1|..n H" 9.525)? :oxs mavens: TY Ll an IIIIIIIIIIIIIIIIIIIIII IIIII 3 1293 01564 IIIIIIII This is to certify that the dissertation entitled BIOTRANSFORMATION OF NON-VOLATILE ORGANOFLUORINE COMPOUNDS presented by Blake Douglas Key has been accepted towards fulfillment of the requirements for Ph.D. degreein Environmental Eng. M;;%C/o/[~_ Major professor Date {4/7//7é MSU is an Affirmative Action/Equal Opportunity Institution 0-12771 LIBRARY Mlchtgan State Unlverslty PLACE III RETURN BOX to romovo this checkout from your record. TO AVOID FINES return on or botoro onto duo. DATE DUE DATE DUE DATE DUE Jim 0 6 1995 I ‘ -- -' _.. 15L— ll MSU to An Affirmative AotloNEquol Opportunity Institution mm: BIOTRANSFORMATION OF NON-VOLATILE ORGANOFLUORINE COMPOUNDS By Blake Douglas Key A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Civil and Environmental Engineering 1996 ABSTRACT BIOTRANSFORMATION OF NON-VOLATILE ORGANOFLUORINE COMPOUNDS By Blake Douglas Key This research addresses the potential for biotransformation of non-volatile fluorinated organics, with a focus on sulfonates and carboxylates. Biodegradation of the following model sulfonates was evaluated: difluoromethane sulfonate (DFMS, CHFQSO3Na), trifluoromethane sulfonate (TFMS; CF3SO3Na), 2,2,2-trifluoroethane sulfonic acid (TES; CF3CH2803H), perfluorooctane sulfonate (PFOSA; C3F17SO3K), and lH,lH,2H,2H-perfluorooctane sulfonic acid (H-PFOSA; C5F13C2H4SO3H). Those compounds with hydrogen at the alpha-carbon degraded under aerobic and sulfur- limiting conditions. DFMS , TF3, and H-PFOSA were transformed by a variety of bacteria, including Pseudomonads, Bacillus subtilis, and Escherichia coli. However, E. coli was not capable of transforming H-PFOSA. DFMS was completely defluorinated with a stoichiometric yield of fluoride. TFMS and H-PFOSA were only partially defluon’nated with equimolar fluoride release from TES and 1-1.4 moles fluoride release from H-PFOSA. Eight volatile and fluorinated byproducts of H-PFOSA were detected by mass spectrometry and atomic emission detection. Oxygen was required for growth on and transformation of DFMS. Complete shut down of transformation occurred with the removal of glucose or ammonium. Inhibition studies with other sulfur sources suggest that the sulfur-containing byproduct of DFMS transformation is assimilated by existing sulfur assimilating pathways. Non-competitive inhibition was observed with sulfate, sulfite, methanesulfonate, cystine, and methionine. These results suggest that transformation of DFMS is linked to a sulfur-scavenging system in which DFMS is desulfonated then defluorinated. Monofluoroacetate (MFA), difluoroacetate (DFA), and trifluoroacetate (TFA) were selected as model compounds for study of anaerobic degradation of fluorinated carboxylates. Denitrifying and sulfate-reducing enrichment cultures defluorinated MFA. A bacterial isolate, designated strain M7, was capable of growth on and defluorination of MFA under aerobic and denitrifying conditions. Strain M7 was unable to degrade DFA or TFA. Phylogenetic analysis of the 16S rRNA sequence shows a 96.0 to 97.5% similarity between strain M7 and the Bradyrhizobium species. Two other bacteria with the ability to utilize MFA under aerobic conditions were also capable of MFA degradation under denitrifying conditions, suggesting that this capability is widespread in nature. Dedicated to my wonderful friend, lover, and lifetime companion Lynda Rae whom I cherish with all of my heart, mind, body, and soul. iv ACKNOWLEDGMENTS I would like to express my deepest appreciation to all of the people who made the completion of this research and dissertation possible. A very special thanks goes to my wife Lynda for her loving support and unconditional faith in my abilities without which I might have given up long ago. I would also like to thank my parents, Rudy and Karolynn Key, for their support and encouragement throughout my life. I would like to thank my friend, mentor, and advisor Dr. Craig Criddle. His enthusiasm for science has influenced me in many ways. I thank him for his unending encouragement, intellectual guidance, and friendship. I am honored to have worked with a scientist of such high caliber. Thank you also to Dr. Robert Howell for his intellectual input and direction in several aspects of my research. Dr. Howell's role as liaison between 3M and Michigan State University has created a c00perative link between industry and academia. Other 3M associates that I would like to thank are Dr. James D. Johnson, Gregory N. Maisel, James B. Drake, and Dr. James K. Lundberg of 3M Environmental Laboratory, for their analytical help. A special thanks goes to Dr. Susan Masten, Dr. Larry Fomey, and Dr. Patrick Oriel for critiquing my work and for providing guidance and direction to my research. Without their help I would not have done many of the experiments presented here. Thank you to the National Science Foundations, Center for Microbial Ecology (CME) for their financial support early in my research career. I would like to commend all of the participants of CME for creating a collaborative, cross-disciplinary research and educational organization. Many of my colleagues and friends deserve thanks for helping me complete my research. If it were not for Yanlyang Pan, I am certain that I would never have finished. His knowledge and skills to maintain and implement all of the analytical equipment in our lab has helped me more than can be imagined. I appreciate Dr. Mike Dybas and Dr. Greg Tatara's help with experimental design and for always keeping our lab interesting. I would like to thank Dr. Brad Upham for his intellectual input and especially for his encouragement and friendship. Thank you also to Mark Sneathen, Joel Klappenbach, Mike Witt, Debbie Hogan, Xianda Zhao, Sadhana Chauhan, Lycely Sepfiveda-Torres, Mike Apgar, Brad Kach, Janice and Bernard Weidenaar, and Myron Erickson for their help in various aspects of my work and my life. A very special thank you goes to Linda Steinman for graciously putting up with me and all the other "pester flies" that are constantly requesting her help. She has been an invaluable asset in dealing with administrative issues. I thank her for her kind and friendly spirit that has cheered me on many occasions. I would like to thank Dr. Mark Emptage of DuPont Corp for the MFA-degrading strains. I would like to thank Jizhong Zhou for his work on the 16S rRNA sequencing of strain M7. vi TABLE OF CONTENTS LIST OF TABLES ................................ LIST OF FIGURES LIST OF SYMBOLS CHAPTER 1 CHAPTER 2 CHAPTER 3 CHAPTER 4 ooooooooooooooooooooooooooooooo oooooooooooooooooooooooooooooo INTRODUCTION .................... References ........................ FLUORINATED ORGANICS IN THE BIOSPHERE. . Introduction ........................ Fate and Effects of Volatile Fluorinated Organics . . . . Fate and Effects of Non-Volatile Fluorinated Organics . A Case History - Trifluoroacetic Acid .......... Summary and Conclusions ................ References ........................ DEFLUORINATION OF ORGANOFLUORINE SULFUR COMPOUNDS BY PSEUDOMONAS SP. STRAIN D2 ........................ Abstract .......................... Introduction ........................ Methods and Materials .................. Results .......................... Discussion ........................ References ........................ PHYSIOLOGY OF DIFLUOROMETHANE SULFONATE TRANSFORMATION BY PSEUDOMONAS SP. STRAIN D2 ............ Abstract .......................... Introduction ........................ Methods and Materials .................. Results .......................... Discussion ........................ References ........................ vii ix 8 15 19 25 3O 33 45 45 46 47 49 62 62 63 64 68 CHAPTER 5 BIOTRANSFORMATION OF SULFUR-CONTAINING ORGANOFLUORINE COMPOUNDS BY DIVERSE ORGANISMS ....................... 81 Abstract .......................... 81 Introduction ........................ 82 Methods and Materials .................. 82 Results .......................... 86 Discussion ........................ 95 References ........................ 97 CHAPTER 6 BIODEGRADATION OF MONOFLUOROACETATE UNDER DENITRIFYING CONDITIONS ....... 98 Abstract .......................... 98 Introduction ........................ 99 Methods and Materials .................. 100 Results .......................... 105 Discussion ........................ 1 13 References ........................ l 16 CHAPTER 7 CONCLUSIONS AND FUTURE WORK RECOMMENDATIONS ................. 1 19 Conclusions ........................ 1 19 Future Research ...................... 121 APPENDIX A ................................... 122 GC/MS using electron impact with ion trap detector . . . 122 GC/MS using electron impact with quadrapole detector . 134 APPENDIX B ................................... 154 GC/MS using chemical ionization with quadrapole detector .......................... 154 APPENDIX C ................................... 158 GC/Atomic emission detection .............. 158 viii LIST OF TABLES Chapter 1 No Tables Chapter 2 Table 2.1. Examples of fluorinated aliphatic compounds and their applications. Table 2.2. Examples of fluorinated aromatic compounds and their applications. Table 2.3. Fluorinated chemical application for crop protection in US, 1995. Calculated from the Agricultural Chemical Usage - 1995 Field Crops Summary, published by the USDA. (Note: this data only includes applications to corn, cotton, soybeans, and wheat for major producing states). Table 2.4. Fluorocarbon production worldwide. Chapter 3 Table 3.1. Emission wavelengths and plasma gases used for atomic emission detection. Table 3.2. Possible volatile fluorinated products of H-PFOSA transformation by strain D2. Chapter 4 No Tables Chapter 5 No Tables ix Chapter 6 Table 6.1. Kinetic parameters for Bradyrhizobium sp. strain M7 Table 6.2. Glycolate determination in cell free extract of strain M7. Values were obtained by ion exclusion chromatography. Glycolate levels were confirmed by colorimetric assay. Table 6.3. Substrate degradation and growth by Bradyrhizobium sp. strain M7 Chapter 7 No Tables Appendix A Table A]. GC/MS listing for peak #1 of strain D2 incubated with glucose and H- PFOSA. Table A2. GC/MS listing for peak #2 of strain D2 incubated with glucose and H- PFOSA. Table A3. GC/MS listing for peak #3 of strain D2 incubated with glucose and H- PFOSA. Table A.4. GC/MS listing for peak #4 of strain D2 incubated with glucose and H- PFOSA. Table A5. GC/MS listing for peak #5 of strain D2 incubated with glucose and H- PFOSA. Table A6. GC/MS listing for peak #6 of strain D2 incubated with glucose and H- PFOSA. Table A7 GC/MS listing for peak #7 of strain D2 incubated with glucose and H- PFOSA. Table A8. GC/MS listing for peak #8 of strain D2 incubated with glucose and H- PFOSA. Table A9. GC/MS listing for peak #9 of strain D2 incubated with glucose and H- PFOSA. Table A.10. GC/MS listing for peak #10 of strain D2 incubated with glucose and H- PFOSA. Table A.11. PFOSA. Table A.12. PFOSA. Table A.13. PFOSA. Table A.14. PFOSA. Table A.15. PFOSA. Table A.16. PFOSA. Table A.17. PFOSA. Table A.18. PFOSA. Table A.19. PFOSA. GC/MS listing for peak A of strain D2 incubated with glucose and H- GC/MS listing for peak B of strain D2 incubated with glucose and H- GC/MS listing for peak C of strain D2 incubated with glucose and H- GC/MS listing for peak D of strain D2 incubated with glucose and H- GC/MS listing for peak E of strain D2 incubated with glucose and H- GC/MS listing for peak F of strain D2 incubated with glucose and H- GC/MS listing for peak G of strain D2 incubated with glucose and H- GC/MS listing for peak H of strain D2 incubated with glucose and H- GC/MS listing for peak I of strain D2 incubated with glucose and H- Appendix B No Tables Appendix C No Tables xi LIST OF FIGURES Chapter 1 No Figures Chapter 2 Figure 2.1. Examples of aromatic fluorinated compounds. Figure 2.2. Biogeochemical cycling of organofluorine compounds. Chapter 3 Figure 3.1. Defluorination of DFMS by Pseudomonas sp. strain D2. Error bars represent the standard deviation of triplicate samples. Figure 3.2. Growth of Pseudomonas sp. strain D2 on DFMS. Error bars represent the standard deviation of triplicate samples. Figure 3.3. Fluoride release from TES. Numbers represent ratio of moles of fluoride to moles of TES. Triplicate samples were measured, but error bars are not visible due to scale. Figure 3.4. Fluoride release from H-PFOSA. Numbers represent ratio of moles of fluoride to moles of H-PFOSA. Triplicate samples were measured, but error bars are not visible due to scale. Figure 3.5. Growth of Pseudomonas sp. strain D2 on TBS. Error bars represent the standard deviation of triplicate samples. Figure 3.6. Growth of Pseudomonas sp. strain D2 on H-PFOSA. Error bars represent the standard deviation of triplicate samples. Figure 3.7. Atomic Emission Detection of volatile fluorinated byproducts of H-PFOSA biodegradation. xii Chapter 4 Figure 4.1. Growth of Pseudomonas sp. strain D2 with glucose and acetate under aerobic conditions. No growth was observed under anaerobic conditions. Error bars represent the standard deviation of triplicate samples. Figure 4.2. Growth of Pseudomonas sp. strain D2 with DFMS as sole source of sulfur. DFMS cannot be used as carbon source. Error bars represent the standard deviation of triplicate samples. Figure 4.3. Effects of removal of glucose or