added to NBG1y immediately prior to inoculation with Pf-5 or derivative strains. An equivalent volume of methanol was added to negative controls. Except where noted, antibiotics were added to the medium at 4 jig/mi PLT and 12.5 ,.tg/ml 2,4-DAPG. These represent the low end of a range of concentrations typically detected in supernatants of Pf-5 cultures grown under conditions optimizing antibiotic production (10, 43, 57), and were selected to mimic extracellular antibiotic concentrations that Pf-5 cells would encounter in a stationary-phase culture. For experiments assessing transcriptional activity of PLT genes using the inaZ reporter gene, samples were harvested periodically for 24 h and assessed for ice nucleation activity (NA) as described below. At 18 Ii after inoculation, both 2,4DAPG and PLT were present at measurable concentrations in the supernatant of Pf-5 cultures grown in NBG1y. Therefore, for all experiments that were not harvested at multiple timepoints, PLT and 2,4-DAPG concentrations were determined from cultures harvested 18 h after inoculation. Bacterial population densities were determined by spreading serial dilutions of cultures onto KMB and counting CFU on the plates after 48 h incubation at 27°C. In each experiment, triplicate 5 ml cultures were grown and evaluated; each experiment was performed at least twice; and the results from a representative experiment are presented. 51 2.33. Antibiotic quantification. Cultures were centrifuged (5,000 x g, 5 mm) to pellet cells, and supernatants were extracted for recovery of PLT and 2,4-DAPG. Extractions were performed by a modification of the methods of Sarniguet et al. (50). Culture supernatants were acidified to pH 2.0 with 1 M HC1 and extracted twice with 0.4 volume of ethyl acetate. Ethyl acetate was removed from extracts under vacuum, and the resultant residues were resuspended in 100 p.! methanol, from which 10 p.! samples were evaluated by high performance liquid chromatography (HPLC). Samples were separated over a Nova-Pak C18 reversed-phase column (Waters Corporation, Milford, MA) eluted (1 mI/mm) with 90% water + 0.1% acetic acid:10% acetonitrile + 0.1% acetic acid (v/v) for two mm, followed by a linear gradient to 100% acetonitrile + 0.1% acetic acid over 18 mm. Antibiotics were detected with a photodiode array detector (PLT, X = 310 nm and retention time (tr) = 13.1 mm; 2,4-DAPQ,A = 270 nm and tr =15.6 miii). Quantification was done by comparison to a standard curve generated from authentic compounds. The detection limit was 0.02 jig/mI for PLT and 0.01 p.g/ml for 2,4-DAPG. Recovery of compounds using the extraction procedure described above averaged 70% for PLT and 45% for 2,4-DAPG. 2.3.4. Transcriptional activity from PLT biosynthetic genes Transcriptional fusions of a promoterless inaZ gene to three PLT biosynthetic genes of Pf-5 were constructed previously (30). Transcription of the inaZ reporter gene from an upstream promoter results in expression of the membrane-bound InaZ protein, which forms aggregates that nucleate water molecules to form ice crystals at temperatures slightly below 0°C (32). Expression of INA from inaZ transcriptional fusions was quantified using a droplet freezing assay at -5°C as described previously (35) and reported as logio(JNA), where INA = ice nuclei/CFU. 52 2.3.5. Assessment of PLT cross-feeding on roots Assays were performed in pasteurized Newberg fine sandy loam, wetted 24b in advance of planting to water holding capacity. Cucumber seeds (Cucumis sativis L. cv. Marketmore 86) were surface disinfested by immersion in 95% ethanol for 1 mm, followed by a 20% (v/v) sodium hypochiorite solution for 15 mm. Seeds were rinsed thoroughly in sterile distilled water and placed on sterile, wetted filter paper within Petri plates to germinate, and incubated at 27°C in darkness for two days before seedlings were removed and inoculated with bacterial suspensions. Inoculum of inducing strains Pf-5, Pf-5 pltR::aacCl, and Pf-5 rpoS::lacZ was produced in KMB broth to encourage PLT production. Inoculum of reporter strain Pf-5 pltB: :inaZ was produced in KMBG1u to discourage pltB induction. Cultures were grown with shaking (200 rpm) overnight at 27°C. Cells were then harvested by centrifugation (2,500 x g, 5 mm) and resuspended in 20 mM potassium phosphate buffer (pH 7.0). Suspensions of three inducing strains, Pf-5, Pf-5 pltR::aacC] , and Pf-5 rpoS::lacZ were adjusted to an OD6 of 1.0 while a suspension of Pf-5 pltB::inaZ was adjusted to an 0D600 of 0.1. The suspension of reporter strain Pf-5 pltB::inaZ was mixed 1:1 with suspensions of each of the other three strains, providing three inoculum mixtures in which the reporter and inducing strains were present at final cell density ratios of 1:10, respectively. Cucumber seedlings were inoculated by soaking them in the bacterial suspensions for 10-20 mm. Inoculated seedlings were planted individually into wells of plastic punch trays (35 mm diameter; 25 ml of soil per well). Treatments within trays were arranged by a complete randomized design. Punch trays were placed in closed plastic containers, and incubated in darkness at 20°C and 100% humidity. At 0, 10, 24, and 48 h, seedlings were retrieved from ten replicate wells, placed in culture tubes containing 20 mM potassium phosphate buffer (pH 7.0), and sonicated for 5 mm in a bath-style sonicator, to remove bacterial cells from roots of the seedlings. Serial ten-fold dilutions were made in 20 mM potassium phosphate buffer (pH 7.0) for determination of INA and CFU. Total CFU per sample were determined by spreading diluted samples onto KMB agar containing streptomycin, while CFU of the reporter 53 strain Pf-5 pltB::inaZ were enumerated separately on KMB agar amended with streptomycin and kanamycin. 2.3.6. Data analysis Treatment means were compared using a Student's t-test following an analysis of variance. Two-tailed P values are reported. Statistical analyses were performed using JMP, version 3 (SAS Institute, Cary, NC). 2.4. Results 2.4.1. PLT production is autoreuIated To evaluate the possibility that PLT induces its own production in Pf-5, PLT was added to NBG1y prior to inoculation with either Pf-5 pltB.:inaZ or Pf-5 iceC. In preliminary experiments, Pf-5 and Pf-5 iceC were similar in production of PLT and 2,4-DAPG (data not shown), so Pf-5 iceC' filled a dual role as an experimental Plt strain for evaluation of antibiotic production and a positive control for NA. Similarly, Pf-5 pltB: :inaZ provided a negative control for PLT production while serving as a reporter of PLT biosynthetic gene transcription. The concentrations of PLT in PLTamended cultures of Pf-5 pltB::inaZ, a Plt strain, initially reflected the amount detectable in supernatant extracts as a result of PLT amendment, but PLT concentrations declined over time in these cultures, presumably because the compound was degraded (Fig. 2.2a). Amendment of NBG1y with PLT resulted in earlier, enhanced PLT production by Pf-5 iceC' (Fig 2.2a). At 48 h, PLT was detectable in both non-amended and PLT-amended cultures of Pf-5 iceC, but PLT concentrations were 4.5 times greater in amended cultures of Pf-5 iceC than in non-amended cultures, after subtracting from the amount of PLT detected in amended Pf-5 iceC cultures the amount detected in amended cultures of Pf-5 pltB::inaZ. As expected, no PLT was detected in Pf-5 pltB::inaZ cultures grown in non-amended medium. 54 Although some variation occurred in the timing and levels of PLT production across multiple time course experiments that we performed, PLT amendments at 4 jig/mI to initial culture medium always hastened PLT production by Pf-5 or Pf-5 icecF, and enhanced it between 4.5- and 20-fold by late stationary phase. 55 10 a. 14 12 0 10 . E 0 10 0 8 3 6 0 0 0 I. 0 .0 0. Ce Ce . 2 '0 1 0 11 10 U 9 0 8 p Pf-Splt&:inaZ -- Pf-5plt&:inaZ+PLT 7 6 12 18 24 30 36 42 48 Time (hours) Figure 2.2. Influence of exogenous PLT on PLT and 2,4-DAPG production by derivatives of Pf-5. Cultures were grown in nonamended NBGIy or in NBG1y amended with 4 .Lg/ml PLT. At the indicated sampling times following inoculation, PLT concentrations (a), 2,4-DAPG concentrations (b), and population densities of cultures (c) were assessed. Population sizes did not differ significantly between PLTamended and nonamended treatments at any time point. All three panels represent data from the same experiment. Error bars denote one standard error. 56 To estimate the minimum concentration of PLT needed to induce PLT autoregulation, Pf-5 was grown in NBG1y amended with PLT at concentrations ranging from 0.004 to 4,000 ng/ml (Fig. 2.3a). Amendment of NBGIy with PLT at concentrations greater than or equal to 0.4 ng/ml (1.47 nM) enhanced PLT production by Pf-5 (P = 0.002), with PLT production increasing with increasing concentrations of PLT up to 400 ngfml in the initial growth medium. Addition of 4 j.tg!mI PLT to the growth medium did not significantly alter the population sizes of Pf-5 iceC or Pf-5 pltB::inaZ (all P> 0.05; Fig. 2.2c); lower concentrations similarly had no effect on bacterial populations (P 0.53; Fig. 2.3b), indicating that the influence of exogenous PLT on PLT production was not achieved through indirect effects on bacterial growth. 57 16 14 a. . e PLT in Pf-5 due to biosynthesis ' d 0 Pf-5 pltB::inaZ 12 ...... 10 C 8 -=0 6 4 2 0 10x10 b. 8x10' 6x109- 4x10' 2x109 0 0.004 0.04 0.4 4 40 400 4000 ng/mI PLT initially present in medium Figure 2.3. Minimum effective concentration of PLT for autoinduction in Pf-5. PLT concentrations (a) and CFU (b) were measured in Pf-5 and Pf-5 pltB::inaZ cultured for 1 8h in NBGIy that was amended with the concentrations of PLT indicated on the abscissa. Black symbols connected by a solid line represent actual concentrations of PLT recovered from Pf-5 and black symbols connected by a dotted line represent PLT that resulted from biosynthesis by Pf-5, approximated by subtracting concentrations of PLT recovered from Pf-5 pltB::inaZ cultures (P11) amended with identical concentrations of PLT. PLT amendments did not significantly alter population sizes. Differing lefters above symbols denote treatments significantly different from one another (P 0.05; two-sample t-test). Error bars denote one standard error. 2.4.2. PLT and 2.4-DAPG exhibit mutual inhibition Pit- mutants produced more 2,4-DAPG than did Pitt strains, as exemplified by comparing 2,4-DAPG production in cultures of Pf-S iceC and Pf-5 pltB::inaZ (Fig. 2.2b). This observation is consistent with the hypothesis that PLT represses 2,4DAPG production in Pf-5. To test this hypothesis, the influence of PLT on 2,4-DAPG production was determined for Pf-5 iceC and Pf-5 pltB. :inaZ grown inNBGIy, a medium conducive to both PLT and 2,4-DAPG production. Exogenous PLT (4 jig/mi) added to NBG1y repressed 2,4-DAPG production by both strains. At 12 h, when 2,4-DAPG was at peak levels, PLT repressed 2,4-DAPG production in Pf-5 iceC and Pf-5 pltB.:inaZ by a factor of 3.8 and 13.6, respectively (Fig. 2.2b). The reverse scenario was also tested: we examined whether 2,4-DAPG repressed PLT production by Pf-5. Prior to inoculation with Pf-5 iceC and Pf-5 pltB::inaZ, NBGIy was amended with 12.5 jig/mi 2,4-DAPG. Amendment of the medium with 2,4DAPG reduced PLT production by Pf-5 iceC at 18 h after inoculation, when PLT production was first detected, and at each time point thereafter (Fig. 24a). As expected, Pf-5 pltB::inaZ did not produce detectable PLT. Again, despite some variation in timing and levels of PLT production across multiple time course experiments, 2,4-DAPG amendments at 12.5 jig/mI to initial culture medium always slowed PLT production in Pf-5 or Pf-5 late stationary phase. iceC and repressed it between 2- and 5-fold by 16 a. 14 12 10 © 8 4 2 0 11 b. 10 U 9 0 7 0 6 12 18 24 30 36 42 48 Culture age (hours) Pf5 iceC 0 Pf5 iceC + 2,4-DAPG V Pf-5pltB::inaZ -v-- Pf-5 pltB::inaZ + 2,4-DAPG Figure 2.4. Influence of exogenous 2,4-DAPG on PLT production by derivatives of Pf-5. Cultures were grown in nonamended NBG1y or in NBG1y amended with 12.5 jig/mi 2,4-DAPG. At the indicated sampling times following inoculation, PLT concentrations (a) and population densities of cultures (b) were assessed. Population sizes did not differ significantly between 2,4-DAPG-amended and nonamended treatments at any time point. Error bars denote one standard error. Symbols denoting PLT concentrations for unamended Pf-5 pltB::inaZ cultures are occluded by those denoting 2,4-DAPG-amended Pf-5 pltB::inaZ cultures. To estimate the minimum concentration of 2,4-DAPG needed to inhibit PLT production, Pf-5 was grown in NBG1y amended with 2,4-DAPG concentrations ranging from 0.0 125 to 12,500 ng/ml (Fig. 2.5a). Amendment of NBG1y with 2,4- DAPG at concentrations greater than or equal to 125 ng/ml (0.595 jiM) inhibited PLT production by Pf-5, with PLT production decreasing as the initial concentrations of 2,4-DAPG in the growth medium exceeded that level (P 0.04). In contrast, when cultures were amended with very low concentrations (between 0.0125 ng/ml and 0.125 ng/ml, or 59.5 pM to 0.595 nM) of 2,4-DAPG, PLT production was reproducibly enhanced over that in non-amended cultures (P = 0.04). Addition of 12.5 jig/mi 2,4-DAPG to the growth medium did not significantly alter the population sizes of Pf-5 iceC or Pf-5 pltB : : inaZ (P 0.09; Fig. 2.4b); lower concentrations similarly had no effect on growth of Pf-5 or Pf-5 pltB::inaZ (P 0.55; Fig. 2.5b), indicating that differences in PLT production could not be attributed to an impact of 2,4-DAPG amendments on growth. 61 5 a. 4 b b ac e f I 0 I I I I 1Z. I2 10x109 b. 8x109 6x10' 14 4x109 2x109 U U.UIL U.1 I.Z 1,h3U Jh,W ng/mI 2,4-DAPG initaily present in medium Figure 2.5. Minimum effective concentration of 2,4-DAPG for repression of PLT production in Pf-5. PLT concentrations (a) and CFU (b) were measured in Pf-5 after culturing for 18 h in NBG1y that was amended with the concentrations of 2,4-DAPG indicated on the abscissa. 2,4-DAPG amendments did not significantly alter population sizes. Differing letters above symbols denote treatments significantly different from one another (P 0.05; two-sample t-test). Error bars denote one standard error. 62 2.4.3. Effects of PLT and 2.4-DAPG on PLT production occur at the transcriDtional level The effects of exogenous PLT and 2,4-DAPG on PLT production were accompanied by corresponding effects on transcription from PLT biosynthetic gene promoters, as measured by activity of the ice nucleation transcriptional reporter gene (inaZ) fused to three separate chromosomal PLT biosynthesis genes, pltB, pitE, and pltF. Because preliminary experiments indicated that inaZ expression fromplt promoters reached a maximum by 24 h after inoculation, JNA was measured periodically from 0 to 24 h after inoculation (Fig. 2.6). INA was enhanced in all three reporter strains when grown in NBGIy amended with 4 .tg/ml PLT. At 24 h after inoculation, INA expressed by Pf-5 derivatives containing pltB::inaZ,pltE::inaZ, and pltF: :inaZ was enhanced by one to two orders of magnitude by the addition of PLT. Conversely, transcription of PLT biosynthetic genes was reduced when NBGIy was amended with 2,4-DAPG (12.5 pg!ml) prior to inoculation. At 24 h after inoculation, 2,4-DAPG amendment of the culture medium reduced INA expressed by Pf-5 derivatives containing pItB: : inaZ, pitE: :inaZ, and pltF: :inaZ by three to twelve-fold, respectively. These data provide evidence that autoinduction of PLT and repression of PLT production by 2,4-DAPG both occur, at least in part, at the level of transcription from PLT biosynthetic gene promoters. Addition of 4 tg/ml PLT or 12.5 tg!ml 2,4DAPG to the growth medium did not significantly alter the population sizes of any strain at any timepoints sampled (all P> 0.05; Fig. 2.6d). 2 a. plL&:inaZ b. pltE:inaZ * * 0 * I: -8 -10 0 6 12182430 I c. pIth:inaZ 0 612182430 0 6 12182430 0 6 12 18 24 30 Cuiture age (hours) I pltB::inaZ pli&:inaZ+PLT 4 jiL&uuzZ+2,4-DAPG A pfr:inaZ ...A.... pItE:inaZ+PLT pfrE:ina2+2,4-DAPG plLF.:inaZ p1tF:inaZ+ILT 4-- p11F:inuZ+ 2,4-DAPG Figure 2.6. Positive and negative influences by PLT and 2,4-DAPG, respectively, on transcriptional activity within the PLT biosynthetic gene cluster. Transcriptional fusions were previously made of the ice nucleation reporter gene (inaZ) to three PLT biosynthetic genes. Panels a-c depict the response of individual reporter strains to exogenous PLT and 2,4-DAPG. Pf-5 pltB: :inaZ (a, square symbols); Pf-5 pitE: :inaZ (b, triangular symbols); and Pf-5 pltF::inaZ (c, round symbols); were grown in unamended NBG1y (solid lines), or in NBG1y amended with 4 j.ig/ml PLT (dotted lines) or 12.5 jig/mI 2,4-DAPG (dashed lines). Population densities (d) are shown for all strains, and did not differ significantly among treatments at any time point. Asterisks above symbols denote treatment means which differ significantly from the mean of the unamended treatment at the identical timepoint (P 0.05; Student's ttest). Error bars denote one standard error. 64 2.4.4. PLT autoinduction is superseded by eneticaIly and environmentally repressive backgrounds In NBG1u, a medium that is not conducive to PLT production by Pf-5 (30, 43), the addition of 4 j.tg/ml PLT did not result in detectable PLT production by the bacterium (data not shown). Thus, PLT amendments in NBGIy and NBG1u had different effects on PLT production by Pf-5, suggesting that the PLT autoinduction phenomenon was overridden by glucose repression. Similarly, Pf-5 pltR: :aacCl and Pf-5 gacA::aphl derivatives grown in NBG1y failed to produce detectable concentrations of PLT, even in the presence of exogenous PLT (data not shown), suggesting that autoinduction does not supersede control of PLT production by the LysR transcriptional regulator encoded bypltR or the by the GacA/GacS histidine kinas c/response regulator pair. Other workers observed that autoinduction of 2,4DAPG is not entirely subordinate to gacA in CHAO (54), indicating that the mechanisms of autoinduction may differ between the two systems. 2.4.5. PLT autoinduction occurs in the rhizosphere To determine whether PLT could influence PLT biosynthetic gene expression by bacterial populations in the rhizosphere as well as in culture, we performed a PLT cross-feeding assay. The PIf strain Pf-5 pItB::inaZ was used as a reporter strain to detect the presence of PLT produced by other bacteria in the rhizosphere. Cucumber seedlings were coinoculated, at a 1:10 ratio, with Pf-5 pltB::inaZ and one of three other Pf-5 derivatives that differ in PLT production: Pf-5, which produces PLT; Pf-5 rpoS::lacZ, which overproduces PLT; or pltR::aacCl , which does not produce detectable PLT (Fig. 2.7). When coinoculated with pltR::aacCl in the rhizosphere of cucumber, Pf-5 pltB::inaZ produced a mean of -4.0 logio(ice nuclei/CFU) from 10 to 48 h after inoculation, which represented a basal level of pltB::inaZ expression in the absence of PLT cross-feeding. When coinoculated with the PLT-producing strain Pf- 65 5, rhizosphere populations of Pf-5 pltB::inaZ expressed significantly greater INA [-2.5 logio(ice nucleiICFU)}. The greatest level of INA expressed by Pf-5 pltB::inaZ in the rhizosphere, averaging -.1.1 logio(ice nuclei/CFU) from 10 to 48 h after inoculation, occurred when Pf-5 pltB::inaZ was coinoculated with the PLT-overproducer Pf-5 rpoS::lacZ. Thus, PLT produced by P1-5 and the overproducing derivative Pf-5 rpoS::lacZ was perceived by coinoculated cells of Pf-5 pltB::inaZ in the rhizosphere, as evidenced by enhanced transcription of pltB in those coinoculated cells. Population sizes (mean log CFU/seedling) of the inducing strains averaged 6.5 at the time of inoculation, and increased to 6.8 over 48 hours. Log CFU/seedling for the reporter strain coinoculated with each inducer averaged 5.3 at the time of inoculation and increased to 5.9 over 48 hours. While slight differences in the rhizosphere population sizes of Pf-5 pltB::inaZ were observed among treatments at two time points, these differences were not found to significantly affect NA determinations. 0 -1 z -4 -5 -6 -7 0 10 24 48 Hours post-inoculation and planting Pf-5 pltR::aacCl; P1t V Pf-5;Plt Pf-5 rpoS:.iacZ, Plt Figure 2.7. Autoinduction of PLT as a result of cross-feeding in the rhizosphere. A Pif reporter strain (Pf-5 p/tB:: inaZ) was coinoculated onto cucumber seedlings at a cell density ratio of 1:10 with each of the following inducing strains: Pf-5 (Plt), Pf-5 pltR: :aacCl (Pit) and Pf-5 rpoS::IacZ (Plt). After seedlings were grown in pasteurized soil for the durations indicated, bacteria were harvested and evaluated for INA conferred by Pf-5 pftB::inaZ in the presence of each coinoculant. Values represent the mean of ten replicate seedlings and error bars indicate one standard error, which are obscured by the symbols for some timepoints. 67 2.5. Discussion The data presented in this paper provide strong evidence that PLT serves as an autoregulator, positively influencing its own production in P.fiuorescens Pf-5. Three lines of evidence from experiments with Pf-5 and derivative strains support this conclusion: 1) exogenous PLT enhanced PLT production, ii) exogenous PLT enhanced transcriptional activity of three PLT biosynthetic genes, and iii) coinoculation with Plt cells induced expression of PLT biosynthetic gene transcription by a Pit- reporter strain in the rhizosphere. Furthermore, PLT functioned as an inducer of its own production when present in culture medium at an extremely low concentration (1.47 nM). In this respect, PLT is similar to established signaling molecules, such as the A-factor of Streptomyces griseus (19, 22, 27), or AHLs of Vibriofischeri and Serratia liquefaciens (13, 14), which influence cell physiology and gene expression when extracellular concentrations are in the nanomolar range. In addition to its autoregulatory role, PLT influenced the production of at least one other secondary metabolite, 2,4-DAPG. Exogenous PLT repressed 2,4-DAPG production by Pf-5, as was previously reported in the related strain, P.Jluorescens CHAO (54). Thus, in addition to its established extracellular role as an antibiotic, PLT contributes to regulation of at least two metabolic pathways within the bacterial cell. The roles of antibiotics in the ecology and physiology of microorganisms have been debated for decades, and evidence is accruing to support intracellular roles for bacterial antibiotics. For example, P. aeruginosa produces a 4-quinolone compound (PQS, for Pseudomonas guinolone igna1) that, when present in micromolar concentrations, plays an integral role in signal transduction of producing bacterial cells (38, 46). Although toxicity has not been demonstrated for PQS itself (46), related molecules in the 4-quinolone class have antibacterial activity and are used medicinally (55). More recently, it has been shown that exposure to subinhibitory concentrations of several commonly used antibiotics results in altered transcription of as many as 5% of genes in Salmonella typhimurium (15); by extension, one might predict that these antibiotics also have physiological relevance in the cells in which they are produced. The finding that nanomolar concentrations of PLT can enhance PLT production and biosynthetic gene expression by Pf-5 provides another example of a regulatory role for antibiotics. PLT biosynthesis genes were transcribed in early to mid-log phase, but PLT did not accumulate to detectable levels until cells were entering or in stationary phase. The lag between induction of gene expression and accumulation of detectable levels of PLT might be explained by slow rates of PLT biosynthesis, or by posttranscriptional control mechanisms like those affecting the antibiotics 2,4-DAPG and hydrogen cyanide in the related P. fluorescens strains CHAO and Fl 13 (1, 5, 6, 20). Transcripts of genes for biosynthesis of hydrogen cyanide and 2,4-DAPG are sequestered by a small RNA-binding protein (RsmA), but a noncoding RNA, PrrB (RsmZ), whose expression is regulated by GacA/GacS, can in turn bind RsmA, and thus relieve RsmA suppression in these strains. An alternative hypothesis is that low levels of PLT, below our detection limits, may gradually accumulate in cultures until a threshold rate of PLT re-entry into producer cells is achieved, whereupon PLT begins autoinduction. Such a mechanism is reminiscent of the prototypic LuxlILuxR autoinduction systems, which require accumulation of an autoinducer in the extracellular milieu for induction of gene expression (reviewed in 39). Established positive autoregulation systems, such as LuxIILuxR, rely upon diffusion or secretion and re-uptake of the signal molecule across the membrane, and include an intracellular signal receptor(s) that mediates signal transduction within the cell. Neither a PLT receptor nor a mechanism for PLT transport across the bacterial membrane is known in Pf-5, but candidates for both are found within the PLT biosynthetic gene cluster. The nucleotide sequence of the p/tHu cluster, which is directly adjacent to the PLT biosynthesis genes p/tLABCDEFG, includes genes encoding members of the ABC transporter family (7). Transcription of p/ti and pltf is enhanced by exogenous PLT (7), indicating that regulation of the genes is coordinated with PLT production, as would be expected if the region encodes proteins that function in transport of PLT into or out of the cell. The possibility also remains that the small hydrophobic PLT molecule could diffuse across the bacterial cell membrane. A candidate PLT receptor(s) is PItR, a LysR transcriptional regulator required for transcriptional activity of pltB,pltE, and pith' and for the accumulation of detectable levels of PLT in culture supernatants (42). LysR. family members generally require a cofactor for activity (e.g. reviewed in 51). The findings of this study demonstrate that regulation of 2,4-DAPG and PLT production is interrelated in Pf-5; both compounds were repressive of the other's production. In P. fluorescens strain CHAO, 2,4DAPG autoinduction and repression by PLT, salicylate and a flingal metabolite, fusaric acid, are mediated by PhIF (54), whose repressive binding to the ph1ACBD promoter is directly stabilized by salicylate but alleviated by 2,4-DAPG (2). Whether the mechanism of 2,4-DAPG autoregulation and repression is similar in Pf-5 remains to be determined; we have not ruled out the possibility that precursor competition (e.g. for malonyl-CoA needed to extend the growing polyketide chains) contributes to PLT repression of 2,4-DAPG in Pf-5. The concentration of 2,4-DAPG (595 nM) required for repression of PLT production in Pf-5 was much higher than the concentration of PLT (1.47 nM) required for autoinduction. Furthermore, the bimodal response of PLT production to increasing concentrations of 2,4-DAPG suggests a concentration-dependent role reversal of 2,4- DAPG from activator to inhibitor. The reason for this shift is unknown. However, similar concentration-dependent role reversals have been observed in other systems. For example, in Vibriofischeri MJ-1, AHL concentrations between 0.01 and 5 jig/mi enhanced luciferase activity, but higher concentrations were inhibitory (13). Similarly, in Bacillus subtilis, a small peptide (CSF for competence and porulation factor) controls gene expression in a cell-density dependent fashion. The response of srfA, a CSF-regulated gene for surfactin biosynthesis, was upregulated at low (1-5 nM) CSF concentrations, but inhibited at higher (>20 nM) concentrations of the signal peptide (56). 70 To test the level of PLT autoinduction in the regulatory hierarchy surrounding antibiotic production in Pf-5, exogenous PLT was added to Pf-5 in an environmental condition and in a genetic background repressive to PLT production. The positive effect of exogenous PLT does not supersede the negative effect of a gacA mutation on PLT production, nor can it override repression of PLT production in the presence of glucose. Because low concentrations of PLT are still detectable in Pf-5 cultures grown in the presence of glucose, lack of PLT does not explain the inability for autoinduction. Rather, this result implies that PLT autoinduction requires a factor(s) which is lacking during glucose repression, or is inhibited by a factor(s) produced in the presence of glucose. For the gacA mutant, which produces no detectable pyoluteorin, this observation implies that factors necessary for PLT production or autoinduction are lacking. This study established that PLT functions as signal molecule triggering the transcription of a PLT biosynthesis gene in bacterial cells in the rhizosphere, as well as in culture. When Pf-5 pltB::inaZ was co-inoculated with PLT-producing strains, expression of p/tB, monitored with the inaZ reporter, was enhanced. Thus, PLT served as an intercellular signal between distinct populations of bacterial cells co- inhabiting the rhizosphere. This observation implies that, in the rhizosphere, PLT biosynthetic genes are upregulated by exogenous PLT, and PLT autoinduction contributes significantly to PLT regulation by Pf-5. To our knowledge, this is the first report demonstrating that an antibiotic functions as a signal molecule in a natural environment. PLT is an antibiotic toxic to Pythium ultimum and other plant pathogenic Oomycetes, and has been shown to be an important factor in biological control of such pathogens (24, 37). The finding that PLT serves as signal in the rhizosphere, influencing the expression of a gene required for its biosynthesis by neighboring bacterial cells, has important implications for biological control of plant disease. PLT 71 can now be added to the growing list of microbial metabolites shown to function as signal molecules influencing the expression of biocontrol traits by Pseudomonas spp. in the rhizosphere. Other examples include AHLs (48, 59), which influence the expression of phenazine biosynthesis genes by P. aureofaciens; fusaric acid (40), which represses the expression of 2,4-DAPG biosynthesis genes by P.fiuorescens CHAO; and siderophores (33), which influence the expression of pyoverdine biosynthesis genes by P. putida. When taken together, these examples provide convincing evidence that the expression of biocontrol traits by Pseudomonas spp. is influenced profoundly by signal molecules produced by their coinhabitants in the rhizosphere. It seems likely that secreted metabolites with various functions in the extracellular environment, retrieved or intercepted by producing or neighboring cells, supply information about the ecological and physiological status of the surrounding community. Therefore, the composition and physiological status of the microbial community is likely to be an important factor influencing the expression of biocontrol traits by bacterial antagonists in the rhizosphere, and consequently, their success as biological control agents. 2.6. Acknowledgements We thank Brian Nowak-Thompson and Regina Notz for providing pure PLT and 2,4-DAPG, respectively. We are grateful to Meadow Anderson, Walter Mahaffee, Dmitri Mavrodi, Caroline Press, and Brenda Shaffer and Virginia Stockwell for critical reviews of the manuscript. This research was supported in part by a fellowship to M.B. from the U.S. Environmental Protection Agency Science to Achieve Results program (U-91582801-0). 72 2.7. References 1. Aarons, S., A. Abbas, C. Adams, A. Fenton, and F. O'Gara. 2000. A regulatory RNA (PrrB RNA) modulates expression of secondary metabolite genes in Pseudomonasfluorescens Fl 13. J. Bacteriol. 182:3913-391 9. 2. Abbas, A., J. P. Morrissey, P. Carnicero Marquez, M. M. Sheehan, I. R. Delany, and F. O'Gara. 2002. Characterization of interactions between the transcriptional repressor Ph1F and its binding site at the ph/A promoter in Pseudomonasfiuorescens Fl 13. J. Bacteriol. 184:3008-3016. 3. Bailey, D., R. E. Johnson, and U. J. Salvador. 1973. Pyrrole antibacterial agents. 1. Compounds related to pyoluteorin. J. Med. Chem. 19:1298-1300. 4. Bangera, M. G., and L. S. Thomashow. 1999. Identification and characterization of a gene cluster for synthesis of the polyketide antibiotic 2,4- diacetylphloroglucinol from Pseudomonasfluorescens Q2-87. J. Bacteriol. 181:3155-3163. 5. Blumer, C., and D. Haas. 2000. Multicopy suppression ofagacA mutation by the infC operon in Pseudomonasfluorescens CHAO: competition with the global translational regulator RsmA. FEMS Microbiol. Lett. 187:53-58. 6. Blumer, C., S. Heeb, G. Pessi, and D. Haas. 1999. Global GacA-steered control of cyanide and exoprotease production in Pseudomonasfluorescens involves specific ribosome binding sites. Proc. Nat!. Acad. Sci. USA 96:14073-14078. 7. Brodhagen, M. 2003. Ph.D. thesis. Oregon State University. 8. Cook, R. J.,, L. S. Thomashow, D. M. Weller, D. Fujimoto, M. Mazzola, G. Bangera, and D. -S. Kim. 1995. Molecular mechanisms of defense by rhizobacteria against root disease. Proc. Nat!. Acad. Sci. USA 92 :41974201. 9. Corbell, N. 1999. Ph.D. thesis. Oregon State University. 10. Corbell, N., and J. E. Loper. 1995. A global regulator of secondary metabolite production in Pseudomonasfluorescens Pf-5. J. Bacteriol. 177:6230-6236. 73 11. Cuppels, D. A., C. R. Howell, R. D. Stipanovic, A. Stoessi, and J. B. Stothers. 1986. Biosynthesis of pyoluteorin: a mixed polyketidetricarboxylic acid cycle origin demonstrated by [1 ,2-13C2acetate incorporation. Z. Naturforsch. 41c:532-536. 12. Dowling, D. N., and F. O'Gara. 1994. Metabolites of Pseudomonas involved in the biocontrol of plant disease. Trends Biotech. 12:133-14. 13. Eberhard, A., A. L. Burlingame, C. Eberhard, G. L. Kenyon, K. H. Nealson, and J. J. Oppenheimer. 