ammonium, from growth medium, on the transformation of DFMS in stationary phase cells. Error bars represent the standard deviation of triplicate samples. Figure 4.4. Non-competitive inhibition of DFMS transformation rates by sodium sulfate. Error bars represent the standard deviation of triplicate samples. Figure 4.5. Non-competitive inhibition of DFMS transformation rates by sodium sulfite. Error bars represent the standard deviation of triplicate samples. Figure 4.6. Non-competitive inhibition of DFMS transformation rates by sodium methane sulfonate. Error bars represent the standard deviation of triplicate samples. Figure 4.7. Non-competitive inhibition of DFMS transformation rates by cystine. Error bars represent the standard deviation of triplicate samples. Chapter 5 Figure 5.1. Defluorination of DFMS (32 1.1M) by various bacteria and yeast under aerobic and sulfur-limiting conditions. Error bars represent the standard deviation of triplicate samples. Figure 5.2. Defluorination of 2,2,2-trifluoroethane sulfonic acid (32 uM) by various bacteria and yeast under aerobic and sulfur-limiting conditions. Error bars represent the stande deviation of triplicate samples. Figure 5.3. Defluorination of lH,lH,2H,2H-perfluorooctane sulfonic acid (32 uM) by various bacteria and yeast under aerobic and sulfur-limiting conditions. Error bars represent the standard deviation of triplicate samples. Figure 5.4. GC/ECD chromatogram of the natural area incubation with glucose and H- PFOSA. Figure 5.5. GC/ECD chromatogram of strain D2 incubated with glucose and H- PFOSA. Figure 5.6. GC/MS chromatogram of the natural area incubation with glucose and H- PFOSA. xiii Figure 5.7. GC/MS spectrum number 604 of the natural area incubation with glucose and H-PFOSA. Figure 5.8. GC/MS chromatogram of strain D2 incubated with glucose and H-PFOSA. Figure 5.9. GC/MS spectrum number 604 of strain D2 incubated with glucose and H- PFOSA. Chapter 6 Figure 6.1. Defluorination of MFA by denitrifying and sulfate-reducing enrichments. Error bars represent the standard deviation of triplicate samples. Figure 6.2. Growth of Bradyrhizobium sp. strain M7 on MFA under denitrifying conditions: (a) consumption of MFA and production of fluoride, (b) consumption of nitrate and growth of cells. Error bars represent the standard deviation of triplicate samples. Figure 6.3. Phylogenetic tree showing the location of Bradyrhizobium sp strain M7 in relation to other Bradyrhizobium species. Figure 6.4. Comparison of specific substrate utilization rates for aerobically and anaerobically grown Bradyrhizobium sp. strain M7. Error bars represent the standard deviation of triplicate samples. Chapter 7 No Figures Appendix A Figure A.1. GC/MS chromatogram of strain D2 incubated with H-PFOSA and glucose. Figure A.2. GC/MS spectrum for peak #1 of strain D2 incubated with glucose and H- PFOSA. Figure A.3. GC/MS spectrum for peak #2 of strain D2 incubated with glucose and H- PFOSA. Figure A.4. GC/MS spectrum for peak #3 of strain D2 incubated with glucose and H- PFOSA. Figure A.5. GC/MS spectrum for peak #4 of strain D2 incubated with glucose and H- PFOSA. xiv Figure A.6. PFOSA. Figure A.7. PFOSA. Figure A.8. PFOSA. Figure A.9. PFOSA. Figure A.10. PFOSA. Figure A.11. H-PFOSA. Figure A.12. glucose. Figure A.13. PFOSA. Figure A.14. PFOSA. Figure A.15. PFOSA. Figure A.16. PFOSA. Figure A.17. PFOSA. Figure A.18. PFOSA. Figure A.19. PFOSA. Figure A20. PFOSA. Figure A.21. PFOSA. GC/MS spectrum for peak #5 of strain D2 incubated with glucose and H- GC/MS Spectrum for peak #6 of strain D2 incubated with glucose and H- GC/MS spectrum for peak #7 of strain D2 incubated with glucose and H- GC/MS spectrum for peak #8 of strain D2 incubated with glucose and H- GC/MS spectrum for peak #9 of strain D2 incubated with glucose and H- GC/MS spectrum for peak #10 of strain D2 incubated with glucose and GC/MS chromatogram of strain D2 incubated with H-PFOSA and GC/MS spectrum for peak A of strain D2 incubated with glucose and H- GC/MS spectrum for peak B of strain D2 incubated with glucose and H- GC/MS spectrum for peak C of strain D2 incubated with glucose and H- GC/MS spectrum for peak D of strain D2 incubated with glucose and H- GC/MS spectrum for peak E of strain D2 incubated with glucose and H- GC/MS spectrum for peak F of strain D2 incubated with glucose and H- GC/MS spectrum for peak G of strain D2 incubated with glucose and H- GC/MS spectrum for peak H of strain D2 incubated with glucose and H- GC/MS spectrum for peak I of strain D2 incubated with glucose and H- XV Appendix B Figure B.l. GC/MS chromatogram of strain D2 incubated with glucose and H-PFOSA. Figure 8.2. GC/MS spectrum for peak #1 of strain D2 incubated with glucose and H- PFOSA. Figure B.3. GC/MS spectrum for peak #2 of strain D2 incubated with glucose and H- PFOSA. Figure B.4. GC/MS spectrum for peak #3 of strain D2 incubated with glucose and H- PFOSA. Figure B.5. GC/MS spectrum for peak #4 of strain D2 incubated with glucose and H- PFOSA. Figure 8.6. GC/MS spectrum for peak #S‘Of strain D2 incubated with glucose and H- PFOSA. ' Figure 8.7. GC/MS spectrum for peak #6 of strain D2 incubated with glucose and H- PFOSA. Figure B.8. GC/MS spectrum for peak #7 of strain D2 incubated with glucose and H- PFOSA. Figure B.9. GC/MS spectrum for peak #8 of strain D2 incubated with glucose and H- PFOSA. Appendix C Figure C.1. AED chromatogram for fluorine (at 690 nm) strain D2 incubated with glucose and H—PFOSA. Figure C.2. AED chromatogram for sulfur (at 181 nm) strain D2 incubated with glucose and H-PFOSA. Figure C.3. AED chromatogram for hydrogen (at 486 nm) strain D2 incubated with glucose and H-PFOSA. Figure C.4. AED chromatogram for carbon (at 496 nm) strain D2 incubated with glucose and H-PFOSA. Figure C.5. AED chromatogram for oxygen (at 777 nm) strain D2 incubated with glucose and H-PFOSA. xvi I k protein/hr) Ks Ki LIST OF SYMBOLS specific substrate utilization coefficient (umoles /mg protein/hr) rate-limiting substrate concentration (uM) inhibitor concentration (uM) maximum specific substrate utilization coefficient (umoles /mg half-velocity coefficient (uM) inhibition constant (uM) total protein concentration (mg/L) initial protein concentration (mg/L) specific growth rate (d'l) maximum specific growth rate (d‘l) endogenous decay rate (d‘l) time in hours (hr) xvii CHAPTER 1 INTRODUCTION Environmental Significance of Organofluorine Compounds The commercial and domestic use of organofluorine compounds has dramatically increased in recent years. These compound are used as propellants, surfactants, agrochemicals, lamprey larvicide, insecticides, adhesives, refrigerants, fire retardants, and medicines [2, 3, 5, 7, 8, 9, 10, 17, 20, 21, 23, 27, 29]. Many of these compounds are used because they are chemically stable and are perceived as more inert biologically and therefore less likely to have an impact on human health or the environment. However, inert molecules tend to persist and accumulate, and they are more difficult to remediate. In addition, several fluorinated organics are subject to at least limited biotransformation under appropriate environmental conditions. Moreover, as discussed in Chapter 2, organofluorine molecules actually do exhibit significant biological effects, as inhibitors of enzymes, cell-cell corrrmunication, membrane transport, and processes for energy generation [3, 4, 6, 11, 22, 24, 28]. Choice of Model Compounds Two categories of organofluorine compounds with particularly useful properties are the fluorinated sulfonates and the fluorinated carboxylates. Perfluorinated sulfonates and perfluorinated carboxylates are used as industrial surfactants and as catalysts in synthetic chemistry. Perfluorooctane sulfonate (PFOSA; C8F17SO3Na) has excellent 2 chemical and thermal stability and is important commercially as a surfactant and as a precursor of other fluorinated surfactants and pesticides [l]. Shorter chained perfluorinated compounds, such as trifluoromethane sulfonate (TFMS; CF3SO3Na), are used as oligomerization or polymerization catalyst. TFMS is one of the strongest acids known, has great thermal stability, does not release fluoride in the presence of strong nucleophiles, and resists both oxidation and reduction [25]. Fluorinated carboxylates of industrial significance include perfluorooctanoic acid (PFOA), perfluorodecanoic acid (PFDA), and trifluoroacetic acid (TFA). In addition, monofluoroacetate (MFA) is used as a potent pesticide throughout the world. PFOSA, TFMS and their partially fluorinated analogues difluoromethane sulfonate (DFMS; CHFst3Na), lH,lH,2H,2H-perfluorooctane sulfonic acid (H-PFOSA; C5F13C2H4SO3H), and 2,2,2-trifluoroethane sulfonic acid (TES', CF3CH2503H) were selected as model compounds to represent the fluorinated sulfonates. MFA, difluoroacetate (DFA), and TFA were chosen as model compounds for the fluorinated carboxylates. Hypotheses The primary objective of this research was to study the biodegradability of non-volatile fluorinated compounds. Two major classes of fluorinated molecules were investigated: fluorinated sulfonates and fluorinated carboxylates. The following hypotheses were evaluated: 1. Perfluorinated sulfonates and carboxylates are refractory. 2. Hydrogen-substituted fluorinated sulfonates can be utilized as sulfur sources under aerobic and sulfur-limiting conditions. 3 3. Hydrogen-substituted fluorinated carboxylates can be utilized as carbon sources under aerobic and denitrifying conditions. 4. Biodegradation of hydrogen-substituted fluorinated sulfonates is linked to sulfur assimilation. 5. The capacity for transformation of hydrogen-substituted fluorinated sulfonates and carboxylates is widely dispersed in nature. Organization of Thesis Chapter 2 provides a comprehensive review of research on the fate and effects of organofluorine compounds. The review summarizes research on the fate of various fluorinated compounds and the effects that these compounds and their metabolic byproducts have on animals, plants, and microorganisms. In addition, the review identifies areas that need additional attention. Chapter 3 evaluates the biodegradability of representative fluorinated sulfonates using Pseudomonas sp. strain D2, an isolate capable of completely defluorinating DFMS under sulfur-limiting and aerobic conditions. Chapter 4 investigates the physiology of transformation of sulfur- containing organofluorine compounds in greater detail using Pseudomonas sp. strain D2. Results from Chapter 3 established that a structural or molecular limitation to transformation occurs when the alpha-carbon of fluorinated sulfonates is completely fluorinated (TFMS and PFOSA), while complete transformation (DFMS) or at least partial transformation occurs when the alpha-carbon has a hydrogen atom (TES and H- PFOSA). Chapter 4 provides additional insight into factors that affect transformation of these compounds. In whole cell experiments oxygen was required for growth and transformation of DFMS. While DFMS is not utilized as a source of carbon and energy, it is used as a source of sulfur under sulfur-limiting conditions. Defluorination of DFMS and other fluorinated sulfonates is linked to a sulfur-scavenging system that is 4 active when both a carbon source (glucose) and a nitrogen source (ammonium) are present in excess. Inhibition studies with other sulfur sources establish a non- competitive pattern of inhibition, suggesting that the sulfur released by DFMS transformation is assimilated through existing sulfur assimilation pathways. To generalize the above findings, the defluorination ability of other bacteria and yeast were evaluated under aerobic and sulfur-limiting conditions (Chapter 5). Several phylogenetically related Pseudomonads, Bacillus subtilis, and Escherichia coli degraded at least one of the tested sulfur-containing organofluorine compounds. Saccharomyces cerevisiae was not able to transform any, suggesting that the mechanism may be specific to prokaryotes. In addition, H-PFOSA degraded in three different soil types, but not in aquatic samples. MFA was selected as a model compound for this work. Pseudomonads and other bacteria, as well as some fungi, grow with MFA as a carbon source aerobically, but transformation under anaerobic conditions was not previously evaluated [12, 13, 14, 15, 16, 18, 19, 26, 30]. The first step in aerobic degradation of MFA is hydrolytic attack of the carbon-fluorine bond yielding glycolic acid [12]. Chapter 6 provides evidence that many of the same organisms are also capable of growth on and defluorination of MFA under denitrifying conditions. In addition, aerobic and anaerobic degradation of MFA is demonstrated for a new MFA-degrading isolate, Bradyrhizobium sp. strain M7. A hydrolytic mechanism is proposed for both aerobic and anaerobic conditions based upon production of glycolic acid intermediate. REFERENCES Abe, T., and S. Nagase. 1982. Electrochemical fluorination (Simons process) as a route to perfluorinated organic compounds of industrial interest, pp. 19-44. In R. E. Banks (ed.), Preparation, properties, and industrial applications of organofluorine compounds. John Wiley & Sons, New York. Banitt, E. H., W. E. Coyne, K. T. McGurran, and J. E. Robertson. 1974. Monofluoromethanesulfonanilides. A new series of bronchodilators. J. Med. Chem. 17:116-120. Cartwright, D. 1994. Recent developments in fluorine-containing agrochemicals, pp. 237-257. In R. E. Banks, B. E. Smart, and J. C. Tatlow (ed.), Organofluorine chemistry: Principles and commercial applications. Plenum Press, New York. Clarke, D. D. 1991. Fluoroacetate and fluorocitrate: Mechanism of action. Neurochemical Research 16: 1055-1058. Commission, Great Lakes Fishery 1985. TFM vs. the sea lamprey: A generation later. Great Lakes Fish. Corn. Spec. Pub. 85-6. Deocampo, N. D., B. L. Upham, and J. E. Trosko. 1996. The role of gap junction communication in the toxicity of quadricyclane, perfluorodecanoic acid, perfluorooctanoic acid and jet fuels JP-8 and JP-4. Fundam. Appl. Toxicol. (suppl) 30: 1065. Elliot, A. J. 1994. Chlorofluorocarbons, pp. 145-157. In R. E. Banks, B. E. Smart, and J. C. Tatlow (ed.), Organofluorine chemistry: Principles and commercial applications. Plenum Press, New York. EPA. 1989. Pesticide Fact Sheet on Sulfluramid. EPA. 205 (540/FS-89-060). Fielding, H. C. 1979. Organofluorine surfactants and textile chemicals, pp. 214- 232. In R. E. Banks (ed.), Organofluorine chemicals and their industrial applications. Ellis Horwood Ltd., Chichester. 10. ll. 12. l3. 14. 15. 16. 17. 18. 19. 20. 6 Filler, R. 1993. Fluoromedicinal chenristry - An overview of recent developments, pp. 1-23. In R. Filler, Y. Kobayashi, and L. M. Yagupolskii (ed.), Organofluorine compounds in medicinal chemistry and biomedical applications. Elsevier, Amsterdam. Gadelhak, G. G., Dissertation, Michigan State University (1992). Goldman, P. 1965. The enzymatic cleavage of the carbon-fluorine bond in fluoroacetate. J. Biol. Chem. 240:3434-3438. Goldman, P. 1971. Enzymology of carbon-halogen bonds, pp. 147-165. In (ed.), Degradation of synthetic organic molecules in the biosphere. National Academy of Sciences, Washington DC. Goldman, P., and G. W. A. Milne. 1966. Carbon-fluorine bond cleavage; 11. Studies on the mechanism of the defluorination of fluoroacetate. J. Biol. Chem. 241:5557-5559. Goldman, P., G. W. A. Milne, and D. B. Keister. 1968. Carbon-halogen bond cleavage; III. Studies on bacterial halidohydrolases. J. Biol. Chem. 243:428-434. Kelly, M. 1965. Isolation of bacteria able to metabolize fluoroacetate or fluoroacetamide. Nature (London) 208:809-810. Kissa, E. 1994. Fluorinated surfactants: synthesis, properties, and applications. Marcel Dekker, Inc, New York. Meyer, J. J. M., N. Grobbelaar, and P. L. Steyn. 1990. Fluoroacetate- metabolizing Pseudomonad isolated from Dichapetalum cymosum. Appl. Environ. Microbiol. 56:2152-2155. Meyer, J. J. M., and D. O'Hagan. 1992. Conversion of 3-fluoropyruvate to fluoroacetate by cell-free extracts of Dichapetalum cymosum. Phytochemistry 31:2699-2701. Moore, G. G. I. 1974. Sulfonamides with antiinflarnmatory activity, pp. 159-176. In R. A. Scherrer, and M. W. Whitehouse (ed.), Antiinflammatory agents: Chemistry and pharmacology. Academic Press, New York. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 7 Moore, G. G. I. 1979. Fluoroalkanesulfonyl Chlorides. J. Org. Chem. 14:1708- 1711. Peters, R. 1972. Some metabolic aspects of fluoroacetate especially related to fluorocitrate, pp. 55-76. In (ed.), Carbon-fluorine compounds: Chemistry, biochemistry, and biological activities (A Ciba Foundation Symposium). Associated Scientific Publishers, Amsterdam. Rao, N. S., and B. E. Baker. 1994. Textile finishes and fluorosurfactants, pp. 321-336. In R. E. Banks, B. E. Smart, and J. C. Tatlow (ed.), Organofluorine chemistry: Principles and commercial applications. Plenum Press, New York. Schnellmann, R. G., and O. M. Randall. 1990. Perfluorooctane sulfonamide: a structurally novel uncoupler of oxidative phosphorylation. Biochemica Et Biophysica Acta 1016:344-348. Stang, P. J., and M. R. White. 1983. Triflic acid and its derivatives. Aldrichimica Acta 16:15-22. Tonomura, K., F. Futai, O. Tanabe, and T. Yamaoka. 1965. Defluorination of monofluoroacetate by bacteria; Part 1. Isolation of bacteria and their activity of defluorination. Agr. Biol. Chem. 29:124-128. Trepka, R. D., et al. 1974. Synthesis and herbicidal activity of fluorinated N- phenylalkanesulfonamides. J. Agr. Food Chem. 22: 1111-1119. Upham, B. L., 1996, personal communication. Vander Meer, R. K., C. S. Lofgren, and D. F. Williams. 1985. Fluoroaliphatic sulfones: a new class of delayed-action insecticides for control of Solenopsis invicta. J. Econ. But. 78: 1 190-1 197. Walker, .1. R. L., and B. C. Lien. 1981. Metabolism of fluoroacetate by a soil Pseudomonas sp. and Fusarium solani. Soil Biol. Biochem. 13:231-235. CHAPTER 2 F LUORINATED ORGANICS IN THE BIOSPHERE INTRODUCTION Research investigating the environmental fate of halogenated compounds has largely focused on brominated and chlorinated organics. Fluorinated organics have received less attention because fewer are regulated, their measurement in environmental samples is generally more difficult, and they are perceived as more inert biologically and therefore less likely to have an impact on human health or the environment. Of course, the perception of "inertness" and its environmental significance are debatable: inert molecules tend to persist and accumulate, and they are more difficult to remediate. In addition, several fluorinated organics are subject to at least limited biotransformation under appropriate environmental conditions. Moreover, organofluorine molecules actually do exhibit significant biological effects, as inhibitors of enzymes, cell-cell communication, membrane transport, and processes for energy generation [16, 19, 23, 37, 77, 85, 105]. The chemistry of organofluorine molecules is unique because of the properties of fluorine. The fluorine atom has a van der Waals radius of 1.47 A, a size more comparable to that of oxygen (1.52 A) than to that of the other halogens (chlorine, 1.8 A; bromine, 1.95 A; and iodine, 2.15 A). Fluorine was once thought to be similar in size to hydrogen (1.2 A), but it is now considered isosterically similar to a hydroxyl group [90]. Compared to other halogens, fluorine is extremely electronegative having an electronegativity of 4.0 compared to an electronegativity of 3.0 for chlorine, 2.8 for 9 bromine, and 2.5 for carbon [76]. This high electronegativity confers a strong polarity to the carbon-fluorine bond [40]. The carbon-fluorine bond also has one of the largest bond energies in nature (105.4 kcal/mol); other carbon-halogen bond energies are similar to those of common metabolic intermediates [76]. The "unnatural" strength of carbon-fluorine bonds confers unusual stability. In fact, many fluorinated agrochenricals capable of enzyme inhibition are fluorine-stabilized analogues of the natural enzyme substrate. A dramatic illustration of the strength and stability of the C-F bond is monofluoroacetate, which can withstand boiling with 100% sulfuric acid without any defluorination [82]. For many man-made fluorinated organics, such as the perfluorinated organics, stability is probably related to the fact that the molecular structure is unlike anything currently known in nature. Fluorine is the most abundant halogen in the earth’s crust and ranks 13th in abundance among all elements [75]. This may explain instances of natural organofluorine production. The best known of these natural organofluorine compounds is monofluoroacetate (MFA). MFA is produced by plants in the genus Dichapetalum, as well as Palicourea marcgravii, Acacia georginae, Gastrolobium grandiflorum, and Oxylobium species [48]. The West African plant, Dichapetalum toxicarium, also produces m-fluorooleic acid, w—fluoropalmitic acid, and possibly w-fluorocaprate and m—fluoromyristate [93]. Certain fungi also produce fluorinated organics; Streptomyces clavus and Streptomyces cattleya produce the fluorine-containing antibiotic nucleocidin and 4-fluorothreonine respectively [48, 93, 94]. Streptomyces cattleya is also capable of producing monofluoroacetate [94]. Finally, production of CFC-11, CFC-12, CFC-113, HCFC-21, HCFC-22, tetrafluoroethylene, and chlorotrifluoroethylene has been reported in volcanic gases and drill wells [44, 51]. All of the known biologically-produced fluorinated organics contain only one carbon- 10 fluorine bond. This contrasts with many man-made fluorinated organics, which often contain many fluorine substituents and may even be fully fluorinated. Because of their many useful properties, the number of man-made fluorinated organics has dramatically increased over the past few years. According to a report from Business Communications Company, Inc., sales of fluoropolymers (including surfactants, textile finishes, fluoroelastomers, and polymer resins) were expected to increase from $1.35 billion in 1994 to $1.76 billion by 1999 [12]. Tables 2.1 and 2.2 list several examples of aliphatic and aromatic fluorinated compounds. Representative structures of some aromatic fluorinated compounds are illustrated in Figure 2.1. These or related compounds have been used in aerosol propellants, surfactants, refrigerants, plastics, anesthetics, herbicides, insecticides, rodenticides, lampricides, plant growth regulators, medicines, adhesives, fire retardants, and even blood substitutes [4, 7, 15, 16, 26, 31, 32, 52, 54, 63, 67, 68, 69, 78, 92, 98, 99, 106, 115]. Synthesis is accomplished using classical chemical or electrochemical processes, although production of novel fluorochemicals using microbial pathways is also possible [79]. 11 Cl Diflubenzuron CF3 CHzCHNHCH2CH3°HCl or ReduxTM (dexfenfluramine hydrochloride) NHCOCH3 H 3C NO2 NHS02CF3 CH3 Mefluidide H CF3 N02 3-Trifluoromethyl-4-nitrophenol “1.13 ,C=C co CHr‘Q’ F C Ffi-Q HCHzCHzNHCH3-HC1 Prozac® (fluoxetine hydrochloride) NO2 1: c ,CHZCHZCH3 —< '>— It 3 crtzcuzcn3 no2 Trifluralin moo-SF dichloro(trifluoromethyl) difluorodiphenylmethane Series of chlorinated homologues (1-5 chlorines) Tefluthrin Figure 2.1. Examples of aromatic fluorinated compounds. 12 The environmental fate of fluorinated organics depends upon the structure of the molecule. Although some generalizations are possible, research is needed before reliable predictions of fate will be possible for many compounds. Broadly speaking, fluorinated organics can be classified as either volatile or nonvolatile. Volatile fluorinated organics consist primarily of partially or completely fluorinated alkanes, ethers, or amines. Fully halogenated fluorinated organics have very long lifetimes in the atmosphere and can migrate to the stratosphere where they are destroyed by ozone and photolysis [1, 66, 80]. By contrast, volatile fluorinated organics containing one or more hydrogen atoms are susceptible to oxidation by hydroxyl radicals in the troposphere, yielding fluoride, chloride, and partially oxidized organic species, the most significant of which is trifluoroacetic acid (TFA) [100, 109]. TFA is also produced industrially, as are many other commercially important nonvolatile fluorinated organics. 13 Table 2.1. Examples of fluorinated aliphatic compounds and their applications. Compound Molecular Formula CAS # Application Volatile CFC-l l CFCl3 75-69-4 Refrigerant CFC-12 Cl32(313 75-71-8 Refrigerant HFC-134a CF3CH2F 81 1-97-2 Refrigerant HCFC-22 CHF2C1 75-45-6 Refrigerant Methoxyflurane CHC13CF20CH3 76-38-0 Anaesthetic Halothane® CF3CHClBr 151-67-7 Anaesthetic Perfluorotributyl-amine (C4F9)3N 311-89-7 Foam blowing agent Nonvolatile Carboxylic Acid Monofluoroacetic acid CHZFC02H 62-74-8 Pesticide Trifluoroacetic acid CF3C02H 76-05-1 Reagent Perfluorooctanoic acid C71:15C02H 335-67-1 Surfactant Sulfonic Acid Trifluoromethane CF3SO3H 1493-13-6 Catalyist/Reagent sulfonic acid Perfluorooctane C8F17SO3H 2795-39-3 Surfactant sulfonic acid 1H, 1 H,2H,2H- C6F13CH 2CH2$O3H 27619-97-2 Surfactant perfluorooctane sulfonic acid Sulfonamide N -acetic-N-ethy1 (C 8F 17SO2N(CH2COOH) 2991 -50-6 Surfactant perfluorooctane (CH2CH3)) sulfonamide Sulfluramid (C8F17802NH(CH2CH3)) 4151-50-2 Insecticide Miscellaneous Polytetrafluoroethylene (~(CF2CF2)n-) 9002-84-0 Teflon® polymer Perfluoropolyether (-(CF(CF3)CF20)n-) 60164-51-4 Lubricant Zony1® Alcohol (C8F17CH2CH20H) 678-39-7 Surfactant 14 Table 2.2. Examples of fluorinated aromatic compounds and their applications. Compound CAS # Application Acifluorfen 50594-66-6 Herbicide; CF3-type Diflubenzuron 35367-38-5 Insecticide Fluometuron 2164-17-2 Herbicide; CF3-type Fluorobenzoates N/A Pharmaceutical and agricultural Fluorouracil 51-21-8 Chemotherapeutic agents Fluotrimazole 3 125 1 -03-3 Fungicide Fluoxydine 671-35-2 Chemotherapeutic agents Flurprimidol 56425-91-3 Plant growth regulator; OCF3-type Flutriafol 76674-21-0 Fungicide Mefluidide 53780-34-0 Herbicide; CF3SO3-type Nucleocidin 2475 1 -69-7 Antibiotic Tefluthrin 79538-32-2 Insecticide Trifluralin 1582-09-8 Herbicide; CF3-type Trifluorobenzoates N/A Pharmaceutical and agricultural Over the past 15 years, the number of fluorine containing agricultural chemicals have grown from 4% to approximately 9% of all agrochemicals and have increased in number faster than non-fluorinated agrochemicals [16]. These compounds are primarily used as herbicides (48%), insecticides (23%), and fungicides (18%) [16]. A summary of the fluorinated agrochemical usage for 1995 in the United States is provided for in Table 2.3. Given the widespread production and use of fluorocarbons, it is perhaps not surprising that organofluorine has been detected in the blood of individuals from the general public as well as industrial workers [8, 38]. For workers handling fluoroorganics, organofluorine levels of 1.0 to 71 ppm have been reported in the blood serum [8]. Individuals who have not been exposed to industrial fluorochemicals have had organic fluorine concentrations from 0.0-0.13 ppm. However, it is unclear whether trace amounts of organic fluorine compounds found in human blood samples are from natural or industrial sources. 