1981. Structural identification of autoinducer ofPhotobacteriumfischeri luciferase. Biochemistry 20:24442449. 14. Eberl, L., M. K. Winson, C. Sternberg, G. S. A. B. Stewart, G. Christiansen, S. R. Chhabra, B. Bycroft, P. Williams, S. Molin, and M. Givskov. 1996. Involvement of N-acyl-L-homoserine lactone autoinducers in controlling the multicellular behaviour of Serratia liquefaciens. Mol. Microbiol. 20:127-136. 15. Goh, E. -B., G. Yim, W. Tsui, J. McClure, M. G. Surette, and J. Davies. 2002. Transcriptional modulation of bacterial gene expression by subinhibitory concentrations of antibiotics. Proc. Nati. Acad. Sci. USA. 99:17025-17030. 16. Gutterson, N., J. S. Ziegle, G. J. Warren, and T. J. Layton. 1988. Genetic determinants for catabolite induction of antibiotic biosynthesis in PseudomonasJluorescens HV37a. J. Bacteriol. 170:380-385. 17. Haas, D., C. Blumer, and C. Keel. 2000. Biocontrol ability of fluorescent pseudomonads genetically dissected: importance of positive feedback regulation. Cun. Opin. Biotech. 11:290-297. 18. Haas, D., and C. Keel. 2003. Regulation of antibiotic production in rootcolonizing Pseudomonas spp. and relevance for biological control of plant disease. Annu. Rev. Phytopathol. 10.1 146/annurev.phyto.41 .052002.095656 19. Hara, 0., and T. Beppu. 1982. Mutants blocked in streptomycin production in Streptomyces griseus - the role of A-factor. J. Antibiot. 35:349-358. 20. Heeb, S., C. Blumer, and D. Haas. 2002. Regulatory RNA as mediator in GacA/RsmA-dependent global control of exoproduct formation in Pseudomonasfiuorescens CHAO. J. Bacteriol. 184:1046-1056. 74 21. Heeb, S., and D. Haas. 2001. Regulatory roles of the GacS/GacA twocomponent system in plant-associated and other gram-negative bacteria. Mol. Plant-Microbe Interact. 14:1351-1363 22. Horinouchi, S. 1999. y-butyrolactones that control secondary metabolism and cell differentiation in Streptomyces. p. 193-207. In G. M. Dunny and S. C. Winans (ed.), Cell-cell signaling in bacteria. ASM Press, Washington, D.C. 23. Howell, C. R., and R. D. Stipanovic. 1979. Control of Rhizoctonia solani on cotton seedlings with Pseudomonasfiuorescens and with an antibiotic produced by the bacterium. Phytopathology 69:480-482. 24. Howell, C. R., and R. D. Stipanovic. 1980. Suppression of Pythium ultimum-induced damping-off of cotton seedlings by Pseudomonas fluorescens and its antibiotic, pyoluteorin. Phytopathology 70:712-715. 25. Howie, W. J., and T. V. Suslow. 1991. Role of antibiotic biosynthesis in the inhibition of Pythium ultimum in the cotton spermosphere and rhizosphere by Pseudomonasfluorescens. Mol. Plant-Microbe Interact. 4:393-399. 26. Keel, C., U. Schnider, M. Maurhofer, C. Voisard, J. Laville, U. Burger, P. Wirthner, D. Haas, and G. Defago. 1992. Suppression of root diseases by Pseudomonasfluorescens CHAO: importance of the bacterial secondary metabolite 2,4-diacetyiphioroglucinol. Mo!. Plant-Microbe Interact. 5:4-13. 27. Khokhlov, A. S. 1988. Results and perspectives of actinomycete autoregulators studies, p. 338-345. In Y. Okami, T. Beppu, and H. Ogawara (ed.). Biology of Actinomycetes '88. Japan Scientific Societies Press, Tokyo, Japan. 28. King, E. 0., M. K. Ward, and D. E. Raney. 1954. Two simple media for the demonstration of pyocyanin and fluorescein. J. Lab. Clin. Med. 44:301-307. 29. Kraus, J., and J.E. Loper. 1992. Lack of evidence for a role of antifungal metabolite production by Pseudomonasfluorescens Pf-5 in the biological control of Pythium damping-off of cucumber. Phytopathology 82:264-271. 75 30. Kraus, J., and J. E. Loper. 1995. Characterization of a genomic region required for the production of the antibiotic pyoluteorin by the biological control agent Pseudomonasfluorescens Pf-5. Appi. Environ. Microbiol. 61: 849-854. 31. Lamont, I. L., P. A. Beare, U. Ochsner, A. I. Vasil, and M. L. Vasil. 2002. Siderophore-mediated signaling regulates virulence factor production in Pseudomonas aeruginosa. Proc. Nati. Acad. Sci. USA 99:7072-7077. 32. Lindgren, P. B., R. Frederick, A. G. Govindarajan, N. J. Panopoulos, B. J. Staskawicz, and S. E. Lindow. 1989. An ice nucleation reporter gene system: identification of inducible pathogenicity genes in Pseudomonas syringae pv. phaseolicola. EMBO J. 8:1291-1301. 33. Loper, J. E., and M. D. Henkels. 1999. Utilization of heterologous siderophores enhances levels of iron available to Pseudomonasputida in the rhizosphere. Appi. Environ. Microbiol. 65:5357-5363. 34. Loper, J. E., and S. E. Lindow. 1994. A biological sensor for iron available to bacteria in their habitats on plant surfaces. Appi. Environ. Microbiol. 60:1934-1941. 35. Loper, J. E., and S. E. Lindow. 2002. Reporter gene systems useful in evaluating in situ gene expression by soil and plant-associated bacteria, p. 627-637. In C. J. Hurst, R. L. Crawford, G. R. Knudsen, J. J. Mclnemey, and L. D. Stetzenbach (ed.), Manual of environmental microbiology, 2' edition. ASM Press, Washington, D.C. 36. Loper, J. E., B. Nowak-Thompson, C. A. Whistler, M. J. Hagen, N. A. Corbell, M. D. Henkels, and V. 0. Stockwell. 1997. Biological control mediated by antifungal metabolite production and resource competition: an overview. In A. Ogoshi, K. Kobayashi, Y. Homma, F. Kodama, N. Kondo and S. Akino (ed.), Proceedings of the Fourth International Workshop on Plant Growth Promoting Rhizobacteria. Sapporo, Japan. 37. Maurhofer, M., C. Keel, U. Schnider, C. Voisard, D. Haas, and G. Defago. 1992. Influence of enhanced antibiotic production in Pseudomonasfluorescens strain CHAO on its disease suppressive capacity. Phytopathology 82:190-195. 38. McKnight, S. L.,, B. H. Iglewski, and E. C. Pesci. 2000. The Pseudomonas quinolone signal regulates rhi quorum sensing in Pseudomonas aeruginosa. J. Bacteriol. 182:2702-2708. 76 39. Nealson, K. H. 1999. Early observations defining quorum-dependent gene expression. In: G.M. Dunny and S.C. Winans (ed.), Cell-cell signaling in bacteria. ASM Press, Washington, D.C. 40. Notz, R., M. Maurhofer, H. Dubach, D. Haas, and G. Defago. 2002. Fusaric acid-producing strains of Fusarium oxysporum alter 2,4diacetyiphioroglucinol biosynthetic gene expression in Pseudomonas fluorescens CHAO in vitro and in the rhizosphere of wheat. App. Environ. Microbiol. 68:2229-2235. 41. Nowak-Thompson, B. 1997. Ph.D. thesis. Oregon State University. 42. Nowak-Thompson, B., N. Chaney, S. J. Gould, and J. E. Loper. 1999. Characterization of the pyoluteorin biosynthetic gene cluster of Pseudomonasfluorescens Pf-5. J. Bacteriol. 181:2166-1274. 43. Nowak-Thompson, B., S. J. Gould, J. Kraus, and J. E. Loper. 1994. Production of 2,4-diacetyiphioroglucinol by the biocontrol agent Pseudomonasfluorescens Pf-5. Can. 3. Microbiol. 40:1064-1066. 44. Nowak-Thompson, B., S. J. Gould, and J. E. Loper. 1997. Identification and sequence analysis of the genes encoding a polyketide synthase required for pyoluteorin biosynthesis in Pseudomonasfluorescens Pf-5. Gene 204:17-24. 45. Ornston, L. N., and R. Y. Stanier. 1966. The conversion of catechol and protocatechuate to -ketoadipate byPseudomonasputida. J. Biol. Chem. 241:3776-3786. 46. Pesci, E. C., J. B. Milbank, J. P. Pearson, S. McKnight, A. S. Kende, K P. Greenberg, and B. H. Iglewski. 1999. Quinolone signaling in the cellto-cell communication system of Pseudomonas aeruginosa. Proc. Natl. Acad. Sci. USA 96:11229-11234. 47. Pfender, W. F., J. Kraus, and J. E. Loper. 1993. A genomic region from Pseudomonasfiuorescens Pf-5 required for pyrrolnitrin production and inhibition of Pyrenophora tritici-repentis in wheat straw. Phytopatho logy 83:1223-1228. 48. Pierson, E. A., D. W. Wood, J. A. Cannon, F. M. Blachere, and L. S. Pierson III. 1998. Interpopulation signaling via N-acyl-homoserine lactones among bacteria in the wheat rhizosphere. Mol. Plant-Microbe Interact. 11:1078-1084. ":4 49. Reimmann, C., L. Serino, M. Beyeler, and D. Haas. 1998. Dihydroaeruginoic acid synthetase and pyochelin synthetase, products of the pchEF genes, are induced by extracellular pyochelin in Pseudomonas aeruginosa. Microbiol. 144:3135-3148. 50. Sarniguet, A., J. Kraus, M. D. Henkels, A. M. Muehichen, and J. E. Loper. 1995. The sigma factor a' affects antibiotic production and biological control activity of Pseudomonasfluorescens Pf-5. Proc. Nati. Acad. Sci. USA 92:1225-12259. 51. Schell, M. A. 1993. Molecular biology of the LysR family of transcriptional regulators. Annu. Rev. Microbiol. 47:597-626. 52. Schnider, U., C. Keel, C. Blumer, J. Troxier, G. Defago, and D. Haas. 1995. Amplification of the housekeeping sigma factor in Pseudomonas fluorescens CHAO enhances antibiotic production and improves biocontro! abilities. J. Bacteriol. 177:5387-5392. 53. Schnider, U., C. Keel, C. Voisard, G. Defago, and D. Haas. 1995. Tn5directed cloning of pqq genes from Pseudomonasfluorescens CHAO: mutational inactivation of the genes results in overproduction of the antibiotic pyoluteorin. App!. Environ. Microbiol. 61:3856-3864. 54. Schnider-Keel, U., A. Seematter, M. Maurhofer, C. Blumer, B. Duffy, C. Gigot-Gonnefoy, C. Reimmann, R. Notz, G. Défago, D. Haas, and C. Keel. 2000. Autoinduction of 2,4- diacetylphioroglucinol biosynthesis in the biocontrol agent Pseudomonasfiuorescens CHAO and repression by the bacterial metabolites salicy!ate and pyoluteorin. J. Bacteriol. 182:12151225. 55. Smith, J. T., and H. -J. Zeiler. 1998. History and introduction, p. 1-11. In J. Kuhlmann, A. Dalhoff, and H.-J. Zeiler (ed.), Quinolone antibacterials. Springer, Berlin, Germany. 56. Solomon, J. M., B. A. Lazazzera, and A. D. Grossman. 1996. Purification and characterization of an extracel!ular peptide factor that affects two different developmental pathways in Bacillus subtilis. Genes Dev. 10:2014-2024. 57. Whistler, C. A., N. A. Corbell, A. Sarniguet, W. Ream, and J. E. Loper. 1998. The two-component regulators GacS and GacA influence accumulation of the stationary-phase sigma factor a' and the stress response in Pseudomonasfluorescens Pf-5. J. Bacteriol. 180:6635-6641. 58. Whistler, C. A., V. 0. Stockwell, and J. E. Loper. 2000. Lon protease influences antibiotic production and UV tolerance of Pseudomonas fluorescens Pf-5. Appl. Environ. Microbiol. 66:2718-2725. 59. Wood, D. W., F. Gong, M. M. Daykin, P. Williams, and L. S. Pierson III. 1997. N-acyl-homoserine lactone-mediated regulation of phenazine gene expression by Pseudomonas aureofaciens 30-84 in the wheat rhizosphere. J. Bacteriol. 179:7663-7670. 79 Chapter 3. Preliminary evidence for effects of pyrrolnitrin, 2,4diacetyiphioroglucinol, and pyoluteorin on secondary metabolism in Pf-5. 3.1. Introduction Because pyoluteonn (PLT) affects its own production as well as that of another polyketide antibiotic, 2,4-diacetyiphioroglucinol (2,4-DAPG) (Chapter 2), I hypothesized that PLT influences levels of other antibiotics produced by Pf-5. PLT is synthesized by a polyketide synthase complex (14,16) and is secreted from cells of Pf- 5. PRN is remarkably similar in structure to PLT (Fig. 1.1), but PRN is derived from tryptophan (8), and is largely retained within the bacterial cell. Because of the similarity in structures, it was postulated that PLT and PRN might compete for the active site on the hypothetical PLT receptor protein that mediates PLT autoinduction. Experiments were performed to identify potential effects of pyrrolnitrin (PRN) and PLT on one another's production in culture. Pyoverdine is a siderophore compound: it is secreted into the extracellular environment where it chelates ferric iron, and subsequently imported back into the cell to release its bound iron. In P. aeruginosa, pyoverdine uptake is also a signal activating the expression of several virulence factors (10). Because of the pyoverdine- virulence factor connection in P. aeruginosa, I tested whether pyoverdine production might be affected by exogenous PLT. Finally, 2,4-DAPG autoinduces its own production in P. fluorescens CHAO. The results presented below demonstrate that the same is true for 2,4-DAPG production in Pf-5. Together, the results presented herein provide evidence that the secondary metabolites secreted by Pf-5 influence their own and one another's production as a rule, rather than as an exception. 3.2. Methods For all experiments evaluating the effects of PLT, PRN, or 2,4-DAPG amendments on antibiotic production, culture conditions and experimental setup was exactly as described in Chapter 2. Bacterial strains and plasmids used in this study are listed in Table 3.1. Pf-5 and derivative strains were cultured routinely on King's Medium B agar (7) containing streptomycin (100 tg/ml). KMB was supplemented with other antibiotics (kanamycin, 50 .tgIml; gentamycin, 40 pg/ml) when culturing strains with antibiotic resistance. For experiments evaluating antibiotic production or PLT biosynthetic gene transcription, cultures were grown at 20°C with shaking (200 rpm) in nutrient broth (pH 6.8; Difco Laboratories, Detroit, Mich.) without selection and supplemented with 0.5 % (w/v) glycerol for NBG1y, a medium conducive to PLT production (unpublished data). For experiments evaluating antibiotic production or PLT biosynthetic gene transcription in defined medium, bacterial cultures were grown in OS medium (17) amended with 0.5% (w/v) glycerol (OSGIy) at 20°C with shaking (200 rpm). Pure antibiotics dissolved in methanol were added to NBGIy immediately prior to inoculation with experimental cultures; where indicated, an equivalent volume of methanol was added to controls. Amendments were made at the following rates: 4 jtg/ml PLT, 12.5 jtg/ml 2,4-DAPU, and 0.7 .tg/ml PRN. PRN amendments were based on the concentrations of PRN found in stationary phase supernatants, which are similar to concentrations extracted from cells. In non-timecourse experiments, PRN was measured at 48h after inoculation, because it reaches maximum concentrations in late stationary phase cultures of Pf-5. Inoculum of Pf-5 and derivatives was obtained from overnight cultures grown with shaking at 200 rpm at 27°C in KMBGIu, modified from KMB broth by substituting the glycerol with 1% (w/v) glucose to repress PLT production. Cells from twice-washed overnight seed cultures grown in KMBG1u were suspended in NBG1y or OSG1y to an OD6 of 0.1. In all experiments, triplicate 5 ml E31 cultures were grown and evaluated; each experiment was performed at least twice; and the results from a representative experiment are presented. For experiments assessing transcriptional activity of PLT genes using the inaZ reporter gene, samples were harvested periodically and assessed for ice nucleation activity (INA) as described in Chapter 2. Antibiotic extractions and HPLC analyses were performed as described in Chapter 2. Cultures were centrifuged (5,000 X g, 5 mm) to pellet cells, and supematants were extracted as described for recovery of PLT and 2,4-DAPG, which are secreted into the culture medium. Because PRN is found both intracellularly and in the supematant, cell pellets were also extracted for PRN by suspending in 5 ml acetone, vortexing for 10 sec, placing in a sonicating bath for 30 sec, and subsequently centrifuging (10,000 X g, 10 mm) to separate the cell debris from the organic extract. Acetone was removed from extracts under reduced pressure, and the resultant residues were resuspended in 100 tl methanol, from which 10 HPLC as described. PRN was detected at ?. ILl samples were evaluated by 254 nm and t1 = 17.9 mm. The detection limit for PRN was 0.01 j.tglml, and the efficiency of recovery for PRN was 77%. Values for other antibiotics are listed in Chapter 2. 82 Table 3.1. Bacterial strains and plasmids used in this study Strain Strain Number Relevant characteristics" Reference or Source Pf-5 JL4474 Pseudomonasfluorescens Pf-5 rhizosphere isolate; Plt, Ina 5 Pf-5 iceC JL4303 Pf-5 (pJEL1 703); jr Plt, Ina 12 Pf-5 pvd-inaZ JL4304 Pf-5 (pJEL17O1); Rif Kmr Plt, 1na 12 Pf-5 JL4389 P. fluorescens pltB..inaZ Pf-5 pltR.:aacCl pltB:Tn3 -nice derivative of Pf-5, PIt, Ina, 9 KmT JL4563 pltR.:aacCl derivative of Pf-5; P1t, Gm', 14 mi Plasmids pJEL1 696 4 kb fragment containing inaZ cloned into pVSP61, in opposite orientation to lac promoter P. Mukeiji pJEL17O1 8 kb fragment containingpvd genes from P. syringae 31R1 cloned into pJEL1 696, upstream of the promoterless inaZ 12 pJEL1 703 iceC from P. syringae cloned into pVSP61 (nucleotide sequences of iceC and inaZ are nearly identical) 12 pVSP61 Stable plasmid vector containing pVS1 replicon and mobilization region and pl5A replicon; Mob, KmT W. Tucker "Plt, produces pyoluteorin; Pif, does not produce detectable pyoluteorin; 1na, expresses ice nucleation activity at -5°C; maT, does not express ice nucleation activity at -5°C; KmT and GmT, resistant to kanamycin and gentamycin, respectively. 83 33. Results and Discussion 3.3.1. Lack of mfluence by exogenous PLT on PRN In multiple PRN amendment experiments, I failed to see an effect of exogenous PLT on the production of PRN. Similar concentrations of PRN were produced in NBG1y by Pf-5, Pf-5 p/tB ::inaZ, and Pf-5 pltR::aacCl , whether or not the culture medium was amended with 4 pg/ml PLT (Table 3.2). PLT amendment also had no reproducible, statistically significant effect on PRN production by Pf-5 iceC grown in defined OSGly medium (P O.14) (Fig. 3.1). Table 3.2. Influence of exogenous PLT on PRN production by derivatives of Pf-5. Amendment of culture medium PRN concentrations No amendment + 4j.tg/ml PLT Pf-5 1.85±0.12 2.07± 0.21 Pf-5 pltB::inaZ 2.02 ± 0.21 2.24 ± 0.05 Pf-5pltR::aacCl 1.81 ± 0.27 2.27 ± 0.06 (.tg/m1 ± SE) 84 1.6 a. 1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 0 12 24 36 48 60 -Ju.0 ' -. 9.5 lb. 9.0 0 pf_5 iceCF S L) 8.0 Pf-5 keCE + PLT 7.5 s Pf-5 iceC + PRN 70 0 12 24 36 48 60 Culture age at harvest (hours) Fig. 3.1. Influence of exogenous PLT and PRN on PRN produced by Pf-5 iceC. Cultures were grown in nonamended NBG1y, in NBGIy amended with 4 .tg/ml PLT, or in NBG1y amended with 0.7 .tg/ml PRN. At the indicated sampling times following inoculation, PRN concentrations (a), and population sizes of cultures (b) were assessed. Error bars denote one standard error. 85 3.3.2. Minimal influence by exogenous PRN on PLT production The opposite scenario was also tested by evaluating the influence of exogenous PRN (0.7 p.g/ml) on PLT production by Pf-5 iceC grown in OSGIy. In contrast to the lack of an effect of PLT on PRN production, amendment of OSG1y with PRN repressed PLT production by Pf-5 iceC slightly in one trial (14.2% reduction; P = 0.01) although the repressive effect was not significant in another experiment (data not shown). Amendment of OSG1y with PRN also reduced INA in Pf-5 pltB::inaZ between 10 and 20% at all timepoints after 24 h in two separate experiments, but the repressive effect was only significant at 22 h (12.7%) in one experiment, and at 44 h (13.9%) in another experiment (both P 0.04) (Fig. 3.2). PRN amendments also affected the production of PRN itself. PRN amendments caused a reproducible increase in PRN production by late stationary phase (67 hours) (P = 0.013). It remains to be determined whether this positive feedback regulation occurred at the level of transcription. The mutually repressive impacts of 2,4-DAIPG and PLT on one another's production occurs, at least in part, at the level of transcription (Chapter 2; Press and Loper, unpublished data). The slight inhibitory effect of PRN on PLT gene transcription and production cannot be explained by precursor competition, as it is unlikely that these compounds share precursors (8,14). 1 0 0 -1 -2 -3 -4 -5 v -6 -7 0 6 12 18 Ff-5pltB::inaz Pf-5 pltB::inaz + PLT Pf-5pltB::inaz + PRN 24 30 36 42 48 Culture age (hours) Fig. 3.2. Positive and negative influences by PLT and PRN, respectively, on transcriptional activity of the PLT biosynthetic gene, p/tB. A transcriptional fusion was previously made of the ice nucleation reporter gene (inaZ) to the chromosomal p/tB gene, to create Pf-5 pltB::inaZ. The transcriptional reporter strain was grown in unamended NBGIy (round symbols), in NBGIy amended with 4 tg/ml PLT (triangular symbols) or in NBG1y amended with 0.7 j.tg/ml PRN (square symbols). Error bars denote one standard error, and are occluded by symbols at many points. 3.3.3. Autoinduction by 2,4-DAPG in P. fluorescens Pf-5 In P. fluorescens CHAO, 2,4-DAPG autoinduces its own production. The results of this study demonstrate that the same is true for 2,4-DAPG production in Pf- 5. 12.5 ig/ml 2,4-DAPG was added to NBG1y prior to inoculation; this concentration represents the low end of concentrations of 2,4-DAPG normally detected in NBG1u, the medium typically used for 2,4-DAPG production (4, 15, 22). 2,4-DAPG concentrations were enhanced in the 2,4-DAPG amended cultures, although the timing and magnitude of 2,4-DAPG production and autoinduction by Pf-5 were variable among experiments under the culture conditions used in these studies. Two selected experiments are shown (Fig. 3.3). In Fig. 3.3a, at 12 hours, amended cultures contained 3 to 29-fold more 2,4-DAPG than nonamended cultures, even after accounting for the original 2,4-DAPG amendment (by subtracting 2,4-DAPG concentrations detectable at the earliest timepoint). In contrast, in Fig. 3.3c, amended cultures contained 2 to 8-fold more 2,4-DAPG than nonamended cultures, after accounting for the original 2,4-DAPG amendment. The average recovery for 2,4DAPG using the extraction method described in Chapter 2 is 45%. Thus, approximately 5.6 jig/mI was expected from extractions of 2,4-DAPU in amended cultures at the earliest timepoint. However, 2,4-DAPG concentrations in amended cultures of the experiment in Fig. 3.3.a were less than 2 tg/m1, suggesting that 2,4- DAPG recovery was lower than normal. Thus, inconsistent extraction efficiencies may explain some of the experiment-to-experiment variation in 2,4-DAPG autoinduction. As noted by members of our laboratory as well as those of other laboratories (L. Thomashow and G. Phillips, personal communication), detection of 2,4-DAPG from culture supernatants is quite variable, and 2,4-DAPG productionmay even appear to cease altogether for certain laboratory strains. ilo : = 0 10 8 0 U 6 0 P9.4 0. 4 U ' 2 I 0 0 11 10 .9 6 12 18 24 30 36 42 48 7,-== 8 Pf5keC 0 Pf5 weC+ 2A-DAPG ,-- Pf.5plzB::inaZ v-- Pf-5pltB::inaZ+ 2,4-DAPG 6 12 26 -- Pf 5 icecF 0-- P15 iceC+ 24-DAPG -- Pf5pftB::iaZ v Pf.5pltB::inaZ+2,4-DAPG Fig. 3.3. Influence of exogenous 2,4-DAPG on production of 2,4-DAPG by derivatives of Pf-5. Pf-5 iceC (round symbols) is a Plt strain, and Pf-5 pltB::inaZ (triangular symbols) is a Plf strain. Cultures were grown in nonamended NBGIy (filled symbols) or in NBGIy amended with 12.5 tg/ml 2,4-DAPG (open symbols). At the indicated sampling times following inoculation, 2,4-DAPG concentrations (a,c) and population densities of cultures (b,d) were assessed. Population sizes did not differ significantly between 2,4-DAPG-amended and nonamended treatments at any time point. Two separate experiments (a,b and c,d) are shown to demonstrate the variation in observations. Error bars denote one standard error, and are occluded symbols at some time points. Confirming that 2,4-DAPG exerts positive autoregulation in Pf-5 as it does in CHA0 was expected, because of the genetic (e.g. 13, 19) and biochemical similarity of these strains. However, a regulatory role for a particular compound may not necessarily be extrapolated to every strain in which that compound is produced. For instance, pyoverdine exhibits evidence of positive autoregulation in P. fluorescens Ml 14, P. putida WCS358, and Pseudomonas strain BlO (1,3, 21), but does appear to do so in P. aeruginosa PAO1, based on studies in BlO and PAO1 comparing the effects of pyoverdine amendments upon activity from structurally similar promoters preceding genes that encode functionally interchangeable gene products (1). 3.3.4. Exogenous PLT represses expression of pyoverdine biosynthesis genes Although it is not known whether pyoverdine autoregulates its production in Pf-5, a serendipitous observation suggests that it, too, may be under regulation by another secondary metabolite. Strain Pf-5 pvd-inaZ is a reporter strain for inaZ expression from the promoter precedingfpvA, a gene encoding the outer membrane ferric-pyoverdine receptor from P. syringae. Pf-5 pvd-inaZ is used as an iron reporter strain. I discovered that adding PLT at 4 tg/ml to NBG1y prior to inoculation with Pf5 pvd-inaZ reduced transcription offpvA. INA from Pf-5 pvd-inaZ remained similar between 0 and 48 hours in the presence of exogenous PLT, whereas expression increased approximately 150-fold over 48 hours in unamended cultures, in two separate experiments (Fig. 3.4). Because these results are from an unexpected observation rather than a planned experiment, several conditions throw doubt on the results. First, thepvd genes used to construct the pvd-inaZ reporter are from another species, P. syringae. Secondly, the pvd-inaZ reporter gene is plasmid-bome and therefore present in multiple copies. Thus, the expression of INA from the pvd::inaZ fusion is likely to be several-fold higher than expression from the chromosomalpvd genes. Third, the medium used for this experiment was NBGIy, an undefined medium that is not an optimal choice for measuring the expression of iron-responsive genes. Fourth, PLT autoinduction occurred as a result of the PLT addition (data not shown). It is possible that the excessive amount of PLT produced by these cultures placed a metabolic load on Pf-5 (pvd::inaZ) so that the response was a decrease in the synthesis of other energetically expensive metabolites, such as pyoverdine. Fifth, it is unclear whether the effect of PLT on pyoverdine gene expression is due to direct effects of one compound upon the other's production, or whether PLT might alter the iron availability to Pf-5 cells and thus indirectly affect pvd gene expression. A follow-up experiment to adthess these issues and test the verity of the results presented here might include the following parameters: 1) The real-time RT-PCR, rather than the Pf-5 pvd-inaZ reporter strain, to assess the effect of PLT on the fpvA homolog in Pf-5, as well as pyoverdine biosynthesis genes 2) The use of a defined medium, such as RSM, with or without a source of iron, such as ferric citrate (e.g. as in reference 11) 3) The use of a Pf-5 pvd::inaZ strain that was also P11, to circumvent the issue of metabolic load by PLT autoinduction Recent findings have linked ferric-pyoverdine uptake by FpvA to regulation of multiple virulence factors in P. aeruginosa. The binding of ferric-pyoverdine to FpvA generates a signal which is passed to the antisigma factor FpvR, and then to either FpvI or PvdS (both xtraçytopIasmic family, or ECF, sigma factors). Binding of FpvI and PvdS to the RNA polymerase results in transcription of thefpvA gene (FpvI) or of genes for biosynthesis of at least three exoproducts: pyoverdine, exotoxin A, and PrpL endoprotease (PvdS) (2, 10, 18, 20). Therefore, follow-up of a possible analogous link between FpvA and exoproducts of Pf-5 could be very exciting. 91 0.0 Exp.1 -0.5 -1.0 -1.5 -2.0 0. -2.5 -3.0 0 I I 10 20 Pf-5 (pvd-inaZ) Pf-5 (pvd-inaZ) 30 PLT 40 0 -a I:: -2. -3, Culture age (hours) Fig. 3.4. Negative influence by PLT on transcriptional activity of a ferric-pyoverdine uptake gene. A transcriptional fusion of a pyoverdine biosynthesis and uptake region to the promoterless ice nucleation reporter gene (inaz) was made previously. Pf-5 pvd- inaZ contains the transcriptional fusion on the stable plasmid pVSP6 1. Pf-5 pvdinaZ was grown in unamended NBG1y (solid symbols), or in NBG1y amended with 4 pg/ml PLT (open symbols). Error bars denote one standard error. 3.4. Conclusions Results from Chapter 2 demonstrated that exogenous PLT suppresses 2,4- DAPG production. The results from this chapter expand those results to suggest the involvement of two more secondary metabolites (PRN and pyoverdine) of Pf-5 in cross-regulation of one another's production. Although the repression of the pyoverdine reporter by exogenous PLT was observed under non-optimal experimental conditions, this result was recently confirmed in the parental strain, P. fluorescens Pf-5 using both DNA hybridization analysis (microarrays) and real-time RT-PCR. Recent experiments (Caroline Press, personal communication) showed that expression of pvsA, a pyoverdine synthetase gene, is repressed in Pf-5 in the presence of exogenous PLT. Notably, I was unable to demonstrate a significant effect of exogenous PLT upon a structurally similar compound, PRN. Thus, although PLT affects the production of at least three (PLT, 2,4-DAPG, and pyoverdine) secreted metabolites of Pf-5 under the conditions tested, its influence is not so global that it influenced all secondary metabolites that I tested in Pf-5. In conclusion, I provide evidence for positive feedback of three secondary metabolites produced by Pf-5: PLT (Chapter 2), 2,4-DAPG, and PRN. Each of these compounds is important for biological control because of its antibiotic properties. Pyoverdine exhibits autoinduction in several related strains; it is not known whether this siderophore also undergoes autoinduction in Pf-5. From examples in the literature, it appears that negative feedback control for secondary metabolites is a more common phenomenon than positive feedback control. The most familiar example of the latter is positive autoregulation by N-acyl-homoserine lactones (AHLs), as described in Chapter 1. The rationale generally given for autoinduction of AHLs is that they provide a mechanism for rapidly ramping up a response to a stimulus. In many cases, the response involves the production of exoproducts secreted into the extracellular environment. Could the antibiotics of Pf-5 be signals for the ramping up 93 of such a response - a response that includes the production of more of the same compound that acted as a signal? What is the significance of the repression, by certain secondary metabolites, of the production of others? Recent DNA hybridization experiments, using an array comprising a limited set of Pf-5 genes important to biological control, have shown that exogenous PLT affects the transcription of a number of these genes. Thus, during PLT autoinduction, increased extracellular PLT concentrations will be expected to wreak profound effects on the metabolism of the cell. The imminent availability of the entire genome sequence of Pf-5, and the possibility for an unabridged Pf-5 microarray, will permit further exploration of the transcriptional responses of Pf-5 genes to autoinducmg antibiotics, and will build a better understanding of the interrelatedness of these transcriptional responses. 94 3.5. References Ambrosi, C., L. Leoni, and P. Visca. 2002. Different responses of pyoverdine genes to autoinduction in Pseudomonas aeruginosa and the group Pseudoinonasfiuorescens-Pseuclomonas putida. App!. Environ. Microbiol. 68(8): 4122-4126. 2. Beare, P. A., R. J. For, L. W. Martin, and I. L. Lamont. 2003. Siderophore-mediated cell signalling in Pseudomonas aeruginosa: divergent pathways regulate virulence factor production and siderophore receptor synthesis. Mo!. Microbiol. 47:195-207. 3. Callanan, M., N.. Sexton, D. N. Dowling, and F. O'Gara. 1996. Regulation of the iron uptake genes in Pseudomonasfluorescens Ml 14 by pseudobactin Ml 14: the pbrA sigma factor does not mediate the siderophore regulatory response. FEMS Microbiol. Left. 144: 61-66. 4. Corbell, N., and J. E. Loper. 1995. A global regulator of secondary metabolite production in Pseudomonasfluorescens Pf-5. J. Bacteriol. 177:6230-6236. 5. Howell, C. N.., and R. D. Stipanovic. 1979. Control of Rhizoctonia solani on cotton seedlings with Pseudomonasfluorescens and with an antibiotic produced by the bacterium. Phytopathology 69:480-482. 6. Keel, C., D. Weller, A. Natsch, G. Défago, R. J. Cook, and L. S. Thomashow. 1996. Conservation of the 2,4-diacetylphloroglucinol biosynthesis locus among fluorescent Pseudomonas strains from diverse geographic locations. App!. Environ. Microbiol. 62: 552-563. 7. King, E. 0., M. K. Ward, and D. E. Raney. 1954. Two simple media for the demonstration of pyocyanin and fluorescein. J. Lab. Clin. Med. 44: 301-307. 8. Kirner, S., P. E. Hammer, B. S. Hill, A. Altmann, I. Fischer, L. J. Weislo, M. Lanahan, K.-H. van Pee, and J. M. Ligon. 1998. Functions encoded by pyrrolnitrin biosynthetic genes from Pseudomonasfluorescens. J. Bact. 180:1939-1943. 9. Kraus, J., and J. E. Loper. 1995. Characterization of a genomic region required for the production of the antibiotic pyoluteorin by the biological control agent Pseudoinonasfluorescens Pf-5. Appi. Environ. Microbiol. 61(3): 849-854. 95 10. Lamont, I. L., P. A. Beare, U. Ochsner, A. I. Vasil, and M. L. Vasil. 2002. Siderophore-mediated signaling regulates virulence factor production in Pseudomonas aeruginosa. Proc. Nati. Acad. Sci. USA 99:7072-7077. 11. Loper, J. E., and M. D. Henkels. 1999. Utilization of heterologous siderophores enhances levels of iron available to Pseudomonas putida in the rhizosphere. AppI. Environ. Microbiol. 65:5357-5363. 12. Loper, J. E., and S. E. Lindow. 1994. A biological sensor for iron available to bacteria in their habitats on plant surfaces. AppI. Environ. Microbiol. 60:1934-1941. 13. McSpadden Gardener, B. B., K. L. Schroeder, S. E. Kalloger, J. M. Raaijmakers, L. S. Thomashow, and D. M. Weller. 2000. Genotypic and phenotypic diversity ofphlD-containing Pseudomonas strains isolated from the rhizosphere of wheat. Appi. Environ. Microbiol. 66: 1939-1946. 14. Nowak-Thompson, B., N. Chaney, S. J. Gould, and J. E. Loper. 