15 In the following sections volatile and nonvolatile fluorinated compounds are discussed and a current assessment of the fate and the effect of these molecules in the biosphere is provided. Table 2.3. Fluorinated chemical application for crop protection in US, 1995. Calculated from the Agricultural Chemical Usage - 1995 Field Crops Summary, published by the USDA. (Note: this data only includes applications to corn, cotton, soybeans, and wheat for major producing states) Fluorinated Agrochemical Total Applied Type of Pesticide (1000 kg) Acifluorfen 675 Herbicide Bifenthrin 43 Insecticide Cyfluthrin 59 Insecticide ‘ Diflubenzuron 56 Insecticide Ethalfluralin l 19 Herbicide Fluazifop-P-butyl 200 Herbicide Flumetsulam 48 Herbicide Fluometuron 1273 Herbicide Fomesafen 286 Herbicide Lactofen 1 20 Herbicide Norflurazon 470 Herbicide Oxyfluorfen 20 Herbicide Primisulfuron 1 9 Herbicide Tefluthrin 1 3O Insecticide Trifluralin 6570 Herbicide Total 10,088 FATE AND EFFECTS OF VOLATILE FLUORINATED ORGANICS Chlorofluorocarbons Volatile fluorinated organics include the chlorofluorocarbons (CFCs), the hydrochlorofluorocarbons (HCFCs), the hydrofluorocarbons (HFCs), halothane, fluorinated ethers, and fluorinated amines. Chemical formulas, atmospheric lifetime 16 estimates, and production values for some of these compounds are provided in Table 2.4. CFCs have long been used as refrigerants and aerosols in industrial processes and domestic products. In 1974, CFCs were implicated as agents of depletion of stratospheric ozone by Molina and Rowland [66] and more recently as contributors to global warming [33]. As a result, worldwide production of the CFCs is being phased out under the terms of the Montreal Protocol and its amendments. Nevertheless, CFCs continue to be released into the environment due to past production and continued use. The major sink for CFCs is expected to be stratospheric oxidation. In aerobic aquatic environments, CFCs are recalcitrant, but they are transformed in anaerobic soils and sediments, as well as anaerobic aquatic environments [22, 50, 55, 56, 57, 58, 88, 91]. The expected anaerobic degradation products are hydrochlorofluorocarbons (HCFCs) and hydrofluorocarbons (HFCs). Hydrochlorofluorocarbons and hydrofluorocarbons The phase out of CFC production and use has inspired a major research effort to assess the environmental fate of the CFC alternatives. The HCFC and HFC alternatives are one- and two-carbon aliphatics, similar in structure and physical properties to the CFCs, but containing one or more hydrogen atoms. Tropospheric oxidation is expected to be the most significant sink for the HCFCs and the HFCs. The presence of hydrogen makes HCFCs and HFCs more susceptible to tropospheric oxidation than the CFCs and therefore less likely to enter the stratosphere. In the troposphere, HCFCs and HFCs are oxidized by hydroxyl radicals, yielding HF, C02, and HCl (in the case of HCFCs) and in some cases trifluoroacetic acid (TFA) [1, 36]. Some volatile fluorinated organics (HCFC-134a, HFC-123, HFC 143a, and halothane) are oxidized by hydroxyl radicals to TFA [62, 109, 110]. It is likely that TFA is produced biochemically when these compounds are oxidized by monooxygenase activities, such as cytochrome P450 in the 17 liver and methane monooxygenase [17, 59, 70]. Rainfall is believed to be the primary mechanism for removal of TFA from the atmosphere [87, 100, 101, 110]. A minor sink for HCFCs and HFCs is biochemical reduction or oxidation in aquatic systems. Lesage et al. [56] reported reductive transformation of HCFC-123a to chlorotrifluoroethane (HCFC-133 and HCFC-l33b) under methanogenic conditions. Oremland et al. [73] have also reported reductive dechlorination of HCFC-123 (CF3CHC12) to chlorotrifluoroethane under anaerobic conditions and degradation of HCFC-21 (CHFC12) under both aerobic and anaerobic conditions. The products of HCFC-21 oxidation in these experiments were not investigated [73]. Oxidative transformations mediated by monooxygenases are known, but these reactions proceeded slowly compared to the reactions with analogous chlorinated compounds. Oxidative pathways are hypothesized for the transformation of HFC-134a (1,1,1,2- tetrafluoroethane) by rat liver microsomes [70]. DeFlaun et a1. [21] reported oxidative transformation of HCFC-21 (CHFC12), HCFC-l4lb (CFC12CH3), HCFC-131 (CFC12CH2C1), and HFC-143 (CHF2CH2F) by Methylosinus trichosporium OB3b. HCFC-123 (CF3CHC12), HCFC-142b (CF2C1CH3), HFC-134a (CF3CH2F), HFC-134 (CHFzCHFz), and CFC-ll (CFC13) were not degraded. Chang and Criddle [17] observed slow and limited oxidation of HCFC-123, HFC-134a, and HCFC-142b by a defined methanotrophic consortium. They also reported oxidation of HCFC-22 (CHF2C1), with indirect evidence of product toxicity for HCFC-22 transformation and production of TFA from HFC-134a. 18 Table 2.4. Fluorocarbon production worldwide Compound Molecular Estimated Total Cumulative Cumulative Formula Atmospheric World Production for Lifetime21 (metric tons)b years (years) HCFC-22 CHFzCl 6.7 3,602,365 1 14,536 1970 to 1994 HFC-134a CF3CFH2 14 85,717 1 1,194 1990 to 1994 HCFC-l41b CFClzCH3 7.1 139,382 1 1,479 1990 to 1994 HCFC-142b CF2C1CH3 17.8 193,324 1 1,177 1981 to 1994 CFC-11 CFC13 60 8,583,266 1 15,398 1931 to 1994 CFC-12 CF2C12 105 11,195,032 1 19,542 1931 to 1994 CFC-113 CF2C1CFC12 90 2,244,779 1 8,797 1980 to 1994 CFC-114 CF2C1CF2C1 185 185,111 1 923 1980 to 1994 CFC-115 CF3CF2C1 380 167,459 1 684 1980 to 1994 a Wallington EST 28:320A-326A, and [26]. b AFEAS Report: Production, sales, and atmospheric release of fluorocarbons through 1994 [3]. Fluorinated Anesthetics Methoxyflurane (2,2—dichloro-1,l-difluoroethyl methyl ether) is used very little in human medicine, but it is used in some veterinary applications [30]. The environmental fate of methoxyflurane is not known, but likely involves tropospheric oxidation by hydroxyl radicals. In rats, methoxyflurane is metabolized by two different pathways [61]. The first is an o-demethylation reaction with release of fluoride to form dichloroacetic acid. The second is hydroxylation at the B-carbon with release of chloride and the formation of methoxydifluoroacetic acid. Halothane (l-bromo-l-chloro-2,2,2-trifluoroethane), a bromofluoroalkane is used as a human anesthetic and as a fire extinguishing agent. It is oxidized in the troposphere to l9 TFA. In fact, TFA currently detected in the environment is thought to have largely originated from tropospheric oxidation of halothane [62]. FATE AND EFFECTS OF NONVOLATILE F LUORINATED ORGANICS Trifluoromethyl substituted aromatics The majority of the fluorinated organics used in agricultural applications are trifluoromethyl-substituted aromatics (54.5%) [16]. Of these, trifluralin (1,1,1-trifluoro- 2,6-dinitro-N,N-dipropyl-p-toluidine), a pre-emergent herbicide, is the most commonly used in the United States. Williams et al. [111] observed reduction of the nitro groups of trifluralin by rumen microflora, but no loss of the trifluoromethyl group and no cleavage of the ring was observed. Trifluralin had little effect on rumen microbe populations, as determined by volatile fatty acid production and endogenous gas evolution [1 1 1]. The pre-emergent herbicides norflurazon and fluridone, both trifluoromethyl-containing aromatics, inhibit carotenoid biosynthesis [16]. This inhibition causes overoxidation of chlorophyll and subsequently the loss of the ability to photosynthesize. Flurprimidol, a fluorine containing plant growth regulator, interferes with Gibberellin (plant growth hormone) biosynthesis [16]. Fluotrimazole and the structurally similar flutriafol, are fungicides that weaken the cell membrane by blocking the carbon-1,4 alpha demethylation step in ergosterol biosynthesis [114]. Several trifluoromethyl-substituted aromatics have been detected in sediments and in fish from the Niagara River and Lake Ontario [52]. Various trifluoromethyl-substituted polychlorinated biphenyls, dichloro(trifluoromethyl)-benzophenone, and 20 dichloro(trifluoromethyl)-difluorodiphenylmethane were found. These compounds originated at a dump site containing 55,000 tons of halogenated waste, of which 10% was from the production of 4-chloro-(trifluoromethyl)benzene [52]. Dichloro(trifluoromethyl)-difluorodipheny1methane was present in fish at concentrations as high as 850 ng/g and in sediments from a creek near the dump site at concentrations as high as 35,000 ng/g. The trifluoromethyl-substituted PCBs are believed to bioaccumulate and partition into the sediment more effectively than non- fluorine containing PCBs based on octanol/water partition coefficients (log Kow from 6.8 to 9.0) [52]. The larnpricide, 3-trifluoromethyl-4-nitrophenol, has been used since 1958 to combat the sea lamprey problem in the Great Lakes basin. It is released into rivers and streams containing lamprey larvae. From 1991 through 1995 the five year average use of 3- trifluoromethyl-4-nitrophenol was 40,823 kg/year (active ingredient) [20]. Carey and Fox [14] observed defluorination of 3-trifluoromethyl-4-nitrophenol by photolysis, but reported that only 15% was degraded assunring that complete defluorination would yield 3 moles of fluoride. Fluorinated aromatics Diflubenzuron (1-(4-chlorophenyl)-3-(2,6-difluorobenzoyl) urea), a urea-based larvicide, inhibits chitin synthesis and the molting process in a broad spectrum of insects. Four fungal isolates (F usarium sp., Penicillium sp., Rhodotoruia sp., and Cephalosporium sp.) are capable of degrading diflubenzuron [89]. The proposed pathway for this transformation is through 4-chlorophenylurea and 2,6-difluorobenzoic acid. Although 4-chlorophenylurea is completely metabolized, 2,6-difluorobenzoic acid is persistent and lethal to soil microbes [89]. Several investigators have also shown that diflubenzuron affects non-target aquatic organisms [112]. 21 Several bacterial isolates defluorinate fluorobenzoic acids [28, 29, 40, 71, 83, 84, 86]. The general pathway for degradation of fluorobenzoic acids is attack by a 1,2- dioxygenase followed by decarboxylation to yield fluorocatechols. These catechols are then subject to ring cleavage followed by defluorination. Fluorinated sulfonamides Sulfluramid (N-ethylperfluorooctane sulfonamide), a fluorinated insecticide used to control cockroaches and ants, is deethylated to perfluorooctane sulfonamide in rats, dogs, and rabbit renal mitochondria [5, 45, 60, 85]. Perfluorooctane sulfonamide has not been shown to undergo further transformation, but will most likely be converted to perfluorooctane sulfonic acid (PFOSA; C3F17SO3H) which is believed to be highly recalcitrant. Schnellman et al. [85] demonstrated that perfluorooctane sulfonamide and sulfluramid are potent uncouplers of oxidative phosphorylation in rabbit renal mitochondria. They also reported that the metabolite perfluorooctane sulfonamide was three times more potent than sulfluramid at uncoupling oxidative phosphorylation. Other fluorinated sulfonamides have demonstrated delayed action toxicity in red imported fire ants [106]. This delayed action toxicity allows the insecticide to be applied in baits which are taken back to the colony by foraging members that distribute it throughout the ant colony, allowing for better control of ants. Fluorinated sulfonic acid Perfluorinated sulfonic acids are used as industrial surfactants and as catalysts depending on their chain length. Trifluoromethane sulfonic acid (triflic acid; CF 3SO3H) is an excellent oligomerization/polymerization catalyst. Triflic acid is one of the strongest organic acids known, has great thermal stability, does not release 22 fluoride in the presence of strong nucleophiles, and resists both oxidation and reduction [92]. Perfluorooctane sulfonic acid (PFOSA) also has excellent chemical and thermal stability. PFOSA is important commercially as a surfactant and as a precursor of other fluorinated surfactants [2]. Unfortunately, production values for most of the perfluorinated compounds are considered proprietary information by manufacturers and are not disclosed to the public [34]. Perfluorooctane sulfonic acid and triflic acid are resistant to biological attack. However, a surfactant sirrrilar to PFOSA, lH,lH,2H,2H- perfluorooctane sulfonic acid (H-PFOSA), was partially degraded by a Pseudomonad under aerobic and sulfur-limiting conditions yielding 1-2 moles fluoride per mole of H- PFOSA (Chapter 3). The degradation of H-PFOSA produced several volatile fluorinated compounds that have not yet been identified. 