1999. Characterization of the pyoluteorin biosynthetic gene cluster of Pseudomonasfluorescens Pf-5. J. Bacteriol. 181:2166-1274. 15. Nowak-Thompson, B., S. J. Gould, and J. E. Loper. 1997. Identification and sequence analysis of the genes encoding a polyketide synthase required for pyoluteorin biosynthesis in Pseudomonasfluorescens Pf-5. Gene 204:17-24. 16. Nowak-Thompson, B., S. J. Gould, J. Kraus, and J. E. Loper. 1994. Production of 2,4-diacetylphloroglucinol by the biocontrol agent Pseudomonasfluorescens Pf-5. Can. J. Microbiol. 40:1064-1066. 17. Ornston, L.N. and R. Y. Stanier. 1966. The conversion of catechol and protocatechuate to -ketoadipate by Pseudomonas putida. J. Biol. Chem. 241: 3776-3786. 18. Redly, G. A., and K. Poole. 2003. Pyoverdine-mediated regulation of FpvA synthesis in Pseudomonas aeruginosa: involvement of a probable extracytoplasmic-function sigma factor, FpvI. J. Bacteriol. 185:1261-1265. 19. Sharifi-Tehrani, A., M. Zala, A. Natsch, Y. Moënne-.Loccoz, and G. Defago. 1998. Biocontrol of soilbome fungal plant diseases by 2,4diacetylphloroglucinol-producing fluorescent pseudomonads with different restriction profiles of amplified 16S rDNA. Eur. J. Plant Pathol. 104: 631643. 20. Shen, J., A. Meidrum, and K. Poole. 2002. FpvA receptor involvement in pyoverdine biosynthesis in Pseudomonas aeruginosa. J. Bacteriol. 184:3268-3275. 21. Venturi, V., Ottewanger, C., Bracke, M., and P. J. Weisbeek. 1995. Iron regulation of siderophore biosynthesis and transport in Pseudomonas putida WCS358: involvement of a transcriptional activator and of the Fur protein. Mo!. Microbiol. 17: 603-610. 22. Whistler, C. A., N. A. Corbell, A. Sarniguet, W. Ream, and J. E. Loper. 1998. The two-component regulators GacS and GacA influence accumulation of the stationary-phase sigma factor c? and the stress response in Pseudomonasfluorescens Pf-5. J. Bacteriol. 180:6635-6641. 97 Chapter 4. Membrane transport genes play a role in pyoluteorin production by Pseudomonasfluorescens Pf-5 4.1. Introduction Pyoluteorin (PLT), a chlorinated polyketide antibiotic secreted by Pseudomonasfluorescens strain Pf-5 suppresses the plant pathogen Pythium ultimum. A region of the Pf-5 chromosome required for PLT production was identified (34) and sequenced (49, 51) and found to comprise nine biosynthetic genes and a transcriptional regulator. The predicted functions of these genes are sufficient to account for PLT biosynthesis (49). Also within the PLT biosynthetic gene cluster is pltR, which encodes a LysR-type transcriptional regulator (49). ApltR derivative of Pf-5 does not exhibit transcriptional activity from promoters within the PLT biosynthetic region, and produces no detectable PLT. Additional loci found to affect PLT production in Pf-5 or the related strain P. fluorescens CHAO contain global regulatory genes or genes encoding proteins that are responsive to the physiological status of the cell. These include the histidine protein kinase/response regulator pair GacS/GacA (15), the protease Lon (82), sigma factors and aD (62, 64), and pyrrolquinoline quinone, a cofactor required by glucose and alcohol dehydrogenases (65). The pit biosynthetic genes (Fig. 4.1) are contained on at least three separate operons. The divergent transcription of pltL and pltR implies the existence of separate promoters between these genes. Although the proximity ofp/tR and pltM suggests that these genes may form a single operon, RT-PCR experiments repeatedly failed to demonstrate a correlation between the presence of apitR transcript and that of the pltM transcript (data not shown). At least two promoters drive transcription of the pltLABCDEFG gene cluster. One of these has been shown experimentally to reside upstream of pltB (34), presumably in the pltR-pltL intergenic region; another has been shown, via activation in trans of a promoterless reporter gene, to precede either p/tD orpltE (43). 98 Genes encoding enzymatic functions for biosynthesis of secondary metabolites, like PLT, are often clustered with genes for immunity to, or transport of the metabolite (reviewed by 39). Moreover, regulatory genes are frequently included in such regions, whose products coordinate expression of genes within the cluster. For example, the gene cluster encoding 2,4-diacetyiphioroglucinol (DAPG) biosythesis in P. fluorescens contains phiE, a gene encoding a putative integral membrane permease (6) with similarity to genes encoding transporters of sugars and other substrates. Intact phlE is required for optimal DAPG production (7). Also within the DAPG gene cluster isphlF, which encodes a transcriptional repressor of DAPG biosynthetic genes, and whose repression is lifted by binding of its cofactor, DAPG itself (6, 1). In P. syringae pv. glycinea PG4 180, the coronatine biosynthetic gene cluster contains genes encoding regulators of coronatine production (77), additionally, a possible membrane permease is encoded by the linked gene cmaU although its function remains unknown (76). The carbapenem biosynthetic gene cluster from Erwinia carotovora subspecies carotovora also includes linked genes encoding regulatory and resistance functions (41,42). In P. syringae pv. syringae, the syringomycin biosynthetic region contains a gene for regulation of syringomycin production (85) and an ABC transporter predicted to be involved in syringomycin secretion (57). ABC transporters, which are ATPpowered, multi-subunit substrate transporters, frequently afford antibiotic producing organisms (prokaryotic and eukaryotic) a mechanism by which to resist selfintoxication (44). Figure 4.1. Map of PLT biosynthetic gene cluster and relationship to subclones used as templates for sequencing templates. All elements in the region are drawn to size. Lettered abbreviations designate the following restriction sites: H, HindUl; B, EcoRI; B, BamHI; P, PstI; S. Smal. 643 E ortR E E E p2tM pltR E 0 pitA E pitB E H 0 pitC 0 E 0 PS BE 0B p.ItD pitE pitF0pItG BEP H rtI p1.tJ pitH ortI< 1kb 6120 Hi 5569 PP P limit of sequence hifomiafion 6247 H 6130 E H 6246 BP P 6198 S H 6274 Fig. 1 C C 101 Like the examples above, the PLT biosynthetic gene cluster in Pf-5 contains a linked transcriptional regulator necessary for PLT gene expression, pltR (49). I hypothesized that the PLT region might contain additional, unknown regulatory functions. PLT is broadly toxic to Gram negative and Gram positive bacteria (5) but similar concentrations have no significant effect on growth of Pf-5 (11), therefore, I hypothesized that self-resistance or transport functions might be encoded by additional genes flanking the known PLT biosynthetic genes. This report provides evidence for new regulatory and transport functions encoded within the PLT biosynthetic gene region. Adjacent topltG is a region containing pitH, a gene encoding a putative transcriptional regulator, and pill and pltJ, two genes encoding a membrane fusion protein and a subunit of an ABC transporter, respectively. I demonstrate transcriptional regulation of p/tI and p/ti by exogenous PLT. I also provide evidence that p/tI, p/ti, and/or adjacent genes encoding transport functions are necessary for PLT production. The results herein provide the basis for the testable hypothesis that p/tI and plt.J are involved in PLT transport and PLT autoinduction. 4.2. Methods 4.2.1. Bacterial strains, plasmids, and culture conditions Bacterial strains and plasmids used in this study are listed in Table 4.1. E. co/i strains were cultured routinely in Luria-Bertani (LB) medium (61) and maintained on LB agar. Pf-5 and derivative strains were cultivated routinely on Kings Medium B agar (KMB; 33). The following antibiotics were included in growth media as required: spectinomycin, 50 jig/mi; streptomycin, 100 jig/mi; kanamycin, 50 jig/mi; gentamycin, 40 jig/mI (P. fluorescens) or 15 jig/mi (E. co/i); tetracycline, 200 jig/mi (P. fluorescens) or 20 jig/mi (E. coli). 102 Table 4.1. Bacterial strains and plasmids used in this study Strain Strain Number Reference Descriptiona P. fluorescens Pf-5 Pf-5pitB::inaZ JL4474 JL4389 Pf-5 pitR..aacCl JL4563 Pf-5 pItF::aacCl JL4629 Rhizosphere isolate. pitB::Tn3 -nice derivative of Pf5, PIf, that, Kmr pltR:: aacCl derivative of Pf-5, Pit-, Gm' pltF..- aacCl derivative of Pf-5, 31 34 49 This study Pit-, G' Pf-5 pit.!: .inaZ JL4669 Pf-5 p/ti: .inaZ JL4671 pltJ::inaZ derivative of Pf-5, Ina, Kmt. pltJ:.inaZ derivative of Pf-5, This study This study Ina, Kmr Pf-5 pltI..inaZ JL4673 pltI..inaZ derivative of Pf-5, This study Jna, Kmr. Pf-5 pitR:: aacCl, pltJ::inaZ JL4675 Pf-5pltR.: aacCl, JL4677 pitR:: aacCl, pltJ:.inaZ This study derivative of Pf-5, Ina, Gm', plt.J::inaZ pltR:. aacCl,pitJ:.inaZ denvative of Pf-5, ma+ Gm, This study pltR:. aacCl pltI:.inaZ This study r - , Pf-5 pltR:: aacCl, pltI::inaZ JL4679 Pf-5 pitF:. aacCl, pltJ::inaZ JL468 1 , + denvative of Pf-5, ma Gm, I. , pitF:. aacCl pit. J: .inaZ This study , derivative of Pf-5, 1na, Gm', Kmr Pf-5 p/tF.. aacCl, pltJ::znaZ JL4683 pitF:: aacCl pltJ:.inaZ + denvative of Pf-5, ma Gm, , This study I. , KmT Pf-5pltF:: aacCl, pitI:.inaZ JL4685 pltF:: aacCl,pitI::inaZ This study derivative of Pf-5, Jna, Gmr, KmT E. coli S 17-1 DH5cx Res, Mod, recA, Tra V endAl hsdRl7 (rKmK) supE44 thi-1 recAl gyrA96 re/Al 48OdiacZA Ml 5 -. 69 61 103 Table 4.2, continued Plasmids pRK415 pRK2O13 JncPl replicon, polylinker of pUC19, Mob, Tc'. Mobilizing plasmid, Tra, Kmr pUC19 pMGm ColE 1 replicon, Apr. ColE 1 replicon, source of pJEL6O33 32 21 61 47 aacCl (Gm') cassette on a 2.0kb Smal fragment, GmT, Tc'. pJEL1938 pJELI939 pTn3 -nice pJEL55O8 pJEL5569 29kbofPf-5DNAthat complements a Plt mutant cloned into pLAFR3, Tcr Ca. 24 kb of Pf-5 DNA that complements a Pit- mutant cloned into pLAFR3, Tc' pTn3-nice; pMB8 replicon, contains Tn3 -nice, Ap', Kmr 12 kb HindIlI fragment from pJEL1938, cloned into pUC18, 34 34 34 S. Carnegie Apr pJEL6 198 pJEL6O27 pJEL6I 19 pJEL6 120 pJIEL6 199 pJEL6200 1.7 kb PstI fragment from pJEL1 938 containing pltF, cloned into pUC19, Apr 8.1 kb HindIII-BamHI fragment from pJEL1 939 containing orJN, cloned into pUC19, Ap' 5.8 kb EcoRil fragment from pJEL1 939 containing orfN, cloned into pUC19, Apr 1.0 kb EcoRI fragment from pJEL1 939 containing pltN, cloned into pUC19, Apr 3.7 kb PstI fragment containing p/IF disrupted by aacCl cassette at an internal BsmI site, contained in pUC19, ApT, Gmr 3.7 kb KpnI - Hindlil fragment containing pltF disrupted by aacCl cassette at an internal BsmI site, cloned into pRK4l5, TCr. GmT. Mob This study N. Chaney This study This study This study This study 104 Table 4.2, continued Plasmids pJEL6274 pJEL6277 pJEL6279 3.4 kb Smal - HindIH fragment from pJEL1938 encompassing pitI and pltJ, cloned into pUC19, Apr 4.6 kb BamHI - PstI fragment spanning inaZ gene from pTn3nice, cloned into pUC19, Apr, lila 1.2 kb PstI fragment containing nptl gene from pJEL5SO8, cloned into PstI site of This study This study This study pJEL6277, Apr, [r Jna pJEL6283 6.2 kb BamHI - AfihII fragment containing inaZ and nptl from pJEL6279, inserted at the NruI site internal to pltJ in This study 6.2 kb BamHI - AflhII fragment containing inaZ and nptl from pJEL6279, inserted at the SphI site internal to pltJ in r Apr pEJL6274, 6.2 kb BamHI - AJ1lII fragment containing inaZ and nptl from pJEL6279, inserted at the BsaAI site internal to pitI in pJEL6274, Kmr Ap' pJEL6283 linearized at NdeI site and fused to pRK415 linearized at EcoRI site, Kmr, Tcr, Mob pJEL6285 linearized at NdeI site and fused to pRK4l5 linearized at EcoRI site, Kmr, Tc', Mob pJEL6287 linearized at NdeI site and fused to pRK4l5 This study pJEL6274, jr Apr pJEL6285 pJEL6287 pJEL6298 pJEL6299 pJEL6300 This study This study This study This study linearized at EcoRI site, Kf Tc', Mob aApr Gmr, Kr and Tcr indicate resistance to ampicillin, gentamycin, kanamycin, and tetracycline, respectively. Pit indicates lack of pyoluteorin production. Ina indicates ice nucleation activity at -5°C. 105 For experiments evaluating antibiotic production or gene transcription, cultures were grown at 20°C with shaking (200 rpm) in nutrient broth (pH 6.8; Difco Laboratories, Detroit, Mich.) without selection and supplemented with either 0.5% (wlv) glycerol for NBGIy, a medium conducive to PLT production, or with 2% (w/v) glucose for NBGIu, a medium repressive to PLT production by Pf-5 (34, 50). Jnoculum for experimental cultures of Pf-5 and derivative strains was derived from overnight cultures grown with shaking at 200 rpm at 27°C in KMB broth, modified by omitting glycerol and substituting 1% (w/v) glucose (KBG1u) to repress PLT production before inoculation of experimental media. Inoculum was washed twice in sterile KBG1u prior to inoculation, and resuspended in 5 mis of NBGIy or NBGIu to 0D600 of 0.1. Where indicated in the Results, pure PLT dissolved in methanol was added to the culture medium immediately prior to inoculation with Pf-5 or derivative strains. An equivalent volume of methanol was added to negative controls. Bacterial population densities were determined by spreading serial dilutions of cultures onto KMB and counting CFUs on the plates after 48 h incubation at 27°C. In each experiment, triplicate 5 ml cultures were grown and evaluated; each experiment was performed at least twice; and the results from a representative experiment are presented. 4.2.2. DNA manipulations Restriction digests, phenol chloroform extraction and ethanol precipitation of DNA, blunting of DNA fragment ends with the Kienow fragment or T4 polymerase, dephosphorylating of DNA fragment ends with calf intestinal alkaline phosphatase, and agarose gel electrophoresis were performed according to standard procedures (4). Plasmids were isolated from E. coli DHSa using alkaline lysis (4) or the PerfectPrep Plasmid Mini kit (Eppendorf AG, Hamburg). DNA fragments were extracted from agarose gels using the QlAquick gel extraction kit (QIAGEN Inc., USA). E. co/i DH5a and E. coli S 17-1 cells were made chemically competent by treatment with CaCl2 prior to transformation (4). Southern analysis was performed using DNA probes labeled by random priming with either digoxigenin- 11 -dUTP using the Genius kit (Roche Applied Science, Indianapolis, IN) or biotin-14-dCTP using the BioPrime DNA Labeling System (Jnvitrogen, Carlsbad, CA). Hybridizations were performed according to the manufacturers instructions. Custom oligonucleotides for sequencing or for PCR were synthesized by Operon (Alameda, CA), Macromolecular Resources (Colorado State University, Fort Collins, CO) or the Central Services Laboratory (Oregon State University, Corvallis, OR). PCRs were performed with Taq polymerase (Promega, Madison, WI), using the primers listed in Table 4.2. Reaction mixtures were heated for 3 mm at 94°c, and then subjected to 35 cycles of 45 sec at 94°C, 45 sec at the indicated annealing temperature (Table 4.2), and 1.5 to 2.25 mm at 72°C (depending on the product size), prior to a final elongation for 10 mm at 72°C. Table 4.2. PCR primers used in this study Target gene pitA Forward primer (5'-3') cgagaaagccaacttccc pltG ctgttcacctttccctatgcc pitH atctatcaccgccagcatc pitI gtgtucctggtttctcc plt.J gggaactgtggagcatcatcg orjK atgcgtgacaagagcaacc odP attgctgttcggctacgg orfQ ctggtggcaggtgtttgg oijU cgaaatgctcggctacacc oijX gcccacacctattccaacc orfY ccaacgcagacaaccttcc pltI::inaZ gtgtttcctggtttctcc pitJ::inaZ Reverse primer Product (5'-3') size (l'p 951 ttcaggcggtaatagagcg 668 agcactgttttgacccgc 517 acgcagaaccagcagaag 943 aggtcttccgtttccacc 1039 atcgtcgtgccgctttgc 866 aagaggaaccccgagagc 857 gcagcggatcgagatagg 1083 cgctggtattcgttgtgg aaaacagcggtccagaaacc 401 1000 tccctgctcgacgtactgc ggcgtectcatacatcactagc 795 ccaacgccttgtcgagattcat 691 2139 Annealing Tm(°C) 55 51 51 51 51 51 51 51 51 51 51 48 48 4.2.3. RT-PCR For RT-PCR experiments, RNA was extracted from Pf-5 and derivative strains during mid-log phase. RNA was extracted from cultures in two different experiments 107 using the RNAqueous kit from Ambion (Austin, TX). RNA templates were LiCl2 precipitated and then treated twice with DNaseI according to the manufacturer's recommendations for the DNA-free kit (Ambion, Austin, TX). Samples from each experiment were subjected to reverse transcription (RT) to derive cDNA, using SuperScript 11 RNaseH reverse transcriptase (Invitrogen, Carlsbad, CA) according to the manufacturer's recommendations. Amplification from 100 .tg of eDNA template was performed on a Robocycler Gradient 96 Temperature Cycler (Stratagene, La Jolla, CA), using conditions described above. PCRs were also performed on genomic DNA templates for positive controls, and on RNA templates not treated with reverse transcriptase, as controls to indicate contamination of RNA and subsequent cDNA samples with genomic DNA. RNA and RT-PCR products were stored at 80°C. 4.2.4. Nucleotide sequencing of pitH, pitl, and pit.! Two regions flanking the PLT biosynthetic genes in Pf-5 were represented by pJEL6 130, pJLE6246, and pJEL6247, and by pJEL6O27, pJEL61 19, and pJEL6I2O, respectively (Fig. 4.1). Sequencing of the regions was initiated using primers complementary to the pUC19 multiple cloning site, and was then continued by primer walking: information from each consecutive sequencing run in a particular direction was used to design a the sequencing primer to begin the next sequencing run, until there were runs spanning the entire insert in the desired direction. Nucleotide sequencing reactions were carried out at the Central Services Laboratory (Oregon State University, Corvallis, OR), using ABI Prism BigDye Terminator Cycle Sequencing chemistry with Amplilaq DNA polymerase, and separated on ABI model 377 and 373A sequencers (Perkin-Elmer, Applied Biosciences Division). Sequence information for the region was considered reliable when a minimum of 3X coverage, with sequencing runs in both directions, was available for the entire region. Assembly, contig joining, and contig editing were performed using the Staden package (71). Finished consensus sequences were analyzed using the algorithms of the University of Wisconsin Genetics Computer Group (version 10). Predicted amino ID1 acid sequences of each gene were searched for patterns, profiles, and motifs listed in the Pfam, ProSite, and InterPro databases (8, 68, 84) for predictions about functionality. Transmembrane regions were predicted by the Dense Alignment Surface method (17) and the "positive inside" rule (14). 4.2.5. Derivation of chromosomal pltI::inaZ and ylt.J::inaZ transcriptional fusions by marker exchanie mutaenesis Strains bearing chromosomal transcriptional fusions of the promoterless ice nucleation reporter gene, inaZ (38) within pltI and p1LJwere constructed as follows (also see Fig. 4.2 for an example of the construction of pJEL6300, used to create Pf-5 pltI::inaZ via marker-exchange mutagenesis). The plasmid pJEL5 569 is pUC 18 containing a 12 kb HindIll fragment from pJEL 1938 and spanning the right end (as depicted in Fig. 4.1) of the PLT biosynthetic gene cluster. The 3.4 kb SmaI-HindIII fragment from pJEL5569 was excised and cloned into pUCI9 linearized with SmaI and Hindlil. The resultant plasmid, pJEL6274, was checked for directionality of inserts by restriction digest mapping. A 4.6 kb BainHI-PstI fragment from pTn3-nice that contains inaZ was excised and cloned directionally into pUC 19 linearized with BamHI and PstI, to yield pJEL6277. The 1.2 kb PstI fragment containing the nptl gene originally derived from Tn903 was also excised from pTn3 -nice, and cloned into pJEL6277 that had been linearized with Pstl. The resultant plasmid, pJEL6279, was checked for directionality of the nptl insert by restriction mapping. Before proceeding, pJEL6279 was also tested for ice nucleation activity (INA) conferred by inaZ fused to the lacZ promoter. A 6.2 kb BamHI - AfthII fragment from pJEL6279 encompassing the inaZ and nptl genes was excised, and cloned via blunt-ended ligation into the BsaAl site of pill in pJEL6274 and into the SphI site of pltJ in pJEL6274. Resultant plasmids (pJEL6287 and pJEL6285, respectively) were checked for directionality of the insert by restriction digest mapping and by PCR using a primer complementary to the first 22 nt of the inaZ ORF in combination with a primer comprising nt 53-70 of the phI gene (Table 4.2). The pltI::inaZ and pltJ::inaZ transcriptional fusions were introduced into the chromosome of P1-5 by marker 109 exchange mutagenesis. For this purpose, constructs were moved into pRK41 5, which is unstable in Pf-5. Briefly, plasmids pJEL6285 and pJEL6287 were linearized by digesting a unique Ndel site. The resultant fragments were blunted with T4 DNA polymerase, and ligated to EcoRI-linearized, T4 DNA polymerase-blunted, and dephosphorylated pRK415. The resultant plasmids (pJEL6299 and pJEL6300) were transformed into chemically competent E. co/i Si 7-1 cells. Putative transformants were verified by restriction digest mapping of the re-extracted plasmids. pJEL6299 and pJEL6300 were mobilized from E. co/i S 17-1 into Pf-5, Pf-5 pltR::aacCl and Pf- 5 pltF::aacCl. Putative transconjugants were screened by plating on KMB containing streptomycin (to which Pf-5, but not the E. coli donor, is naturally resistant) and kanamycin and tetracycline (to select for plasmid transfer to Pf-5 derivatives). Verification that putative transconjugants were Pf-5 derivatives and notE. co/i strains was achieved by patching putative transconjugants onto KMB containing spectinomycin, to which Pf-5 is naturally resistant. Single colonies were then inoculated into KMB broth with streptomycin and kanamycin, but no tetracycline, to maintain selection for the nptl insertion but not for the pRK415 derivative plasmid. After three days, serial dilutions were plated onto KMB containing streptomycin and kanamycin; single colonies from these platings were patched onto KMB containing streptomycin and kanamycin, and onto replica plates containing streptomycin, kanamycin, and tetracycline. Genomic DNA was extracted from strains that were resistant to streptomycin and kanamycin but sensitive to tetracycline, and insertions were further verified via PCR using the p111-F and inaZ-R primers described above (Table 4.2), and by Southern analysis using PCR-generated (Table 4.2) probes complimentary to p/tI and pltJ, labeled as described above. 110 Figure 4.2. Cloning strategy used to construct pJEL6300, which was used in markerexchange mutagenesis to derive Pf-5 pltI::inaZ. Numbers in parentheses indicate distance of restriction sites from 5' end of insert (in pJEL5569) or from bp 1 ofpUCl9 or a pUC19 derivative, when known. Block arrows or thickened lines represent open reading frames and denote directionality, when known. Checkered spaces indicate noncoding regions. Elements within each plasmid are drawn to size with respect to that individual plasmid, but plasmids are not drawn to size with respect to other plasmids. $aml-H (4099) saAI (4224) SmaI (3034) BamHl (578) PstJ (2369) EcoRl (32) $amHI (4) Smal (4cI) SphI (74) \\\ i coRI (4311) (4845) \ rul (4869) $amHI (4981) PstI (224 ) SmaI) ,PstI (5310) / SphI (5620) HiudIII (6470) I I [[ ii 4 S pUCl 9 EcoRJ (3) ma I (4i) amHI (418) \\PstI (440) \$phl (446) HindIIL (448) region flanking PLT gene cluster (3' end of inse in pJEL5569) i/HindIii digest of pUCI9 ligation SmaiIHindUi digest of pJEL5569 TSmaI(415) I pJEL6274 JBamHI (1480) BsaAJ (1605) 6089 bp EcoRI (1692) oriksp (2226) Hu1III (3851) NruI (2250) BamHI (2362) PstI (2691) Sh 1(3001) Fig. 2 Continued on next page NdeI BsaAl digest of pJEL6274 On Two fragments from pJEL55O8 cloned into pUCI9: I 4.6 kb and BamHlIPstI fragment containing inaZ Pstl fragment containing npt F'Iac ligation 1.2 kb pJEL6287 12.2kb ,BamHI Digest pJEL6279 with BamHl, AfIlIl to yield 6.2 kb cassette containing inaZ and nptl. Blunt ends. tIüxlHI Ori/'''_ ti PstI 8.4 kb IIlkLac HindIH ligation digest to linearize. Blunt ends. Tr N 1415 10.5kb / I \ IlindIH PstI BamHI EcoRI Fig. 2, continued 113 4.2.6. Derivation of chromosomal pltF:: aacCl mutant by marker exchan2e mutaenesis The constructs described above for assessing transcriptional activity of p/tI and plt.J include an nptl gene that confers kanamycin resistance. To assess the activity of these transcriptional reporter genes in a Pf-5 derivative bearing a mutation in aplt structural gene, it was necessary to disrupt the desired pit structural gene with a cassette other than one conferring kanamycin resistance. The pltF gene, which is predicted to encode an acyl-CoA synthetase necessary for precursor activation during induction of PLT biosynthesis (49), was mutagenized by insertion of a gentamycin resistance-conferring gene, aacCl. A 1.7 kb PstI fragment containing the pltF gene and flanking sequences was cloned into pUC19 to yield pJEL6 198, and the aacCl gene was transferred on a 2.0 kb SmaI fragment from pMGm (47) into a blunted BsmI site located between nt 937 - 942 of the pltF gene, to yield pJEL6 199. The insert disrupts the predicted sequence of PIItF five amino acid residues beyond the conserved core D sequence shared by adenylate-forming enzymes (49). The integrity of the constructs was verified by restriction digest mapping and by nucleotide sequence analysis. A 3.7 kb HindIII KpnI fragment containing the resultant pltF::aacCl was cloned into pRK4I5 to make pJEL6200. pJEL6200 was mobilized from E. co/i DH5a into Pf-5 via triparental mating, using pRK2Ol3 as a helper plasmid, for marker- exchange mutagenesis of the chromosomal pltF in Pf-5. Transconjugants were grown on KMB containing gentamycin (to select for pJIEL6200 or for the subsequent incorporation ofpltF:: aacCl into the chromosome) and streptomycin (to select against the E. co/i strains) and replica plated on KMB containing tetracycline (200 .tg/ml) to test for loss of the pJEL6200. Chromosomal mutations of gentamycin resistant, tetracycline sensitive strains were confirmed by Southern analysis as described below, and evaluated for PLT production. 114 4.2.7. Antibiotic guantification Cultures were centrifuged (5,000 X g, 5 mm) to pellet cells, and supematants were extracted for recovery of PLT and 2,4-DAPG while cell pellets were extracted separately for pyrrolnitrin (PRN). Culture supematants were acidified to pH 2.0 with 1M HCI, extracted twice in 0.4 volume of ethyl acetate, and the extracts were pooled. Cell pellets were suspended in 5ml acetone, vortexed for 10 sec, placed in a sonicating bath for 30 sec, and subsequently centrifuged (10,000 X g, 10 mm) to separate the cell debris from the organic extract. Ethyl acetate and acetone were both removed from extracts under reduced pressure, and the resultant residues were resuspended in 100 tL methanol, from which 10 .tL samples were evaluated by high performance liquid chromatography (HPLC). Samples were separated over a Nova- Pak C18 reversed-phase colunm (Waters Corporation, Milford, MA) eluted (1 mi/mm) with 90% water + 0.1% acetic acid:10% acetonitrile + 0.1% acetic acid (v/v) for two minutes, followed by a linear gradient that changed from 90% water + 0.1% acetic acid:10% acetonitrile + 0.1% acetic acid to 100% acetonitrile + 0.1% acetic acid over 18 mm. Antibiotics were detected with a photodiode array detector (PLT, X = 310 nm and retention time (ti) = 13.1 mm; 2,4-DAPG, ? = 270 mn and t1 =15.6 mm; PRN, 2 254 nm and tr = 17.9 mm). Quantification was done by comparison to a standard curve generated from authentic compounds. The detection limit was 0.02 p.g/ml for PLT, 0.01 tg/ml for 2,4-DAPG, and 0.01 tg/ml for PRN. Recovery of compounds using the extraction procedures described above averaged 70% for PLT, 45% for 2,4DAPG, and 77% for PRN. 4.2.8. Transcriptional activity from PLT biosynthetic genes Transcriptional fusions of a promoterless inaZ gene to several PLT biosynthetic genes of Pf-5 were constructed previously (34). 1NA activity from strains carrying the inaZ reporter gene was assessed as described in Chapter 2 and in reference 11. 115 4.2.9. Statistical analysis Treatment means were compared using a Student's t-test following an analysis of variance. Two-tailed P values are reported. Statistical analyses were performed using JMP, version 3 (SAS Institute, Cary, NC). 4.3. Results 4.3.1. Nucleotide seguence analysis Nucleotide sequence was determined for the region of the Pf-5 genome spanned by pJEL6246, pJEL6247, and pJEL6 130, and by pJEL6 120 and portions of pJEL6I 19 and pJEL6O27 (Fig. 4.1), yielding information extending 3776 bp beyond the 3' end of pltG and 1618 bp beyond the 3' end of pltM. Codon preference analysis (81) was used to identify three open reading frames (ORFs) designated pitH, pitI and, pit..!. Fig. 4.3 describes salient characteristics of features of all three ORFs. 116 pltF pitH pitJ orfP ) pitG pltI orfK ortu art Q artY H-4 orfl ORF size (at) Predicted protein size (aa) Putative protein function pitH 636 211 TetR-tWe transcriptional regulator pill 1014 337 HIyD homolog (membrane fission protein) p11.1 1770 589 Cytoplasmic ATP-binding cassette of ABC transporter orfk 1149 382 Transmembrane subunit of ABC transporter oiJP 1119 372 Transmembrane subunit of ABC transporter oifQ 1497 498 Outer membrane protein or channel orf U 606 201 LysE-type outer membrane channel oiJX 1146 381 Zinc-binding oxidoreductase orfY 879 292 Chloroperoxidase Figure 4.3. Map and features of open reading frames flanking one end of the PLT biosynthetic gene region. pltF and pUG are previously described PLT biosynthetic genes (49). 117 4.3.1.1. orfN I partially sequenced (1491 bp) an open reading frame downstream of pltM that I have designated orJN. The partial sequence of orJN gene reveals a gene whose translated product is highly similar (88% - 93% amino acid identity) to proteins from Pseudomonas spp. and the closely related genera Azotobacter and Raistonia, that belong to both the OsmC/Ohr family (accession number PF02566) and to COG! 944, a family of conserved hypothetical proteins with unknown function. OsmC is an osmotically inducible protein; Ohr is induced by organic hydroperoxides and detoxifies these compounds (3, 46). Because no homologs with predicted protein functions were found for orjN at the time of original analysis, the involvement of orJN in PLT production has not been experimentally tested. 4.3.1.2. pitH. Sequence analysis downstream of pltG revealed a convergently transcribed ORF designated pitH. PSI-BLAST searching (2) revealed sequence similarity (35% identity and 60% similarity at the amino acid level; Fig. 4.4a) to PhIH (GenBank accession number AF207529), a putative TetR transcriptional regulator from the 2,4- diacetlyiphioroglucinol gene cluster of P. fluorescens CHAO. The predicted amino acid sequence contains an N-terminal DNA binding motif (Fig. 4.4b) that typifies the TetR family of transcriptional regulators (Pfam accession number PFOO4O). A potential transmembrane region between amino acid residues 174 and 185 (Fig. 4.4c) suggests that PItH may be anchored within the cytoplasmic membrane. 118 a. PitH: 3 62 T Ph1H: 10 PitH: 63 Ph1H: 70 RT R DG T+ -*-ILE A +LFA+ GYP,NT SK ICE + A+ A+NYHF +++ LY TAARTPRQDGQATRQQILETAGQLFAELGyANTTSKQI CENSRNASVNYHFENKDGLY 69 ICAVLIEGHKQLVSFEALSQLAQSEEPALIKLIJSFIDAIVTRVLDE-QSWQSKVCAREILA 121 +AVL H H LV H + Kb + I ++ + ++ + W XV RE+L+ RAVLREAHDRLVRIETMISLSESQGSPEDKLRAI ITVLIEGLGNQREGWALKVLTRELLS 129 L+S PitH: 122 PTVHFTSLVQEEVMPKFRLLHALISEITGFPIGDPALARCTISIIAPCLMLAVIDRQQPS 181 P+ ++++E+ 91< +++ ++ +1 DP R +S+ APCL L + + Ph1H: 130 PSAVLPAILEEQAFPKAKIVRTILGQIMDLSPEDPVTLRSAVSVFAPCLFLLIAHQPLAR 189 PitH: 182 PLQAVLQHDANALKAHFKLFARSGLAAIAQ 211 + L+ D AL H +A +GL+A+++ Ph1H: 190 NVLPGLELDPPALIEHMMSYALAGLSAI4SR 219 b. Consensus ILdAAiebfaekGydat svrelAkrAGvskgaiYrhFksKeebl lal IL+A++Lfa+Gy+t I ++AG +a+ +hFs e+L+ a+ 31 ILEVAARLFAQHGYANTASKLICEEAGADLAAINYHFGSREALYKIW 77 PitH C. "DAS n4-segnet uredictio I-, 0 S 0) 0 0 ID 0 50 100 150 200 250 Query sequence loose cutoff -- - - strict cutoff Figure 4.4. a. Alignment of predicted amino acid sequence of PitH with that of the PhJH sequence from P.fluorescens CHAO. The conserved helix-turn-helix motif is shown in bold characters. b. Alignment of PItH with the consensus DNA-binding domain for TetR family proteins (8). Capital letters designate residues aligned to match states in the profile hidden Markov model. c. Prediction of a PitH transmembrane domain by the Dense Alignment Surface (DAS) method. 119 4.3.1.3. LL The ORF adjacent to and divergently transcribed from pitH is designated p/tI, and shares similarity at the amino acid level (E= 0.0001) across its entire length with the HIyD family of secretion proteins (Pfam accession number PF00529); Fig. 4.5a). A single transmembrane domain is predicted for PitI, between amino acid residues 6 and 21 (Fig. 4.5b). The predicted PitI amino acid sequence is similar to numerous hypothetical and confirmed membrane fusion proteins, which are components of multidrug efflux pumps of the major facilitator superfamily. One of these is FusE, a fusaric acid resistance protein from E. co/i (GenBank accession no. BAA1 5404.1; Evalue = 0.002). 