2,2,2-Trifluoroethane sulfonate (TES) was also partially degraded with equimolar release of fluoride (Chapter 3). Another fluorinated sulfonate, difluoromethane sulfonate (DFMS; CHFZSO3Na), was completely metabolized by this Pseudomonad yielding stoichiometric amounts of fluoride (Chapter 3). Transformation of DFMS, TES and H-PFOSA was subsequently observed with Bacillus subtilus and Escherichia coli (Chapter 5). However, E. coli was not capable of utilizing H-PFOSA. Evidence of H-PFOSA degradation was also observed in soil incubations. This and other work suggests that the transformation of fluorinated sulfonates requires the presence of hydrogen. Although PFOSA is resistant to metabolism, it is not biologically inactive. For example, PFOSA was shown to inhibit gap junction intercellular communication (GJIC) in rat liver epithelial cells cultured in vitro [105]. In addition, Gadelhak [37] showed that perfluorooctane sulfonic acid was an uncoupler of phosphorylation in rat liver mitochondria. Although PFOSA alone was not as potent of an uncoupler as perfluorooctane sulfonamide, when PFOSA was ion-paired with various monoamines, 23 polyamines, and phospholipids, the effect of uncoupling was in some instances as high as that of perfluorooctane sulfonamide [37]. Fluorinated alcohols While some of the perfluorinated organics undergo limited biotransformation, none undergo extensive defluorination. An example of a highly fluorinated molecule that has shown some limited defluorination is lH,lH,2H,2H-perfluorodecanol. lH,lH,2H,2H- perfluorodecanol was metabolized first to 2H,2H-perfluorodecanoic acid and then to perfluorooctanoic acid (PFOA) in adult male rats [47]. Hagen et a1 [47] suggest that the overall reaction is production of PFOA with the release of 2 moles of fluoride per mole of 1H,1H,2H,2H-perfluorodecanol. PFOA is metabolically stable in rats [72] and has been found in the blood serum of humans [47]. Fluorinated carboxylic acids Monofluoroacetate (MFA) is one of the most toxic substances known, based on a lethal dose (LD50) of 0.7-2.1 mg/kg for man [6]. Its toxicity is due to "lethal synthesis" of fluorocitrate which inhibits the aconitase enzyme of the Kreb's cycle [77] although recent investigations implicate fluorocitrate as a "suicide" substrate instead of a competitive inhibitor [19]. Given that certain plants can produce MFA, it is not surprising that several microorganisms can metabolize MFA. Pseudomonads and other bacteria, as well as some fungi, have been shown to grow with MFA and monofluoroacetamide (a systemic pesticide and rat poison) as a carbon source aerobically [39, 40, 41, 42, 53, 64, 65, 97, 108]. The first step in degradation of MFA has been shown to be a hydrolytic attack of the carbon-fluorine bond yielding glycolic acid [39]. Many of the same organisms are also capable of growth on and 24 defluorination of MFA under denitrifying conditions (Chapter 6). Gregg et al. [43] demonstrated defluorination of MFA by genetically modified rumen bacteria Butyrivibrio fibrisolvens. This was accomplished by transferring the plasmid responsible for MFA defluorination from Moraxella sp. strain B into B. fibrr'solvens. We know of only one other report of anaerobic MFA degradation. Visscher et al. [107] reported reductive dehalogenation of MFA to acetate, under methanogenic conditions, but this transformation was not reproducible in subsequent investigations [74]. MFA has been shown to inhibit methanogenesis in anaerobic digestor sludge, rumen fluid, and freshwater mud [27]. Another fluorinated carboxylate, the herbicide flupropranate (CHFzCFzCOzH), is persistent in soils [113]. Other fluorinated organic acids of industrial significance include perfluorooctanoic acid (PFOA), perfluorodecanoic acid (PFDA), and trifluoroacetic acid (TFA). No evidence of PFOA or PFDA transformation has been reported. Although, PFOA and PFDA are not metabolized they have been found to inhibit gap junction intercellular communication (GJIC) in rat liver epithelial cells and human kidney epithelial cells at concentrations of 100 and 250 uM respectively [23]. The inhibition of GJIC has been implicated in tumor promotion during carcinogenesis [49, 104, 116], teratogenesis [103], and reproductive dysfunction [9, 102, 117]. Significant hydrolytic defluorination has so far been observed only for monofluorinated molecules. Another biotransformation of importance is decarboxylation. Meyer and O'Hagan have demonstrated decarboxylation of 3-fluoropyruvate to MFA by cell-free extracts of D. cymosum [65]. In addition, Chauhan et a1 [18], have demonstrated decarboxylation of TFA by Azoarcus tolulyticus TOL-4. An intensive international effort is currently underway to explore the fate and effects of TFA, and some results of this effort are summarized in the next section. 25 A CASE HISTORY - TRIFLUOROACETIC ACID Three of the chlorofluorocarbon (CFC) alternatives, HCFC-123, HCFC-124, and HFC- 134a are expected to yield TFA as an atmospheric degradation product (Figure 2.2) [25, 100, 109, 110]. As a result, TFA is expected to become a ubiquitous low-level global contaminant. Estimates based on 100% conversion of HCFC-123, and HCFC-124 and 20% conversion of HFC-134a to TFA predict 0.1 rig/L TFA in rainwater [13] although it is possible that TFA levels in urban areas could reach levels as high as 2 to 20 ug/L [87]. Average TFA concentrations of 50 pg/m3 in air, 100 ng/L in rain water, and 140 ng/L in surface waters have been reported [35]. TFA is concentrated in plant leaves and it has the potential to accumulate in ecosystems such as saline lakes, vernal and temporary rain pools, closed basin lakes, playa lakes, and aestival ponds and certain prairie lakes that exhibit high evaporative potential [100, 101]. Environmental factors such as climate, geology, topography, biota, and time are factors affecting evapoconcentration of solutes. TFA adsorbs weakly to most soils, and, therefore, soils are not expected to be a permanent sink for TFA. However, adsorption is favored in low pH soils which are enriched with iron and aluminum oxides [24]. Photo-oxidation of TFA-iron complexes by near-UV radiation in clouds and the ocean was dismissed because of the low concentrations of iron in these environments [13]. Sinks for TFA would include groundwater due to leaching from soils and vegetation due to bioaccumulation [l3]. TFA can be degraded by the enzyme acetyl-CoA synthetase, but rates are low even with high concentrations of enzyme and substrate (> 10 mM) [27]. As a result, this pathway is unlikely in the environment where TFA is at nM levels [27]. Hydrolysis of TFA is also an unlikely mechanism for degradation due to the extreme stability of the carbon-fluorine bond and the shielding effect of three fluorine atoms on the carbon atom. 26 Even small quantities of TFA in solution can influence the pH of unbuffered solutions because TFA behaves like a strong inorganic acid at environmental pH [13]. Bioaccumulation in animals is thought to be unlikely due to a low octanol/water partition coefficient (log Kow = -4.21) [13], but TFA can accumulate in plants through root uptake. Estimates of the concentration factor give approximately 10-32 times the soil concentration in plant leaves with virtually no degradation in the plant [13, 96]. Vascular plants seem to be affected by TFA when bioaccumulation through roots to leaves occurs, however, little toxicity was shown for seed germination [96]. Toxicological tests have been performed on five species of algae. Three species of microalgae (freshwater diatom Navicula, marine diatom, Skeletonema, and freshwater blue-green Anabaena) showed no toxic effects from TFA concentrations up to 1000 mg/L. The freshwater green alga Chlorella sp. also showed no toxicity up to 1200 mg/L. A second freshwater green alga, Selenastrum capricornutum , exhibited toxicity at concentrations as low as 0.36 mg/L [96]. Bott and Standley reported that TFA did not significantly affect the metabolism of acetate by microbial communities at environmentally expected concentrations [11]. However, Visscher et al. reported inhibitory effects on methanogenic activity at TFA concentrations 2 luM [107]. TFA appears to be non-mutagenic in bacteria [10, 81]. The LD 50 of TFA is reported to be between 200-400 mg/kg (oral exposure to rats), and sodium TFA is only slightly toxic when administered intraperitoneally to mice with no deaths at doses up to 5000 mg/kg [81]. Trifluoroacetic acid does not interfere with homeostasis of rat liver epithelial cell by inhibiting gap junction intercellular communication (GJIC) [105]. 27 Although defluorination of monofluoroacetate by bacteria is well established, few reports of biodegradation of trifluoroacetic acid are available. Visscher et al. [107] reported reductive defluorination of TFA under methanogenic and sulfate-reducing conditions. They report that TFA is sequentially defluorinated to difluoroacetic acid (DFA), monofluoroacetic acid, and finally to acetic acid. Under aerobic conditions, Visscher et al [107] observed production of fluorofonn (CHF3). However, these results with TFA were not reproducible in subsequent experiments leading the investigators to conclude that TFA is widely recalcitrant to biodegradation [74]. Chauhan et al. [18] have demonstrated that Azoarcus tolulyticus TOL—4 is capable of decarboxylating TFA. Cells were grown under denitrifying conditions with toluene as the carbon source and then incubated aerobically without nitrate. No fluoride or volatile fluorinated products were detected [18]. The fate of the trifluoromethyl moiety of TFA is unknown. This is also true of most other trifluoromethyl-substinrted compounds. Photolysis of the trifluoromethyl moiety has been reported for trifluoromethyl-benzoate and 3- trifluoromethyl-4-nitrophenol [14, 95]. Taylor et al. [95] observed equimolar concentrations of fluoride from trifluoromethyl-benzoate. Carey and Fox [14] observed defluorination of 3-trifluoromethyl-4-nitrophenol by photolysis, but reported that only 15% was degraded assuming that complete defluorination would yield 3 moles of fluoride. In both of these cases there appears to be only partial defluorination of the trifluoromethyl moiety. Of particular interest is the fate of these trifluoromethyl grbups and their photo-degradation products and the role of photolysis as a sink for these compounds. In the absence of photolysis, it seems plausible that the trifluoromethyl group remains intact during biotransformation of trifluoromethyl-substituted aromatics, possibly contributing to the TFA load on the environment. 28 Figure 2.2. Biogeochemical cycling of organofluorine compounds. 29 cozmsfizoomco mESomtsm 8.885305 A: 85552 can: + ... n .. r . 86382 a. M. M 28.8ch 8855:: w_mo_Eo;o m_mo_Eocoocom 0 mm: m w: . a _ E 9: 8555:: 88:52“. am‘ Do mm: ozmchn cozmecoawcmtofi mOuI 6 903993 5: m mOuo n 92889. O SON A--- o + 8 as NO + ..o A--- O + .06 NO + .06 A--- m0 + ..o Ono + ._o A--- >3 + mono 2023993 30 SUMMARY AND CONCLUSIONS It is clear that fluorinated molecules are unique with regard to their physical, chemical, and biological properties and they do not fit with the usual paradigm of chemicals of environmental concern. For example, the fluorinated surfactants have distinctly different characteristics to those of their hydrocarbon counterparts. Fluorinated surfactants are simultaneously oleophobic and hydrophobic, but solubilizing moieties such as carboxylic acids, sulfonic acids, phosphates, and quaternary ammonium groups can change their solubility in water [4, 54]. In addition, the perfluorinated alkyl chain is more "rigid" due to fluorine atoms on the molecule [46]. This rigidness almost certainly interferes with molecule/enzyme interactions, protecting fluorocarbon molecules from biological attack. Currently, little is known of the sorption/partitioning prOperties of fluorinated organics. Instances in which the carbon-fluorine bond is ruptured by direct attack are only known for a rarely observed reductive defluorination of tri-, di-, and monofluoroacetate [107] and for hydrolytic defluorination of monofluorinated organics [39]. Reductive defluorination seems to require extreme and uncommon reducing conditions, and so far has been observed only under methanogenic conditions that proved impossible to replicate [74]. In addition, hydrolytic defluorination of carbon atoms with two or more fluorine substituents appears to be too slow to be of environmental significance [27]. More often, transformation of highly fluorinated organics requires attack at functional groups or bonds attached to the fluorinated moiety. Attack on adjacent functional groups can be hydrolytic, oxidative, or reductive, and can result in decarboxylation, desulfonation, deamination, and fluoride elimination. Although there have been instances of partial defluorination of perfluoroalkyl chains, complete mineralization in the biosphere has not been reported. The products of these 3] partial transformations are of interest with regard to their lifetimes in the environment and their effects on the biosphere directly or indirectly (see Figure 2.2). For example, it has been shown that several nonvolatile compounds such as sulfluramid and 1H,lH,2H,2H-perfluorooctane sulfonic acid (Chapter 3 and 5) are transformed to volatile fluorinated compounds [5]. In addition, volatile compounds such as HCFC- 123, HCFC-124, and HFC-134a are expected to yield TFA (a nonvolatile fluorinated compound) as a tropospheric degradation product [25, 100, 109, 110]. The CFCs are degraded in the stratosphere and are responsible for ozone depletion and global warming. While several of the fluorinated agrochemicals are reported to "dissipate" in soils, mineralization of these chemicals has not been demonstrated in most cases. Similarly, the environmental fate of the trifluoromethyl group, which is utilized in many fluorinated organic compounds, is largely unknown. The potential for TFA production from these