120 a. consensus pitI rskevqpqvsGiVieilvkeGdrvkaGdvLvrlflptpaeaala. r+++ s ++ + ++eGd+V GvL+1D ++++ + ++++ 42 consensus RQVSL.FDGSERIAALYEEGDLVQPGQVLAELDTRTLRLEINrska 88 raeaelalaqadaaqarlaaekldrlqalvlpaaiq. asaqeaaaarral r a+ +++ +++ r +++ ++ +a++ +a++q 4--i-ag ++r 89 RIGAQEQALLRLKNGTRPQEV- -EQSKARFDAAQAQtnQLAQLHMQRLR- p1 tI consensus 134 realaasqrealaaskaqleaardnleaakatlpasiak ......... ig r-i-a + +s+i-+1+-i-a -i--i-i-i- aka+++ ++ + + +++ 135 RIADDTQG- - - KGVSQQRLDQAAARLQVAKAQSQEQRESwt lakigprNE 181 p1 tI consensus risasqatlgaqltntllrklqaqglaakaqldeaylqlaetaaellrli +i++++a Li-a ++ id + la+++ + 182 DIAQATADLQASRAD ---------------- LDLLEHYLARAQ ------ L 209 pitI consensus pitI rapaqgervsleakersvqvgqevgsevgypLiugvvpfndvevdanl eea +ap g+ ++++r+-4-+g+ ++ + p-I-+++ +v a++ e 210 KPTQA ----- RIRTRLLEPGDMAS--PQRPVFALALTDPKWITRAYVNER 252 consensus qlrlvrsgiraplsggvyvlvvhteGgvvgagei. . lmpivpvdrllpve + a++ + +++ G+v i++ +p++ + ++ 253 QLGHIRADQMMWYTDSFPD-QAIDGKVGY- --Is5VAEFTPKSVETED- 297 ql +-i-r pitI consensus pitI 4 aqvavvnidkvvqglpVeikldpgafnqtttplvpGevviv +++ + ++ ++ 1+-i- + V +++++++ 298 --LRTSL ------ VYEIRVLVK---DPDDRL.RLGMPATVYL 327 b. TM-sequent ztrediction DAS 0 Pa Pt 0Pa I- I-. S 0 Pt S 0 50 100 150 200 250 300 350 Query sequence loose cutoff strict cutoff -- - Figure 4.5. a. Alignment of predicted amino acid sequence of PitT with HIyD consensus domain (8). Capital letters designate residues aligned to match states in the profile hidden Markov model. An arrow designates site of insertion of the inaZ cassette in Pf-5 pltI::inaZ derivatives. b. Prediction of a PitI transmembrane domain by the DAS method. 121 4.3.1.4. itL Adjacent to pill and transcribed from the same strand is pltL The pltJORF overlaps the last four nucleotides of p/tI, and also lacks a consensus Shine-Dalgarno sequence preceding the ATG start codon. Two separate consensus ABC transporter ATP-binding motifs (Pfam accession number PF00005) reside inpltf (Fig. 4.6). Expect (E) values for the consensus with the model were 1.44 e55 and 3.8 e55 for the N-terminal and C-terminal ABC transporter motifs, respectively. The nucleotide binding site comprises a Walker A box, a Walker B box, and a switch region (28, 48, 63, 79). Both ABC transporter motifs in pltJ include the y-phosphate-binding Walker A box (accession number PD0000017) and magnesium-binding Walker B box, as well as a switch region thought to be important for ATP hydrolysis. Also present in each ABC transporter motif ofpltj are ABC transporter signature motifs (previously termed linker peptides) next to the Walker B boxes. For ABC transporters involved in import of substrates, the signature motif is proposed to be involved in interactions with periplasmic substrate binding proteins (9). A single switch region (9) follows the Walker B box in the C-terminal ABC transporter motif. There were no apparent transmembrane domains, indicating thatpltjis likely to be cytoplasmic. 122 N-terminal ATP-binding domain consensus pltJ Gev1aIv3pNGaGKSrLLk1isG11ppteGt illdGardlsdlsklk G++ alvfp GaGK+IrLL+iG11+ +eG +++ G d++ + GKLSALVPDGAGK*LLR4IAGLLKADEGLLKVLG-LDVAADP- 79 37 Walker A consensus pltJ erleslrknigvvfQdptlfpnpeltvreniafglrlslglskdeqddrl ++ + i++ +Q +1+ lt++en+ ++ 1 + ++++ 80 --QQVQDL-ISYMPQKFGLYE--DLTIQENLELYADL ------ HGVRAQQ 118 consensus p1 tJ kkagaee1Ler1g1gydd11drrptLsGGqkQRvafiARaL1tkpk.L1L 4+4+ IL i-+ 1+ +++dr ++LsGG+kQ44+A4 L++ p LlL 119 REERFSRLLQMMDLT- RFRDRLAdOLSGGMKpKL(ACTLVRSPQtLLL_166 Signature motif consensus p1 tJ DEPTagLDasraq11e11re1rqq . ggTvllitHdldlldrlaDrilvl DEPT g D sr++1+ ++ +1 q++ Tvl+t ++d+ ++ + ++vl 167 DEPTVGVD'LSRRELWSIIEQLIEQeNLTVLISTAYMDE-AQRCAQVFVL 215 Walker B edG pltJ 216 YQG 218 C-terminal ATP-bindiug domain consensus pltJ Gevlal pNGaGKS LLklisGllppteGtilldGardlsdlsklk Ge+ +1+ pNGaGK+ ++++Gllp G + + G 1++ GEIFGL PNG GK FRMLCGLLPASSGYLEVAG-VNLRTA---- 408 Walker A 367 consensus p1 tJ erleslrknigvvfQdptlfpnpeltvreniafglrlslglskdeqddrl ++ r++ig+v+Q 1+ +1+++en++f++ ++++r gi 409 -RAAARRKIGYVSQKFALYA- -NLSALENLRFFGGA-YGLHGAKLKAR- 452 consensus p1 tJ 453 kkagaeelLerlglgyddlldr + +Le+ 1+ + + r - -VALMLEQFDLD- -EHKHLRS LSGGq QRV ARaL1t3cp4LI L GG kQR+ A aLl++p L+L LPGGYK R VALLHEPQJ Signature motif 496 consensus p1 tJ DEPTagLDasraq11e11re1rqqgTv1l itRdldlldrilaDrilvle DEPT+g+E4 +r++++ + +1 q +T+++tH +4+ +.4Dri+ + 497 DEPTSGIDLARRAFWRRITALAQSTTIVITTHFMEE-AE'4CDRIVIQD 545 Walker B Switch region consensus pltJ dG 546 AG 547 Figure 4.6. Alignment of predicted amino acid sequence of P1tJ with the consensus ABC-binding cassette of ABC transporter proteins (8). Capital letters designate residues aligned to match states in the profile hidden Markov model. Regions important for nucleotide binding are boxed and identified in bold typeface. The conserved histidine residue in the switch region is boWed. An arrow designates site of insertion of the inaZ cassette in Pf-5 pltJ::inaZ derivatives. 123 4.3.1.5. Analysis of adjoining sequence Analysis of preliminary data from the unfinished genome of Pf-5 (http://www.tigr.org) revealed that four more ORFs reside downstream ofpltf and are transcribed from the same strand (Fig. 4.3). These include, in order of increasing distance from pltJ, two consecutive ABC transporter transmembrane proteins (orfK and orfP), an outer membrane efflux protein (orfQ), and a LysE-type efflux protein, orf U, with homology to cmaU (GenBank accession number U 14657) from the coronatine biosynthesis pathway in P. syringae pv. glycinea. The closest homologs of pitH, p/tI, pltj, orJK, and orJP (but not pltG or orfQ) in the extant databases reside in a cluster of genes with identical synteny, from Azotobacter vinelandii (GenBank accession numbers ZP_00090839.1 through ZP 00090843.1). Expect (H) values for similarity gene pairs in the Pf-5 cluster and the A. vinelandii cluster ranged from 3e-41 (pitH! ZP 00090839); 45% identical and 62% similar at the amino acid level, to 0.0 (pltJ/ ZP00090841.l); 65% identical and 76% similar at the amino acid level. 4.3.2. Phenotypic characterization of pill, pltJ mutants Pf-5 derivatives bearing chromosomalpltl::inaZ orpitf::inaZ disruptions accumulated approximately 4.7 times less PLT in their culture supernatants than did wild-type Pf-5 at 48 hours (P = 0.011) (Fig. 4.7a). PLT concentrations in cell extracts ofpltl and pkJ derivatives were, similarly, 5.2 times lower (P = 0.0085) than PLT concentrations in cell extracts of Pf-5 (Fig. 4.7b). Population sizes (not shown) did not differ significantly among strains at any time point (P 0.43). 124 2.0 PLT detected in supernatant extracts 13 1.0 0.5 0.0 0.16 lb. PLT detected in cell extracts 0.14 0.12 0.10 0.08 I- 0.06 0.04 0.02 0.00 :1 F 1' Figure 4.7. PLT production by Pf-5 and bypltl::inaZ,plt.J::inaZ, and pltB:.inaZ derivatives of Pf-5. Cultures were grown for 48 h in NBGIy before harvesting and extracting supernatants (a) and cells (b) to determine PLT concentrations. Both panels represent data from the same experiment. Error bars denote one standard error. 125 I examined whether the effects on PLT production could be attributable to polar effects on transcription caused by the inaZ insertion mutations in p/ti and pltJ. RT-PCR was utilized to test transcription of pltJ in apltl mutant and in wild-type Pf-5. Amplification pkl cDNA occurred reproducibly in the wild type but not the pitI mutant strain (Fig. 4.8). This result raises the possibility that the inaZ reporter gene used to disrupt pill and pltJ has polar effects on transcription of downstream genes. Although p/tI: :inaZ and plt.J::inaZ derivative strains have utility as transcriptional reporter strains, the phenotypes of these strains cannot be attributed to the p/ti and plt.J mutations without further analysis. 126 Pf-5 Pf-5 pltI::inaZ I- z z z 12,000 5,000 2,000 1,600 1,000 4 plt.J 850 650 500 Figure 4.8. RT-PCR products obtained from pltJ forward and reverse primers (Table 4.2) and the templates indicated above each lane. Templates were: lane 2, Pf-5 genomic DNA; lane 3, Pf-5 cDNA; lane 4, Pf-5 RNA; lane 5, none; lane 6, Pf-5 pltI::inaZ genomic DNA; lane 7, Pf-5 pill: :inaZ cDNA; lane 8, Pf-5 p/tI: :inaZ RNA. Lane 1 contains a 1 kb MW ladder, with band sizes indicated on the left margin. The expected migration destination for the pltJ PCR product is depicted by an arrow. 127 4.3.2.1. PLT auloinduction is inhibited in a pill derivative of P1-5 Recently, PLT has been shown to positively regulate its own production (11). Although pitI and pltJ mutants produce PLT, they do so at a reduced level. Therefore, I hypothesized that reduced PLT production inpltl and pltJmutants might be due to defective PLT autoinduction. To test this hypothesis, I compared the capacity of Pf-5 plt.J::inaZ for PLT autoinduction with that of Pf-5 (Fig. 4.9). Pf-5 and Pf-5 pltJ::ina.Z were grown in NBGIy without amendment and in NBG1y to which 4 .tg/ml PLT had been added. The non-PLT producer Pf-5 pltB::inaZ was included in each treatment to demonstrate the fraction of PLT from the original amendment detectable in culture supernatants. In Pf-5, PLT production (calculated as {PLTJ in supematant minus [PLT} detectable in supernatants of Pf-5 pltB::inaZ) in PLT-amended cultures were 2.8 times greater than those in non-amended cultures by 48 hours. However, no such enhancing effect by PLT was observed on PLT production in apltJ mutant. Pf-5 pltJ::inaZ produced a small amount of PLT (0.53 .Lg/ml at 48 h), but PLT amendment did not increase PLT production by Pf-5 pltJ::inaZ. Rather, PLT detected in culture supernatants of PLT-amended Pf-5 plt.J::inaZ was less than or equal to that detected in the P1t strain Pf-5 pltB::inaZ at all time points. There were no differences in population sizes between PLT-amended or non-amended cultures at any time points. 128 10 a. 8 16 I- 2 0 le+1O b. le+9 E U v U 0 v-- Pf-5 plt&:inaZ le+8 le+7 0 6 12 18 24 30 36 Pf-SpltB::inaZ + PLT Pf-5 pItJ.:inaZ Pf-5pltJ::inaZ + PLT 42 48 Culture age (hours) Figure 4.9. Influence of exogenous PLT on PLT production by Pf-5, and by Pf-5 pltJ::inaZ and Pf-5 pltB::inaZ. Cultures were grown in nonamended NBGIy or in NIBG1y amended with 4 tg/ml PLT. At the indicated sampling times following inoculation, PLT concentrations (a) and population densities of cultures (b) were assessed. Population sizes did not differ significantly between PLT-amended and nonamended treatments at any time point. Error bars denote one standard error. 129 4.3.2.2. Expression ofpltl and pltJ is inhibited in Pl( strains Within the PLT biosynthetic gene cluster ispltR, a gene encoding a predicted LysR-type transcriptional regulator that is necessary for PLT biosynthetic gene transcription. A shared feature of the promoter targets for LysR-type transcriptional regulators is an inverted repeat built around the core sequence T-N11-A. These dyad sequences are most often located 50-60 bp upstream of the regulated gene (23). One such dyad (Fig. 4.10) exists 32 bp upstream of pltR in the pltR-pltL intergenic region, and has potential to form a stem-loop structure (G -2.9). Thep/tH-pltlintergenic region is 111 bp in size and contains an inverted repeat sequence similar to one found in the pltR-pltL intergenic region (Fig. 4.10), which was previously proposed to be a P1tR-binding site (49). The potential PItR-binding site in the pitH-p itI intergenic region resides 54 bp upstream of pitH and 40 bp upstream of the divergently transcribed piti. The dyad has no predicted stem-loop potential surrounding the TN1 1-A core sequence, unlike that of its counterpart in the pltR-pltL intergenic region (G = -2.9), but rather forms one side of a larger stem-loop (G -7.5). The possibility of PltR binding to the pitH-pill intergenic region led me to test the effect of pltR on transcription of p/ti and pltJ. p1 tR -p1 tL: TGTCTTATTTTTGTAAAGGCAAATTACAAATAGATCATGAC p1 tH-pl tI: TCTGTTATTT TCTAATTTAAATTCAAATTGAATTTTAATT IIIIiIIlIIII ,I.11 II I I Figure 4.10. Possible PItR binding sites from the pltR-pltL and pitH-p itI intergenic regions. Dyad regions are depicted in bold, and the center of symmetry is in large font. The T-N11-A core sequence is underlined. Each complete predicted stem-loop is indicated by converging arrows. Identical nucleotides are indicated by vertical lines. 130 I evaluated chromosomal transcriptional reporter gene fusions to pill and pltJ in wild-type (Plt) Pf-5, and inpltR and pltF backgrounds (both Plt)(Fig. 4.11). After 18 h of growth, expression of 1NA from both pltl::inaZ and pltJ::inaZ was approximately one order of magnitude higher in the Pf-5 wild type background than INA from the same fusions in apltR orpltF background (P 0.025). The expression of 1NA from Pf-5 pltB::inaZ was assayed simultaneously to compare activity from the pltI::inaZ and pltJ::inaZ fusions with activity of a similar fusion to a PLT biosynthetic gene. Expression of INA from pltB::inaZ was at least three orders of magnitude greater than INA from eitherpltl::inaZ orpltJ::inaZ at all times. INA expressed from pltI::inaZ was lower than that expressed from pltJ::inaZ, exceeding background levels (P < 0.05) in only case (Pf-5 pltI.:inaZ at I 8h). At no time were population sizes significantly different among strains; thus, growth differences among strains do not account for differences in INA. 131 -L -h *-- Pf-5pltB::inaZ -3 -4 -5 -6 -7 -8 -9 -In C.? 1e4-1O le-+-1O Ie4 le+9 Ie+8 le+8 -I le+7 Ie+7 J Ie+1O 19 Ie+8 P111 d. Ie-I-6 6 12 18 le+7 I 1e-1-6 -j -1 pitt Ie-I-6 0 I 24 0 6 12 18 pit.! 24 0 6 12 18 Culture age (hours) Figure 4.11. NA activity from pltB::inaZ (a),pltI.:inaZ (b), and pltJ::inaZ (c) transcriptional fusions in wild-type (circles), pltF::aacCl (triangles) orpltR::aacCl (squares) backgrounds. The parental Pf-5 (Ina; open circles) was included as a negative control (a). Cultures were grown in NBG1y and harvested at 1, 6, and 18 hours for assessment of NA (a-c) and population densities (d-e). Population sizes did not differ significantly among strains at any time point. Error bars denote one standard error. 24 132 To evaluate the possibility that the difference between INA expression of Pf-5 pltI::inaZ and Pf-5 pltJ::inaZ derivatives was due to positional effects of the inaZ insertion, a second pltJ::inaZ fusion, bearing the inaZ insertion at the upstream Nrul site ofpltf (Fig. 4.2), was tested. INA expressed by Pf-5 bearing the inaZ insertion at the NruI site of pit..! lower than that expressed by Pf-5 bearing the inaZ insertion at the SphI site ofplt.J and essentially identical to 1NA expressed from Pf-5 pltI::inaZ. These results indicate that the difference in expression levels between Pf-5 pltI::inaZ and Pf-5 plt.J::inaZ are likely to be artefactual, resulting from posttranscriptional effects due to placement of the reporter construct within the transcript (see Section 4.4.5). Only results from pltJmutants bearing the inaZ insertion at the SphI site are reported in all experiments, for brevity. 4.3.2.3. Expression ofpltl and pit..! in Pl( strains is restored by exogenous FL T Because pltR and pltF mutants both exhibit a P11 phenotype, I hypothesized that the effects of pltR and pltF mutations on pitI and pltJ transcription might be due to a requirement for PLT. When cells were cultured in NBGIy to which 4 .tg/m1 PLT had been added at the time of inoculation, INA from both plt.J::inaZ (Fig. 4.12) and pltI::inaZ (Fig. 4.13) was enhanced in wild type, pltR, andp/tF mutant backgrounds. 1NA from pltJ::inaZ (Fig. 4.12) was enhanced by PLT amendment as early as two hours after inoculation, by more than three orders of magnitude, in all backgrounds (P 0.000l). In contrast, Pf-5 pltB::inaZ showed no difference in 1NA expression at 2 hours (P 0.95) between PLT-amended and non-amended treatments. By 48 h, INA expressed bypltf::inaZ derivative strains in PLT-amended medium was at least 3.6 log units higher (P 0.0001) than that expressed in non-amended medium, and 1NA reached similar levels among strains in PLT-amended medium regardless of genetic background (P = 0.96). In this and several other experiments, NA expressed by Pf-S pltI::inaZ and Pf-5 plt.J::inaZ peaked at 12 hours, and was stable between 12 and 48 hours. Therefore, in subsequent experiments, NA expressed by Pf-5 pltI::inaZ and derivatives was assessed at 0,2, and 48 hours to evaluate the induction of, and 133 maximum levels of, expression (Fig. 4.13). Like pltJ::inaZ derivatives,pltI::inaZ derivatives grown in PLT-amended medium showed more than 10-fold enhanced INA in all backgrounds, as early as two hours after inoculation (P O.0004). By 48 h, INA expressed from pltI::inaZ was at least 3.4 log units higher (all P 0.004) in PLT-amended cultures than in non-amended cultures, and did not differ (P = 0.78) regardless of genetic background. Slight differences in population sizes were observed between PLT-amended and non-amended cultures ofPf-5pltF::aacCl, plt.J::inaZ at 12 hours, and of Pf-5 pltR::aacCl, pltI::inaZ at each timepoint. Because 1NA values are normalized for CFU, differences in population size did not account for differences in INA. 134 Figure 4.12. Positive influence by PLT on transcriptional activity from pltB::inaZ in a wild type Pf-5 background (a); plt.J::inaZ in a wild type Pf-5 background (b); pltJ::inaZ in apltF background (c), and pltJ::inaZ in apltR background (d). The parental Pf-5 (mi; triangles) was included as a negative control (a). Cultures were grown in NBG1y with (open symbols) or without (filled symbols) an amendment of 4 pg/mI PLT added at the time of inoculation and were harvested at 0, 2, 12, 24, and 48 hours for assessment of INA (a-d) and population densities (e-h). Error bars denote one standard error. 0 0 0 0 -2 -2 -2 -2 -4 -4 -4 I: -6 -6 -6 -8 -8 -8 -8 -10 -10 le+10 -lO- le+10 le+1O le+8 7l1fl / .f Pf-5 pltR::inaZ + PLT le+8 . I I I5p1:in 0 I U 0 6 12 18 2430 3642 48 le-4-7 i I I 'plt:aa1, pftJ:inaZ le+8 0-- Pf-5 pltF::aacCl, Pf-5 plt.J.:inaZ + PLT V Pt$ + PLT le-$-7 le+10 g. e. le+9 -10 ii. le+9 1pIcCI, pltJ.:inaZ le+8 0 Pf-5 pIeR::aacCl, pltJ::inaZ + PLT le-1-7 I 0 6 12 18 24 30 3642 48 le-I-7 0 6 12 18 24 303642 48 pltJ::inaZ + PLT I I I 0 6 12 18243036 42 48 Culture age (hours) Fig. 4.12 (-'I 136 Figure 4.13. Positive influence by PLT on transcriptional activity from pltB::inaZ in a wild type Pf-5 background (a); pltI::inaZ in a wild type Pf-5 background (b); pltI::inaZ in apltF background (c), and pill: :inaZ in apltR background (d). The parental Pf-5 (Ina; triangles) was included as a negative control (a). Cultures were grown in NBG1y with (open symbols) or without (filled symbols) an amendment of 4 jtg/ml PLT added at the time of inoculation and were harvested at 0, 2, and 48 hours for assessment of 1NA (a-d) and population densities (e-h). Population sizes did not differ significantly between treatments at any time point. Error bars denote one standard error. E U U U U -2 -2 -2 -2 -4 -4 -4 -4 -6 -6 -6 -8 -8 -8 -8 _1 (I _1 ft ft -10 e le+11 le+10 le+11 e. le-f-11 1e+9 le+9 L? le+8 le+7 le+7 0 12 24 36 48 60 le+11 g. J f. 19 19 le+8 le+8 le-I-7 0 12 24 36 48 60 0 Culture age (hours) Fig. 4.13 12 I I 24 36 -i le-+-7 48 60 0 12 24 36 48 60 138 4.4. Discussion 4.4.1. PLT production requires functionality of transport genes flanking the PLT biosynthetic gene cluster This study provides evidence that genes with predicted transport functions are necessary for PLT production. The P11 phenotype of Pf-5 pltJ::inaZ and Pf-5 pltJ::inaZ could result from polar effects of the inaZ insertions on any or all of four downstream ORFs, each of which is predicted to encode a protein with a transport function. Thus, regardless of the specific gene(s) responsible for the phenotype of reduced PLT production, blockage of expression of a transport function is correlated with this phenotype. The mechanism by which PitI, PItJ, and/or flanking gene products affect PLT production remains to be explored. Low concentrations of PLT produced by pitI and pltf derivatives of Pf-5 could be a consequence of enhanced PLT degradation or reduced PLT production. This study hints that apklmutant degrades PLT more than apltB mutant. In all experiments, Pf-5 pltf: :inaZ produced a small amount of PLT (e.g. 0.53 tg/ml at 48 h; Fig. 4.9), whereas Pf-5 pltB::inaZ produced none. However, PLT recovered from supematants of PLT-amended Pf-5 pltJ::inaZ cultures was less than or equal to that recovered from cultures of the completely P1t strain Pf-5 pltB::inaZ at all time points (Fig. 4.9). Although these results suggest PLT degradation by Pf-5 pltJ: :inaZ, the rates are very small compared to the differences in PLT production between Pf-5 and Pf-5 plt.J::inaZ. Thus, PLT degradation might account for some, but not all, of the reduction in PLT accumulation bypltl orplt.J derivatives (Fig. 4.7). Therefore, PLT degradation cannot fully account for the Plt phenotypes of pitI and pit..! derivatives. Alternatively, it is possible that expression of PLT biosynthetic genes is affected at the transcriptional or posttranscriptional level in pitI and pit..! mutants. This hypothesis remains to be explored. The influence of transport genes in regulation of PLT production suggests that PLT production may be coordinated with transport of a substance to or from the cell. 139 The substrate for PitI and P1tJ is unknown, as is the directionality of transport. PLT export is a possible function of PitI and P1tJ, but PLT does not appear to accumulate in cells of pill and p11.1 strains. Rather, PLT concentrations in p/tI and p/U strains are similarly reduced relative to the wild type, in both cell and supernatant extracts. PLT may, however, be able to utilize an alternative mechanism for escape from cells even if p/tI and pitJ represent its primary export apparatus. PLT import is another potential function for PltI and PltJ. Support for the notion of PLT import by PItI and PltJ comes from the observation thatp/tlandpld derivatives of Pf-5 are unable to undergo PLT autoinduction, which is presumed to require either an extracellular sensor of PLT or the transport of extracellular PLT into cells of Pf-5 for perception by a cytoplasmic PLT sensor. Therefore, loss of PLT transport to the inside of the cell could explain lack of autoinduction in p/tI and p/ti mutant derivatives. 4.4.2. Exo2enOuS PLT induces expression of pitI and pltJ This study demonstrates that pill and pltJ are transcriptionally regulated by PLT. The rapid, positive effect of PLT on 1NA expressed by Pf-5 pltI::inaZ and Pf-5 pli.J::inaZ (Figs. 4.12 and 4.13) suggests that transcription of p/tI and p/tJ is tightly regulated by the presence of PLT. Although PLT also positively affected transcription of PLT biosynthetic genes (e.g. pltB; Chapter 2), the magnitude of the response to exogenous PLT by Pf-5 pltB::inaZ was much greater than that of Pf-5 pltI::inaZ or Pf-5 plt.J::inaZ, suggesting that the mechanism by which PLT induced transcription of pilB differed from the mechanism by which PLT induced transcription of pill and pilJ. This difference could reflect separate proteins to mediate the effect of exogenous PLT on promoters, or subtle differences in promoter structures of pltB, pill, and p11.1 causing differential activation by a single transcriptional regulator. In the absence of exogenous PLT, levels of p/tB expression were greater than those of pill orplti (Figs. 4.12 and 4.13). Thus, it is also possible 140 that basal pltB expression is already near a threshold which pitI and pltJ approach only in the presence of PLT. Regardless of the mechanism, it is clear that the effect of PLT on transcription of p/tI and p/ti is qualitatively, but not quantitatively, similar to the positive effect of PLT onpltB transcription. The accumulation of PLT both influences, and is influenced by, Piti, PItJ, and/or downstream gene products. PLT induced pltI and pltJ transcription, and mutations in pill or pltf reciprocally repressed PLT accumulation in cells and culture supematant. Precedents exist in the literature for regulation by substrate levels, and of substrate metabolism, with regard to genes encoding transport machinery for the same substrate. For example, transcription of genes involved in transport and genes for assimilation of inorganic phosphate (P1) in Escizerichia co/i both require an intact ABC transporter encoded bypstABC and the binding protein gene pstS. Mutations that functionally disable the periplasmic binding protein PstS or the ATP-binding component of the ABC transporter PstB simultaneously obviate P1 control of the Pho regulon. Reciprocally, the Pst system itself is under P1 of the Pho regulon, which is regulated by extracellular control. The pst genes are part P1 concentrations (reviewed in 80). Another example is that of MaIFGK2, an ABC maltose transporter in E. co/i and part of the ma! regulon (see references 9 and 10 for reviews). A mutation in the ATPbinding component, MaIK, results in the loss of maltose transport and also in constitutive expression of genes involved in maltose metabolism (30, 67); conversely, MaIK overexpression has a repressive effect on ma! gene expression (59). The influence of Ma1FGK2 on genes for maltose metabolism is mediated by a physical interaction with the transcriptional regulator MalT (54). Again, reciprocal regulation is evident: maIEFGK, encoding the maltose transporter, are among those genes under MalT regulation. Examples also exist of non-ABC transporters which play roles in substrate-regulated gene expression, and are themselves under substrate regulation. Such "circular" regulation also appears to be common to the phosphoenolpyrouvate:sugar phosphotransferase system (PTS) for sugar uptake in bacteria. The glucose-specific permease of the phosphotransferase system, PtsG, 141 binds to the transcriptional regulator Mic, and thereby controls Mic-regulated expression a number of glucose-regulated genes, including ptsG itself (36, 73). Similarly, depending on presence of substrate, the -glucoside transporter BgIF from E. coli directly alters the phosphorylation state of the transcriptional regulator BgIG, which in turn regulates the expression of the f-glucoside transporter gene, blgF, and other genes (13, 24). Bg1G is the prototype of a family of transcriptional regulators that share this type of direct transporter/regulator interaction (72, 75). Transport of compounds other than those involved in primary metabolism can also experience transporter substrate-dependent regulation of gene expression for transport and biosynthetic genes. The secondary metabolite pyoverdine, an iron-binding siderophore of P. aeruginosa, regulates expression of genes for pyoverdine biosynthesis (35) and transporter synthesis (22, 56, 58) in such a manner, via the ferripyoverdine receptor and associated proteins (35, 58). 4.4.3. p1111 The influence of PLT on activity from the promoter(s) of pitI and pltJ suggests the existence of a PLT receptor which directly or indirectly activates transcription from the pltlandp/tfpromoter(s). One possible candidate is the TetR transcriptional regulator encoded bypitH. TetR homologs are often involved in cofactor-dependent regulation of genes, frequently linked, encoding biosynthesis or export functions for hydrophobic compounds (e.g. 18, 29, 66, 83). Although the cofactor may be the very compound whose biosynthesis or transport is under regulation (e.g. 18, 29, 66), regulation of non-linked genes, whose product is not a cofactor for the TetR regulator, also occurs (52, 53). In the 2,4-DAPG gene cluster of P. fluorescens CHAO, two TetR family members, phiF and phlH, have been identified. In strain CHAO, PhlF is required for 2,4-DAPG autoinduction of a 2,4-DAPG biosynthetic gene (66). Ph1F binds repressively to the phiA CBD promoter in the absence of 2,4-DAPG, but dissociates from its target sequence in the presence of 2,4-DAPG (1). The amino acid sequence of PItH closely resembles that of PhIH in the 2,4-DAPG gene cluster of 142 strain CHAO. Pf-5 also produces the polyketide 2,4-DAPG, and a search of the incomplete genomic sequence of Pf-5 from The Institute for Genomic Research (http://www.tigr.org) revealed an ORF in Pf-5 with 99% similarity to phlH at the nucleotide level. Thus,pltHandphlHare predicted to be paralogous genes in Pf-5. The function of either pitH orphlH has yet to be revealed for either strain. A testable hypothesis is that PItH, like other TetR homologs, acts as a transcriptional repressor. The promoters of PLT biosynthetic genes, of the adjacent transport genes, and of pitH itself are candidate targets for regulation. In keeping with this line of thought, PLT might act as a cofactor. Cofactor binding in the case of the prototype TetR protein not only derepresses transcription of tetA, a gene encoding a tetracycline efflux protein, but also derepresses transcription of the tetR gene itself (60). The observation that transcription ofpitH itself is upregulated in the presence of exogenous PLT (Caroline Press, unpublished data), supports the hypothesis that PltH may be involved in regulatory functions involving the perception of PLT. Support for the hypothesis of a functional relationship between PltH and the products of pitl and plt.J lies in the colinearity of pitH, piti, pit.J, orJK, and orfP with a homologous cluster from A. vinelandii. The tight homology of the five genes in this cluster between two bacterial species suggests evolutionary conservation of a mechanism for transport and/or regulation. It is not known whether bacterial species other than A. vinelandii and Pf-5 regulate or transport exoproducts by this mechanism, or whether A. vinelandii produces PLT. PitH may be anchored in the cytoplasmic membrane, as suggested by the prediction of a transmembrane domain using DAS analysis. Transmembrane domains are typically 15-30 residues in length (17); the predicted PitH transmembrane domain is only 12 residues in length. However, the ends of such domains are only approximated by homology- and hydophilicity-based algorithms. The predicted domain (TISIIAPCLMLA) can be extended by including one polar (C) and one charged (R) residue that separate the predicted transmembrane domain from a stretch of four additional hydrophobic residues (PALARCTISIIAPCLMLAVI). The 143 modified transmembrane domain prediction includes 20 amino acids, resulting in a more typical length for a transmembrane domain. The validity of the transmembrane prediction can only be determined experimentally; it is possible that this hydrophobic region has an alternative function, such as substrate binding. Could PItH function as a transcriptional regulator while membrane-bound? Two well-studied transcriptional activators, TcpP and ToxR from Vibrio choierae, are anchored in the cytoplasmic membrane but have N-terminal DNA-binding domains (27,45). Additionally, it has been shown that association with the cytoplasmic membrane does not impede proteinnucleic acid interactions which take place at a discreet location in the cytoplasm: BglG, a transcriptional antiterminator, acts on nascent mRNA which is indirectly anchored to the DNA strand via RNA polymerase. BgIG functions effectively while artificially tethered to the cytoplasmic membrane (25). 4.4.4. yltI The occurrence of pill in a cluster of genes encoding putative ABC transporter subunits is enigmatic. The piti gene bears homology with the membrane fusion protein HJyD. H1yD homologs are found in ABC transporters of Gram negative bacteria specifically associated with Type I secretion of protein toxins, proteases, or lipases from the cell - such as the protein toxin hemolysin A in E. coil (reviewed in 86), or metalloprotease in Erwinia chrysanthemi (reviewed in 19). Membrane fusion proteins are also associated with the major facilitator superfamily (MFS) class of substrate transporters (20, 43), which are quite distinct from ABC transporters in form and function. MFS transporters are single-polypeptide permeases that use protonmotive force to drive transport of small molecules across the membrane, while ABC transporters are multi-subunit assemblies that use ATP hydrolysis to drive substrate transport. In Gram positive bacteria, ABC transporters whose substrates are small molecules (bacteriocins and peptides) utilize membrane fusion proteins as accessory factors (26); however, I am aware ofno examples in which Gram negative bacteria incorporate membrane fusion proteins into the ABC transporter machinery except for 144 the case of Type I secretion mentioned above. Still, a shared function for INtl and P1tJ is suggested by the transcriptional activation ofbothpltlandpltfby exogenous PLT, and the occurrence of pill in the cluster of colinear genes also found to reside in A. vinelandjj. Because the substrate for PItI and PItJ is not known, it remains possible that a protein or peptide substrate is transported by Type I apparatus including both Pill and Pill, under the tight regulation of extracellular PLT. I also speculate that PLT may be transported by an unusual transmembrane complex of which PitT and PIIJ are both subunits. The predicted amino acid sequence of PitT exhibits a predicted N-terminal transmembrane domain conserved among H1yD homologs. N-terminal transmembrane domains may sometimes be difficult to distinguish from signal sequences that are cleaved upon passage of the nascent polypeptide through the cell membrane (78). It remains possible that PitT is not an integral membrane protein, but rather a periplasmic protein that uses the general secretory pathway (Type II secretion) to insert into the membrane, and whose N-terminal signal sequence is subsequently cleaved. Because periplasmic binding proteins and HlyD homologs both associate with periplasmic portions of ABC transporters (9, 37, 74), it is not altogether unlikely that certain motifs might be shared between these proteins. This rationale is seductive because ABC transporters involved in substrate import require a periplasmic binding protein to deposit substrate into the transport apparatus. Although it is possible that pitI and p11.1 are involved in import rather than export, no such binding protein gene is apparent in the region flanking the PLT gene cluster. Another viable hypothesis is that a binding protein might be encoded by a gene within another regulon, perhaps one in which the binding protein performs additional functions. Substrate-laden binding proteins may interact with multiple proteins in the periplasmic space; for example, ChvE from Agrobacterium tumefaciens is a monosaccharide-binding protein that, when complexed with substrate, interacts with a monosaccharide transport system, with chemosensory proteins necessary for chemotaxis, and with VirA, a histidine kinase member of a two-component regulatory system governing virulence (12, 16). 