trifluoromethyl-substituted aromatics should be investigated. The biogeochemical cycling of fluorinated compounds is just now being discovered and is not yet clearly understood. The stability that makes fluorinated compounds desirable for commercial use also makes them potentially significant environmental contaminants due to their persistence. Very little research into the toxicological effects of these compounds and their byproducts on animals and plants has been done. Therefore, caution is warranted when evaluating these compounds because although they may not be chemically or even biochemically reactive, they may still be biologically active. To successfully study the fate and the effects of fluorinated compounds in the environment, sufficiently sensitive analytical methods need to be developed and published. Currently, very few methods are available to measure or detect nonvolatile fluorinated organics in environmental samples. In addition, many of the perfluorinated 32 organics are produced as mixtures of straight and branched homologues making analysis even more challenging. Although traditional analytical methods for determining hydrocarbon analogues have been used, most environmental samples have co-eluting ions and metabolic products that confound analysis. Inorganic fluoride measurement, by ion-selective electrode or ion chromatography, has been an effective means of indirectly detecting transformation of these compounds, but this method requires defluorination which may not occur with many perfluorinated compounds. Destructive conversion of organofluorine compounds to inorganic fluoride ions can provide information on the ratio of fluorine in the samples, but this method cannot distinguish between different homologues. Another indirect method is monitoring of oxygen consumption during aerobic incubation. Methane and carbon dioxide production can be measured with anaerobic incubations. Again these methods fail to answer key questions about the fate of fluorinated compounds. Real progress on these fate issues may require synthesis of radiolabelled compounds. It is likely that industrial, agricultural, domestic, and medical applications for fluorine containing compounds will continue to increase in the coming years. As we continue to use and manufacture molecules containing carbon-fluorine bonds, several questions arise pertaining to the fate of such compounds. Are they recalcitrant in the environment or do they possess substituents, such as the trifluoromethyl group, which persists as mobile, recalcitrant functional groups? In which environments do they accumulate? What are the products of biological or abiotic reactions? What is the fate of reaction products? What is the toxicity and mutagenicity of such reaction products? How can we accurately measure and detect these compounds in environmental samples? Clearly, considerable research is needed to better understand the fate and effects of fluorinated organics in the biosphere. REFERENCES World Meteorological Organization. 1989. Scientific Assessment of Stratospheric Ozone. Global Ozone Research and Monitoring Project, Report No. 20, Vol I, 11. Geneva, Switzerland. Abe, T., and S. Nagase. 1982. Electrochemical fluorination (Simons process) as a route to perfluorinated organic compounds of industrial interest, pp. 19-44. In R. E. Banks (ed.), Preparation, properties, and industrial applications of organofluorine compounds. John Wiley & Sons, New York. AFEAS. 1996. Production, sales, and atmospheric release of fluorocarbons through 1994. Alternative Fluorocarbons Environmental Acceptability Study (AFEAS). Allison, M. C. 1987. Fluorochemical Surfactants, In (ed.), Industrial Applications of Surfactants (Special Publication 59). Royal Society of Chemistry, London. Arrendale, R. F., J. T. Stewart, R. Manning, and B. Vitayavirasuk. 1989. Determination of GX-071 and its major metabolite in rat blood by cold on- column injection capillary GC/ECD. J. Agr. Food. Chem. 37: 1 130-1 135. Atzert, S. P. 1971. A review of sodium monofluoroacetate (compound 1080) its properties, toxicology, and use in predator and rodent control. US. Fish and Wildlife Service, Special Science Report on Wildlife No. 146. Banitt, E. H., W. E. Coyne, K. T. McGurran, and J. E. Robertson. 1974. Monofluoromethanesulfonanilides. A new series of bronchodilators. J. Med. Chem. 17:116-120. Belisle, J. 1981. Organic Fluorine in Human Serum: Natural versus industrial sources. Science 212:1509-1510. Bergoffen, J., et al. 1993. Connexin mutations in X-linked Charcot-Marie- Tooth disease. Science 262:2039-2042. 33 10. ll. 12. 13. 14. 15. 16. l7. 18. 34 Blake, D. A., M. C. DiBlasi, and G. B. Gorden. 1981. Absence of mutagenic activity of trifluoroethanol and its metabolites in Salmonella typhimurium. Fundam. Appl. Toxicol. 1:415-418. Bott, T. L., and L. J. Standley, The effect of TFA on biological processes in streams, Proceedings of the 1994 AFEAS Workshop on the environmental fate of trifluoroacetic acid (AFEAS Administrative Organization: SPA-AFEAS, Inc., Miami Beach, Florida, 1994), Business Communications Company Inc. 1995. RC-193 Performance fluorine chemicals and polymers: Future prospects. Business Communications Company, Inc., Norwalk, CT. Calamari, D., Preliminary assessment of environmental distribution and fate and environmental toxicity of the HFC/HCFC degradation product: Trifluoroacetic acid, Proceedings of the 1994 AFEAS Workshop on the environmental fate of trifluoroacetic acid (AFEAS Administrative Organization: SPA-AFEAS, Inc., Miami Beach, Florida, 1994), Carey, J. H., and M. E. Fox. 1981. Photodegradation of the lampricide 3- trifluoromethyl-4-nitrophenol (TFM) 1. Pathway of the direct photolysis in solution. J. Great Lakes Res. 7:234-241. Carey, J. H., M. E. Fox, and L. P. Schleen. 1988. Photodegradation of the lampricide 3-trifluoromethyl-4-nitrophenol (TFM) 2. Field confirmation of direct photolysis and persistence of formulation impurities in a stream during treatment. J. Great Lakes Res. 14:338-346. Cartwright, D. 1994. Recent developments in fluorine-containing agrochemicals, pp. 237-257. In R. E. Banks, B. E. Smart, and J. C. Tatlow (ed.), Organofluorine chemistry: Principles and commercial applications. Plenum Press, New York. Chang, W. K., and C. S. Criddle. 1995. Biotransformation of HCFC-22, HCFC-142b, HCFC-123 and HFC-134a by methanotrophic mixed culture MMl. Biodegradation 6: 1-9. Chauhan, S., J. M. Tiedje, and C. S. Criddle. Decarboxylation of trifluoroacetic acid by Azoarcus tolulyticus TOL-4. Manuscript in preparation. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 35 Clarke, D. D. 1991. Fluoroacetate and fluorocitrate: Mechanism of action. Neurochemical Research 16: 1055-1058. Great Lakes Fishery Commission. 1996, personal communication. DeFlaun, M. F., B. D. Ensley, and R. J. Steffan. 1992. Biological oxidation of hydrochlorofluorocarbons (HCFC's) and Hydrofluorocarbons (HFC's) by a methanotrophic bacterium. Biotechnology 10:1576-1578. Denovan, B. A., and S. E. Strand. 1992. Biological degradation of chlorofluorocarbons in anaerobic environments. Chemosphere 24:935-940. Deocampo, N. D., B. L. Upham, and J. E. Trosko. 1996. The role of gap junction communication in the toxicity of quadricyclane, perfluorodecanoic acid, perfluorooctanoic acid and jet fuels JP-8 and JP-4. Fundam. Appl. Toxicol. (suppl) 30: 1065. Driscoll, C. T., The transport and fate of trifluoroacetate (TFA) in the environment, Proceedings of the 1994 AFEAS Workshop on the environmental fate of trifluoroacetic acid (AFEAS Administrative Organization: SPA-AFEAS, Inc., Miami Beach, Florida, 1994), Edney, E. 0., B. W. Gay, and D. J. Driscoll. 1991. Chlorine initiated oxidation studies of hydrochlorofluorocarbons: results for HCFC-123 and HCFC-l4lb. J. Atmospheric Chemistry 12: 105-120. Elliot, A. J. 1994. Chlorofluorocarbons, pp. 145-157. In R. E. Banks, B. E. Smart, and J. C. Tatlow (ed.), Organofluorine chemistry: Principles and commercial applications. Plenum Press, New York. Emptage, M., Biological and chemical degradation of fluoroacetates, Proceedings of the 1994 AFEAS Workshop on the environmental fate of trifluoroacetic acid (AFEAS Administrative Organization: SPA-AFEAS, Inc., Miami Beach, Florida, 1994), Engesser, K.-H., M. A. Rubio, and H.-J. Knackmuss. 1990. Bacterial metabolism of side-chain-fluorinated aromatics: unproductive meta-cleavage of 3-trifluoromethylcatechol. Appl. Microbiol. Biotechnol. 32:600-608. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 36 Engesser, K. H., G. Auling, J. Busse, and H.-J. Knackmuss. 1990. 3- Fluorobenzoate enriched bacterial strain FLB 300 degrades benzoate and all three isomeric monfluorobenzoates. Arch. Microbol. 153:193-199. Ferstandig, L. L. 1995. Fluorinated anesthetics, pp. 1133-1137. In M. Hudlicky, and A. E. Pavlath (ed.), Chemistry of Organic Fluorine Compounds 11; A critical review. American Chemical Society, Washington DC. Fielding, H. C. 1979. Organofluorine surfactants and textile chemicals, pp. 214- 232. In R. E. Banks (ed.), Organofluorine chemicals and their industrial applications. Ellis Horwood Ltd., Chichester. Filler, R. 1993. Fluoromedicinal chemistry - An overview of recent developments, pp. 1-23. In R. Filler, Y. Kobayashi, and L. M. Yagupolskii (ed.), Organofluorine compounds in medicinal chemistry and biomedical applications. Elsevier, Amsterdam. Fisher, D. A., C. H. Hales, W. C. Wang, M. K. W. Ko, and N. I). Sze. 1990. Model calculations of the relative effects of CFCs and their replacements on global warming. Nature 344:513-516. Flynn, R. M. 1994. Fluoroethers and fluoroamines, pp. 525-534. In J. I. Kroschwitz, and M. Howe-Grant (ed.), Kirk-Othmer Encyclopedia of Chemical Technology. John Wiley & Sons, New York. Frank, H., A. Klein, and D. Renschen. 1996. Environmental Trifluoroacetate. Nature 382:34. Franklin, J. 1993. The atmospheric degradation and impact of l,1,1,2- tetrafluoroethane (hydrofluorocarbon 134a). Chemosphere 27:1565-1601. Gadelhak, G. G., Dissertation, Michigan State University (1992). Gilliland, F. D., and J. S. Mandel. 1996. Serum perfluorooctanoic acid and hepatic enzymes, lipoproteins, and cholesterol: A study of occupationally exposed men. Am. J. Ind. Med. 29:560-568. Goldman, P. 1965. The enzymatic cleavage of the carbon-fluorine bond in fluoroacetate. J. Biol. Chem. 240:3434-3438. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 37 Goldman, P. 1971. Enzymology of carbon-halogen bonds, pp. 147-165. In (ed.), Degradation of synthetic organic molecules in the biosphere. National Academy of Sciences, Washington DC. Goldman, P., and G. W. A. Milne. 1966. Carbon-fluorine bond cleavage; 11. Studies on the mechanism of the defluorination of fluoroacetate. J. Biol. Chem. 241:5557-5559. Goldman, P., G. W. A. Milne, and D. B. Keister. 1968. Carbon-halogen bond cleavage; III. Studies on bacterial halidohydrolases. J. Biol. Chem. 243:428-434. Gregg, K., et al. 1994. Detoxification of plant toxin fluoroacetate by a genetically modified rumen bacterium. Bio/Technology 12:1361-1365. Gribble, G. W. 1994. The natural production of chlorinated compounds. Environ. Sci. Technol. 28: 310A-3 19A. Grossman, M. R., M. E. Mispagel, and J. M. Bowen. 1992. Distribution and tissue elimination in rats during and after prolonged dietary exposure to a highly fluorinated sulfonamide pesticide. J. Agr. Food. Chem. 40:2505-2509. Guo, W., T. A. Brown, and B. M. Fang. 1991. Micelles and aggregates of fluorinated surfactants. J. Phys. Chem. 95:1829-1836. Hagen, D. F., J. Belisle, J. D. Johnson, and P. Venkateswarlu. 1981. Characterization of fluorinated metabolites by gas chromatographic-helium microwave plasma detector- The biotransformation of 1H, 1H, 2H, 2H- perfluorodecanol to perfluorooctanoate. Anal. Biochem. 118:336-343. Harper, D. B., and D. O'Hagen. 1994. The fluorinated natural products. Natural Product Reports 11:123-133. Holder, J. W., E. Elmore, and J. C. Barrett. 1993. Gap junction function and cancer. Cancer Res. 53: 3475—3485. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. 38 Hur, H.-G., M. J. Sadowsky, and L. P. Wackett. 1994. Metabolism of chlorofluorocarbons and polybrorrrinated compounds by Pseudomonas putida G786(pHG-2) via an engineered metabolic pathway. Appl. Environ. Microbiol. 60:4148-4154. Isidorov, V. A. 1990. Organic Chemistry of the Earth's Atmosphere. Springer- Verlag, Berlin. Jafi'e, R., and R. A. Hites. 1985. Identification of new, fluorinated biphenyls in the Niagara River-Lake Ontario area. Environ. Sci. Technol. 19:736-740. Kelly, M. 1965. Isolation of bacteria able to metabolize fluoroacetate or fluoroacetamide. Nature (London) 208:809-810. Kissa, E. 1994. Fluorinated surfactants: synthesis, properties, and applications. Marcel Dekker, Inc, New York. Krone, U. E., and R. K. Thauer. 1992. Dehalogenation of trichlorofluoromethane (CFC-11) by Methanosarcina barkeri. FEMS Microbiol. Lett. 90:201-204. Lesage, S., S. Brown, and K. R. Hosler. 1992. Degradation of chlorofluorocarbon-l 13 under anaerobic conditions. Chemosphere 24:1225- 1243. Lovely, D. R., and J. C. Woodward. 1990. Consumption of Freons CFC-ll and CFC-12 in methane producing aquatic sediments. Eos, American Geophysical Union 71: 1236. Lovely, D. R., and J. C. Woodward. 1992. Consumption of Freons CFC-11 and CFC-12 by anaerobic sediments and soils. Environ. Sci. Technol. 26:925- 929. Ma, T. G., Y. H. Ling, G. D. McClure, and M. T. Tseng. 