145 By analogy, a hypothetical PLT-binding protein serving multiple functions might be encoded by a regulon outside of the PLT biosynthetic gene cluster. 4.4.5. pltJ The p/ti ORF lacks a typical ribosome binding site upstream from the start codon. Consensus Shine-Dalgarno sites are not always found preceding the start codon of prokaryotic genes. Indeed, Shine-Dalgamo site anchoring appears to be the final step in a series of subtle interactions between the 30S subunit and the mRNA strand (40). Thus, the ribosomal binding information inherent in the plt.J sequence may simply elude detection. It is also possible that p/ti translation is coupled to that of phi via ribosome frameshifting. Ribosomal frameshifting normally requires a slippery region for slippage at the P site and an mRNA pseudoknot to stall the progress of the ribosome on the mRNA strand (55). At least one niRNA pseudoknot was found between nucleotides 8 and 502 ofplti(AG = -329.98; mfold program, Zuker et al. [87}). The proximity of the pseudoknot to the start codon ofpltfmight imply a role in ribosomal stalling; however, no consensus (XXXYYYZ) slippery region (55) preceding the start codon was apparent although a near-consensus sequence (GGGGCCA) begins at position -11 from the start codon. The consequence of such a frameshifted translational product would be a fused polypeptide, comprising PltI and PltJ. A protein expression system could be employed to determine the molecular mass of the translated product of the p/ti gene alone, compared to that of p/tI adjoining p/t.J. The predicted P1tJ polyp eptide is homologous to the ATP-binding subunit of ABC transporters; the single open reading frame contains two fused ATP-binding domains. This arrangement is not uncommon for ABC transporters, whose subunits occur in nearly every possible combination of fused and individual subunits (9). As ABC transporters comprise paired cytoplasmic ABC binding domains and paired transmembrane domains, the predicted PltJ polypeptide represents only half of a FE!i putative ABC transporter. Transmembrane domains, but no ATP-binding motifs, are found in the adjacent genes orf K and orfP. Thus, a reasonable hypothesis is that orJK and orfP together, or a dimer of either gene product, form(s) the transmembrane portion of the ABC transporter of which PItJ is part. Because no other ORFs with homology to ABC transporter genes appear adjacent to orJP, it is likely that p/tJ orJK, and orjP encode subunits of the same ABC transporter. In conclusion, I have provided preliminary evidence for the existence ofa new mechanism for regulation of PLT production, which involves substrate transfer. The substrate for the transporter(s) encoded by genes flanking the PLT region remains unknown, but the transport machinery itself is necessary for optimal PLT production. In turn, the expression of transport genes pitI and p11,1 is regulated by extracellular PLT. Whether the substrate for PitI and P1tJ is PLT remains to be explored experimentally. Iii at least one example (prodigiosin in Serratia sp. ATCC 39006), there is genetic evidence that nutrient (P1) transport may regulate antibiotic production (70). However, to my knowledge this is the first report of coordinate regulation of antibiotic biosynthetic genes and a transport apparatus by the antibiotic itself. Biochemical characterization of the mechanism for PLT regulation of p/tI and pltf, and of the reciprocal influence of PitI, PltJ, and/or adjacent transport gene products on PLT production, is likely to reveal a novel mode of regulation of antibiotic production in Gram negative bacteria. 147 4.5. References 1. Abbas, A., J. P. Morrissey, P. Carnicero Marquez, M. M. Sheehan, I. R. Delany, and F. O'Gara. 2002. Characterization of interactions between the transcriptional repressor Ph1F and its binding site at the phiA promoter in Pseudomonasfluorescens Fl 13. J. Bacteriol. 184:3008-3016. 2. Altschul, S. F., T. L. Madden, A. A. Schäffer, J. Zhang, Z. Zhang, W. Miller, and D. J. Lipman. 1997. Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res. 25:33893402. 3. Atichartpongkul S., S. Loprasert, P. Vattanaviboon, W. Whangsuk, J. D. Helmann, and S. Mongkolsuk. 2001. Bacterial Ohr and OsmC paralogues define two protein families with distinct functions and patterns of expression Microbiol. 147:1775-82. 4. AusubeL, F., R. Brent, R. E. Kingston, B. E. Moore, J. C. Seidman, J. A. Smith, and K. Struhl (eds.). 1998. Current Protocols in Molecular Biology. Wiley and Sons, Inc. New York, NY. 5. Bailey, B. M., R. E. Johnson, and U. J. Salvador. 1973. Pyrrole antibacterial agents. 1. Compounds related to pyoluteorin. J. Med. Chem. 16: 1298-1300 6. Bangera, G. M., and L. S. Thomashow. 1999. Identification and characterization of a gene cluster for synthesis of the polyketide antibiotic 2,4diacetylphioroglucinol from Pseudomonasfluorescens Q2-87. J. Bacteriol. 181:3155-3163. 7. Bangera, M. G., and L. S. Thomashow. 1996. Characterization of a genomic locus required for synthesis of the antibiotic 2,4diacetylphloroglucinol by the biological control agent Pseudomonas fluorescens Q2-87. Mol. Plant-Microbe Interact. 9:83-90. 8. Bateman, A., E. Birney, L. Cerruti, R. Durbin, L. Etwiller, S. It Eddy, S. Griffiths-Jones, K. L. Howe, M. Marshall, and E. L. L. Sonnhammer. 2002. The Pfam protein families database. Nucleic Acids Res. 30:276-280. 9. Boos, W. and T. Eppler. 2001. Prokaryotic binding protein-dependent ABC transporters. p. 77-114, In: G. Winkelmann (ed.), Microbial transport systems. Wiley VCH, Weinheim, Germany. 148 10. Boos, W. and Shuman, H. 1998. Maltose/maltodextrin system of Escherichia co/i: transport, metabolism, and regulation. Microbiol. Mol. Biol. Rev. 62:204-229. 11. Brodhagen, M., M. D. Henkels, and J. E. Loper. 2003. Positive autoregulation of the antibiotic pyoluteonn in the biological control organism Pseudomonas fluorescens Pf-5. Submitted. 12. Cangelosi, G. A., R. G. Ankenbauer, and E .W. Nester. 1990. Sugars induce the Agrobacterium virulence genes through a periplasmic binding protein and a transmembrane signal protein. Proc. Nati. Acad. Sci. USA 87:6708-6712. 13. Chen, Q., A. Nüssbaum-Shochat, and 0. Amster-Choder. 2001. A novel sugar-stimulated covalent switch in a sugar sensor. 3. Biol. Chem. 276:4475 144756. 14. Claros, M. G., and G. von Heijne. 1994. TopPred II: an improved software for membrane protein structure predictions. Comput. Appl. Biosci. 10:685686. 15. Corbel!, N., and J. E. Loper. 1995. A global regulator of secondary metabolite production in Pseudomonasfluorescens Pf-5. J. Bacteriol. 177:6230-6236. 16. Cornish, A., J. A. Greenwood, and C. W. Jones. 1989. Binding-proteindependent sugar transport by Agrobacterium radiobacter and A. tumefaciens grown in continuous culture. J Gen Microbiol. 135:3001-3013 17. Cserzo, M., E. Wallin, I. Simon, G. von Heijne, and A. Elofsson. 1997. Prediction of transmembrane alpha-helices in procariotic membrane proteins: the Dense Alignment Surface method. Prot. Eng. 10:673-676. 18. Delaney, I., M. M. Sheehan, A. Fenton, S. Bardin, S. Aarons and F. O'Gara. 2000. Regulation of production of the antifungal metabolite 2,4diacetylphioroglucinol in Pseudomonasfluorescens F 113: genetic analysis of ph/F as a transcriptional repressor. Microbiology 146:537-546. 19. Delepaire, P. 1998. Erwinia metalloprotease pennease: aspects of secretion pathway and secretion functions. Meth. Enzymol. 292:67-8 1. 20. Dinh, T., I. T. Paulsen, and M. H. Saier, Jr. 1994. A family of extracytoplasmic proteins that allow transport of large molecules across the outer membranes of Gram-negative bacteria. J. Bacteriol. 176:3825-3831. 149 21. Figurski, K. H. and D. R. Helinski. 1979. Replication ofan origincontaining derivative of plasmid RK2 dependent on a plasmid function provided in trans. Proc. Nati. Acad. Sci. USA 76:1648-1652. 22. Gensberg, K., K. Hughes, and A. W. Smith. 1992. Siderophore-specific induction of iron uptake in Pseudomonas aeruginosa. J. Gen. Microbiol. 138(Pt 11):2381-2387. 23. Goethals, K., M. van Montagu, and M. Holsters. 1992. Conserved motifs in a divergent nod box of Azorhizobium caulinodans 0RS57 1 reveal a common structure in promoters regulated by LysR-type proteins. Proc. Natl. Acad. Sci. USA 89:1646-1650. 24. Görke, B., and B. Rak. 1999. Catabolic control of Escherichia co/i regulatory protein Bg1G activity by antagonistically acting phosphorylations. EMBO J. 18:3370-3379. 25. Görke, B., and B. Rak. 2001. Efficient transcriptional antitermination from the Escherichia co/i cytoplasmic membrane. J. Mol. Biol. 308:131-145. 26. Harley, K. T., G. M. Djordjevic, T. T. Tseng, and M. H. Saier, Jr. 2000. Membrane fusion protein homologues in Gram-positive bacteria. Mo!. Microbiol. 36:516-517. 27. HAse, C. C., and J. J. Mekalanos. 1998. TcpP protein is a positive regulator of virulence gene expression in Vibrio cholerae. Proc. Natl. Acad. Sci. USA 95:730-734. 28. Higgins, C. F. 1992. ABC transporters: from microorganisms to man. Annu. Rev. Cell. Biol. 8:67-113. 29. Hillen, W., and C. Berens. 1994. Mechanisms underlying expression of TnlO encoded tetracycline resistance. Ann. Rev. Microbiol. 43:345-369. 30. Hofnung, M., D. Hatfield, and M. Schwartz. 1974. maiB region in Escherichja co/i K-12: characterization of new mutations. J. Bacteriol. 117:40-47. 31. Howell, C. R. and R. D. Stipanovic. 1979. Control of Rhizoctonia solani on cotton seedlings with Pseudomonasfluorescens and with an antibiotic produced by the bacterium. Phytopathology 69:480-482. 32. Keen, N. T., S. Tamaki, D. Kobayashi, and D. Trollinger. 1988. Improved broad-host-range plasmids for DNA cloning in Gram-negative bacteria. Gene 70:191-197. 150 33. King, E. 0., M. K. Ward, and D. E. Raney. 1954. Two simple media for the demonstration of pyocyanin and fluorescein. J. Lab. Clin. Med. 44:301307. 34. Kraus, J., and J. E. Loper. 1995. Characterization of a genomic region required for production of the antibiotic pyoluteorin by the biological control agent Pseudomonasfluorescens Pf-5. App!. Environ. Microbiol. 61 :849-854. 35. Lamont, I. L., P. A. Beare, U. Ochsner, A. I. Vasil, and M. L. Vasil. 2002. Siderophore-mediated signaling regulates virulence factor production in Pseudomonas aeruginosa. Proc. Nati. Acad. Sci. USA 99:7072-7077. 36. Lee, S.-J., W. Boos, J.-P. Bouché, and J. Plumbridge. 2000. Signal transduction between a membrane-bound transporter, PtsG, and a soluble transcription factor, Mic, of Escherichia coil. EMBO J. 19:5353-5361. 37. Létoffé, S., P. Delepaire, and C. Wandersman. 1996. Protein secretion in gram-negative bacteria: assembly of the three components of ABC proteinmediated exporters is ordered and promoted by substrate binding. EMBO J. 15:5804-5811. 38. Lindgren, P. B., R. Frederick, A. G. Govindarajan, N. J. Panopoulos, B. J. Staskawicz, and S. E. Lindow. 1989. An ice nucleation reporter gene system: identification of inducible pathogenicity genes in Pseudomonas syringae pv. phaseolicola. EMBO J. 8:1291-1301. 39. Martin, J. F., and P. Liras. 1989. Organization and expression genes involved in the biosynthesis of antibiotics and other secondary metabolites. Annu. Rev. Microbiol. 43:173-206. 40. McCarthy, J. E. G. and R. Brimacombe. 1994. Prokaryotic translation: the interactive pathway leading to initiation. Trends Genet. 10:402-407. 41. McGowan, S. J., M. Sebaihia, S. O'Leary, K. R. Hardie, P. Williams, G. S. A. B. Stewart, B. W. Bycroft, and G. P. C. Salmond. 1997. Analysis of the carbapenem gene cluster of Erwinia carotovora: definition of the antibiotic biosynthetic genes and evidence for a novel -1actam resistance mechanism. Mol. Microbiol. 26:545-556. 42. McGowan, S. J., M. Sebaihia, L. E. Porter, G. S. A. B. Stewart, P. Williams, B. W. Bycroft, and G. P. C. Salmond. 1996. Analysis of bacterial carbapenem antibiotic production genes reveals a novel -lactam biosynthesis pathway. Mo!. Microbiol. 22:415-426. 43. McKeegan, K. S., M. I. Borges-Walmsley, and A. R. Walmsley. 2003. The structure and function of drug pumps: an update. Trends Microbiol 11:21-29. 151 44. Mendez, C., and J. A. Salas. 2001. The role of ABC transporters in antibiotic-producing organisms: Drug secretion and resistance mechanisms. Res. Microbiol. 152:341-50. 45. Miller, V. L., R. K. Taylor, and J. J. Mekalanos. 1987. Cholera toxin transcriptional activator ToxR is a transmembrane DNA binding protein. Cell 48:271-279. 46. Mongkolsuk, S., W. Praituan, S. Loprasert, M. Fuangthong, and S. Chamnongpol. 1998. Identification and characterization of a new organic hydroperoxide resistance (ohr) gene with a novel pattern of oxidative stress regulation from Xanthomonas campestris pv.phaseoli. J. Bacteriol. 180:26362643. 47. Murillo, J., H. .1. Shen, D. Gerhold, A. Sharma, D. A. Cooksey, and N. T. Keen. 1994. Characterization of pPT23B, the plasmid involved in syringolide production byPseudomonas syringae pv. tomato PT23. Plasmid 31:275-287. 48. Nikaido, H., and J. A. Hall. 1998. Overview of bacterial ABC transporters. Methods Enzymol. 292:3-20. 49. Nowak-Thompson, B., N. Chaney, J. S. Wing, S. J. Gould, and J. E. Loper. 1999. Characterization of the pyoluteorin biosynthetic gene cluster of Pseudomonasfluorescens Pf-5. J. Bacteriol. 181:2166-2174. 50. Nowak-Thompson, B., S. J. Gould, J. Kraus, and J. E. Loper. 1994. Production of 2,4-diacetyiphioroglucinol by the biocontrol agent Pseudomonas fluorescens Pf-5. Can. J. Microbiol. 40:1064-1066. 51. Nowak-Thompson, B., S. J. Gould, and J. E. Loper. 1997. Identification and sequence analysis of the genes encoding a polyketide synthase required for pyoluteorin biosynthesis in Pseudomonasfluorescens Pf-5. Gene 204:17-24. 52. Obnishi, Y., S. Kameyama, H. Onaka, and S. Horinouchi. 1999. The Afactor regulatory cascade leading to streptomycin biosynthesis in Streptomyces griseus: identification of a target gene of the A-factor receptor. Mo!. Microbiol. 34:102-111. 53. Onaka, H., N. Ando, T. Nihira, Y. Yamada, T. Beppu, and S. Horinouchi. 1995. Cloning and characterization of the A-factor receptor gene from Streptomyces griseus. J. Bacteriol. 177:6083-6092. 54. Panagiotidis, C. H., W. Boos, and H. A. Shuman. 1998. The ATP-binding cassette subunit of the maltose transporter MalK antagonizes MalT, the activator of the Escherichia co/i mal regulon. Mo!. Microbiol. 30:535-546. 152 55. Plant, E. P., K. L. Muldoon Jacobs, J. W. Harger, A. Meskauskas, J. L. Jacobs, J. L. Baxter, A. N. Petroy, and J. D. Dinman. 2003. The 9-A solution: How mRNA pseudoknots promote efficient programmed -1 ribosomal frameshifting. RNA 9:168-174. 56. Poole, K., S. Neshat, and D. Heinrichs. 1991. Pyoverdine-mediated iron transport in Pseudomonas aeruginosa: involvement of a high-molecular-mass outer membrane protein. FEMS Microbiol. Lett. 78:1-6. 57. Quigley, N. B., Y.-Y. Mo, and D. C. Gross. 1993. SyrD is required for syringomycin production by Pseudomonas syringae patbovar syringae and is related to a family of ATP-binding secretion proteins. Mol. Microbiol. 9:787801. 58. Redly, G. A., and K. Poole. 2003. Pyoverdine-mediated regulation of FpvA synthesis in Pseudomonas aeruginosa: involvement of a probable extracytoplasmic-function sigma factor, FpvI. J. Bacteriol. 185:1261-1265. 59. Reyes, M. and H. A. Shuman. 1988. Overproduction of MaIK protein prevents expression of the Escherichia coli mal regulon. J. Bacteriol. 170:4598-4602. 60. Saenger, W.., P. Orth, C. Kisker, W. Hillen, and W. I-linrichs. 2000. The tetracycline repressor a paradigm for a biological switch. Angew. Chem. mt. Ed. 39:2042-2052. 61. Sambrook, J. and D. W. Russell. 2001. Molecular Cloning: A Laboratory Manual. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY. 62. Sarniguet, A., J. Kraus, M. D. Henkels, A. M. Muehichen, and J. E. Loper. 1995. The sigma factor a affects antibiotic production and biological control activity of Pseudomonasfiuorescens Pf-5. Proc. Natl. Acad. Sci. USA 92:12255-12259. 63. Schneider, E. and S. Hunke. 1998. AlP-binding-cassette (ABC) transport systems: functional and structural aspects of the ATP-hydrolyzing subunits/domains. FEMS Microbiol. Rev. 22:1-20. 64. Schnider, U., C. Keel, C. Blumer, J. Troxier, G. Defago, and D. Haas. 1995. Amplification of the housekeeping sigma factor in Pseudomonas fluorescens CHAO enhances antibiotic production and improves biocontrol abilities. J. Bacteriol. 177:5387-5392. 153 65. Schnider, U., C. Keel, C. Voisard, G. Defago, and D. Haas. 1995. Tn5directed cloning ofpqq genes from Pseudomonasfluorescens CHAO: mutational inactivation of the genes results in overproduction of the antibiotic pyoluteonn. App!. Environ. Microbiol. 61:3856-3864. 66. Schnider-Keel, U., A. Seematter, M. Maurhofer, C. Blumer, B. Duffy, C. Gigot-Gonnefoy, C. Reimmann, R. Notz, G. Défago, D. Haas, and C. Keel. 2000. Autoinduction of 2,4- diacetyiphloroglucinol biosynthesis in the biocontrol agent Pseudomonas fluorescens CHAO and repression by the bacterial metabolites salicylate and pyoluteorin. Journal of Bacteriology 182:1215-1225. 67. Schwartz, M. 1966. Location of the maltose A and B loci on the genetic map of Escherichia co/i. J. Bacteriol. 92:1083-1089. 68. Sigrist, C. J., L. Cerutti, N. Hub, A. Gattider, L Falquet, M. Pagni, A. Bairoch, and P. Bucher. 2002. PROSITE: a documented database using patterns and profiles as motif descriptors. Brief Bioinform. 3:265-274. 69. Simon, R., U. Priefer, and A. Puhler. 1983. A broad host range mobilization system for in vivo genetic engineering: transposon mutagenesis in Gram-negative bacteria. Bio/Technology 1: 784-791. 70. Slater, H., M. Crow, L. Everson, and G. P. C. Salmond. 2003. Phosphate availability regulates biosynthesis of two antibiotics, prodigiosin and carbapenem, in Serratia via both quorum-sensing-dependent and -independent pathways. Mo!. Microbiol. 47:303-320. 71. Staden, R., K. F. Beal, and J. K. Bonfield. 2000. The Staden package, 1998. Methods Mo!. Biol. 132:115-130. 72. Stülke, J., I. Martin-Verstraete, V. Charrier, A. Klier, J. Deutscher, and G. Rapoport. 1995. The HPr protein of the phosphotransferase system links induction and catabolite repression of the Bacillus subtiis levanase operon. J. Bacteriol. 177:6928-6936. 73. Tanaka, Y., K. Kimata, and H. Aiba. 2000. A novel regulatory role of glucose transporter of Escherichia co/i: membrane sequestration of a global repressor Mlc. EMBO J. 19:5344-5352. 74. Thanabalu, T., E. Koronakis, C. Hughes, and V. Koronakis. 1998. Substrate-induced assembly of a contiguous channel for protein export from E. co/i: reversible bridging of an inner-membrane translocase to an outer membrane exit pore. EMBO J. 17:6487-6496. 154 75. Tortosa, P., S. Aymerich, C. Lindner, M. II. Saier, Jr., J. Reizer, and D. Le Coq. 1997. Multiple phosphorylation of SacY, a Bacillus subtilis transcriptional antiterminator negatively controlled by the phosphotransferase system. J. Biol. Chem. 272:17230-17237. 76. Ullrich, M. and C. L. Bender. 1994. The biosynthetic gene cluster for coronamic acid, an ethylcyclopropyl amino acid, contains genes homologous to amino acid-activating enzymes and thioesterases. J. Bactenol. 176:75747586. 77. Ulirich, M., A. Penaloza-Vazquez, A.-M. Bailey, and C. L. Bender. 1995. A modified two-component regulatory system is involved in temperaturedependent biosynthesis of the Pseudomonas syringae phytotoxin coronatine. J. Bacteriol. 177:6160-6169. 78. von Heijne, G. 1999. Recent advances in the understanding of membrane protein assembly and structure. Quarterly Reviews of Biophysics 32:285-307. 79. Walker, J. E., M. Saraste, M. J. Runswick, and J. J. Gay. 1982. Distantly related sequences in the alpha- and beta-subunits of ATP synthase, myosin, kinases and other ATP-requiring enzymes and a common nucleotide binding fold. EMBO J. 1:945-95 1. 80. Wanner, B. L. 1995. Signal transduction and cross regulation in the Escherichia coil phosphate regulon by PhoR, CreC, and acetyl phosphate. p. 302-22 1, In: Hoch, J. A. and T. J. Silliavy (ed.). Two-component signal transduction. ASM Press, Washington, D.C. 81. West, S. E. H. and B. H. Iglewski. 1988. Codon usage in Pseudomonas aeruginosa. Nucleic Acid Res. 16:9323-9335. 82. Whistler, C. A., V.0. Stockwell, and J. E. Loper. 2000. Lon protease influences antibiotic production and UV tolerance ofPseudomonasfluorescens Pf-5. Appi. Environ. Microbiol. 66:2718-2725. 83. Yang, K., L. Han, and L. C. Vining. 1995. Regulation ofjadomycin B production in Streptomyces venezueiae 1SP5230: involvement of a repressor gene,jadR2. J. Bacteriol 177:6111-6117. 84. Zdobnov, E. M. and R. Apweiler. 2001. InterProScan -an integration platform for the signature-recognition methods in InterPro. Bioinformatics 17:847-848. 155 85. Zhang, J.-II.., N. B. Quigley, and D. C. Gross. 1997. Analysis of the syrP gene, which regulates syringomycin synthesis by Pseudomonas syringae pv. syringae. Appi. Environ. Microbiol. 63:2771-2778. 86. Zhang, F., J. A. Sheps, and V. Ling. 1998. Structure-function analysis of hemolysin B. Meth. Enzymol. 292:51-66. 87. Zuker, M., D. H, Mathews, and D. H. Turner. 1999. Algorithms and thermodynamics for RNA secondary structure prediction: a practical guide. p. 11-43. In: Barciszewski, J. and B. F. C. Clark (eds.), RNA biochemistry and biotechnology. NATO ASI Series, Kiuwer Academic Publishers. 156 Chapter 5. Concluding Remarks A goal of this study was to describe autoinduction of the polyketide pyoluteorin (PLT) in P. fluorescens Pf-5, and to explore possible cross-regulatory relationships between PLT and two other secondary metabolites [2,4diacetyiphioroglucinol (2,4-DAPG) and pyrrolnitrin (PRN)J important in biological control by Pf-5. I showed that concentrations of exogenous PLT as low as 1.5 nM initiated autoinduction. These low concentrations are comparable to those of many molecules proposed to function in signaling roles. Moreover, I demonstrated PLT autoinduction as an ecologically relevant phenomenon. In rhizosphere soil, derivatives of Pf-5 that carried a reporter gene fusion to a PLT biosynthetic gene, but which were unable to produce PLT, expressed the reporter gene at higher rates when co-inoculated as a mixed population with PLT-producing strains. This observation i) demonstrates that PLT can serve as a signal between distinct populations of cells within the rhizosphere, and ii) implies that PLT biosynthetic genes are upregulated in the rhizosphere by exogenous PLT, and therefore, by extension, PLT autoinduction is a significant component of in situ PLT regulation by Pf-5. This, and other new information presented in this thesis, prompts the prediction of several repercussions of PLT production by Pf-5 on fellow members of a rhizosphere community. 5.1. Repercussions of PLT production for the rhizosphere community PLT produced and secreted by some Pf-5 cells within a community will stimulate PLT production by others nearby, via autoinduction. This autoinduction will result in an intensification of the other community effects listed below. 2. PLT will adversely affect Oomycete plant pathogens, inhibiting their growth. 157 3. The data from Chapter 3 demonstrate that PLT can affect expression from the promoter offpvA, the ferric-pyoverdine receptor gene from P. syringae. FpvA is highly conserved among Gram negative bacteria, so it is likely that this result can be extrapolated tofivA of P.fluorescens Pf-5. In support of this assumption, PLT also affects expression of pvsA, a gene for pyoverdine biosynthesis, in Pf-5 (Caroline Press, personal communication). These results together raise the question of whether secreted PLT in the rhizosphere has potential to affect the iron status of conspecific cells, andlor that of other species that utilize the pyoverdine system, provided that PLT is taken up into, or sensed by, those cells. 4. High concentrations of PLT in the microenvironment will inhibit production of 2,4-DAPG by cells of P. fluorescens within that environment. Therefore, while Pf-5 can inhibit Pythium, it is not a good choice for biocontrol in a situation dominated by pathogens better controlled by 2,4-DAPG (e.g. Fusarium oxysporum, Gaeumannomyces gramini var. tritici, or Thielaviopsis basicola. In such a situation, a 2,4-DAPG producer that doesn't produce PLT may be a better biocontrol agent. As a matter of fact, this idea has experimental support (9). Large collections exist of P.fluorescens 2,4DAPG-producers, collected from rhizospheres in suppressive soils worldwide (e.g. 4, 8). Phylogenetic grouping of these strains by amplified spacer ribosomal DNA restriction analysis (ARDRA) placed them into several groups. ARDRA1 strains had the phenotype Plt, 2,4-DAPG, while ARDRA2 strains were Pif, 2,4-DAPG. The latter group was better in rhizosphere assays at inhibiting Fusarium oxysporum (9). One explanation might be antibiotic antagonism, hindering the production of 2,4-DAPG by ARDRA1 strains. On the other hand, ARDRA1 strains are closely related, and therefore might share another characteristic that lowers biocontrol efficacy. 158 5.2. Functionality of bioactive compounds in a soil environment The soil environment presents a challenging environment through which to relay a biochemical signal. For example, Howell and Stipanovic (3) demonstrated the inactivation or adsorption of PLT by passage through non-sterile soils, although pure PLT applied to cotton seeds in a carrier of diatomaceous earth retained the ability to protect seeds in non-sterile soil from Pythium ultimum. In truly natural field soils, PLT might be carried by soil water away from producing cells, or be degraded or utilized by heterologous bacteria, or compete with other substrates for binding to the PLT receptor hypothesized to mediate autoinduction (Chapter 2; Appendix 1). In a similar system, indirect evidence for receptor competition exists (6). In CHAO, 2,4DAPG autoinduction and repression of 2,4-DAPG by PLT, salicylate and a fungal metabolite, fusaric acid, are all mediated by PhlF, (10), which is directly activated by 2,4-DAPG and repressed by salicylic acid (1). Notz et al. (6) provided evidence for the inhibition of 2,4-DAPG gene expression in the presence of fusaric-acid producing strains of Fusarium oxysporum in soil. Although indirect mechanisms are possible, one explanation for this effect is direct competition between 2,4-DAPG and fusaric acid for a PhlF binding site. 5.3. Secondary metabolite production by Pf-5 is affected by the profile of metabolites already present In addition to PLT autoinduction, I demonstrated several other effects of Pf-5 metabolites. These are briefly listed below. 1. 2,4-DAPU positively regulates its own production (Chapter 3). 2. Exogenous PLT suppresses 2,4-DAPG production (Chapter 2). 3. Exogenous 24,-DAPG inhibits PLT production (Chapter 2). 4. Exogenous PLT suppresses expression of a heterologous pyoverdine gene in nutrient broth (Chapter 3). 159 5. Exogenous PRN exerts a slight inhibitory effect on PLT gene transcription and production (Chapter 3). I failed to see a significant effect of exogenous PLT upon PRN, despite its structural similarity (Chapter 3). Thus, PLT affects the production of at least three (PLT itself, 2,4-DAPG, and pyoverdine, see above), but not all, secondary metabolites in Pf-5 under the conditions I tested. The mutually repressive impacts of PLT and 2,4DAPG on one another's production occurs, at least in part, at the level of transcription (Chapter 2; Press and Loper, unpublished data). The slight inhibitory effect of PRN on PLT gene transcription and production (Chapter 3) cannot be explained by precursor competition, as it is unlikely that these compounds share precursors (5, 7). 5.4. Site of perception of the PLT signal may be cytoplasmic or extracytoplasmic. The data presented herein imply that PLT has, in addition to its role as an extracellular toxin, a role with intracellular signaling repercussions. PLT could be synthesized and perceived (during autoinduction) in either the cytoplasm, or the periplasmic space. Some researchers argue that the membrane-bound periplasmic space of Gram-negative bacteria is functionally akin to membrane-bound eukaryotic organelles, and that the periplasmic "organelle" functions principally to isolate the cytoplasm from molecular traffic entering from, or leaving for, the extracellular milieu (Mario Pantoja, personal communication). Although this idea is appealing, there is scant evidence to support such a broad proposal. However, it is noteworthy that all of the genes in the PLT biosynthetic gene cluster are predicted by DAS analysis to have transmembrane domains (data not shown), suggesting that PLT biosynthesis is a membrane-associated process. Which side of the membrane is responsible for the biosynthesis of PLT is not known. There is evidence of at least one instance (albeit in a Gram-positive bacterium) in which an enzymatic reaction resulting in polyketide synthesis occurs outside of the cytoplasmic membrane (2). 160 5.5 PLT accumulation requires a set of flanking transporter genes, which are in turn regulated by PLT. Circumstantial evidence supports a cytoplasmic location for PLT production. A set of transport functions encoded by genes flanking the PLT biosynthetic gene appears to be tightly associated with PLT production. Such a transport system would seem unnecessary if PLT were produced in the periplasm. The transport system may facilitate the passage of PLT across the membrane; alternatively, its role may involve another substrate entirely. I identified two genes, pill and pltJ, that appear to be part of a larger cluster (pitH, pill, pitf, orJK, and orfP) with potential for either i) PLTregulated PLT transport or ii) PLT-regulated transport of a different substrate. Mutations inpitlorpklinhibited PLT production by Pf-5 approximately five-fold. Moreover, apit.J mutation abolished PLT autoinduction. Optimal transcription of piti and pltJ, in turn, requires the production of PLT by Pf-5 or its addition to the medium. The upregulation of pill and pit.J expression by exogenous PLT was nearly immediate (2 hours). Thus, it is evident that, besides affecting its own production and that of other secondary metabolites, PLT also exerts tight control over the machinery for transport of a substrate, possibly itself, across the cytoplasmic membrane. The integrity of that substrate transport machinery, in return, is required for optimal PLT production and for the PLT autoinduction phenomenon. A hypothesis awaiting testing is that PLT autoinduction requires transport of PLT from the external medium across the outer and inner membrane, into the cytoplasm, and that pitI and pilf are part of an ABC transport apparatus that is involved in PLT import. In such a scenario, pitI and pil.J mutations would be expected to lower, but not abolish, PLT production by interrupting PLT autoinduction. This hypothesis is consistent with the data collected thus far. 5.6. Significance The data from the studies described herein demonstrate that the production of PLT by Pf-5 can affect the production of PLT by neighboring cells. PLT and other 161 exogenous secondary metabolites were shown to have both autoregulatory and cross- regulatory effects in vitro. Because Pf-5 derivatives engaged in PLT cross-feeding in the rhizosphere, it is likely that cross-feeding occurs for other secondary metabolites as well. If cross-feeding occurs, then production of PLT as well as other secondary metabolites by one cell will profoundly affect the production of secondary metabolites in neighboring cells. New evidence hints that PLT influences the iron status of cells that perceive it. Thus, a new variable in the efficacy of biological control is now apparent: inhibition of the production of certain antibiotics by others produced by the biocontrol agent. For instance, when environmental conditions and the physio1ogica status of Pf-5 favor the production of 2,4-DAPG, the production of PLT is inhibited, and vice versa. Because the different antibiotics produced by Pf-5 do not have the same spectrum of activity against target pathogens, (e.g. reference 3), knowledge of environmental conditions that affect antibiotic production during application of a biological control agent is essential. Strategies should be employed to maintain expression of traits that are most important in the suppression of the target pathogen. 162 5.7 References Abbas, A., J. P. Morrissey, P. Carnicero Marquez, M. M. Sheehan, I. R. Delany, and F. O'Gara. 2002. Characterization of interactions between the transcriptional repressor PhIF and its binding site at the ph/A promoter in Pseudomonasfluorescens Fl 13. J. Bacteriol. 184:3008-3016. 2. Hesketh, A. R., G. Chandra, A. D. Shaw, J. J. Rowland, D. B. Kell, M. J. Bibb, and K. F. Chater. 2002. Primary and secondary metabolism, and post-translational protein modifications, as portrayed by proteomic analysis of Streptomyces coelicolor. Mol. Microbiol. 46:917-932. 3. Howell, C. R., and R. D. Stipanovic. 1980. Suppression of Pythium ultimum-induced damping-off of cotton seedlings by Pseudomonas fluorescens and its antibiotic, pyoluteorin. Phytopathology 70:712-7 15. 4. Keel, C., D. M. Weller, A. Natsch, G. Défago, R. J. Cook, and L. S. Thomashow. 1996. Conservation of the 2,4-diacetylphioroglucinol biosynthesis locus among fluorescent Pseudomonas strains from diverse locations. Appi. Environ. Microbiol. 