1990. Effects of trifluoroacetic acid, a halothane metabolite, on C6 Glioma cells. J. Toxicol. Environ. Health 31: 147-158. 60. 61. 62. 63. 64. 65. 66. 67. 68. 69. 39 Manning, R. 0., J. V. Bruckner, M. E. Mispagel, and J. M. Bowen. 1991. Metabolism and dispostion of sulfluramid, a unique polyfluorinated insecticide, in the rat. Drug Metab. Dispos. 19:205-211. Mazze, R., and B. Hill. 1978. Effects of phenobarbital and 3- methylcholanthrene on anesthetic defluorination in Fischer 344 rats. Dmg Metab. Dispos. 6:680-681. McCulloch, A., The mechanisms of atmospheric degradation of HCFCs and HFCs and the formation of trifluoroacetic acid, Proceedings of the 1994 AFEAS Workshop on the environmental fate of trifluoroacetic acid (AFEAS Administrative Organization: SPA-AFEAS, Inc., Miami Beach, Florida, 1994), Mead, R. J., M. G. Feldwick, and J. T. Bunn. 1991. 1,3-Difluoro-2-propanol: a potential replacement for 1080 in fauna management programs in Australia. Wildl. Res. 18:27-37. Meyer, J. J. M., N. Grobbelaar, and P. L. Steyn. 1990. Fluoroacetate- metabolizing Pseudomonad isolated from Dichapetalum cymosum. Appl. Environ. Microbiol. 56:2152-2155. Meyer, J. J. M., and D. O'Hagan. 1992. Conversion of 3-fluoropyruvate to fluoroacetate by cell-free extracts of Dichapetalum cymosum. Phytochemistry 31:2699-2701. Molina, M. J., and F. S. Rowland. 1974. Stratospheric sink for chlorofluoromethanes: chlorine atom-catalysed destruction of ozone. Nature 249:810-812. Moore, G. G. I. 1979. Fluoroalkanesulfonyl Chlorides. J. Org. Chem. 14: 1708- 1711. Moore, G. G. 1., and J. K. Harrington. 1975. Antiinflarnmatory fluoroalkanesulfonanilides. 3. Other fluoroalkanesulfonarnido diaryl systems. J. Med. Chem. 18:386-391. Moore, G. G. 1., L. R. Lappi, J. E. Bauchhuber, and A. C. Conway, Anticonvulsant fluoroalkanesulfonamides, 106th National Meeting of the American Chemical Society (MEDI Abstracts, Chicago, 111., 1970), 70. 71. 72. 73. 74. 75. 76. 77. 78. 79. 40 Olson, M. J., C. A. Reidy, J. T. Johnson, and T. C. Pederson. 1990. Oxidative defluorination of l,1,1,2-tetrafluoroethane by rat liver microsomes. Drug Metab. Dispos. 18:992-998. Oltmanns, R. H., R. Muller, M. K. Otto, and F. Lingens. 1989. Evidence for a new pathway in the bacterial degradation of 4-fluorobenzoate. Appl. Environ. Microbiol. 55:2499-2504. Ophaug, R. H., and L. Singer. 1980. Metabolic handling of perfluorooctanoic acid in rats (40715). Proceedings of the Society for Experimental Biology and Medicine 163:19-23. Oremland, R. S., D. J. Lonergan, C. W. Culbertson, and D. R. Lovely. 1996. Microbial degradation of hydrochlorofluorocarbons (CHCle and CHCIZCF3) in soils and sediments. Appl. Environ. Microbiol. 62: 1818-1821. Oremland, R. s., L. J. Matheson, J. a. Guidetti, J. K. Schaefer, and P. T. Visscher. 1995. Summary of research results on bacterial degradation of trifluoroacetate (TFA), November, 1994-May,1995. USGS. OF 95-0422. Paul, E. A., and P. M. Huang. 1980. Handbook of Environmental Chemistry, In 0. Hutzinger (ed.), Springer Verlag, Berlin. Pauling, L. 1960. The nature of the chemical bond. 3rd. Cornell University Press, Ithaca, New York. Peters, R. 1972. Some metabolic aspects of fluoroacetate especially related to fluorocitrate, pp. 55-76. In (ed.), Carbon-fluorine compounds: Chemistry, biochemistry, and biological activities (A Ciba Foundation Symposium). Associated Scientific Publishers, Amsterdam. Rao, N. S., and B. E. Baker. 1994. Textile finishes and fluorosurfactants, pp. 321-336. In R. E. Banks, B. E. Smart, and J. C. Tatlow (ed.), Organofluorine chemistry: Principles and commercial applications. Plenum Press, New York. Ribbons, D. W., et al. 1987. Biotransformations of fluoroaromatic compounds. J. Fluor. Chem. 37:299-326. 80. 81. 82. 83. 84. 85. 86. 87. 88. 89. 41 Rowland, F. S., and M. J. Molina. 1975. Chlorofluoromethanes in the environment. Rev. Geophys. Space Phys. 13:1-35. Rusch, G. M., Toxicity of trifluoroacetic acid, Proceedings of the 1994 AFEAS Workshop on the environmental fate of trifluoroacetic acid (AFEAS Administrative Organization: SPA-AFEAS, Inc., Miami Beach, Florida, 1994), Saunders, B. C. 1972. Chemical characteristics of the carbon-fluorine bond, pp. 9-32. In K. Elliott, and J. Birch (ed.), Carbon-fluorine compounds; chemistry, biochemistry and biological activities. (A Ciba Foundation Symposium). Associated Scientific Publishers, Amsterdam. Schennen, U., K. Braun, and H.-J. Knackmuss. 1985. Anaerobic degradation of 2-fluorobenzoate by benzoate-degrading, denitrifying bacteria. J. Bacteriol. 161:321-325. Schlomann, M., P. Fischer, E. Schmidt, and H.-J. Knackmuss. 1990. Enzymatic formation, stability, and spontaneous reactions of 4- fluoromuconolactone, a metabolite of bacterial degradation of 4-fluorobenzoate. J. Bacteriol. 172:5119-5129. Schnellmann, R. G., and O. M. Randall. 1990. Perfluorooctane sulfonanride: a structurally novel uncoupler of oxidative phosphorylation. Biochemica Et Biophysica Acta 1016:344-348. Schreiber, A., M. Hellwig, E. Dom, W. Reineke, and H. J. Knackmuss. 1980. Critical reactions in fluorobenzoic acid degradation by Pseudomonas sp. B13. Appl. Environ. Microbiol. 39:58-67. Schwarzbach, S. E. 1995. CFC alternatives under a cloud. Nature 376:297-298. Semprini, L., G. D. Hopkins, P. V. Roberts, and P. L. McCarty. 1990. In-situ biotransformation of carbon tetrachloride, 1,1,l-trichloroethane, Freon-11, and Freon-113 under anoxic conditions. Eos, American Geophysical Union 71:1324. Seuferer, S. L., H. D. Braymer, and J. J. Dunn. 1979. Metabolism of diflubenzuron by soil microorganisms and mutagenicity of the metabolites. Pestic. Biochem. Physiol. 10:174-180. 90. 91. 92. 93. 94. 95. 96. 97. 98. 99. 42 Smart, B. E. 1994. Characteristics of C-F systems, pp. 57-88. In R. E. Banks, B. E. Smart, and J. C. Tatlow (ed.), Organofluorine Chemistry: Principles and Commercial Applications. Plenum Press, New York. Sonier, D. N., N. L. Duran, and G. B. Smith. 1994. Dechlorination of trichlorofluoromethane (CFC-11) by sulfate-reducing bacteria from an aquifer contaminated with halogenated aliphatic compounds. Appl. Environ. Microbiol. 60:4567-4572. Stang, P. J., and M. R. White. 1983. Triflic acid and its derivatives. Aldrichinrica Acta 16:15-22. Suida, J. F., and J. F. DeBernardis. 1973. Naturally occurring halogenated organic compounds. Lloydia 36:107-143. Tamura, T., M. Wada, N. Esaki, and K. Soda. 1995. Synthesis of fluoroacetate from fluoride, glycerol, and beta-hydroxypyruvate by Streptomyces cattleya. J. Bacteriol. 177:2265-9. Taylor, B. F., J. A. Amador, and H. S. Levinson. 1993. Degradation of meta- trifluoromethylbenzoate by sequential microbial and photochemical treatments. FEMS Microbiol. Lett. 110:213-216. Thompson, R. 8., Effects of TFA on algae and higher plants, Proceedings of the 1994 AFEAS Workshop on the environmental fate of trifluoroacetic acid (AFEAS Administrative Organization: SPA-AFEAS, Inc., Miami Beach, Florida, 1994), Tonomura, K., F. Futai, O. Tanabe, and T. Yamaoka. 1965. Defluorination of monofluoroacetate by bacteria; Part 1. Isolation of bacteria and their activity of defluorination. Agr. Biol. Chem. 29: 124-128. Trepka, R. D., et al. 1974. Synthesis and herbicidal activity of fluorinated N- phenylalkanesulfonamides. J. Agr. Food Chem. 22:1111-1119. Trepka, R. D., J. K. Harrington, J. E. Robertson, and J. T. Waddington. 1970. Fluoroalkanesulfonanilides, a new class of herbicides. J. Agr. Food Chem. 18:1176-1177. 100. 101. 102. 103. 104. 105. 106. 107. 108. 109. 110. 43 Tromp, T. K., M. K. W. Ko, J. M. Rodriguez, and N. D. Sze. 1995. Potential accumulation of a CFC-replacement degradation product in seasonal wetlands. Nature 376:327—330. Tromp, T. K., J. M. Rodriguez, M. K. W. Ko, C. W. Heisey, and N. D. Sze, Scenarios for delivery of TFA to the global environment: Model predictions of environmental load, Proceedings of the 1994 AFEAS Workshop on the environmental fate of trifluoroacetic acid (AFEAS Administrative Organization: SPA-AFEAS, Inc., Miami Beach, Florida, 1994), Trosko, J. E., C. Chang, and B. V. Madhukar. 1994. The role of modulated gap junctional intercellular communication in epigenetic toxicology. Risk Analysis 14:303-312. Trosko, J. E., C. Chang, and M. Netzloff. 1982. The role of inhibited cell-cell communication in teratogenesis. Teratog. Carcinog. Mutagen. 2:31-45. Trosko, J. E., C. Jone, and C. C. Chang. 1987. Inhibition of gap junction- mediated intercellular communication in vitro by aldrin, dieldrin, and toxaphene: A possible cellular mechanism for their tumor-promoting and neurotoxic effects. Mol. Toxicol. 1:83-93. Upham, B. L., 1996, personal communication. Vander Meer, R. K., C. S. Lofgren, and D. F. Williams. 1985. Fluoroaliphatic sulfones: a new class of delayed-action insecticides for control of Solenopsis invicta. J. Econ. Ent. 78:1190-1197. Visscher, P. T., C. W. Culbertson, and R. S. Oremland. 1994. Degradation of trifluoroacetate in oxic and anoxic sediments. Nature 369:729-731. Walker, J. R. L., and B. C. Lien. 1981. Metabolism of fluoroacetate by a soil Pseudomonas sp. and Fusarium solani. Soil Biol. Biochem. 13:231-235. Wallington, T. J., M. D. Hurley, J. C. Ball, and E. W. Kaiser. 1992. Atmospheric chemistry of hydrofluorocarbon 134a: Fate of the alkoxy radical CF3CFHO. Environ. Sci. Technol. 26:1318-1324. Wallington, T. J., et al. 1994. The environmental impact of CFC replacements- HFC's and HCFCs. Environ. Sci. Technol. 28:320A-326A. 111. 112. 113. 114. 115. 116. 117. 44 Williams, P. P., and V. J. Feil. 1971. Identification of trifluralin metabolites from rumen microbial cultures. Effects of uifluralin on bacteria and protozoa. J. Agr. Food. Chem. 19: 1 198-1204. Wilson, J. E. H., P. A. Cunningham, D. W. Evans, and J. D. Costlow. 1995. Using grass shrimp embryos to determine the effects of sediment on the toxicity and persistence of diflubenzuron in laboratory microcosms, In J. S. Hughes, G. R. Biddinger, and E. Mones (ed.), Environmental toxicology and risk assessment. American Society for Testing and Materials, Philadelphia. Worthington, C. R., and J. H. Raymond. 1991. The pesticide manual. Worthington, C. R., and J. H. Raymond (ed.). Unwin Brothers Ltd., Old Woking, Surrey. Worthington, P. A. 1987. Synthesis and Chemistry of Agrochemicals, pp. 302. In D. R. Baker, J. G. Fenyes, W. K. Moberg, and B. Cross (ed.), ACS Symposium Series No. 355, Washington DC. Yamanouchi, K., and C. Heldebrandt. 1992. Perfluorochemicals as a blood substitute. Chemtech 22:354-359. Yamasaki, H., and C. C. G. Naus. 1996. Role of connexin genes in growth control. Carcinogenesis 17:1 199-1213. Ye, Y., et al. 1990. The modulation of gap junctional communication by Gossypol in various mammalian cell lines in vitro. Fundam. Appl. Toxicol. 14:817-832. CHAPTER 3 DEFLUORINATION OF ORGANOFLUORINE SULFUR COMPOUNDS BY PSEUDOMONAS SP. STRAIN D2 ABSTRACT Little is known of the potential for biodegradation of sulfur containing fluorinated surfactants. Difluoromethane sulfonate (DFMS, CHF2803Na) was chosen as a representative model compound. A Pseudomonad (strain D2) was isolated that completely defluorinated DFMS under aerobic and sulfur-limiting conditions in minimal medium. Strain D2 was subsequently used to evaluate the potential for biotransformation of the following fluorinated sulfonates: trifluoromethane sulfonate (TFMS; CF3SO3Na), 2,2,2-trifluoroethane sulfonic acid (TES; CF3CHzSO3H), perfluorooctane sulfonate (PFOSA; CgF] 7803 K), and 1H,]H,2H,2H-perfluorooctane sulfonic acid (H-PFOSA; C6F13C2H4803H). Strain D2 was capable of utilizing those compounds containing hydrogen (TES and H-PFOSA), but only partial defluorination was observed. When TES was the sole sulfur source, one mole of fluoride was released per mole of TES transformed. Transformation of H-PFOSA yielded one mole of fluoride per mole of H-PFOSA when incubations were open to the atmosphere and 1.4 moles per mole when incubations were closed to the atmosphere. Eight volatile and fluorinated byproducts of H-PFOSA were detected by GC/mass spectrometry, but volatile transformation products were not detected for TES or DFMS. This is the first report of biotic transformation of fluorinated sulfonates resulting in defluorination. These findings suggest that the initial steps of transformation are linked to sulfur metabolism 45 46 with subsequent defluorination of the molecule. The results also suggest that hydrogen substitution is required for transformation. INTRODUCTION The commercial use of organofluorine compounds has dramatically increased in recent years. Many of these compound are used as propellants, surfactants, agrochemicals, adhesives, refrigerants, fire retardants, and medicines [2, 3, 4, 5, 6, 7, ll, 14, 15, 16, 22, 25]. One category with particularly useful properties is the fluorinated sulfonates. Perfluorinated sulfonates are used as industrial surfactants and as catalysts in synthetic chemistry. Perfluorooctane sulfonate (PFOSA; C3F17SO3Na) has excellent chemical and thermal stability and is important commercially as a surfactant and as a precursor of other fluorinated surfactants and pesticides [1]. Shorter chained perfluorinated compounds, such as trifluoromethane sulfonate (TFMS; CF3SO3H), are used as oligomerization or polymerization catalyst. TFMS is one of the strongest acids known, has great thermal stability, does not release fluoride in the presence of strong nucleophiles, and resists both oxidation and reduction [19]. Difluoromethane sulfonate (DFMS; CHFZSO3N a), 1H,1H,2H,2H-perfluorooctane sulfonic acid (H-PFOSA; C6F13C2H4803H), and 2,2,2-trifluoroethane sulfonic acid (TES; CF3CH2803H) are partially fluorinated analogues of the perfluorinated sulfonates. Because of the apparent stability of these compounds and their potential for accumulation in the environment, it is important to understand their environmental fate and the mechanisms by which they are degraded. The biodegradability of TFMS, TBS, PFOSA, and H-PFOSA was evaluated using a Pseudomonad that is capable of completely defluorinating DFMS under aerobic and sulfur-limiting conditions. 