62:552-563. Kirner, S., P. E. Hammer, D. S. Hill, A. Altmann, I. Fischer, L. J. Weislo, M. Lanahan, K-H. van Pee, and J. M. Ligon. 1998. Functions encoded by pyrrolnitrin biosynthetic genes from Pseudomonasfluorescens. J. Bact. 180(7): 1939-1943. 6. Notz, R., M. Maurhofer, H. Dubach, D. Haas, and G. Defago. 2002. Fusaric acid-producing strains of Fusarium oxysporum alter 2,4diacetyiphioroglucinol biosynthetic gene expression in Pseudomonas fluorescens CHAO in vitro and in the rhizosphere of wheat. App. Environ. Microbiol. 68:2229-2235. 7. Nowak-Thompson, B., N. Chaney, S. J. Gould, and J. E. Loper. 1999. Characterization of the pyoluteorin biosynthetic gene cluster of Pseudomonasfluorescens Pf-5. J. Bactenol. 181 :2166-1274. 8. Raaijmakers, J. M., D. M. Weller, and L. S. Thomashow. 1997. Frequency of antibiotic-producing Pseudomonas spp. in natural environments. Appi. Environ. Microbiol. 63:881-887. 163 9. Sharifi-Tehranj, A., M. Zala, A. Natsch, Y. Moënne-Loccoz, and G. Défago. 1998. Biocontrol of soil-borne fungal plant diseases by 2,4diacetyiphioroglucinol-producing fluorescent pseudomonads with different restriction profiles oof amplified 16S rDNA. Eur. J. Plant Pathol. 104:63 1643. 10. Schnider-Keel, U., A. Seematter, M. Maurhofer, C. Blumer, B. Duffy, C. Gigot-Gonnefoy, C. Reimmann, R. Notz, G. Défago, D. Haas, and C. Keel. 2000. Autoinduction of 2,4- diacetylphloroglucinol biosynthesis in the biocontrol agent Pseudomonasfluorescens CHAO and repression by the bacterial metabolites salicylate and pyoluteorin. J. Bacteriol. 182:12151225. 164 Bibliography Aarons, S., A. Abbas, C. Adams, A. Fenton, and F. O'Gara. 2000. A regulatory RNA (PrrB RNA) modulates expression of secondary metabolite genes in Pseudomonasfluorescens P113. J. Bacteriol. 182:3913-3919. Abbas, A., J. P. Morrissey, P. Carnicero Marquez, M. M. Sheehan, I. R. Delany, and F. O'Gara. 2002. Characterization of interactions between the transcriptional repressor Ph1F and its binding site at the phiA promoter in P. fluorescens F 113. J. Bacteriol. 184:3008-3016. Agrios, G. N. 1988. Plant Pathology. Academic Press, San Diego, CA. Altschul, S. F., T. L. Madden, A. A. Schäffer, J. Zhang, Z. Zhang, W. Miller, and D. J. Lipman. 1997. Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res. 25 :3389-3402. Altuvia, S., A. Zhang, L. Argaman, A. Tiwari, and G. Storz. 1998. The Escherichia coli OxyS regulatory RNA repressesflulA translation by blocking ribosome binding. EMBO J. 17:6069-6075. Ambrosi, C., L. Leoni, and P. Visca. 2002. Different responses of pyoverdine genes to autoinduction in Pseudomonas aeruginosa and the group Pseudomonas fluorescens-Pseudomonas putida. Appl. Environ. Microbiol. 68(8): 4122-4 126. Anjaiah, V., N. Koedam, B. Nowak-Thompson, J. E. Loper, M. Höfte, J. T. Tambong, and P. Cornelis. 1998. Involvement of phenazines and anthranilate in the antagonism of Pseudomonas aeruginosa PNA1 and Tn5 derivatives against Fusarium spp. and Pythium spp. Mo!. Plant-Microbe Interact. 11:847-854. Atichartpongkul S., S. Loprasert, P. Vattanaviboon, W. Whangsuk, J. D. Helmann, and S. Mongkolsuk. 2001. Bacterial Ohr and OsmC paralogues define two protein families with distinct functions and patterns of expression. Microbiol. 147:1775-82. Ausubel., F., R. Brent, R. E. Kingston, D. E. Moore, J. G. Seidman, J. A. Smith, and K. Struhi (eds.). 1998. Current Protocols in Molecular Biology. Wiley and Sons, Inc. New York, NY. Bailey, D., R. E. Johnson, and U. J. Salvador. 1973. Pyrrole antibacterial agents. 1. Compounds related to pyoluteorin. J. Med. Chem. 19:1298-1300. 165 Bangera, G. M., and L. S. Thomashow. 1996. Characterization of a genomic locus required for synthesis of the antibiotic 2,4-diacetyiphioroglucinol by the biological control agent Pseudomonasfluorescens Q2-87. Mol. Plant-Microbe Interact. 9:83-90. Bangera, G. M., and L. S. Thomashow. 1999. Identification and characterization of a gene cluster for synthesis of the polyketide antibiotic 2,4-diacetyiphioroglucinol from Pseudomonasfluorescens Q2-87. J. Bacteriol. 181:3155-3163. Bateman, A., E. Birney, L. Cerruti, R. Durbin, L. Etwiller, S. R. Eddy, S. Griffiths-Jones, K. L. Howe, M. Marshall, and E. L. L. Sonnhammer. 2002. The Pfam protein families database. Nucleic Acids Res. 30:276-280. Beare, P. A., R. J. For, L. W. Martin, and I. L. Lamont. 2003. Siderophoremediated cell signalling in Pseudomonas aeruginosa: divergent pathways regulate virulence factor production and siderophore receptor synthesis. Mol. Microbiol. 47:195-207. Becker, J. 0., and F. J. Schwinn. 1993. Control of soil-borne pathogens with living bacteria and fimgi: status and outlook. Pestic. Sci. 37:355-363. Benbrook, C. M., E. Groth, J. M. Halloran, M. K. Hansen, and S. Marquardt. 1996. Pest management at the crossroads. 272 pp. Consumers Union, Yonkers, NY. Bertani, I., G. Devescovi, and V. Venturi. 1999. Controlled specific expression and purification of 6xllis-tagged proteins for Pseudomonas. FEMS Microbiol. Left. 179:101-106. Blumer, C., and D. Haas. 2000. Iron regulation of the hcnABC genes encoding hydrogen cyanide synthase depends on the anaerobic regulator ANR rather than on the global activator GacA in Pseudomonasfluorescens CHAO. Microbiology 146:24 172424. Blumer, C., and D. Haas. 2000. Multicopy suppression of a gacA mutation by the infC operon in Pseudomonasfluorescens CHAO: competition with the global translational regulator RsmA. FEMS Microbiol. Lett. 187:53-58. Blumer, C., S. Heeb, G. Pessi, and D. Haas. 1999. Global GacA-steered control of cyanide and exoprotease production in Pseudomonasfluorescens involves specific ribosome binding sites. Proc. Natl. Acad. Sci. USA 96:14073-14078. Bonsall, R. F., D. M. Weller, and L. S. Thomashow. 1997. Quantification of 2,4diacetylphloroglucinol produced by fluorescent Pseudomonas spp. in vitro and in the rhizosphere of wheat. Appi. Environ. Microbiol. 63:951-955. 166 Boos, W. and T. Eppler. 2001. Prokaryotic binding protein-dependent ABC transporters. p. 77-114, In: G. Winkelmann (ed.), Microbial transport systems. Wiley VCH, Weinheim, Germany. Boos, W. and Shuman, H. 1998. MaltoseImaltodextrin system of Escherichia co/i: transport, metabolism, and regulation. Microbiol. Mo!. Biol. Rev. 62:204-229. Breines, D., and J. Burnham. 1994. Modulation of Escherichia coil type 1 fimbrial expression and adherence to uroepithelial cells following exposure of logarithmic phase cells to quinolones at subinhibitory concentrations. J. Antimicrob. Chemother. 34:205-221. Brodhagen, M., M. D. Henkels, and J. E. Loper. 2003. Positive autoregulation of the antibiotic pyoluteorin in the biological control organism Pseudomonas fluorescens Pf-5. Submitted. Bundy, B. M., L. S. Collier, T. R. Hoover, and E. L. Neidle. 2002. Synergistic transcriptional activation by one regulatory protein in response to two metabolites. Proc. Nat!. Acad. Sci. USA 99:7693-7698. Callanan, M., R. Sexton, D. N. Dowling, and F. O'Gara. 1996. Regulation of the iron uptake genes in Pseudomonasfluorescens Ml 14 by pseudobactin Ml 14: the pbrA sigma factor does not mediate the siderophore regulatory response. FEMS Microbiol. Lett. 144:61-66. Camacho, M. M. 2001. Molecular characterization of the role of Type 1V pili, NDH-I and PyrR in rhizosphere colonization ofPseudomonasfluorescens WCS36S. PhD Thesis. Universiteit Leiden. Cangelosi, G. A., R. G. Ankenbauer, and E .W. Nester. 1990. Sugars induce the Agrobacterium virulence genes through a periplasmic binding protein and a transmembrane signal protein. Proc. Nati. Acad. Sci. USA 87:6708-67 12. Carroll, H., Y. Moënne-Loccoz, D. N. Dowling, and F. O'Gara. 1995. Mutational disruption of the biosynthesis genes coding for the antifungal metabolite 2,4diacetyiphioroglucinol does not influence the ecological fitness of Pseudomonas fluorescens Fl 13 in the rhizosphere of sugarbeets. App!. Environ. Microbiol. 61:3002-3007. Chadwick, D. J., and J. Whelan (eds.). 1992. Secondary metabolites: their function and evolution. John Wiley and Sons, Chichester, England. 318 pp. Chatterjee, A., Y. Cui, Y. Liu, C. K. Dumenyo, and A. K. Chatterjee. 1995. Inactivation of rsmA leads to overproduction of extracellular pectinases, cellulases, and proteases in Erwinia carotovora subsp. carotovora in the absence of 167 starvation/cell density-sensing signal, N-(3-oxohexanoyl)-L-homoserine lactone. Appi. Environ. Microbiol. 61:1959-1967. Chen, Q., A. Nüssbaum-Shochat, and 0. Amster-Choder. 2001. A novel sugarstimulated covalent switch in a sugar sensor. J. Biol. Chem. 276:44751-44756. Chin-A-Woeng, T. F. C., G. V. Bloemberg, A. J. van der Bij, K. M. G. M. van der Drift, J. Schripsema, B. Kroon, R. J. Scheffer, C. Keel, P. A. H. M. Bakker, H. V. Tichy, F. J. de Bruijn, J. E. Thomas-Oates, and B. J. J. Lugtenberg. 1998. Biocontrol by phenazine- 1 -carboxamide-producing Pseudomonas chioraphis PCL1 391 of tomato root rot caused by Fusarium oxysporum f. sp. radicis-lycopersici. Mol. Plant-Microbe Interact. 11:1069-1077. Chin-A-Woeng, T. F. C., D. van den Broek, G. de Voer, D. M. G. M. van der Drift, S. Tuinman, J. E. Thomas-Oates, B. J. J. Lugtenberg, and G. V. Bloemberg. 2001. Phenazine-1-carboxamide production in the biocontrol strain Pseudomonas chiororaphis PCL1391 is regulated by multiple factors secreted into the growth medium. Mol. Plant-Microbe Interact. 14:969-979. Claros, M. G., and G. von Heijne. 1994. TopPred II: an improved software for membrane protein structure predictions. Comput. Appi. Biosci. 10:685-686. Cook, J. IL 1988. Biological control and holistic plant-health care in agriculture. American Journal of Alternative Agriculture 3:51-62. Cook, R. J. 1993. Making greater use of introduced microorganisms for biologica! control of plant pathogens. Annu. Rev. Phytopath. 31:53-80. Cook, IL J., W. L. Bruckart, J. IL Coulson, M. S. Goettel, IL A. Humber, R. D. Lumsden, J.V. Maddox, M. L. McManus, L. Moore, S.F. Meyer, P.C. Quimby, Jr., J.P. Stack, and J. L. Vaughn. 1996. Safety of microorganisms intended for pest and plant disease control: a framework for scientific evaluation. Biol. Control 7:333351. Cook, IL J., L. S. Thomashow, D. M. Weller, D. Fujimoto, M. Mazzola, G. Bangera, and D. -S. Kim. 1995. Molecular mechanisms of defense by rhizobacteria against root disease. Proc. Nati. Acad. Sci. USA 92:4197-4201. Corbell, N. 1999. A two-component regulatory system controlling antibiotic production by Pseudomonasfluorescens Pf-5. PhD Thesis. Oregon State University. Corbell, N., and J.E. Loper. 1995. A global regulator of secondary metabolite production in Pseudomonasfluorescens Pf-5. J. Bacteriol. 177(21): 6230-6236. 168 Cornish, A., J. A. Greenwood, and C. W. Jones. 1989. Binding-protein-dependent sugar transport by Agrobacterium radiobacter and A. tumefaciens grown in continuous culture. J Gen Microbiol. 135:3001-3013. Couzin, J. 2002. Breakthrough of the year #1, the winner: Small RNAs make big splash. Science 298:2296-2297. Cserzo, M., E. Wallin, I. Simon, G. von Heijne, and A. Elofsson. 1997. Prediction of transmembrane aipha-helices in procariotic membrane proteins: the Dense Alignment Surface method. Prot. Eng. 10:673-676. Cui, Y., Chatterjee, A., and A. K. Chatterjee. 2001. Effects of the two-component system comprising GacA and GacS of Erwinia carotovora subsp. carotovora on the production of global regulatory rsmB RNA, extracellular enzymes, and harpin&. Mo!. Plant-Microbe Interact. 14:516-526. Cui, Y., A. Chatterjee, Y. Liu, C. K. Dumenyo, and A. K. Chatterjee. 1995. Identification of a global repressor gene, rsmA, of Erwinia carotovora subsp. carotovora that controls extracellular enzymes, N-(3-oxohexanoyl)-L-homoserine lactone, and pathogenicity in sort-rotting Erwinia spp. J. Bactenol. 177:5108-5115. Cui, Y., A. Mukherjee, C. K. Dumenyo, Y. Liu, and A. K. Chatterjee. 1999. rsmC of the soft-rotting bacterium Erwinia carotovora subsp. carotovora negatively controls extracellular enzyme and harpinE production and virulence by modulating levels of regulatory RNA (rsmB) and RNA-binding protein (RsmA). J. Bacteriol. 181:6042-6052. Cuppels, D.A., C.R. Howell, RD. Stipanovic, A. Stoessi, and J.B. Stothers. 1986. Biosynthesis of pyoluteorin: a mixed polyketide-tricarboxylic acid cycle origin demonstrated by [1 ,2-13C2}acetate incorporation. Z. Naturforsch. 41 C: 532-536. Delaney, I., M. M. Sheehan, A. Fenton, S. Bardin, S. Aarons and F. O'Gara. 2000. Regulation of production of the antifungal metabolite 2,4diacetylphloroglucinol in Pseudomonasfluorescens Fl 13: genetic analysis of phiF as a transcriptional repressor. Microbiology 146:537-546. Delepaire, P. 1998. Erwinia metalloprotease permease: aspects of secretion pathway and secretion functions. Meth. Enzymol. 292:67-81. Dinh, T., I. T. Paulsen, and M. H. Saier, Jr. 1994. A family of extracytoplasmic proteins that allow transport of large molecules across the outer membranes of Gramnegative bacteria. J. Bacteriol. 176:3825-3831. Dowling, D. N., and F. O'Gara. 1994. Metabolites of Pseudomonas involved in the biocontrol of plant disease. Trends Biotech. 12:133-14. 169 Duffy, B. K., and G. Défago. 1997. Zinc improves biocontrol of Fusarium crown and root rot of tomato by Pseudomonasfluorescens and represses the production of pathogen metabolites inhibitory to bacterial antibiotic biosynthesis. Phytopathology 87:1250-1267. Duffy, B. K. and G. Défago. 1999. Environmental factors modulating antibiotic and siderophore biosynthesis by Pseudomonasfluorescens biocontrol strains. Appi. Environ. Microbiol. 65(6'): 2429-2438. Duke, S. 0., Menn, J. J., and J. R. Plimmer. 1993. Challenges of pest control with enhanced toxicological and environmental safety. p. 1-13 In: Duke, S. 0., J. J. Menn, and J. R. Plimmer (eds.). Pest control with enhanced environmental safety. American Chemical Society, Washington, D.C. Eberhard, A., A. L. Burlingame, C. Eberhard, G. L. Kenyon, K. H. Nealson, and J. J. Oppenheimer. 1981. Structural identification of autoinducer of Photobacterium fischeri luciferase. Biochemistry 20:2444- 2449. Eberl, L., M. K. Winson, C. Sternberg, G. S. A. B. Stewart, G. Christiansen, S. it Chhabra, B. Bycroft, P. Williams, S. Molin, and M. Givskov. 1996. Involvement of N-acyl-L-homoserine lactone autoinducers in controlling the multicellular behaviour of Serratia liquefaciens. Mol. Microbiol. 20:127-136. Eisenstark, A., Calcutt, M.J., Becker-Hapak, M., and A. Ivanova. 1999. Role of Escherichia co/i rpoS and associated genes in defense against oxidative damage. Free Rad. Biol. Med. 21: 975-993. Elmholt, S. 1991. Side effects of propiconazole (Tilt 250 EC) on non-target soil fungi in a field trial compared with natural stress effects. Microb. Ecol. 22:99-108. El-Sayed, A. K., J. Hothersall, and C. M. Thomas. 2001. Quorum-sensingdependent regulation of biosynthesis of the polyketide antibiotic mupirocin in Pseudomonasfluorescens NCIMB 10586. Microbiol. 147:2127-2139. Feiton, G. W., and D. L. Dahlman. 1984. Nontarget effects of a fungicide: toxicity of Maneb to the parasitoid Microplitis croceipes (Hymenoptera: Braconidae) Heliothis virescens]. J. Econ. Entomol. 77:842-850. Fenton, A. M., P. M. Stephens, J. Crowley, M. OCa1laghan, and F. O'Gara. 1992. Exploitation of gene(s) involved in 2,4-diacetylphloroglucinol biosynthesis to confer a new biocontrol capability to a Pseudomonas strain. Appl. Environ. Microbiol. 58:3873-3878. IVKI] Figurski, K. H. and D. R. Helinski. 1979. Replication of an origin-containing derivative of plasmid RK2 dependent on a plasmid function provided in trans. Proc. Nat!. Acad. Sci. USA 76:1648-1652. Flaishman, M., Z. Eyal., C. Voisard, and D. Haas. 1990. Suppression of Septoria tritici by phenazine or siderophore-deflcient mutants of Pseudomonas. Curr. Microbiol. 20:121-124. Froyd, J. D. 1997. Can synthetic pesticides be replaced with biologically-based alternatives? - an industry perspective. Journal of Industrial Microbiology and Biotechnology 19:192-195. Gensberg, K., K. Hughes, and A. W. Smith. 1992. Siderophore-speciflc induction of iron uptake in Pseudomonas aeruginosa. J. Gen. Microbio!. 138(Pt 11):2381-2387. Glandorf, D. C. M., P. Verheggen, T. Jansen, J.-W. Jorritsma, E. Smit, P. Leeflang, K. Wernars, L. S. Thomashow, E. Laureijs, J. E. Thomas-Oates, P. A. H. M. Bakker, and L. C. van Loon. 2001. Effect of genetically modified Pseudomonasputida WCS358r on the fungal rhizosphere microflora of field-grown wheat. Appl. Environ. Microbiol. 67:3371-3378. Goethals, K., M. van Montagu, and M. Holsters. 1992. Conserved motifs in a divergent nod box of Azorhizobium caulinodans 0RS571 reveal a common structure in promoters regulated by LysR-type proteins. Proc. Nati. Acad. Sci. USA 89:16461650. Gob, E. -B., G. Yim, W. Tsui, J. McClure, M. G. Surette, and J. Davies. 2002. Transcriptional modulation of bacterial gene expression by subinhibitory concentrations of antibiotics. Proc. Nati. Acad. Sci. USA. 99:17025-17030. Görke, B., and B. Rak. 1999. Catabolic control of Escherichia coli regulatory protein BgIG activity by antagonistically acting phosphorylations. EMBO J. 18:33703379. Görke, B., and B. Rak. 2001. Efficient transcriptional antitermination from the Escherichia co/i cytoplasmic membrane. J. Mol. Biol. 308:131-145. Gottesman, S. 1996. Proteases and their targets in Escherichia co/i. Annu. Rev. Genet. 30: 465-506. Grimwood, K., M. To, H. IL Rabin, and D. E. Woods. 1989. Inhibition of Pseudomonas aeruginosa exoenzyme expression by subinhibitory antibiotic concentrations. Antimicrob. Agents Chemother. 33:41-47. 171 Gurusiddaiah, S., D. M. Weller, A. Sarkar, and R. J. Cook. 1986. Characterization of an antibiotic produced by a strain ofPseudomonasfluorescens inhibitory to Gaeumannomyces graminis var. tritici and Pythium spp. Antimicrob. Agents Chemother. 29:488-495. Gutterson, N., J.S. Ziegle, G.J. Warren, and T.J. Layton. 1988. Genetic determinants for catabolite induction of antibiotic biosynthesis in Pseudomonas fluorescens HV37a. J. Bacteriol. 170(1): 380-385. Haas, D., C. Blumer, and C. Keel. 2000. Biocontrol ability of fluorescent pseudomonads genetically dissected: importance of positive feedback regulation. Curr. Opin. Biotech. 11:290-297. Haas, D., and C. Keel. 2003. Regulation of antibiotic production in root-colonizing Pseudomonas spp. and relevance for biological control of plant disease. Annu. Rev. Phytopathol. Online. doi: 10.1 l46fannurev.phyto.4 1.052002.095656 Handeisman, J., and E. V. Stabb. 1996. Biocontrol of soilbome plant pathogens. Plant Cell 8:1855-1869. Hara, 0., and T. Beppu. 1982. Mutants blocked in streptomycin production in Streptomyces griseus - the role of A-factor. J. Antibiot. 35:349-35 8. Hardy, R. W. F. 1993. Biologically based pest management in managed ecosystems: Increasing its acceptance. p. 2-9, In: Lumsden, R. D. and J. L. Vaughn (eds.). Pest management: Biologically based technologies. American Chemical Society, Washington, D.C. Hardy, R. W. F., Beachy, R. N., Browning, H., Caulder, J. D., Charudattan, R., Faulkner, P., Gould, F. L., Hinkle, M. K., Jaffee, B. A., Knudson, M. K., Lewis, W. J., Loper, J. E., Mahr, D. L., and van Alfen, N. K. 1996. NRC Report. Ecologically based pest management: New solutions for a new century. 144 pp. Nati. Acad. Sci. Press. Harley, K. T., G. M. Djordjevic, T. T. Tseng, and M. H. Saler, Jr. 2000. Membrane fusion protein homologues in Gram-positive bacteria. Mol. Microbiol. 36:5 16-5 17. Harrison, L. A., L. Letendre, P. Kovacevich, E. Pierson, and D. Weller. 1993. Purification of an antibiotic effective against Gaeumannomyces graminis var. tritici produced by a biocontrol agent, Pseudomonas aureofaciens. Soil Biol. Biochem. 25:215-221. 172 Häse, C. C., and J. J. Mekalanos. 1998. TcpP protein is a positive regulator of virulence gene expression in Vibrio cholerae. Proc. NatI. Acad. Sci. USA 95:730734. Heeb, S., and D. Haas. 2001. Regulatory roles of the GacS/GacA two-component system in plant-associated and other gram-negative bacteria. Mol. Plant-Microbe Interact. 14:1351-1363. Heeb, S., C. Blumer, and D. Haas. 2002. Regulatory RNA as mediator in GacAIRsmA-dependent global control of exoproduct formation in Pseudomonas fluorescens CHAO. J. Bacteriol. 183:1046-1056. Hengge-Aronis, R. 2000. The general stress response in Escherichia co/i. pp. 161178. In: Bacterial Stress Responses. Storz, G. and R. Hengge-Aronis (Eds.). ASM Press, Washington, D.C. Hengge-Aronis, R., Lange, R., Henneberg, N., and D. Fischer. 1993. Osmotic regulation of ipoS-dependent genes in Escherichia co/i. J. Bacteriol. 175: 259-265. Herbert, S., P. Barry, and R. P. Novick. 2001. Subinhibitory clindamycin differentially inhibits transcription of exoprotein genes in Staphylococcus aureus. Infect. Immun. 69:2996-3003. Hesketh, A. R., G. Chandra, A. D. Shaw, J. J. Rowland, D. B. Kell, M. J. Bibb, and K. F. Chater. 2002. Primary and secondary metabolism, and post-translational protein modifications, as portrayed by proteomic analysis of Streptomyces coelicolor. Mol. Microbiol. 46:917-932. Higgins, C. F. 1992. ABC transporters: from microorganisms to man. Annu. Rev. Cell. Biol. 8:67-113. Hill, D. S., J. I. Stein, M. R. Torkewitz, A. M. Morse, C. IL Howell, J. P. Pachlatko, J. 0. Becker, and J. M. Ligon. 1994. Cloning of genes involved in the synthesis of pyrrolnitrin from Pseudomonasfluorescens and role of pyrrolnitrin synthesis in biological control of plant disease. Appl. Environ. Microbiol. 60:78-85. Hillen, W., and C. Berens. 1994. Mechanisms underlying expression of TnlO encoded tetracycline resistance. Annu. Rev. Microbiol. 48:345-369. Hofnung, M., D. Hatfield, and M. Schwartz. 1974. ma/B region in Escherichia coli K-12: characterization of new mutations. J. Bacteriol. 117:40-47. Holden, M. T. G., S. R. Chhabra, R. de Nys, P. Stead, J. J. Bainton, P. J. Hill, M. Manefield, N. Kumar, M. Labatte, D. England, S. Rice, M. Givskov, G. P. C. Salmond, G. S. A. B. Stewart, B. W. Bycroft, S. Kjelleberg, and P. Williams. 173 1999. Quorum-sensing cross talk: isolation and chemical characterization of cyclic dipeptides from Pseudomonas aeruginosa and other Gram-negative bacteria. Mo!. Microbjol. 33:1254-1266. Homma, Y. 1994. Mechanism of biological control focussed on the antibiotic pyrrolnitrin. p. 100-103 In: M. H. Ryder, P. M. Stephens, and G. D. Bowen (eds.), Improving plant productivity with rhizobacteria. CSIRO Division of Soils, Adelaide, Australia. Horinouchi, S. 1999. y-butyrolactones that control secondary metabolism and cell differentiation in Streptomyces. p. 193-207. In G. M. Dunny and S. C. Winans (ed.), Cell-cell signaling in bacteria. ASM Press, Washington, D.C. Howell, C. R. (Agricultural Research Station, U.S. Department of Agriculture, College Station, Tex.). 1993. Personal communication. Howell, C. R. and R. D. Stipanovic. 1979. Control ofRhizoctonia solani on cotton seedlings with Pseudomonasfluorescens and with an antibiotic produced by the bacterium. Phytopathology 69:480-482. Howell, C. R., and R. B. Stipanovic. 1980. Suppression of Pythium ultimuminduced damping-off of cotton seedlings by Pseudomonasfluorescens and its antibiotic, pyoluteorin. Phytopathogy 70:712-715. Howie, W. J. and T. V. Suslow. 1991. Role of antibiotic biosynthesis in the inhibition of Pythium ultimum in the cotton spermosphere and rhizosphere by Pseudomonasfluorescens. Mol. Plant-Microbe Interact. 4:393-399. Ishimaru, C. A., E. J. Klos, and R. R. Brubaker. 1988. Multiple antibiotic production by Erwinia herbicola. Phytopathology 78:746-750. Keel, C.., U. Schnider, M. Maurhofer, C. Voisard, J. Laville, U. Burger, P. Wirthner, D. Haas, and G. Défago. 1992. Suppression ofroot diseases by Pseudomonasfluorescens CHAO: importance of the bacterial secondary metabolite 2,4-diacetylphloroglucinol. Mol. Plant-Microbe Interact. 5:4-13. Keel, C., D. M. Weller, A. Natsch, G. Defago, R. J. Cook, and L. S. Thomashow. 1996. Conservation of the 2,4-diacetylphloroglucinol biosynthesis locus among fluorescent Pseudomonas strains from diverse locations. Appi. Environ. Microbiol. 62:552-563. Keel, C., P. Wirthner, T. Oberhänsli, C. Voisard, Burger, D. Hass, and G. Défago. 1990. Pseudomonads as antagonists of plant pathogens in the rhizosphere: role of the antibiotic 2,4-diacetylphloroglucinol in the suppression of black root rot of tobacco. Symbiosis 9:327-341. 174 Keen, N. T., S. Tamaki, D. Kobayashi, and D. Trollinger. 1988. Improved broadhost-range plasmids for DNA cloning in Gram-negative bacteria. Gene 70:191-197. Khokhlov, A. S. 1988. Results and perspectives of actinomycete autoregulators studies, p. 338-345. In Y. Okami, T. Beppu, and H. Ogawara (ed.). Biology of Actinomycetes '88. Japan Scientific Societies Press, Tokyo, Japan. Khvorova, A., Y.-G. Kwak, M. Tamkun, I. Majerfield, and M. Yarns. 1999. RNAs that bind and change permeability of phospholipid membranes. Proc. Nati. Acad. Sci. USA 96:10649-10654. Kim, B. S., S. S. Moon, and B. K. Hwang. 1999. Isolation, identification, and antifungal activity of a macrolide antibiotic, oligomycin A, produced by Streptomyces libani. Can. J. Bot. 77:850-858. King, E. 0., M. K. Ward, and D. E. Raney. 1954. Two simple media for the demonstration of pyocyanin and fluorescein. J. Lab. Clin. Med. 44:30 1-307. Kirner, S., P. E. Hammer, D. S. Hill, A. Aitmaun, I. Fischer, L. J. Weislo, M. Lanahan, K-H. van Pee, and J. M. Ligon. 1998. Functions encoded by pyrrolnitrin biosynthetic genes from Pseudomonasfluorescens. J. Bact. 180:1939-1943. Kraus, J., and J.E. Loper. 1992. Lack of evidence for a role of antifungal metabolite production by Pseudomonasfiuorescens Pf-5 in the biological control of Pythium damping-off of cucumber. Phytopathology 82: 264-271. Kraus, J., and J.E. Loper. 1995. Characterization of a genomic region required for the production of the antibiotic pyoluteorin by the biological control agent Pseudomonasfluorescens Pf-5. App!. Environ. Microbiol. 61(3): 849-854. Lamont, I. L., P. A. Beare, U. Ochsner, A. I. Vasil, and M. L. Vasil. 2002. Siderophore-mediated signaling regulates virulence factor production in Pseudomonas aeruginosa. Proc. Nati. Acad. Sci. USA 99:7072-7077. Lease, R. A., and M. Belfort. 2000. A trans-acting RNA as a control switch in Escherichia coli: DsrA modulates function by forming alternative structures. Proc. Nati. Acad. Sci. USA 97:9919-9924. Lease, R. A., M. E. Cusick, and M. Belfort. 1998. Riboregulation in Escherichia coli: DsrA RNA acts by RNA:RNA interactions at multiple loci. Proc. Nati. Acad. Sci USA 95:12456-12461. 175 Lee, S.-J., W. Boos, J.-P. Bouché, and J. Plumbridge. 2000. Signal transduction between a membrane-bound transporter, PtsG, and a soluble transcription factor, Mic, of Escherichia coli. EMBO J. 19:5353-5361. Létoffé, S., P. Delepaire, and C. Wandersman. 1996. Protein secretion in gramnegative bacteria: assembly of the three components of ABC protein-mediated exporters is ordered and promoted by substrate binding. EMBO J. 15:5804-5811. Liang, T. C., W. Luo, J.-T. Hsieh, and S.-H. Lin. 1996. Antibody binding to a peptide but not the whole protein by recognition of the C-terminal carboxy group. Arch. Biochem. Biophys. 329:208-214. Lindgren, P. B., R. Frederick, A. G. Govindarajan, N. J. Panopoulos, B. J. Staskawicz, and S. E. Lindow. 1989. An ice nucleation reporter gene system: identification of inducible pathogenicity genes in Pseudomonas syringae pv. phaseolicola. EMBOJ. 8:1291-1301. Liu, Y., Y. Cui, A. Muklierjee, and A. K. Chatterjee. 1998. Characterization of a novel RNA regulator of Erwinja carotovora ssp. carotovora that controls production of extracellular enzymes and secondary metabolites. Mo!. Microbiol. 29:219-234. Liii, Y., G. Jiang, Y. Cui, A. Mukherjee, W. 0. Ma, and A. K. Chatterjee. 1999. negatively regulates genes for pectinases, cellulase, protease, harpin&, and a global RNA regulator in Erwinia carotovora subsp. carotovora. J. Bacteriol. 181:2411-2422. /cdgRECC Loper, J. E., and M. D. Henkels. 1999. Utilization of heterologous siderophores enhances levels of iron available to Pseudomonas putida in the rhizosphere. App!. Environ. Microbiol. 65:5357-5363. Loper, J. E., and S. E. Lindow. 1994. A biological sensor for iron available to bacteria in their habitats on plant surfaces. App!. Environ. Microbiol. 60:1934-1941. Loper, J. E., and S. E. Lindow. 2002. Reporter gene systems useful in evaluating in situ gene expression by soil and plant-associated bacteria, p. 627-637. In C. J. Hurst, R. L. Crawford, G. R. Knudsen, J. J. Mclnemey, and L. D. Stetzenbach (ed.), Manual of environmental microbiology, 2 edition. ASM Press, Washington, D.C. Loper, J. E., B. Nowak-Thompson, C. A. Whistler, M. J. Hagen, N. A. Corbell, M. D. Henkels, and V. 0. Stockwell. 1997. Biological control mediated by antifungal metabolite production and resource competition: an overview. In A. Ogoshi, K. Kobayashi, Y. Honima, F. Kodama, N. Kondo and S. Akino (ed.), Proceedings of the Fourth International Workshop on Plant Growth Promoting Rhizobacteria. Sapporo, Japan. 176 MartIn, J. K, and P. Liras. 1989. Organization and expression genes involved in the biosynthesis of antibiotics and other secondary metabolites. Annu. Rev. Microbiol. 43:173-206. Mathre, D. E., R. J. Cook, and N. W. Callan. 1999. From discovery to use: Traversing the world of commercializing biocontrol agents for plant disease control. Plant Dis. 83:972-983. Maurhofer, M., C. Keel, D. llaas, and G. Défago. 1994. Pyoluteorin production by Pseudomonasfluorescens strain CHAO is involved in the suppression of Pythium damping-off of cress but not of cucumber. Eur. J. Plant Pathol. 100:221-232. Maurhofer, M., C. Keel, D. Haas, and G. Défago. 1995. Iniluence of plant species on disease suppression by Pseudomonasfluorescens strain CHAO with enhanced antibiotic production. Plant Pathol. 44:40-50. Maurhofer, M., C. Keel, U. Schnider, C. Voisard, D. Haas, and G. Défago. 1992. Influence of enhanced antibiotic production in Pseudomonasfluorescens strain CHAO on its disease suppressive capacity. Phytopathology 82:190-195. Mazzola, M., R. J. Cook, L. S. Thomashow, D. M. Weller, and L. S. Pierson III. 1992. Contribution of phenazine antibiotic biosynthesis to the ecological competence of fluorescent pseudomonads in soil habitats. App!. Environ. Microbiol. 48:26162624. McCarthy, J. E. G. and R Brimacombe. 1994. Prokaryotic translation: the interactive pathway leading to initiation. Trends Genet. 10:402-407. McGowan, S. J., M. Sebaihia, S. O'Leary, K. R. Hardie, P. Williams, G. S. A. B. Stewart, B. W. Bycroft, and G. P. C. Salmond. 1997. Analysis of the carbapenem gene cluster of Erwinia carotovora: definition of the antibiotic biosynthetic genes and evidence for a novel -lactam resistance mechanism. Mol. Microbiol. 26:545-556. McGowan, S. J., M. Sebaihia, L. E. Porter, G. S. A. B. Stewart, P. Williams, B. W. Bycroft, and G. P. C. Salmond. 1996. Analysis of bacterial carbapenem antibiotic production genes reveals a novel 13-lactam biosynthesis pathway. Mol. Microbiol. 22:415-426. McKeegan, K. S., M. I. Borges-Walmsley, and A. R. Walmsley. 2003. The structure and function of drug pumps: an update. Trends Microbiol 11:21-29. McKnight, S. L., B. H. Iglewski, and E. C. Pesci. 2000. The Pseudomonas quinolone signal regulates rhi quorum sensing in Pseudomonas aeruginosa. J. Bacteriol. 182:2702-2708. 177 McSpadden Gardener, B. B., and D. R. Fravel. 2002. Biological control of plant pathogens: Research, commercialization, and application in the USA. Online. Plant Health Progress doi: 10.1 094[PHP-2002-05 10-01 -RV. McSpadden Gardener, B. B., K. L. Schroeder, S. E. Kalloger, J. M. Raaijmakers, L. S. Thomashow, and D. M. Weller. 2000. Genotypic and phenotypic diversity of phlD-contaimng Pseudomonas strains isolated from the rhizosphere of wheat. App!. Environ. Microbiol. 66: 1939-1946. Mendez, C., and J. A. Salas. 2001. The role of ABC transporters in antibioticproducing organisms: Drug secretion and resistance mechanisms. Res. Microbiol. 152:341-50. Miller, J. H. 1992. A short course in bacterial genetics: A laboratory manual and handbook for Escherichia coli and related bacteria. Cold Spring Harbor Laboratory Press, Plainview, NY. Miller, V. L., R. K Taylor, and J. J. Mekalanos. 1987. Cholera toxin transcriptional activator ToxR is a transmembrane DNA binding protein. Cell 48:27 1279. Mo, Y. Y.., M. Geibel, R. F. Bonsall, and D. C. Gross. 1995. Analysis of sweet cherry (Prunus avium L.) leaves for plant signal molecules that activate the syrB gene required for synthesis of the phytotoxin, syringomycin, by Pseudomonas syringae pv. syringae. Plant Physiol. 107:603-612. Mo, Y. Y., and D. C. Gross. 1991. Plant signal molecules activate the syrB gene, which is required for syringomycin production by Pseudomonas syringae pv. syringae. J. Bacteriol. 173:5784-5792. Mongkolsuk, S., W. Praituan, S. Loprasert, M. Fuangthong, and S. Chamnongpol. 1998. Identification and characterization of a new organic hydroperoxide resistance (ohr) gene with a novel pattern of oxidative stress regulation from Xanthomonas campestris pv. phaseoli. J. Bacteriol. 180:2636-2643. Mukherjee, A.., Y. Ciii, W. Ma, Y. Liii, A. Ishihama, A. Eisenstark, and A. K Chatterjee. 1998. RpoS (Sigma-S) controls expression of rsmA, a global regulator of secondary metabolites, harpin, and extracellular proteins in Erwinia carotovora. J. Bacteriol. 180:3629-3634. Murillo, J., H. J. Shen, D. Gerhold, A. Sharma, 0. A. Cooksey, and N. T. Keen. 1994. Characterization of pPT23B, the plasmid involved in syringolide production by Pseudomonas syringae pv. tomato PT23. Plasmid 31:275-287. 178 Nakayama, T., Y. Homma, Y. Hashkloko, .1. Mizutani, and S. Tahara. 1999. Possible role of xanthobaccms produced by Stenotrophomonas sp. strain SB-K88 in suppression of sugar beet damping-off disease. App!. Environ. Microbiol. 65:43344339. Nealson, K. H. 1999. Early observations defining quorum-dependent gene expression. In: G.M. Dunny and S.C. Winans (ed.), Cell-cell signaling in bacteria. ASM Press, Washington, D.C. Nguyen, L.H., Jensen, D.B., Thompson, N.E., Gentry, D.R., and R.R. Burgess. 1993. In vitro functional characterization of overproduced Escherichia coli katF/rpoS gene product. Biochemistry 32: 11112-11117. Nikaido, H., and J. A. Hall. 1998. Overview of bacterial ABC transporters. Methods Enzymol. 292:3-20. Nogueira, T., and M. Springer. 2000. Post-transcriptional control by global regulators of gene expression in bacteria. Curr. Opin. Microbiol. 3:154-158. Notz, R., M. Maurhofer, H. Dubach, D. Haas, and G. Defago. 2002. Fusaric acidproducing strains of Fusarium oxysporum alter 2,4-diacetylphioroglucinol biosynthetic gene expression in Pseudomonasfluorescens CHAO in vitro and in the rhizosphere of wheat. App. Environ. Microbiol. 68:2229-2235. Notz, it, M. Maurhofer, U. Scbnider-KeeI, B. Duffy, D. Ilaas, and G. Défago. 2001. Biotic factors affecting expression of the 2,4-diacetylphioroglucinol biosynthesis gene phlA in Pseudomonasfluorescens biocontrol strain CHAO in the rhizosphere. Phytopathology 91:873-881. Nowak-Thompson, B. 1997. An integrative approach to understanding pyoluteorin biosynthesis in the biological control bacterium Pseudomonasfluorescens Pf-5. PhD Thesis. Oregon State University. Nowak-Thompson, B., N. Chaney, J. S. Wing, S. J. Gould, and J. E. Loper. 1999. Characterization of the pyoluteorin biosynthetic gene cluster of Pseudomonas fluorescens Pf-5. J. Bacteriol. 181 :2166-2174. Nowak-Thompson, B., Gould, S.J., Kraus, J., and J.E. Loper. 1994. Production of 2,4-diacetyiphioroglucinol by the biocontrol agent Pseudomonasfluorescens Pf-5. Can. LI. Microbiol.40: 1064-1066. Nowak-Thompson, B., S. J. Gould, and J. E. Loper. 1997. Identification and sequence analysis of the genes encoding a polyketide synthase required for pyoluteorin biosynthesis in Pseudomonasfluorescens Pf-5. Gene 204:17-24. 179 O'Sullivan, D.J. and F. O'Gara. 1992. Traits of fluorescent Pseudomonas spp. involved in suppression of plant root pathogens. Microbiological Reviews 56:662676. Ohnishi, Y., S. Kameyama, H. Onaka, and S. Horinouchi. 1999. The A-factor regulatory cascade leading to streptomycin biosynthesis in Streptomyces griseus: identification of a target gene of the A-factor receptor. Mo!. Microbiol. 34:102-111. Onaka, H., N. Ando, T. Nihira, Y. Yamada, T. Beppu, and S. Horinouchi. 1995. Cloning and characterization of the A-factor receptor gene from Streptomyces griseus. J. Bacteriol. 177:6083-6092. Ornston, L. N., and R. Y. Stanier. 1966. The conversion of catechol and protocatechuate to -ketoadipate by Pseudomonas putida. J. Biol. Chem. 241:37763786. Panagiotidis, C. H., W. Boos, and H. A. Shuman. 1998. The ATP-binding cassette subunit of the maltose transporter MaIK antagonizes MalT, the activator of the Escherichia co/i ma! regulon. Mo!. Microbiol. 30:535-546. Paulitz, T. C., and R. IL Belanger. 2001. Biological control in greenhouse systems. Annu. Rev. Phytopathol. 39:103-133. Pesci, E. C., J. B. J. Milbank, J. P. Pearson, S. McKnight, A. S. Kende, E. P. Greenberg, and B. H. Iglewski. 1999. Quinolone signaling in the cell-to-cell communication system of Pseudomonas aeruginosa. Proc. Nat!. Acad. Sci. USA. 96:11229-11234. Pessi, G., F. Williams, Z. Hindle, K. Heurlier, M. T. G. Holden, M. Camara, D. Haas, and P. Williams. 2001. The global posttranscriptional regulator RsmA modulates production of virulence determinants and N-acylhomoserine lactones in Pseudomonas aeruginosa. J. Bacteriol. 183:6676-6683. Pfender, W. F., J. Kraus, and J. E. Loper. 1993. A genomic region from Pseudomonasfluorescens Pf-5 required for pyrrolnitrin production and inhibition of Pyrenophora tritici-repentis in wheat straw. Phytopathology 83:1223-1228. Pidoplichenko, V. N., and A. D. Garagulya. 1974. Effect of antagonistic bacteria on the development of wheat root rot. Microbilogische 1.36:599-602. Pierson, L. S. III, and E. A. Pierson. 1996. Phenazine antibiotic production in Pseudomonas aureofaciens: role in rhizosphere ecology and pathogen suppression. FEMS Microbiol. Lett. 136:101-108. Pierson, E. A., D. W. Wood, J. A. Cannon, F. M. Blachere, and L. S. Pierson III. 1998. Interpopulation signaling via N-acyl-homoserine lactones among bacteria in the wheat rhizosphere. Mol. Plant-Microbe Interact. 11:1078-1084. 180 Plant, E. P., K. L. Muldoon Jacobs, J. W. Harger, A. Meskauskas, J. L. Jacobs, J. L. Baxter, A. N. Petrov, and J. D. Dinman. 2003. The 9-A solution: How mRNA pseudoknots promote efficient programmed -1 ribosomal frameshifting. RNA 9:168174. Poole, K., S. Neshat, and D. Ileinrichs. 1991. Pyoverdine-mediated iron transport in Pseudomonas aeruginosa: involvement of a high-molecular-mass outer membrane protein. FEMS Microbiol. Left. 78:1-6. Quigley, N. B., and D. C. Gross. 1994. Syringomycin production among strains of Pseudomonas syringae pv. syringae: conservation of the syrB and syrD genes and activation of phytotoxin production by plant signal molecules. Mol. Plant-Microbe Interact. 7:78-90. Quigley, N. B., Y.-Y. Mo, and D. C. Gross. 1993. SyrD is required for syringomycin production by Pseudomonas syringae pathovar syringae and is related to a family of ATP-binding secretion proteins. Mo!. Microbiol. 9:787-801. Raaijmakers, J. M., R. F. Bonsall, and D. M. Weller. 1999. Effect of population density of Pseudomonasfluorescens on production of 2,4-diacetylph!oroglucinol in the rhizosphere of wheat. Phytopathology 89:470-475. Raaij makers, J. M., D. M. Weller, and L. S. Thomashow. 1997. Frequency of antibiotic-producing Pseudomonas spp. in natural environments. App!. Environ. Microbiol. 63:881-887. Redly, G. A., and K. Poole. 2003. Pyoverdine-mediated regulation of FpvA synthesis in Pseudomonas aeruginosa: involvement of a probable extracytoplasmicfunction sigma factor, FpvI. J. Bacteriol. 185:1261-1265. Reimmanu, C., L Serino, M. Beyeler, and D. Haas. 1998. Dihydroaeruginoic acid synthetase and pyochelin synthetase, products of the pchEF genes, are induced by extrace!!u!ar pyochelin in Pseudomonas aeruginosa. Microbiol. 144: 3135-3148. Reyes, M. and H. A. Shuman. 1988. Overproduction of Ma1K protein prevents expression of the Escherichia coli mal regulon. J. Bacteriol. 170:4598-4602. Rodriguez, F., and W. F. Pfender. 1997. Antibiosis and antagonism of Scierotinia homoeocarpa and Drechslera poae by Pseudomonasfluorescens Pf-5 in vitro and in planta. Phytopathology 87:614-621. Romeo, T. 1998. Global regulation by the small RNA-binding protein CsrA and the non-coding RNA molecule CsrB. Mo!. Microbiol. 29:1321-1330. 181 Russell, P. E. 1995. Fungicide resistance: occurrence and management. J. Agric. Sci. 124:317-323. Saenger, W., P. Orth, C. Kisker, W. Rillen, and W. llinrichs. 2000. The tetracycline repressor - a paradigm for a biological switch. Angew. Chem. hit. Ed. 39:2042-2052. Sambrook, J. and D. W. Russell. 2001. Molecular Cloning: A Laboratory Manual. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY. Sarniguet, A. (INRA, Centre de recherches de Rennes, Station de patbologie vegetale, Le Rheu cedex, France). 1997. Personal communication. Sarniguet, A., J. Kraus, M. D. Henkels, A. M. Muehichen, and J. E. Loper. 1995. The sigma factor & affects antibiotic production and biological control activity of Pseudomonasfluorescens Pf-5. Proc. Nati. Acad. Sci. USA 92:12255-12259. Schell, M. 1993. Molecular biology of the LysR family of transcriptional regulators. Annu. Rev. Microbiol. 47:597-626. Schneider, E. and S. Hunke. 1998. ATP-binding-casstte (ABC) transport systems: functional and structural aspects of the ATP-hydrolyzing subunits/domains. FEMS Microbiol. Rev. 22:1-20. Schnider, U., C. Keel, C. Blumer, J. Troxier, G. Defago, and D. Haas. 1995. Amplification of the housekeeping sigma factor in Pseudomonasfluorescens CHAO enhances antibiotic production and improves biocontrol abilities. J. Bacteriol. 177:5387-5392. Schnider, U., C. Keel, C. Voisard, G. Defago, and D. Haas. 1995. Tn5-directed cloning ofpqq genes from Pseudomonasfluorescens CHAO: mutational inactivation of the genes results in overproduction of the antibiotic pyoluteorin. Appi. Environ. Microbiol. 61:3856-3864. Schnider-Keel, U., A. Seematter, M. Maurhofer, C. Blumer, B. Duffy, C. GigotGonnefoy, C. Reimmann, R. Notz, G. Defago, D. Haas, and C. Keel. 2000. Autoinduction of 2,4-diacetyiphioroglucinol biosynthesis in the biocontrol agent Pseudomonasfluorescens CHAO and repression by the bacterial metabolites salicylate and pyoluteorin. J. Bacteriol. 182: 1215-1225. Schumann, G. L. 1991. Plant diseases: Their biology and social impact. APS Press, St. Paul, MN. 397 pp. Schwartz, M. 1966. Location of the maltose A and B loci on the genetic map of Escherichia coiL J. Bacteriol. 92:1083-1089. 182 Shah-Smith, D. A., and R. G. Burns. 1997. Shelf-life ofa biocontrol Pseudomonas putida applied to sugar beet seeds using commercial coatings. Biocontrol. Sci. Technol. 7:65-74. Shang, H., J. Chen, J. Handeisman, and R. M. Goodman. 1999. Behavior of Pythium torulosum zoospores during their interaction with tobacco roots and Bacillus cereus. Curr. Microbiol. 38:199-204. Sharifi-Tehrani, A., M. Zala, A. Natsch, Y. Moënne-Loccoz, and G. Défago. 1998. Biocontrol of soilborne fungal plant diseases by 2,4- diacetyiphloroglucinolproducing fluorescent pseudomonads with different restriction profiles of amplified 16S rDNA. Eur. J. Plant Pathol. 104: 631-643. Shen, J., A. Meidrum, and K. Poole. 2002. FpvA receptor involvement in pyoverdine biosynthesis in Pseudomonas aeruginosa. J. Bacteriol. 184:3268-3275. Sigrist, C. J., L. Cerutti, N. Hub, A. Gattider, L Faiquet, M. Pagni, A. Bairoch, and P. Bucher. 2002. PROSITE: a documented database using patterns and profiles as motif descriptors. Brief Bioinform. 3:265-274. Sibo-Suh, L. A., B. J. Lethbridge, S. J. Raffel, H. He, J. Clardy, and J. Handlesman. 1994. Biological activities of two fungistatic antibiotics produced by Bacillus cereus UW85. AppI. Environ. Microbiol. 60:2023-2030. Simon, R, U. Priefer, and A. Puhier. 1983. A broad host range mobilization system for in vivo genetic engineering: transposon mutagenesis in Gram-negative bacteria. Bio/Technology 1: 784-791. Slater, H., M. Crow, L. Everson, and G. P. C. Salmond. 2003. Phosphate availability regulates biosynthesis of two antibiotics, prodigiosin and carbapenem, in Serratia via both quorum-sensing-dependent and -independent pathways. Mo!. Microbiol. 47:303-320. Smith, J. T., and H. -J. Zeiler. 1998. History and introduction, p. 1-11. In J. Kuhlmann, A. Daihoff, and H.-J. Zeiler (ed.), Quinolone antibacterials. Springer, Berlin, Germany. Solomon, J. M., B. A. Lazazzera, and A. D. Grossman. 1996. Purification and characterization of an extracellular peptide factor that affects two different developmental pathways in Bacillus subtilis. Genes Dev. 10:2014-2024. Spaink, H. P., R. J. H. Okker, C. A. Wijffelman, E. Pees, and B. J. J. Lugtenberg. 1987. Promoters in the nodulation region of the Rhizobium leguminosarum Syin plasmid pRL1J1. Plant Mol. Biol. 9:27-39. 183 Staden, R., K. F. Beat, and J. K. Bonfield. 2000. The Staden package, 1998. Methods Mol. Biol. 132:115-130. Stierle, A. C., J. H. Cardellina, and G. A. Strobel. 1988. Maculosin, a hostspecific phytotoxin for spotted knapweed from Alternaria alternata. Proc. Nati. Acad. Sci. USA 85: 8008-8011. Stülke, J., I. Martin-Verstraete, V. Charrier, A. Klier, J. Deutscher, and G. Rapoport. 1995. The HPr protein of the phosphotransferase system links induction and catabolite repression of the Bacillus subtilis levanase operon. J. Bacteriol. 177:6928-6936. Tanaka, K., Takayanagi, Y., Fugita, N., Ishihama, A., and H. Takahashi. 1993. Heterogeneity of the principal c factor in Escherichia co/i: the rpoS gene product, 38 is a second principal c factor of RNA polymerase in stationary-phase Escherichia co/i. Proc. Nat!. Acad. Sci. USA 90: 3511-35 15. Tanaka, Y., K. Kimata, and H. Aiba. 2000. A novel regulatory role of glucose transporter of Escherichia co/i: membrane sequestration of a global repressor Mic. EMBO J. 19:5344-5352. Tateda, K., R. Comte, J.-C. Pechere, T. KOhler, K. Yamaguchi, and C. van Delden. 2001. Azithromycin inhibits quorum sensing in Pseudomonas aeruginosa. Antimicrob. Agents Chemother. 45:1930-1933. Thanabalu, T., E. Koronakis, C. Hughes, and V. Koronakis. 1998. Substrateinduced assembly of a contiguous channel for protein export from E. coli: reversible bridging of an inner-membrane translocase to an outer membrane exit pore. EMBO J. 17:6487-6496. Thomashow, L. S., and D. M. Weller. 1988. Role of a pbenazine antibiotic from Pseudomonasfluorescens in biological control of Gaeumannomyces graminis var. tritici. J. Bacteriol. 170:3499-3508. Thomashow, L. S., and D. M. Weller. 1996. Current concepts in the use of introduced bacteria for biological disease control: mechanisms and antifungal metabolites. p. 236-27 1 In: Plant-Microbe interactions. Stacey, G. and N. T. Keen (eds.). Chapman and Hall, NY. Thomashow, L. S., D. M. Welter, R. F. Bonsall, and L. S. Pierson Ill. 1990. Production of the antibiotic phenazine-1 -carboxylic acid by fluorescent Pseudomonas species in the rhizosphere of wheat. App!. Environ. Microbiol. 56:908-91 2. 184 Thrane, C., T. H. Neilsen, M. N. Neilsen, J. Sørensen, and S. Olsson. 2000. Viscosinamide-producing Pseudomonasfluorescens DR54 exerts a biocontrol effect on Pythium ultimum in sugar beet rhizosphere. FEMS Microbiol. Ecol. 33:139-146. Tortosa, P., S. Aymerich, C. Lindner, M. H. Saier, Jr., J. Reizer, and D. Le Coq. 1997. Multiple phosphorylation of SacY, a Bacillus subtilis transcriptional antiterminator negatively controlled by the phosphotransferase system. J. Biol. Chem. 272:17230-17237. Trancassini, M., A. Giordano, A. Magni, and P. Cipriani. 1993. Subinhibitory concentrations of antibacterial drugs and Pseudomonas aeruginosa virulence factors. New Microbiol. 16:275-279. Ulirich, M. and C. L. Bender. 1994. The biosynthetic gene cluster for coronamic acid, an ethylcyclopropyl amino acid, contains genes homologous to amino acidactivating enzymes and thioesterases. J. Bacteriol. 176:7574-7586. Ullrich, M., A. Penaloza-Vazquez, A.-M. Bailey, and C. L. Bender. 1995. A modified two-component regulatory system is involved in temperature-dependent biosynthesis of the Pseudomonas syringae phytotoxin coronatine. J. Bacteriol. 177:6160-6169. Venturi, V., Ottewanger, C., Bracke, M., and P. J. Weisbeek. 1995. Iron regulation of siderophore biosynthesis and transport in Pseudomonasputida WCS358: involvement of a transcriptional activator and of the Fur protein. Mol. Microbiol. 17: 603-610. Voisard, C., C. Keel, D. Haas, and G. Défago. 1989. Cyanide production by Pseudomonasfluorescens helps suppress black root rot of tobacco under gnotobiotic conditions. EMBO J. 8:35 1-358. von Heijne, G. 1999. Recent advances in the understanding of membrane protein assembly and structure. Quarterly Reviews of Biophysics 32:285-307. Walker, J. E., M. Saraste, M. J. Runswick, and J. J. Gay. 1982. Distantly related sequences in the alpha- and beta-subunits of ATP synthase, myosin, kinases and other ATP-requiring enzymes and a coimnon nucleotide binding fold. EMIBO J. 1:945-951. Wanner, B. L. 1995. Signal transduction and cross regulation in the Escherichia coli phosphate regulon by PhoR, CreC, and acetyl phosphate. p. 302-221, In: Hoch, J. A. and T. J. Silhavy (Cd.). Two-component signal transduction. ASM Press, Washington, D.C. Weller, D.M. 1988. Biological control of soilbome plant pathogens in the rhizosphere with bacteria. Annu. Rev. Phytopatbol. 26: 379-407. 15 West, S. E. H. and B. H. Iglewski. 1988. Codon usage in Pseudomonas aeruginosa. Nucleic Acid Res. 16:9323-9335. Whipps, J. M. 2001. Microbial interactions and biocontrol in the rhizosphere. J. Exp. Bot. 52:487-511. Whistler, C. A. 2000. Regulation of antibiotic production and stress response by the biological control organism Pseudomonasfluorescens Pf-5. PhD Thesis. Oregon State University. Whistler, C. A., Corbell, N. A., Sarniguet, A., Ream, W., and J. E. Loper. 1998. The two-component regulators GacS and GacA influence accumulation of the stationary-phase sigma factor aS and the stress response in Pseudomonasfluorescens Pf-5. J. Bacteriol. 180(24): 6635-6641. Whistler, C. A., V. 0. Stockwell, and J. E. Loper. 2000. Lon protease influences antibiotic production and UV tolerance ofPseudomonasfluorescens Pf-5. Appl. Environ. Microbiol. 66:27 18-2725. Whiteley, M., K. M. Lee, and E. P. Greenberg. 1999. Identification of genes controlled by quorum sensing in Pseudomonas aeruginosa. Proc. Nati. Acad. Sci. USA 96:13904-13909. Wood, D. W., F. Gong, M. M. Daykin, P. Williams, and L. S. Pierson III. 1997. N-acyl-homoserine lactone-mediated regulation of phenazine gene expression by Pseudomonas aureofaciens 30-84 in the wheat rhizosphere. J. Bactenol. 179:76637670. Yang, K., L. Han, and L. C. Vining. 1995. Regulation ofjadomycin B production in Streptomyces venezuelae 1SP5230: involvement of a repressor gene,jadR2. J. Bacteriol 177:6111-6117. Zdobnov, E. M. and R. Apweiler. 2001. InterProScan -an integration platform for the signature-recognition methods in InterPro. Bioinformatics 17:847-848. Zhang, A., S. Altuvia, A. Tiwari, L. Argaman, R. Hengge-Aronis, and G. Storz. 1998. The OxyS regulatory RNA represses rpoS translation and binds the Hfq (HF-I) protein. EMBO J. 17:6061-6068. Zhang, F., J. A. Sheps, and V. Ling. 1998. Structure-function analysis of hemolysin B. Meth. Enzymol. 292:51-66. Zhang, J.-H., N. B. Quigley, and D. C. Gross. 1997. Analysis of the syrP gene, which regulates syringomycin synthesis by Pseudomonas syringae pv. syringae. Appi. Environ. Microbiol. 63:2771-2778. Zuker, M., D. H. Mathews, and D. H. Turner. 1999. Algorithms and thermodynamics for RNA secondary structure prediction: a practical guide. p. 11-43. Barciszewski, J. and B. F. C. Clark (eds.), RNA biochemistry and biotechnology. NATO ASI Series, Kiuwer Academic Publishers. 187 APPENDICES 188 Appendix 1. Overexpression of PItR in Escherichia coli Al.!. Introduction. In Pf-5 and related Pseudomonas species, complex regulatory cascades govern the production of secondary metabolites. Participants in such regulatory cascades that affect the production of pyoluteorin (PLT) were identified by transposon mutagenesis (6) and have been found to encode proteins with global roles in regulation (5, 14, 18, 19). Biosynthetic genes have also been identified by transposon mutagenesis (6). Sequencing of the region revealed the PLT biosynthetic genes to be clustered (11, 12). Sequencing of the region also revealed a pathway-linked transcriptional regulator. The gene, designated pltR, encodes a LysR transcriptional regulator, and is necessary for the transcription of PLT biosynthetic genes from all known promoters (11). LysR homologues are typically associated with, and activate transcription from, regulons responsible for secondary metabolite biosynthesis. Many LysR proteins enhance transcription from their own promoters. Many LysR proteins also are activated by cofactor binding, and the cofactor is often a compound whose synthesis is regulated by the LysR protein. Thus, the binding of cofactor can indirectly and simultaneously induce transcription of genes for cofactor biosynthesis and the cofactor's receptor. This scenario implies amplification of the cofactor and the LysR transcriptional activator via autoregulation (15). I found that a Pf-5 derivative bearing a mutation in pltR was unable to produce PLT, and, therefore, also unable to undergo PLT autoinduction (data not shown). Thus, 1 hypothesized that P1tR could act as the receptor for PLT during autoinduction (Fig. A1.1). transcription PLT . . PItR pltR I -+$_ pltR pitA pltL pltR pltL PItM orfN 4_9 pltJ rn pltB pitC pltD pit? pita Fig. A1.1. Proposal for P1tRIPLT interaction at PLT promoters. In this scenario, P1tR does not activate transcription when not bound to cofactor (PLT). However, cofactor (PLT) binding is thought to induce P1tR activation, perhaps by a conformational change in the protein. In P. fluorescens strain CHAO, 2,4-diacetylphloroglucinol (2,4-DAPG) autoinduction and repression by PLT, salicylate and a fungal metabolite, fusaric acid, are mediated by PhIF, a TetR type transcriptional regulator encoded by a gene linked to the 2,4-DAPG biosynthetic gene cluster (16). Binding by Ph1F to the ph1ACBD promoter inhibits transcription; PhlF binding is directly stabilized by salicylate but alleviated by 2,4-DAPG (1). Other examples of dual regulation of a single transcriptional regulatory protein by two metabolites also exist. For example, BenM, a LysR-type transcriptional regulator from Acinetobacter sp. ADP1, regulates genes for aromatic compound degradation. BenM is activated by synergistic binding of benzoate, an aromatic compound, together with a breakdown product, cis,cis- muconate (4). Based on such reports, I further hypothesized that PItR could act as a receptor for both PLT (as a co-activator) and for 2,4-DAPG (as a co-repressor). If this were the case, PltR would mediate the positive effects of exogenous PLT and the negative effects of exogenous 2,4-DAPG at PLT promoters (Fig. Al .2). 190 O PItR° pitH j-. ro° #} .. piti pltL .-f-4- orfN . llvu.I1lIuI pitR pitA p1W pltC P1W or K pltF pito Fig. A1.2. Proposal for dual binding by PItR of PLT (co-activator) and 2,4DAPG (co-repressor). In this scenario, PItR does not activate transcription when bound to corepressor (2,4-DAPG). However, PLT binding may activate PItR, perhaps by a conformational change in the protein. P1tR is an attractive focus of study because, in contrast to several global regulatory proteins studied in our laboratory that affect PLT production but have undesirable pleiotropic effects (5, 14, 18, 19), P1tR bears potential to act as the final member of regulatory cascades affecting PLT production. Therefore, PltR was envisioned as a target protein that might be manipulated to overproduce PLT without affecting other aspects of the physiology of Pf-5. Because PLT at high concentrations is phytotoxic (9), I also reasoned that P1tR might eventually be engineered to respond to a specific environmental signal such as Pythium exudates, so that PLT could be selectively produced in the rhizosphere as appropriate for biocontrol. A1.2. Characterization of P1tR PltR had been characterized previously by mutating the pltR gene and evaluating the phenotype of the mutant strain (11). The obvious phenotype was a complete lack of transcription of PLT biosynthetic genes and resultant PLT production. Overexpression of PltR in Pf-5 was also attempted. A 2.2 kb KpnIHin dill fragment containing the entire pltR gene was cloned into pME6000 (Brian Nowak-Thompson, personal communication), an Escherichia coli -Pseudomonas 191 shuttle vector of the IncP incompatibility group, stable in P. fluorescens at a copy number of 16-20 (Stephan Heeb, persona! communication). Mating this plasmid into Pf-5 or Pf-5 pltR strains was unsuccessful, as colonies displayed unstable phenotypes and became sectored on plates (Brian Nowak-Thompson, personal communication, M. Brodhagen, unpublished data). When cultured in nutrient broth + 1% glycerol, a medium conducive to PLT production, the complemented strains produced no PLT or produced only trace amounts (data not shown), although they presumably carried multiple copies of a transcriptional activator for PLT biosynthetic genes. Moreover, nearly half of the strains I tested appeared to have acquired mutations in the gacS/gacA two component regulatory genes during storage. Thus, I reasoned that overproduction of P1tR in Pf-5 was not a favorable route to overproduce PLT. However, I continued to characterize PItR as a potential model of pathway-specific regulation and possible eventual target for protein engineering. Two separate strategies were attempted in order to further characterize PltR. These involved 1) attempts to generate a transcriptional fusion to pltR in order to test various factors which might upregulate or repress its expression, and ii) attempts to obtain pure P1tR protein. The goal of the latter work was to use PItR in protein-DNA studies to identify P!tR binding sites. A second goal was to add PLT and 2,4-DAPG to pure P1tR in DNA-binding assays, to test their potential as effector molecules by assessing their ability to alter binding of PItR to its target promoter(s). Al.2.1. A transcriptional reporter for pltR The pltR gene was amplified using PCR, with primers designed to place EcoRI and XbaI restriction sites at the 5' and 3' ends, respectively, of the open reading frame. The primers pltREcoR-F (gtctagaatcctcaataagaatgcact) and pltRXba-R (ggtctagatcattgacaacgcctag) were used to amplify a 475 bp fragment containing the pltR promoter region. The fragment was then cloned into pMP22O, a broad host-range plasmid of the IncP incompatibility group (17). The pMP22O plasmid contains a 192 promoterless lacZ gene downstream of the multiple cloning site, and thepltR promoter was expected to drive its expression. The promoter probe vector containing the pltR promoter mated into Pf-5 and derivatives, and tested for lacZ expression. Disappointingly, Pf-5 containing the pltR promoter probe exhibited no difference in 13galactosidase activity, as determined by standard assays spectrophotometrically measuring the conversion of o-nitrophenyl--D-galactopyranoside into the yellow onitrophenol (10), when grown on PLT-conducive medium (nutrient broth + 1% glycerol) as opposed to PLT-repressive medium (nutrient broth + 1% glucose). However, this was not entirely unexpected. The expression of PLT biosynthetic genes had previously been measured using transposon insertion mutants bearing the promoterless ice nucleation reporter gene (inaZ). The inaZ reporter system is sensitive enough to detect expression of several transcripts per cell. In contrast, the lacZ system is 100,000 times less sensitive. Thus, genes that have low expression levels may be detected using the inaZ system while escaping detection by the lacZ system (8). PLT biosynthetic genes were within the detection range of the inaZ reporter gene, and recent DNA hybridization analysis has confirmed that the pltR transcript and the PLT biosynthetic gene transcripts are present in low abundance when compared to those of other genes in Pf-5 (Caroline Press, personal communication). Interestingly, the genes for 2,4-DAPG biosynthesis presented the opposite problem. Whereas the highly expressed phiA gene from P. fluorescens CHAO was detectable using the lacZ reporter gene, aphlA::inaZ fusion was expressed so highly that detection was never below the saturation limit (Regina Notz, personal communication). Future experiments to assess transcriptional activity of pltR might take advantage of real time RT-PCR technology, which can measure transcript abundance directly without the technical difficulties associated with reporter genes. 193 A1.22. Overexpressjon of PItR in E. coli A1.2.2.1. Design of a PltR peptide antibody The protein expression systems (pTYB 1, pET1 5b) described below allow the generation of fusion proteins, so that the fused affinity tag can be used in protein purification. Both expression systems also allow the subsequent cleavage of the affinity tag, leaving the native or near-native protein. Native protein was desirable in assays to examine the DNA-binding and cofactor-binding activities of PItR, because a native protein was more likely to retain such activities than a fusion protein. However, subsequent detection of the PItR protein in Western analysis, or "supershifts" in electromobility shift assays, required an antibody. Unfortunately, cleavage of the affinity tag simultaneously cleaves a useful antigen for readily available commercial antibodies. Therefore, an antibody to P1tR was needed. A synthetic peptide comprising the final 17 amino acids (CSALSRATTLQDKSQTS) of the predicted PItR protein was synthesized, and used to raise polyclonal antibodies in rabbits (Alpha Diagnostic International, San Antonio, TX). The C-terminal region of P1tR was chosen because of its hydrophilicity (implying surface-exposed residues), antigenicity, and accessibility (as assessed by Alpha Diagnostics International, San Antonio, TX). As is often the case, the extreme C-terminus represented the most attractive choice for the generation of a peptide antibody, as it is highly antigenic [e.g. Liang et al. (7)]. Antibodies were raised in two separate rabbits (7074 and 7075) to the PItR peptide conjugated to the highly immunogenic keyhole limpet hemocyanin carrier protein, and were affinity purified by passing the rabbit serum over sepharose beads to which the PltR peptide had been coupled. Purified antisera were tested for activity by standard ELISA procedures; antibody titer for antisera from both rabbits was 1:100,000. 194 A1.2.2.2. Expression ofP1tR inpTYB1 The pltR gene was PCR-amplified with NruI and NdeI restriction sites engineered at the 5' and 3' ends, respectively, of the gene. The following primers were used: pltR_Nde (ggatttacatatgaaggcgctaggcg) and pltR_Nru (tactgattcgcgatgtctggctc). The PCR-amplifiedpltR gene was cloned in-frame into the pTYB1 vector digested with NdeI and Sapi. The resultant construct contained the pItR open reading frame, flanked by a C-terminal fusion to an intein, followed in turn by a chitin-binding domain (CBD) that served as an affinity tag for purification. The integrity of two constructs was verified by sequence analysis (Central Services Laboratory, Oregon State University). The final construct was moved into the expression strain, E. coli ER2566. Protein expression in the pTYB1/ER2566 system (IMPACTTh1-CN, New England Biolabs, Beverly, MA) is driven by a promoter containing the T7 polymerase recognition sequence. A chromosomal copy of bacteriophage 17 RNA polymerase is carried by expression hosts. The 17 RNA polymerase gene is under the control of the lacUV5 promoter. Thus, expression of the T7 RNA polymerase is repressed by Lad until IPTG is added. However, a low level of basal expression remains, which can be eliminated in expression hosts carrying the pLysS plasmid, as discussed below. The T7 polymerase is highly processive and highly specific. It recognizes a small (20 bp) promoter sequence that E. coli RNA polymerase does not recognize. Thus, a target gene behind the T7 promoter is not transcribed until the 17 polymerase gene expression is induced by IPTG addition. Premature expression of the target gene itself, like that of T7 polymerase, is controlled by a lac operator sequence within the promoter; the Lad repressor protein is encoded elsewhere in the pTYB 1 vector. After expression, the fusion protein was affinity-purified from lysed cells by passing over chitin-sepharose beads, which bound the chitin-binding domain tag. Addition of thiols [50 mM dithiothreitol (DTT)} to the resin-bound fusion protein activated the intein, causing on-colunm cleavage of the fusion protein between the last amino acid residue of P1tR and the first amino acid residue of the intein. Thus, native P1tR was released from the column. 195 A1.2.2.3. Expression of FuR in pETl5b The pitR open reading frame was also cloned into pETl5b, a vector that, like pTYB1, drives expression of the target gene from the T7 promoter and controls leaky expression with the Lad repressor encoded on the same plasmid. A fragment containing the p/zR open reading frame was amplified using the primers pltR_Nde (ggatttacatatgaaggcgctaggcg) and pitR_BamHI_1 416 (caggatcccgaaccgacgtgctg), which created NdeI and BamHI sites at the 5' and 3' ends, respectively, of the gene. The pltR gene was then cloned in-frame into pET1 5b digested with NdeI and BamHI, creating a translational fusion to an N-terminal 6X-His tag separated from PItR by a ten amino acid thrombin cleavage site. Although lack of success with this system prevented the purification of large amounts of His-tagged PItR, the goal of this cloning strategy had been to obtain a 6X-His affinity tagged PItR fusion protein, which could be purified from crude cell lysate by passing over a nickel resin, which binds to the His residues. Near-native PItR, containing one extra His residue at the N-terminus, could then be obtained by cleavage from the affinity tag using thrombin. A1.2.2.4. Host strain E. coli BL21(DE3)pLysS E. coil "B" strains are more robust than K-12 derivatives generally used in molecular biology. Strain BL21 is deficient in proteases Lon and OmpT. The pLysS plasmid carried by the expression hosts E. coil BL21(DE3)pLysS and the Rosetta strains (see below) encodes T7 lysozyme, which degrades the T7 polymerase at a low rate and thus prevents basal transcription of the target gene from the T7 promoter until intentional expression is induced by IPTG addition. After induction, expression rates are high and easily overpower the effect of the T7 lysozyme. 196 A1.2.2.5. Host strain E. coil B ER2566 E. coli B ER2566 (New England Biolabs, Beverly, MA) is a derivative of E. co/i BL2 1. Like BL2 1, it carries a chromosomal copy of the T7 polymerase preceded by the lacUV5 promoter, and lacks proteases Lon and OmpT. However, this strain does not carry the pLysS plasmid; therefore, control over leaky expression of the target gene is less stringent. A1.2.2.6. Host strain Rosetta The Rosetta strain is also a BL21 derivative. It carries a pLysS plasmid for stringent control of target gene expression. In addition, it carries plasmid-bome genes for rare tRNAs. The predicted PItR protein contains codons for which the following complementary tRNAs are not frequently found in E. co/i: cua (Leu), cga, aga, and agg (Arg). TheE. coli codon frequencies for these were < 5%; a total of 3.2% of the amino acids in PItR are encoded by rare E. coli codons. Thus, codon usage was predicted to be a factor in the low rates of P1tR expression (described below) in E. coli B ER2566 and E. co/i BL21(DE3)pLysS. Expression different conditions A1.2.2. 7. ofP1tR in djfferent vector - host combinations and under The expression of P1tR was attempted first in the IMPACT-CN system described above. Overnight 5 ml cultures grown at 37°C, with shaking at 200 rpm, in Luria broth (LB) (13) containing 100 pg/mi ampicillin (Ap'°°) were used to inoculate 100 mis LB Ap10° in 500 ml Erlenmeyer flasks. Experimental cultures were grown with shaking at 200 rpm to an OD of 0.6, and then induced with 0.3 mM IPTG. Following induction, cultures were grown under several different combinations of induction times and temperatures. These conditions and the results are summarized in Table Al. 1. PltR was harvested from cells using the protocol recommended in the manufacturer's instruction manual (IMPACTTM-CN, New England Biolabs, Beverly, 197 MA). Because the predicted p1 of PItR lies between 8.0 and 9.0, buffers were made up at pH 6.0, to keep P1tR in solution. After the indicated times, cells were harvested by centrifugation, resuspended in lysing buffer, and sonicated to break open cells. Cell lysates were clarified by centrifugation, and the clarified lysate was loaded onto preequilibrated columns containing a sepharose-chitin affinity resin. Once loaded, the bound proteins were cleaved by activation of the intein in 50 mM DTT. The liberated, native P1tR was eluted. Fractions collected at each step of the protein purification process were separated using SDS-polyacrylamide gel electrophoresis (SDS-PAGE) and visualized by Coomassie staining. Further examination of the fractions was done via Western analysis, using an antibody to the chitin binding domain of the fusion protein (New England Biolabs, Beverly, MA), and the PltR peptide antibody. The expected sizes of the various protein products were as follows: P1tR'-'intein'-'CBD 92 kDa intein'-'CBD 55 kDa PItR 38 kDa. None of these protein products was seen in fractions from uninduced cultures. The 92 kDa fusion protein was not seen on Coomassie-stained gels due to overloading (data not shown), but a band corresponding to the 92 kDa fusion protein was detected with the PltR peptide antibody in fractions of crude cell lysate and clarified lysate. In fractions eluted from the sepharose-chitin bead resin, a 38 kDa band corresponding to the expected size of native P1tR was detected with the PItR peptide antibody, and a 55 kDa band corresponding to the expected size of the residual intein'-'CBD was detected with the CBD antibody in fractions eluted from the sepharose-chitin bead resin. Yield of PltR was low in all treatments, but was optimal at 15°C and 20°C. Table Al.!. Expression of PItR° from pTYB1 in E. coli B ER2566 cells, as visualized by immunoblotting. Temperature Harvest time 15°C 15°C 20°C 20°C 27°C (hours post-induction) 6 hours l6hours +1- + +1- + 'Phis signs designate expression of PltR, as visualized by Western analysis. Minus signs denote the absence of protein. +1- indicates that PItR was observed via Western analysis but in nearly undetectable quantities. Expression of P1tR in pTYB1/ER2566 was repeated, using the same culture and extraction conditions as described above except that 2xYT (13) was used as the culture medium in place of LB. Because yields were low in the first experiment, protein degradation was suspected and cultures were harvested sooner after induction (3.5 and 6 hours). Purification of P1tR from cells as described above yielded approximately 0.4 j.tg of protein per ml of culture. Fractions eluted from the chitin resin were corrected (based on OD6 of cells prior to lysis) by dilution to obtain an approximately equivalent amount of total protein in each sample. Protein samples were separated by SDS-PAGE and visualized by Coomassie staining. Although a band corresponding to the predicted 38 kDa PltR protein was seen, another band of 55 kDa was also seen, indicating that fractions were contaminated with the intein!CBD tag. Because protein yield was poor, the vectors containing pltR were expressed in two other hosts, under a variety of conditions. The results are summarized in Table Al .2, below. 199 Table A1.2. Expression° of PItR in various host/vector combinations's. Temperature Expression host. E. coli BL21 (DE3)pLysS Expression vector pTYB1 Harvest timec Rosetta pTYB1 pETl5b - 6 hours 19 hours pETl5b 15°C 20°C 27°C 30°C 37°C 3 hours 6 hours 19 hours 1 hour 2hours 3hours 4hours Shours 6hours + - + + + * * + * + * + +: + * + * + * + * + : + * + * + * + * + +: + * + * + * + * + +: + * + * + * + * + apis signs designate expression of P1tR, as visualized byCoomassie staining. Minus signs denote the absence of protein. bThe results indicated with asterisks were verified by immunoblotting. dHOS after inducing with 1mM IPTG. Western analysis of proteins expressed from pETl5b was done using an anti- His antibody. No His-tagged protein was visible in E. coli BL21(DE3)pLysS cells. Cell lysates from Rosetta cells carrying pltR in the pET 1 5b vector were separated on SD S-PAGE and visualized with immtrnoblotting. A 38 kDa band that reacted to the anti-His antibody was visible. However, a smear of smaller degradation products was also apparent, increasing in intensity with increasing temperature. The only sample lacking these degradation products were those from cells grown at 15°C for 1 or 2 hours. Protein yield in none of the samples described in this chapter exceeded 0.4 .ig protein per ml of culture. Thus, I did not obtain sufficient quantities of pure PItR to perform DNA-binding and cofactor-binding studies with pure protein. Due to time constraints, the experiments were abandoned for the time being. 200 A1.3. Future recommendations. From the experiments described above, it appears that PItR is not stable when overexpressed in either Pf-5 or E. coli. It is logical from previous studies (19) to hypothesize that PItR may be a target of Lon protease. Because the expression hosts used in this study were all deficient in Lon, it appears that PltR suffers from proteolysis by other routes. Expression in a host with fewer proteases might be useful, although such hosts are not robust and are difficult to culture. In some instances, expression of a target protein with a protein affinity tag is more stable than one that is only tagged with 6xHis, because the larger protein can stabilize the target protein and make it less susceptible to proteolysis. However, the PltR'-'intein'-'CBD fusion protein, although stable in E. co/i B ER2566, was not expressed in E. co/i BL21(DE3)pLysS or the Rosetta strain, so it is unlikely that expression could be improved with a larger affinity tag. An expression vector resulting in a 6xHis tagged target protein, and suitable for use in Pseudomonas, exists (3). The expression of PItR could be attempted using this vector expressed in Pf-5. In that case, proper coclon usage, correct folding, and the availability of necessary prosthetic groups/cofactors would be guaranteed from the parent strain. Because the vectors are inducible through the lac promoter, and the expression vector carries the lad repressor gene, overexpression is controlled, and the unsatisfactory results previously obtained from overexpressing P1tR from pME6000 (described above) might be circumvented using these vectors (pLAH3O, pLAH31, and pLAH32). It is also possible to perform DNA-binding studies using bacterial cell lysates rather than purified proteins. This strategy has been successfully used in both electromobility shift assays (EMSAs) and in DNAse I footprinting studies (2). Cofactor binding during DNAse I footprinting analysis has also been done using crude 201 bacterial cell lysates, resulting in the successful identification of both benzoate and muconate as cofactors for the LysR type transcriptional regulator BenM (Becky Bundy, personal communication). Future attempts to determine the site of PltR binding, and to determine whether PLT and 2,4-DAPG might act as co-effectors, are expected to provide insights about mechanisms by which Pseudomonas antibiotics might exert their intracellular regulatory influences. Al .4. Acknowledgements I am grateful to Dr. Linda Thomashow for inviting me to perform the PItR expression studies in her laboratory, and to members of Dr. Thomashows lab for invaluable assistance and advice. I am especially grateful to Dr. Dmitri Mavrodi and Greg Phillips, who were my mentors in protein expression. 202 Al .5. References Abbas, A., J. P. Morrissey, P. Carnicero Marquez, M. M. Sheehan, I. IL Delany, and F. O'Gara. 2002. Characterization of interactions between the transcriptional repressor PhIF and its binding site at the phiA promoter in Pseudomonasfluorescens Fl 13. J. Bacteriol. 184:3008-3016. 2. Ausubel., F., R. Brent, R. E. Kingston, D. E. Moore, J. G. Seidman, J. A. Smith, and K. Struhi (eds.). 1998. Current Protocols in Molecular Biology. Wiley and Sons, Inc. New York, NY. 3. Bertani, 1., G. Devescovi, and V. Venturi. 1999. Controlled specific expression and purification of 6xllis-tagged proteins for Pseudomonas. FEMS Microbiol. Lett. 179:101-106. 4. Bundy, B. M., L. S. Collier, T. R. Hoover, and E. L. Neidle. 2002. Synergistic transcriptional activation by one regulatory protein in response to two metabolites. Proc. Nati. Acad. Sci. USA 99:7693-7698. 5. Corbell, N., and J. E. Loper. 1995. A global regulator of secondary metabolite production in Pseudomonasfiuorescens Pf-5. J. Bacteriol. 177:6230-6236. 6. Kraus, J., and J.E. Loper. 1992. Lack of evidence for a role of antifungal metabolite production by Pseudomonasfluorescens Pf-5 in the biological control of Pythium damping-off of cucumber. Phytopathology 82:264-271. 7. Liang, T. C., W. Luo, J.-T. Hsieh, and S.-H. Lin. 1996. Antibody binding to a peptide but not the whole protein by recognition of the Cterminal carboxy group. Arch. Biochem. Biophys. 329:208-2 14. 8. Lindgren, P. B., R. Frederick, A. G. Govindarajan, N. J. Panopoulos, B. J. Staskawicz, and S. E. Lindow. 1989. An ice nucleation reporter gene system: identification of inducible pathogenicity genes in Pseudomonas syringae pv. phaseolicola. EMBO J. 8:1291-1301. 9. Maurhofer, M., C. Keel., U. Schnider, C. Voisard, D. Haas, and G. Défago. 1992. Influence of enhanced antibiotic production in Pseudomonasfluorescens strain CHAO on its disease suppressive capacity. Phytopathology 82:190-195. 10. Miller, J. H. 1992. A short course in bacterial genetics: A laboratory manual and handbook for Escherichia coli and related bacteria. Cold Spring Harbor Laboratory Press, Plainview, NY. 203 11. Nowak-Thompson, B., N. Chaney, S.J. Gould, and J.E. Loper. 1999. Characterization of the pyoluteorin biosynthetic gene cluster of Pseudomonasfluorescens Pf-5. Journal of Bacteriology 181 (7):2 166-1274. 12. Nowak-Thompson, B., S. J. Gould, and J. E. Loper. 1997. Identification and sequence analysis of the genes encoding a polyketide synthase required for pyoluteorin biosynthesis in Pseudomonasfluorescens Pf-5. Gene 204:17-24. 13. Sambrook, J. and D. W. Russell. 2001. Molecular Cloning: A Laboratory Manual. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY. 14. Sarniguet, A., J. Kraus, M. D. Ilenkels, A. M. Muehichen, and J. E. Loper. 1995. The sigma factor o affects antibiotic production and biological control activity of Pseudomonasfluorescens Pf-5. Proc. Natl. Acad. Sci. USA 92:1225-12259. 15. Schell, M.A. 1993. Molecular biology of the LysR family of transcriptional regulators. Annual Review of Microbiology 47:597-626. 16. Schnider-Keel, U., A. Seematter, M. Maurhofer, C. Blumer, B. Duffy, C. Gigot-Gonnefoy, C. Reimmann, R. Notz, G. Défago, D. Haas, and C. Keel. 2000. Autoinduction of 2,4- diacetyiphioroglucinol biosynthesis in the biocontrol agent Pseudomonasfluorescens CHAO and repression by the bacterial metabolites salicylate and pyoluteorin. J. Bacteriol. 182:12 151225. 17. Spaink, H. P., R. J. H. Okker, C. A. Wijffelman, E. Pees, and B. J. J. Lugtenberg. 1987. Promoters in the nodulation region of the Rhizobium leguminosarum Sym plasmid pRL1J1. Plant Mo!. Biol. 9:27-39. 18. Whistler, C. A., N. A. Corbell, A. Sarniguet, W. Ream, and J. E. Loper. 1998. The two-component regulators GacS and GacA influence accumulation of the stationary-phase sigma factor c? and the stress response in Pseudomonasfluorescens pf-5. J. Bacterioll. 180:6635-6641. 19. Whistler, C. A., V. 0. Stockwell, and J. E. Loper. 2000. Lon protease influences antibiotic production and UV tolerance of Pseudomonas fluorescens Pf-5. Appi. Environ. Microbiol. 66:2718-2725. 204 Appendix 2. Nucleotide and Amino Acid Sequence for pitH, pill, pit.!, or/K, or/F, orfQ, or/U, orJX, and orfY. A2.1. Predicted amino acid sequences and positions of open reading frames in nucleotide sequence The deduced amino acid sequences for pitH, pit!, pit.J, orfK, orJP, orfQ, orfU, orJX, and orfY are listed below. All genes are transcribed in the same direction as p1tLABCDEFG unless noted as "reverse complement". The numbering scheme for nucleotide sequences is continued from that of published pyoluteorin biosynthetic gene region (accession number AFO8 1920) for convenience. Asterisks designate stop codons. Where sequence information was taken from the incomplete Pf-5 genome sequenced by The Institute for Genomic Research, the open reading frame designation is preceded by "TIGR". pitH, reverse complement Nucleotide 23,340-22,705. An alternative start codon allows an open reading frame from nucleotide 23,376 to nucleotide 22,705; the resultant additional amino acids are underlined. MKQPPAQTTPNIM!JTRRRTSRSDGEHTKIRILEVAARLFAQHGYANTASK LICEEAGADLAAINYHFGSREALYKAVLIEGHKQLVSFEALSQLAQSEEP ALIKLDSFIDAIVTRVLDEQSWQSKVCAREILAPTVHFTSLVQEEVMPKF RLLEALISEITGFPIGDPALARCTISIIAPCLMLA\TI]DRQQPSPLQAVLQ HDZNALIKAHFKLFARSGLAAIAQ * 205 piti Nucleotide 23,452 to 24,465 MKKQLIVGVAVLLVATAGVSWFLLRPEKQNDHLKLHGNVDIRQVSLAFDG SERIAALYAEEGDLVQPGQVLAEIJDTRTLRLEINRSKARIGAQEQALLRL KNGTRPQEVEQSKARFDAAQAQMQLAQLHMQRLRRIADDTQGKGVSQQRL DQAAARLQVAKAQSQEQRESWTLAKIGPRNEDIAQATADLQASRDLDLL EHYLABAQLKAPTQARIRTRLLEPGDMS PQRPVFALALT]DPKWVRAYVN ERQLGHIRADQMARVYTDSFPDQAIDGKVGYI SSVAEFTPKSVETEDLRT SLIVYEIRVLVKDPDDRLRLGMPATVYL]DQAPVAGATP * pltJ Nucleotide 24,462 to 26,231 MSAPAVTVQAVELVKRFRSGTREQLALEQVSLEVPSGKLSALVGPDGAGK TTLLRMIAGLLKADEGLLKVLGLD\TAADPQQVQDL ISYMPQKFGLYEDLT I QENLELYADLHGVRAQQREERFSRLIJQMMDLTRFRDRIJAGQLSGGMKQK LGLACTLVRS PQLLLLDEPTVGVDPLSRRELWSIIEQLIEQENLTVLI ST AYMDEAQRCAQVFVLYQGRLIAQGPPAQLCIjLAQGLCYGVGPEAGTPARL LQARLIJDDNEHI I DAVPEGGMVRFIRSAQVPI1QAI PSLAGQAS PQALEAR LEDGFMVLLRQQVGSVASLDEPAQAPASYGQG]DAASGPAVVI EVRDLVRK FGDFTAVASTSFEVLKGEIFGLLGPNGAGKTTTFRMLCGLLPAS SGYLEV AGVNLRTARAAARRKIGYVSQKFALYANLSALENLRFFGGAYGLHGAKLK ARVALMLEQFDLDEHKHLRSGELPGGYKQRLAMAVALLHE PQILFLDEPT SGIDPLARRAFWRRITALAQSGTTIVITTHFMEEAEYCDRIVI QDAGRLL ALGTPEQVRHQGSESLADMNSAFIAVVERARAQASAPVQ * TIGR orJK Nucleotide 26,235-27,384 MTMTDSNRSASGFRRRLIALTRKELRQLMRDKSNLAIGIVLPIVLILIFG YGLSLDIRNTPLAVVLEDSS PSARSVVAKLASSDYFSVVQVTSMAEARAIi MDKRRVDGIVRVPADFSRRLEQQDARLQLLLHGADANSAATLGRYVRGAL NVWQQQQFDRSRQPQSAGRVLVVERTyrWFNAANS STWYLVPGLIALIMTLV GAFLTSLVLAREWERGTLESLFVTPVRSIEILLAKII PYFLVGLILGLVNC LVSARLLFEVPIQGSLVLLLLSSMLYLLVTLGIGLLI SAKTRNQFLLASQI All FSFLPALMLSGFLFDLR11VPTFIRVVGSI LPATYFMELVKTLIFLAGN NWPLIAKNLAILAAYA\TFLLNAARLCTRKKLD * TIGR orfp Nucleotide 27,389-28,507 MDKLJMJFLRELLFLFQKEFLAIVKDPANRVILIAPAIVQSLLFGYGATYD LNYVPYAVLDQSQGRASTELL,ARFDASGVFARVSTLTSVDEIAPVIDRGS ALJLVLHIAADFDERLSRHENAGLQLILDGRNSTTAGMAAGHIANLVADFN QQFLAVDNAPVRLETPAWFNPNLQTRWNIVPGLIAALSMIQTLMLAALSV AREREQGTF]DQLLVTPLTPFVILLGKALPSVLIGLLQSSLILLIGLFWFK I PMIGSVLDLYLGLFVFTSACVGIGLSI SALSANMQQANVYTFVLI4NPLI LLSGLITPVHSMPEALQLLTYLDPLRFAIDLVRRIYLEGATLADVSWNLV PMIJAVALVTLPLAAWLFRNRLV* TIGR orfQ Nucleotide 28,519-30,016 MKTTHSGYLRNGLTLANLLAITACTVGPDFKAPAVTS PERWTDWHSGPPA S SVPGKAPATTSQTADGNWWQVFGDPVLDQLQAQVREGSPDLHTALLRFA QVRLQRQIVASLETPEVSFSAP.ASRNRQSRYAPNNRMLEALGSDSDALEK VLTDPYSLYQAGFDVSWELDLWGRVSRLGEAAQAEVDGAAATLDDVLLSV SSELARNYFEVRTAQRQTRLLEQEAQI LAkHLQWAVQAREGEEDGFAVE RQEARLASLHAQI PGWKALHTQAGNRIALLLGEHPGALETLLAERPDNAL SRPLPNLQLGLPGDLARERPDIRAAQARLHRATAGIGVAEAELYPSVNLG ADFGFESYKSGQFADWSSRTWSVGpRLDLPLFDRGRRSKTVVLRTLEQQE AAVAFHRTVLAAWQEVDDAMSRYHNEYQRAHLKDSYDSKARTYEWTRVR YAAGEASYLEELEAQRTVLEAQPDIJVNSDSQLRTHLI SIYKSMGGHST * TIGR o,-JU Nucleotide 30,622-31,227 MQMDDLYKIALFS SLILVVPGPSNTLL.LASGFKFGVLRSLPLVLI EMLGY TVS ICTWGWGLVLLSHDYPWLIHLIKLLCALYIAALALKTWKTSTTGAQH TERFE SWPRHLLIATLLNPKGLIFASVIFPAS SFASLGDFLP SLSAFFLL LAPIALLWVSLGAGIRAQYLGRVSGPL,FSRGTAVVLALFSATLTYTVVRA v* 207 TIGR orJX reverse complement Nucleotide 32,065-33,210 MRTPGTAAFS STTRTSWPRSGRSLTTEPHDSNPTQAPPTPI PTDKSRTTM KAFLIDRYGKNSGRMGEAPTPTLGAHDVLIEVHAS SVNVLDSKIBNGEFK LILPYALPLVLGNDLAGVVVEVGPQVQRFKPGPEVYARPPETRIGTFAQW IAVNENAIALKPANTTMTEAAS I PLVALTAWQVLVETAQLKKGQKVLIHZ GAGGVGTL,AI QLAKHLGAFVATTTSThNVEWVKALGADQVIDYKQQNFES L1LHGYDLVLNSLGSDVLEKSLKVLKPGGQLI SI SGPPTVQFAQEQGLAWP LQQVMRLLSFGIRRKARQRDVRYAFVFMIADGAQLQQITALI EAGI IKPV VDRSF SFESTAEALQYVEQGRTKGKVVVRIK * TIGR orfY reverse complement Nucleotide 33,122-34,000 MIWTQTTFPPALQTSFINAANQSIMVRGI PFAYRDTGPAGGVPLVLFNHW GAVLDNFDPAI IDGLAQTRRVITTDYRGIGGSGGTAPLJTVGDMAQDAIEL IQALGFKSVDILGFSLGGFVAQDIALQAPQLVHRLILTGTGPAGGIGIDI( VGSVTWPLMLIKGLLTLRDPKYYLFFTSTPHGRRAASEYLQRLKARSKHRD KGPTPGAFLRQLQAITAWGKQAPQDLGRLIRAPTLIVNGDN]DIMVPSANSH ALAKRVPNAQLVMYEDAGHGGI FQYHADFVAKVRAFLDDRAP * A2.2. Relevant nucleotide sequence The nucleotide sequence for the region including pitH, pitl, pit.1, orJK, orJP, orfQ, orftf, orJX, and orfY is displayed below. For convenient reference, the numbering sequence is continued from that of published pyoluteorin biosynthetic gene region (accession number AFO8 1920). Sequence information derived from this thesis is highlighted in bolded type. Sequence information generously provided by Ian Paulsen (The Institute for Genomic Research, Rockville, MD) comprises the remaining sequence for the section, in normal typeface. 22701 22751 22801 22 851 22901 22951 23001 23051 23101 23151 23201 23251 23301 23351 23401 23451 23501 23551 23601 23651 23701 23751 23801 23851 23901 CCGACTATTG GGCATGGCC GCGAGGCCGG ACCGGCGAA CAGCTTGAAG TGGGCTTTCA GGGCGTTGGC GTCGTGTGC AGG&CGGCT GCkGCGGGCT GGGCTGC!GG CGACTATCA CCGCCAGCAT CAGGCAGGGC GCGATGATGC TGATGGTGCA ACGCGCCAGC GCCGGGTCAC CGATCGGAAA GCCGGTGTC TCGCTGAGA GCGCTCCAG CAGACGGAAT TTGGGATCA CTTCTTCTTG CACAAGGCTG GTGAAkTGGA CAGTGGGCGC CAGTATTTCC CGTGCACAGA CTTTGCTTG CCAGCTCTGC TCATCGGCA CCCGGGTGAC GATGGCATCG ATGAAGCTAT CGAGCTTGAT CAGCGCCGGC TCTTCGCTT GCGCCAGCTG CGACAGGGCT TCGAAGCCA CAGCTGCTT GTGTCCCTCG ATCAGCACCG CCTTGTACAG GGCTTCCCGG CTACCGAGT GGTAGTTGAT GGCGGCGAGG TCGGCGCCGG CTTCTTCACA GATGGCTTG CTCGCGGTGT TGGCATAGCC ATGCTGGGCA AACAGGCGGG CAGCCACTTC GAGAATGCGG ATTTTAGTGT GTTCGCCGTC GCTTCTGCTG GTTCTGCGTC GCGTGTTCAT GATGTTCcTC GTGGTCTGAG CGGGGGGTTG CTTCATGGCC AGTATCTGTT ATTTCTAT TTAA.ATTCAA ATTGAATTTT AkTTAGGCCT TGGCAAACA AGGACTCATG GATGAAGAAG CAGTGTTG TAGGGGTTGC CGTACTGCTG GTGGCCACAG CGGGTGTTTC CTGGTTTCTC CTTCGCCCCG AGAAGCAGA CGAecACCTC AAGCTCCATG GCAACGTCGA CATTCGCCAG GTGTCACTGG CGTTCGACGG CAGTGAGCGA ATCGCCGCGC TGTATGCCGA AGAGGCGAC CTGGGCAGC CAGGCCAGGT CCTGGCGGAA CTCGACkCAC GCACACTGCG CCTGGA.AATA AATCGCTCCA AGGCCAGGAT CGGCGCCCAG GAACAGGCGC TGCTGCGCTT GAAAAATGGC ACACGGCCTC AGAGGTCGA GCAGTCCAAG GCACGTTTG ATGCCGCCCA GGCGCAAATG CAGC'GGCTC AACTGCACAT GCAGCGCCTG CGCAGGATCG CCGACGACAC CCAGGGCAAG GGCGTCAGCC AGCACGCCT GGACCAGGCA GCCGCCCGCT TGCAGTGGC CAAGGCCCAG TCGCAGGAGC 23951 AGCGCGATC CTGGACCCTG GCCAAGATCG GCCCACGCAA TGAGCATC 24001 24051 24 101 24151 24201 24251 24301 24351 24401 24451 24501 24 551 24601 24651 24701 24751 24 801 24851 GCCCAGGCCA CGGCCGACCT GCAAGCCTCC AGGGCCGACC TGGACCTGCT GGIACATTAC CTGGCGCGCG CTCAGCTCAA GGCGCCGACC CAGGCCCGGA CCGCACGCG CCTGCTGGAG CCGGGGGATA TGGCCTCGCC CCAGCGCCCG GTGTTTGCCC TGGCCCTGAC CGACCCCAAG TGGGTACGGG CCTATGTCAA CGAACGCCAG TTGGGCCACA TACGTGCGGA CCAGATGGCG CGGGTCTACA CCGACAGCTT TCCTGACCAG GCCATCGACG GCAAGTCGG CTACATCTCC PCGGTTGCCG AATTCACCCC CAAGTCGGTG GAACGGAAG ACCTGCGCAC CAGCCTGGTC TACGAGATCC GGGTGCTGGT CAAGGACCC GACGATCGCC TGCGGCTCGG TATGCCGGCG ACCGTCTACC TTGACCAGGC ACCGGTGGCC GGGGCCACGC CATGAGTGCG CCAGCGGTAA CGGTTCAGGC TGTTGAACTG GTCAAGCGCT TTCGCAGTGG AACGCGAGAA CAGPTGGCTC TGGAGCAGG TTCTCTGGAG GTGCCCAGTG GCAATTGAG GCCCTGGTG GGGCCGGACG GGGCCGGAA AACCACCCTG TTGCGGATGA TTGCCGGCTT GCTCAAGGCC GACGAAGGGC TGTTGAAGGT GCTGGGGCTG GACGTCGCCG CAGATCCGCA GCAGGTGCAG GACCTCATCA GCTACATGCC CAGAAGTTC GGCCTGTACG AAGACCTGAC GATCCAGGA.A AAccTcGAAc TCTACGCCGP CCTGCATGGC GTGCGCGCCC AGCAGCGCGA GGAGCGCTTT TCCCGGTTGC TGCAGTGAT GGACCTGACG CGTTTCGCG ACCGGCTGGC CGGACAGCTT CCGGCGGCA 24901 TGWICAGAA ACTGGGGTG GCCTGTACCC TGGTCCGTTC 24951 TTGCTGCTGG ACGAGCCCAC CGTGGGGGTG GATCCGCGT 25001 ACTGTGGAGC ATCATCGAAC AACTGATCGA GCAGGAAAAC 25051 TGATCAGCAC TGCCTACATG GACGAGGCCC AGCGTTGCGC 25101 GTGCTATACC AGGGGCGGCT GATCGCCCAG GGGCCACCGG 25151 CCATCTGGCG CAGGGCCTGT GCTATGGCGT CGGCCCCGAG 25201 CGGCCAGGC GTTGCAGGCG CGCTTGCTTG ACGCATGA 25251 GATGCCGTGC CCGAAGGCGG CATGGTTCGC TTTATCCGTA 25301 GCCTCTGCAG GCCAT!CCAT CCCTGGC!GG CCAGGCCTCT 25351 TGGAGGCCCG CCTCGAAGAT GGCTTCATGG TGTTGCTCCG 25401 GGCAGCGTTG CAAGCCTGGA CGAGCCGGCG CAGGCGCCCG ACCGCAACTG CGCGGCGGGA CTCACGGTGC CCAGGTCTT? CGCAGTTGTG GCCGGCACCC ACACATCATC GCGCCCAGGT CCCCAGGCGC GCAACAGGTT CCAGCTACGG 25451 25501 25551 25601 25651 25701 25751 25801 25851 25901 25951 26001 26051 26101 CCAGGGCGAC GCTGCCTCTG GGCGTGCGGT GGTCATCGA GTCCGCGPCC TGGTGCGCAA GTTCGGGGAC TTCACCGCGG TTGCCAGCAC CTCGTTCGAG GTGTTGAAGG GCGAGATATT CGGCTTGCTG GGCCCCAACG GAGCCGGCAA GACCACCACG TTTCGCATGC TCTGCGGTTT GCTCCCGCC TCCAGCGGGT ACCTTGAGGT GGCTGGCGTC AACCTGCGCA CCGCACGGGC GGCGGCGCGG CGCAAGACG GCTACGTCTC GCAGAAGTTC GCCCTGThCG CGAACCTGTC GGCCCTGGAG AACCTGCGCT TCTTCGGCGG CGCCTATGGC CTGCACGGCG CGAAGCTGAA AGCGCGGGTG GCGTTGATGC TCGAACAGTT CGACCCGAC GAACACAAGC ACCTGCGCAG CGGCGAGCTG CCCGGTGGC ACAAGCAGCG CCTGGCCATG GCGGTCGCGC TGCTGCACGA GCCACAGATC CTGTTTCTCG ACGAACCCAC CAGCGGTATC G&CCCCTTGG CCCGCCGGGC GTTCTGGCGC CGTATCACCG CCCTGGCGCA AAGCGGCACG ACGATCGTGA TCACC?CCCA CTTCATGGAA GAGGCCGAGT ACTGCGACCG TATCG!CPLTC CA.GGATGCCG GACGGCTGCT GGCCCTGGGC ACTCCGGA.AC AGGTACGCCA CCAGGGCAGC 26 151 GAGTCGCTG CCGACATGAA CAGCGCCTTC ATTGCCGGG TCGAGCGTGC 2620 1 TCGGGCCCkG GCCAGCGCCC CGGTGCAGTG AGGAATGACC ATGACTGACT 26251 CCATCGATC CGCCTCAGGG TTCAGGCGCA GGCTCATTGC GCTCACCCGA 26301 AAGGAGTTGC GCCAATTGAT GCGTGACAAG AGCAACCTGG CGATAGGCAT 26351 TGTCCTGCCC ACGTACTG TCCTGATAT TGGCTACGGC CTGTCCCTGG 26401 ACATTCGCAA CACGCCGCTG GCCGTGGTCC TGGAAGACAG CTCCCCCAGC 264 51 GCCCGCAGCG TGGGGCGAA GCTTGCGAGT TCGGACTACT TCTCGGTGGT 26501 CCAGGTGACC AGCATGGCCG AGGCCCGAGC GCTGATGGAC AAGCGTCGGG 26551 TCGACGGCAT CGTGCGCGTG CCGGCGGATT TTTCCAGGCG CCTCGAACAG 26601 CAGGATGCGC GCTTGCAATT GCTCCTGCAT GGCGCCGATG CCAACAGCGC 26651 GGCGACCCTT GGACGCTATG TCAGGGGCGC GCTGAJTGTC TGGCAACAGC 26701 AGCAGTTCGA CCGCTCCCGG CAGCCCCAGT CGGCCGGTCG GGTGCTGGTC 26751 GTCGAACGCA TGTGGTTCAA CGCCGCCAAC AGCAGTACCT GGTACCTGGT 26801 GCCCGGGCTC ATCGCCCTGA TCATGACCCT GGTGGGTGCC TTCCTGACCT 26851 CTCTGGTACT GGCCCGGGAG TGGGAGCGGG GCACGCTCGA ATCACTGTTC 26901 GTCACCCCGG TCAGGTCGAT CGAAATACTC CTGGCCAAGA TCATTCCCTA 26951 TTTCCTGGTG GGCCTGCTGG GCCTGGTGAT GTGCCTGGTT TCGGCGCGGT 27001 TGCTGTTCGA GGTACCGATC CAGGGTTCGC TGGTGCTCCT GCTGTTGAGC 27051 TCCATGCTCT ACCTGCTGGP CACCCTGGGC ATTGGCCTGC TGATTTCGGC 27101 GAAGACACGC AACCAGTTCC TGGCCAGCCA GATCGCGATC ATCTTCAGCT 27 151 TCCTGCCGGC GCTGATGCTC TCGGGGTTCC TCTTCGATCT GCGCAATGTG 27201 CCGACGTTCA TCCGCGTGGT GGGCTCGATC CTGCCGGCGA CTTACTTCAT 272 51 GGAACTGGTC AAGACCCTGT TCCTGGCTGG CAACAACTGG CCATTGATCG 27301 CCAAGAACCT CGCAATCCTC GCGGCCTACG CCGTGTTCCT GTTGAACGCT 27351 GCACGCCTGT GCACCCGCAA GAAGCTGGAT TGAAGCCATG GACAAGCTGA 27401 TGAATTTCCT ACGCGAGTTG CTGTTCCTGT TTCAGAAGGA GTTCCTGGCC 27451 ATCGTCAAGG ACCCGGCCAA CCGGGTGATC CTGATTGCCC CGGCGATCGT 27501 CCAGTCATTG CTGTTCGGCT ACGGCGCCAC CTACGACCTG AACTACGTGC 27551 CCTATGCCGT GCTGGATCAG AGCCAGGGGC GCGCTTCGAC AGAGCTTTTG 27601 GCACGCTTCG ATGCCTCCGG GGTATTTGCC CGAGTGTCGA CCCTGACCAG 27651 CGTCGATGAG ATCGCCCCGG TGATCGACCG GGGCAGTGCC TTGCTGGTGC 27701 TGCATATCGC CGCAGACTTC GACGAGCGCC TGAGCCGGCA TGAAAACGCC 27751 GGCTTGCAGT TGATCCTGGA TGGGCGTAAC TCGACGACCG CGGGAATGGC 27801 GGCCGGGCAT ATCGCCAACC TGGTGGCTGA CTTCAACCAG CAGTTCCTGG 27851 CGGTGGACAA CGCGCCGGTA CGCCTGGAGA CCCGTGCCTG GTTCAACCCC 27901 AACCTGCAAA CCCGCTGG? CATAGTGCCG GGCCTGATCG CCGCGCTGAG 27951 CATGATCCAG ACCCTGATGC TGGCGGCGTT GTCGGTAGCC AGGGAGCGCG 28001 AGCAGGGCAC CTTCGACCAA CTGCTGGTGA CTCCGCTGAC GCCTTTCGTG 28051 ATCCTGCTGG GCAAGGCCCT GCCGTCGGTA CTGATCGGTC TCTTGCAGTC 28101 CAGCCTGATC CTCCTCATCG GCCTGTTCTG GTTCAAGATC CCCATGATCG 28151 GCTcCGTGCT GGACCTGTAC CTGGGGCTGT TCGTCTTCAC CAGCGCCTGC 210 28201 28251 28301 28351 28401 28451 28501 28551 28601 28651 28701 28751 28801 28851 28901 28951 29001 29051 29101 29151 29201 29251 29301 29351 29401 29451 29501 29551 29601 29651 29701 29751 29801 29851 2 9901 29951 30001 30051 30101 30151 30201 30251 30301 30351 30401 30451 30501 30551 30601 30651 30701 30751 30801 30851 30901 GTGGGCATCG GGCTGTCGAT CTCGGCGCTG TCGGCCAACA TGCAGCAGGC CATGGTGTAC ACCTTCGTGC TGATGATGCC GCTGATCCTG CTCTCCGGCC TGATCACCCC GGTGCACAGC ATGCCCGAGG CCCTGCAACT GCTGACCTAT CTCGATCCGC TGCGCTTTGC CATCGACCTG GTCAGGAGGA TCTACCTGGA GGGGGCAACG CTGGCGGATG TGTCCTGGAA CCTGGTGCCC ATGCTTGCCG TTGCGCTGGT GACCCTGCCG CTGGCTGCGT GGCTGTTTCG CAATCGATTG GTGTGATAAG GAACGTCATG AAAACTACTC ACTCTGGATA CCTGCGTAAT GGGCTGACGC TGGCCATGCT GCTGGCCATC ACCGCCTGCA CCGTCGGGCC CGATTTCAAG GCGCCGGCAG TTACCAGCCC GGAACGCTGG ACGGATTGGC ACAGTGGCCC GCCAGCGTCT TCGGTGCCGG GCAAAGCCCC GGCGACCACC TCGCAPCCG CAGACGGCAA CTGGTGGCAG GTGTTTGGCG ACCCGGTGCT GGACCAGCTG CAGGCCCAGG TGCGCGAAGG CAGCCCGGAC CTGCACACGG CCTTGCTGCG GTTTGCCCAG GTGCGGTTGC AACGCCAGAT AGTCGCTTCC CTGGAAACGC CCGAAGTGTC CTTTTCCGCG GCGGCTTCGC GAAACCGGCA AAGCCGGTAT GCGCCGAACA ATCGAATGCT TGAAGCCTTG GGCAGCGACA GTGATGCCCT GGAAAAGGTG CTTACCGACC CTTATAGCCT GTACCAGGCC GGGTTCGACG TTTCATGGGA GCTGGATCTC TGGGGGCGGG TGTCACGCCT GGGTGAAGCC GCCCAGGCCG AGGTCGATGG TGCGGCCGCG ACCCTGGACG ACGTGTTGCT CAGTGTCTCC AGCGAGCTGG CGCGAAACTA CTTCGAAGTG CGTACCGCAC AGCGCCAGAC CCGTCTTCTG GAGCAGGAGG CACAGATCCT GGCCGCGCAC CTGCAAGTGG TTGCCGTCCA GGCCCGGGAG GGCGAGGAAG ACGGGTTCGC GGTGGAACGC CAAGAGGCAC GGCTGGCCTC GCTGCATGCA CAGATACCGG GTTGGAAGGC GCTGCACACC CAGGCGGGCA ACCGTATTGC GCTGTTGCTG GGGGAGCATC CCGGAGCCCT GGAAACCTTG CTCGCGGAGC GGCCTGATAA TGCCTTGAGC CGCCCGCTGC CCAACCTGCA GTTGGGCCTG CCTGGCGACC TGGCCCGGGA GCGTCCGGAC ATCCGCGCGG CCCAGGCCCG CTTGCATCGG GCGACGGCGG GCATCGGAGT GGCCGAGGCC GAACTGTACC CGAGCGTGAT GCTGGGGGCC GACTTCGGCT TCGAATCCTA CAAGAGCGGC CAGTTTGCCG ACTGGAGCAG CCGCACCTGG TCGGTAGGGC CGCGCCTGGA CCTGCCGCTG TTCGACCGTG GGCGGCGCAG CAAGACCGTG GTCCTGCGCA CACTCGAACA GCAAGAGGCG GCCGTGGCAT TTCACCGGAC CGTACTGGCA GCCTGGCAGG AAGTGGACGA TGCGATGAGC CGCTACCACA ACGAATACCA GCGTGCCGCG CACCTCAAGG ACAGCTATGA CAGCAAGGCC CGCACCTATG AGTGGACCCG CGTCCGGTAC GCGGCCGGGG AGGCCAGCTA CCTGGAGGAA CTGGAAGCGC AGCGCACCGT GCTGGAGGCG CAGCGCGACC TGGTCAATAG CGACAGTCAA CTGCGTACTC ACCTGATCAG CATTTACAAA TCCATGGGTG GGCATTCCAC CTGAGCCCCT GAAAQGCTGC CTCTGGCAGC ACTCCCCCCA GACCCACTTT CCTAACGGAA TAGGAAGCCA TGCAAATGGA TGACCTCTAC AAGATCGCGC TTTTTTCGTC CCTGATACTG GTAGTGCCGG GGCCCTCAAA CACCTTGCTG CTGGCCTCGG GCTTCAAGTT CGGCGTGTTG CGCTCGCTGC CACTGGTGTT GATCGAAATG CTCGGCTACA CCGTTTCGAT CTGCACCTGG GGCTGGGGGC TTGTGCTGCT GTCCCACGAC TACCCGTGGC TGATCCACCT GATCAAGCTG CTGTGCGCCT TGTACATCGC GGCGCTGGCC TTGAAGACCT GGAAGACCTC CACCACCGGA GCCCAGCACA CCGAGCGCTT TGAGAGCTGG CCACGGCACC TGTTGATTGC GACCTTGCTC AACCCCAAGG GGCTGATCTT CGCCAGTGTG ATCTTCCCGG CCAGCTCATT CGCCAGCCTG GGGGATTTCC TGCCGTCATT GAGCGCTTTC TTCCTGGCAC TGGCGCCAAT AGCCTTGCTC TGGGTCTCGC TGGGGGCGGG CATTCGAGCC CAATACCTGG GGCGGGTTTC TGGACCGCTG TTTTCCCGGG GCACCGCCGT CGTTCTGGCG CTGTTTTCCG CCACCCTCAC TTACACCGTG GTTCGTGCGG TGTAGGTGTT TTTGCCGCGG CTTTCAGCGC CCAAGTGAGG CTGGGGCAAT GGCGCTGCTG AGAGAGGCCC TTGTACCCCC CGTGTTCCAG CACCACTCTG TTGTCGCCAA CACCCGGGGG GGCTCATTCC ACGTAGCGCA ACCACAAATG GGTGCGACCG CGAACAGCCT CGACAACCGT TGAAACCACG GCTTGCCGGT TTTCCTCGGA GGTGTTGTAG AGCGCACAAA AAACGCTCGC CACCGCGCTG GCGACGCAAA GGTCGGCAGG 211 30951 31001 31051 31101 31151 31201 31251 31301 31351 31401 31451 31501 31551 3 1601 31651 31701 31751 31801 31851 31901 31951 32001 32051 32101 32151 32201 32251 32301 32351 32401 32451 32501 32551 32601 32651 32701 32751 32801 32851 32901 32951 33001 33051 33101 33151 33201 33251 33301 33351 33401 33451 33501 33551 33601 33651 CCCTGCACGC AGGAAACCGA GGTAAGACGC GCCAATCCCC ATCATCCGGG ATTTTGCTCG CCAGTGGGGC GGTTTTTGGG CAGGCGCGTA CCCGCGGGCG CTGAGCGGTG GCCTTGATCC GATGAACAGG GCGTTTGGCG GATGTTAATC TGCGCCTCCC GGACAAACGG TGGGCCAGTG GATGCAGAGG ATGTTGAGTG AAGCCGAGCT TCACAAGGTG CTGGAATCGA TCGAAAGCGA CTCCTGCCAT GACCTGAATC TTCTGTATGA CGGGATGCAG ATAGACCGGT GTGCTCTTTT CAATCAAGTC GCCATGGCGG TGGCCCGGCT CTTTATCGAG GGGCAGCGCG ACTTCCATTA CGGCGATGCG GTCATGTGGG TCGCGGCCAG AACCGCCGGC GAACTCACGC GGTTTCGCTA CGGCTCCAGC ATCAAGCGGC GCCAAAGAAT CGGCGGCGCG ACAGCAGCGG CCGGCGCTTG CGGCTCCGGC TTCTCCGGCG CCTGGGCTGA CGTTGCCTCG CGGGGCTCAG GCGCTCTTGT CGACTCGGTC GGGTCATCGG TCGGCGATGG GGTCATTTGA TCCTGACCAC CACCTTGCCC TTGGTTCGTC CCTGCTCGAC GTACTGCAAG GCCTCTGCCG TCGATTCGAA ACTGAAGCTG CGGTCAACCA CAGGCTTGAT GATTCCGGCC TCGATCAGCG CGGTGATCTG TTGCAGCTGG GCGCCATCGG CCCGCATGAA CACAAAGGCG TAGCGGACGT CCCGCTGGCG CGCCTTGCGC CGGATACCGA AGCTCAAGAG GCGCATGACC TGCTGCAACG GCCAGGCCAG CCCTTGCTCC TGGGCGAATT GCACCGTGGG TGGCCCGGAG ATGGAAATCA GCTGTCCGCC GGGCTTGAGG ACCTTGAGGG ACTTTTCCAG TACGTCGCTG CCCAGGCTGT TCAACACCAG GTCGTAGCCA TGCAAGAGGC TTTCGAAGTT CTGCTGCTTG TAGTCGATCA CCTGATCGGC GCCCAGGGCT TTGACCCATT CGACATTCGC CGTGCTGGTG GTGGTCGCGA CGAM.GCGCC GAGGTGTTTG GCCAGCTGGA TGGCAAGGGT GCCGACACCG CCTGCGCCTG CGTGGATCAG CACTTTCTGG CCCTTTTTCA GTTGGGCGGT TTCCACCAGC ACTTGCCAGG CGGTGAGTGC GACCAAGGGG APCGACGCGG CCTCGGTCAT GGTGGTGTTG GCAGGTTTCA GGGCGATTGC GTTTTCGTTC ACGGCGATCC ACTGGGCGAA CGTCCCGATC CGTGTTTCAG GCGGGCGCGC GTAGACTTCG TCTCCCGGCT TGAAGCGTTG CACTTGGGGG CCAACCTCAA CCACGACGCC CGCCAGGTCG TTGCCCAGTA CCAGCGGCAA TGCATAGGGC AGGATCAGCT TGAATTCGCC GTTGCGGATC TTCGAGTCCA GCACGTTGAC GCTGCTGGCG TGGACCTCGA TCAACACGTC GTGGGCACCC AGCGTGGGCG TGGGCGCTTC GCCCATGCGC CCACTGTTCT TGCCATAGCG ATCAATTAGA AAGGCTTTCA TCGTGGTGCG GCTCTTGTCG GTTGGAATAG GTGTGGGCGG TGCTTGGGTG GGGTTGCTGT CATGGGGCTC TGTCGTCAAG GAACGCCCGG ACCTTGGCCA CGAAGTCCGC GTGGTACTGG AAAATGCCGC CGTGCCCGGC GTCCTCATAC ATCACTAGCT GTGCGTTGGG GACACGCTTG GCCAGGGCAT GGGAGTTGGC GCTGGGCACC ATGATGTCGT TGTCGCCATT GACGATCAGC GTCGGCGCCC GCAAGCGCCC GAGATCCTGC GGCGCCTGTT TGCCCCACGC GGTGATGGCT TGCAGCTGGC GCAGGAAAGC GCCGGGCGTT GGGCCCTTGT CGCGATGCTT GCTGCGGGCT TTCAGGCGTT GCAGATATTC CGAGGCGGCT CGACGACCAT GGGGCGTCGA GGTAAAGAAC AGGTAGTACT TGGGATCGCG CAAGGTCAGC AAGCCTTTGA GCATCAAGGG CCAGGTCACC GAGCCGACCT TGTCGATACC GATGCCGCCG GCCGGTCCGG TGCCGGTGAG AATCAAQCGG TGCACCAATT GGGGGGCTTG CAGGGCGATG TCCTGAGCGA CGAAGCCGCC GAGTGAGAAG CCCAGGATAT CGACGCTCTT GAACCCCAGG GCCTGTATCA GCTCAATGGC GTCTTGCGCC ATGTCGCCGA CCGTCAGGGG GGCGGTTCCA CCCGAGCCGC CGATGCCGCG ATAGTCGGTG GTGATGACTC GCCGGGTCTG TGCCAGCCCA TCGATGATTG CCGGGTCGAA GTTGTCCAGC ACCGCGCCCC AGTGGTTAAA CAGCACCAGG GGCACGCCGC CCGCGGGACC GGTATCGCGA TAGGCGAAGG GAATGCCCCT GACCATGATC GACTGGTTCG CCGCGTTGAT GAIICGACGTC TGCAAAGCCG GGGGGAAGGT TGTCTGCGTT GGATTCATTG CTACTCCGAA CAGCGGACCG GATTCGTCGC AGTGCCTGCC GCGAATCCGC CGCGAATGCT CGCTACCTGA AGATCACTCT TGGGTCTGGT AACGCTGGGC AACCTCGGAT TGACGGATGT GCCCCAGCAG GCCCTGACGT JCGCTCTGCA GGTCATCCCA GGGTTCAGCC TGTTGCAGTG GCGGGATGGT CACCGGTTCG CGCCGATCGA AACCGACCAG TGCGGCATCC ACCAGATCGC 212 33701 33751 33801 33851 33901 33951 34001 34051 34101 34151 34201 34251 34301 34351 34401 34451 34501 34551 34601 34651 34701 34751 34801 34851 34901 34951 35001 35051 35101 35151 35201 35251 35301 35351 CCACTTCCAT GATTTCGTCC AGGGTATTGA TGTCGACGCC GGAGCGGTCC CAGATTTCCG TGCGGGTGGC GGCCGGCAGC ACGGCCTGCA CGTACACGCC CCGGGGCGAC AGTTCCAGGC TCAGGCCCTG AGACAGGAAC AGCACGAAAG CCTTGGTTGC ACCATAGACC GTCATGCCGA ACTCCGGCGC CAGGCCCACC ACCGAACCGA TGTTGATGAT CGCGCCGTCA CCGGCTTTGG CCAGGCGCGG GGCAATGGCA CTGGCCAGCC GCACCAGTGC CGTGGTGTTG AGGGCAACCA GTTGCGCTAC GCTGTCGGTG CTTTGCTCGA TGAAGTTGCC GGACAGCGCG GCGCCGGCGT TATTGATGAG GATGCCGATA CGGGCGTCGT CGCGCAGGCG GCTTTCAACG GTTGTCAGAT CGCTGAGTTG GGTCAAATCC GCTTTCATCA CCTCGACGGC GACCCCGTGT TCGCCGCGCA GTCGGGCTGC AAGTGCGTCC AGGCGTGCCC GGTCGCGGGC AACCAGGACC AGATCATGGC CACGTTGCGC GAAGCGCTCG GCGTAGATGG CTCCGATGCC AGTGGAGGCG CCAGTGATGA GAACGGTAGG GCGAGTGTTC ATGGGGTCAT CTCTCTTTCA AAATGGGTGA AAACGGACAG CTCGGTTTCT GCTTGCTCAG CTTGCGCTGG CGCTCGAATC AACCAGCGTC GGGTAGTCGA TGTAGCCTGC CGCGCCGCCG CCGTAGAGGG TGGCGGGATT GAATGCGGAC AAGGCGGCGC CGGTCTTGAG ACGCTCCACC AGATCCGGGT TGCCGATGAA CGGGCGACCG AACGCGATCA GATCGGCCTG GTCTTCGGCG AGCCGTGAGG TTGCCAGCTC CAGGTCATAG CCGTTGTTGG CGATGTAGGT GTTTTTGAAA CGCTGGCGCA GGGCGGTGAA ATCCAATGGC GCCACGTCAC GTGGGCCACC GGTCGCGCCT TCGACCATGT GCAGGTAAAC GACGTCGAGG GCGTCGAGTT GATCGACGAC GTAATTGAAT TGCGCCTGCG GATTGCTGCT GGAGACACCG TTGGCCGGCG ACACCGGCGA CAAGCGCACC CCGGTGCGAT CCGCGCCGAT TTCATTGACC ACGGCAGCGG TGACTTCGAG CAACAGACGC GCACGATTTT CAATCGAGCC GCCGTAAGCA TCGGTGCGTA CGTTGGCGCT GTCCTTGAGG ACTGATCCA GCAPTAGCC GTTCGCGCCG TGAATTTCCA CGCCATCGAA CCCGGCAGCG ATAGCGTTTG CCGCGGCCTG GCGGAAATCG GCGACGATCC CCGGCAGCTC GCTGATGTCC AGTGCGCGGG GCTCGCTGGC GTCTTCAAAG CGGTTGTTGA CGAACACCTT GGTGGCCGCA CGCAGCGCGG AAGGGGCCAC GGGGGCGGCG CCGTTTTCTT GCAGATCAAC GTGGGACACG CGGCCGACAT GCCACAGTTG CACAAAGATC TTTGCGCCCT GGGCGTGGAC CGCGTCGGTC ACCGTGCGCC AGCCATCAAT CTGCGCCTGG GTGTAGATCC CGGGGGTGTC CTGATAGCCC TGGCCCTGTT GGGAAATCTG CGTGGCTTCG CTGATCAGCA AGCCTGCGCT AGCGCGTTGG CTGTAATAGG TGGCGGCGAA TTCGCTGGGA ACAAAGCCTG CGC