47 MATERIALS AND METHODS Media and chemicals. A defined rrrineral medium containing (in grams per liter): glucose, 2.0; KZHPO4, 3.5; KH2P04, 2.0; NH4C1, 1.0; MgC12°6H20, 0.5; 0.15M CaC12°2H20 stock, 1.0 ml/liter; trace elements stock I and II, 1.0 mllliter. Trace elements stock 1 contained (in grams per liter): FeCl3, 1.36; CoC12-6H20, 0.2; MnC12-4H20, 0.122; ZnClz, 0.07; N azMoO4°2H20, 0.036; NiC1206H20, 0.12; B3(OH)3, 0.062; CuC12-2H20, 0.017. Concentrated H2SO4 was added at 2.5 ml/L to the trace elements solution 1. Trace elements solution 11 contained (in grams per liter): NazSeO3°5H20, 0.006; N aWO4-2H20, 0.033; NazMoO4o2H20, 0.024. To this medium was added the appropriate organofluorine sulfonate as the sulfur source. Sodium difluoromethane sulfonate (DFMS) was provided by 3M, St Paul, MN. Potassium perfluorooctane sulfonate (PFOSA) and lH,lH,2H,2H-perfluorooctane sulfonic acid (H-PFOSA) were obtained from ICN Pharmaceuticals Inc., Costa Mesa, CA. 2,2,2-Trifluoroethane sulfonyl chloride was obtained from Sigma Chemical, St. Louis, MO. Hydrolysis of 2,2,2-trifluoroethane sulfonyl chloride by addition to water and autoclaving yielded trifluoromethane sulfonic acid (TES). All other carbon sources and chemicals were obtained from Sigma Chemical. Growth conditions for Pseudomonas sp. strain D2. Plate colonies of strain D2 were inoculated into 5 ml of nutrient broth (Difco, Detroit, MI) and incubated at 30° C for 24 hours. A 1% inoculum of this was used in above medium (pH of 6.9-7.0) with the specified organofluorine compound as the sole sulfur source. Cells were grown aerobically at 30°C and were shaken on a rotorary shaker at 160 rpm. Fluoride and DFMS analysis. Fluoride and DFMS were measured on a Dionex ion chromatography model 2000i/sp fitted with an IonPac AS4A ion exchange column and a 48 Dionex IonPac AG4A guard column. This system utilizes an anion micromembrane suppressor with a Dionex Conductivity Detector-H (CDM). The eluant was a carbonate buffer (1.8 mM N azCO3 and 1.7 mM N aHCO3) with a flow rate of 2 ml/min. Fluoride concentration was used as an indirect method of detecting biotransformation of the fluorinated sulfonates. Fluoride was also measured using an ion selective electrode (Orion 96-09 BN). A total ionic strength adjuster (TISA) was used to stabilize fluoride measurements in samples. TISA was prepared by adding 57 m1 of glacial acetic acid and 58 g reagent grade sodium chloride into 500 ml of deionized water. This mixture was cooled on ice, adjusted to a pH of 5.0-5.5 with 5M sodium hydroxide, and diluted with water to a final volume of one liter. One part TISA to one part sample was used to measure fluoride. Mass spectrometry and atomic emissions analysis. GCIMS data were obtained using a Hewlett-Packard 5995 series H GC/MS. A DB624 capillary column (30 m x 0.25 mm x 1.4 pm) was used for separation of volatile byproducts (J & W Scientific, Inc., Folsom, CA). Operating conditions were as follows: flow of 30 cm/sec linear velocity, initial temperature 40° C for 4 minutes followed by a 10° C/minute ramp to 200° C, injector temperature of 250° C, transfer line temperature of 225° C. Samples were taken by injecting a solid phase microextraction (SPME) fiber assembly with a 100 uM polydimethylsiloxane coating (Supelco Inc., Bellefonte, PA) in through the septum of the sample. This assembly was allowed to equilibrate for 30 rrrinutes within the sample headspace before injecting onto the GC/MS. A Hewlett-Packard 5890 series H GC with operating conditions identical to those above was used with a Hewlett-Packard 5921A atomic emission detector (AED) for elemental analysis of volatile products of H-PFOSA. Table 3.1 lists the emission wavelengths and plasma gases used for the various elements. 49 Table 3.1. Emission wavelengths and plasma gases used for atomic emission detection. Element Monitored Emission Wavelength (nm) Plasma Reagent Gases Sulfur 181 oxygen and hydrogen Carbon 496 oxygen Hydrogen 486 oxygen Fluorine 690 hydrogen Oxygen 777 hydrogen with auxillary gas (10% methane/90% nitrogen) RESULTS Enrichment and identification of bacteria. A Pseudomonad, designated as strain D2, was isolated from a mixed culture of bacteria that fortuitously contaminated a laboratory batch of carbon rich medium containing DFMS as the sole source of sulfur. Strain D2 was isolated by streaking on nutrient agar plates. This isolate is a motile, Gram-negative rod capable of using select fluorinated sulfonates as a source of sulfur for growth under aerobic and sulfur-limiting conditions. It is catalase positive and oxidase positive. Optimal growth was observed at 30° C. Fatty acid profiles for strain D2 were performed by Microbial ID, Inc. (MIDI), Newark, Delaware. This analysis gave a similarity index of 0.788 for Pseudomonas chloroaphis and a similarity index of 0.692 for Pseudomonas fluorescens. Biolog, Inc. (Hayward, CA), identified strain D2 as Pseudomonas fluorescens. Biotransformation of difluoromethane sulfonate. Cells of strain D2 were grown for 24 hours in nutrient broth (Difco, Detroit, MI), and a 1% inoculum of this culture was added to medium containing DFMS. Fluoride release, disappearance of DFMS, and 50 optical density were monitored. Figure 3.1 shows complete defluorination of DFMS with stoichiometric yield of fluoride. Figure 3.2 demonstrates the corresponding growth associated with the utilization of DFMS as the sole sulfur source. Control samples (not shown) did not defluorinate DFMS and showed no change in optical density. 50 _ DFMS —) _ (as fluoride) (— fluoride -....~.—.-—».«..—.._..—.. -....-— 40 Concentration (uM) 41 . \I\*I 4n ‘ 20 30 40 Time (hours) Figure 3.1. Defluorination of DFMS by Pseudomonas sp. strain D2. Error bars represent the standard deviation of triplicate samples. 51 ' fa 0-8 a 50 ‘11-“...1. 0.6 g E" ‘ DFMS —) ‘ optical 'I Q, B 40 - (as fluoride) density E a 0.4 a o 30 'I or 3 , - o 8 20 - 3' , 0.2 '8- O 10 q/ \\ I 0 is 1 r . r 1 r 1 r .. . . 00 0 10 20 30 40 50 60 Time (hours) Figure 3.2. Growth of Pseudomonas sp. strain D2 on DFMS. Error bars represent the standard deviation of triplicate samples. Biotransformation of other fluorinated sulfonates. Pseudomonas sp. strain D2 was used to evaluate the biotransformation of other sulfur containing fluorinated compounds such as trifluoromethane sulfonate (TFMS), 2,2,2-trifluoroethane sulfonic acid (TES), perfluorooctane sulfonate (PFOSA), and 1H,lH,2H,2H-perfluorooctane sulfonic acid (H-PFOSA). Biotransformation experiments were similar to those used for DFMS studies except that TES and H-PFOSA experiments were conducted in bottles sealed with teflon lined septa (1:6 liquid to gas volume). Strain D2 was only capable of utilizing the hydrogen-substituted sulfonates - DFMS, TES, and H-PFOSA . Unlike the transformation of DFMS, TES and H-PFOSA were only partially defluorinated. When TES was the sole sulfur source for growth, one mole of fluoride was released per mole of TES degraded. Transformation of H-PFOSA yielded approximately one mole of fluoride per mole of H-PFOSA transformed when 52 incubations were open to the atmosphere and 1.4 moles per mole when incubations were closed to the atmosphere. Figures 3.3 and 3.4 show fluoride released from various concentrations of TES and H-PFOSA, respectively. As with biotransformation of DFMS, transformation of TES and H-PFOSA correlated with growth of strain D2 indicating that sulfur was released and assimilated. Figures 3.5 and 3.6 show growth of strain D2 with TES and H-PFOSA as the sole sources of sulfur. 400 T I T r r r . r v I. 3:1 300 ...-.._-_....-M. 1.1.-..111--_-_1._---- "-1. -. .--__--.-..M,.,1 m--.-........-1-. .---._.---,_’ I r I I Fluoride x'” I, 23] Released 200 ’," ’,,-"' (”M) I I” I ’ ' ' a ’,-" ,-””' ’ 11 100 ”*’ ,x“ ,,,,,, "" A ” ’ ’ ””””” ‘ I ’ - - ’ ’ _ ‘ ’ - 4 A- - - - 0 n I n I 1 l n l n 20 4O 6O 80 100 120 Initial 2,2,2-trifluoroethane sulfonate (uM) Figure 3.3. Fluoride release from TES. Numbers represent ratio of moles of fluoride to moles of TES. Triplicate samples were measured, but error bars are not visible due to scale. 53 100 V r I I l I U 2:1 ’1 1’ 1” ’1 75 ,’ I,” 1.4:1_ b ’1‘ ””" ‘ Fluoride / x” I’ ” Released so ,4 -- -.1 (HM) ’1’ ’,v’ """" - I ’ ’o I' ’1 ’1’ a' ” I" ””” I I’ a” ...-p" 25 -’ ,- “-4— 0 1 l n I u 1 1 10 20 30 40 50 Inital lH,lH,2H,2H-perfluorooctane sulfonic acid (uM) Figure 3.4. Fluoride release from H-PFOSA. Numbers represent ratio of moles of fluoride to moles of H-PFOSA. Triplicate samples were measured, but error bars are not visible due to scale. 20 ' I ' I ' I ' I ' I ' 0.30 fluoride ‘ ... 15 («Ch 2* o A - 0 20 8 i - ' s. g 10 0.15 g 2 optical density ‘ a m - 0.10 a 5 g j - 0.05 o O n J 1 l n I n l n I 1 0m 0 20 40 6O 80 100 120 Time (hours) Figure 3.5. Growth of Pseudomonas sp. strain D2 on TBS. Error bars represent the standard deviation of triplicate samples. 54 I ‘ _ 0.4 :3, - 0.3 5 A D E . 8 g i . 0.2 .... 3 1% m d .- 55'; . / (———- optical density - 0,1 3 2 8' 0 1 1 ‘ 1 ‘ 1 ‘ l 4 l T 0.0 O 20 40 60 80 100 120 Time (hours) Figure 3.6. Growth of Pseudomonas sp. strain D2 on H-PFOSA. Error bars represent the standard deviation of triplicate samples. Eight volatile and fluorinated byproducts of H-PFOSA were detected with GC/MS and AED. However, no transformation products were detected for TES. As shown in Figure 3.7, all of the volatile products of H-PFOSA contained carbon, oxygen, hydrogen, and fluorine. Sulfur was not detected in any volatile products of H-PFOSA. Complete GC/MS data and spectrums (both electron impact and chemical ionization) can be found in Appendices A and B. In addition, complete AED plots are also found in Appendix C. 55 12000- 100004 8000-1 4 I 6000-1 40001 I carbon 2000 1 .’ hydrogen I. ' '1. - 1"“ j “I I. 0‘ _fluorine A flit“ _ _ | 75.11%. "10" -1,21..1,4- - -1,6.-11.8.--2,0 Time (minutes) Figure 3.7. Atomic Emission Detection of volatile fluorinated byproducts of H-PFOSA biodegradation. DISCUSSION Several researchers have demonstrated the utilization of aliphatic sulfonates as sole sulfur source [9, 10, 12, 18, 20, 21, 23, 24], but there are no reports establishing that fluorinated sulfonates can serve as the sole source of sulfur for growth. In general, fluorinated organics are perceived as refractory in natural environments with the potential for accumulation. Therefore, it is important to understand the details of how these compounds are metabolized and/or degraded. It is demonstrated in this report that specific fluorinated sulfonates are at least partially degraded. Presumably, the growth associated with these transformations is due to the utilization of the sulfur in these 56 molecules (no growth was observed in their absence). As shown in Figures 3.5 and 3.6, growth continues even after the release of fluoride has st0pped. A likely explanation is that the initial transformation and defluorination are rapid and not rate-limiting, while the assimilation of the sulfur byproduct into biomass is slower and rate-limiting. These findings suggest that the initial steps of transformation are linked to sulfur scavenging with subsequent defluorination of the molecule. Kelly et al. [10] have shown that sulfite and formaldehyde are the products of degradation of methanesulfonic acid (MSA) by a methylotrophic bacterium. The methylotroph was not able to utilize sulfonates with more than three carbons when it was pregrown with MSA and grew poorly on compounds containing more than one carbon. An NADl-I- dependent monooxygenase was found to be responsible for the reaction. Based on these results and those of other investigators [21, 24], a possible mechanism for degradation of the fluorinated sulfonates is desulfonation of the molecule, with the production of sulfite or bisulfite and an aldehyde. The sulfite would then be subsequently used as a source of sulfur for growth (no free sulfite was detected in incubation of strain D2). An aldehyde formed from the degradation of H-PFOSA may or may not be stable and could decompose to other products. The discrepancy in fluon‘ned stoichiometry between closed and open incubations of H-PFOSA with strain D2 could be explained by subsequent transformation of one or more of the volatile fluorinated products. Postulation of a mechanism that removes the sulfur moiety from the molecule is supported by atomic emissions spectroscopy data that failed to detect sulfur in any of the volatile fluorinated byproducts of H-PFOSA biodegradation. Table 3.2 illustrates possible structures for the volatile products from H-PFOSA transformation. 57 Table 3.2. Possible volatile fluorinated products of H-PFOSA transformation by strain D2. Possible H-PFOSA products (-0 fluoride) C6F13CH2CH2-O-CH3 C6F13CH2-O-CH2CH3 C5F13CH2CCH3 C6F13CH2CH20H Possible H-PFOSA roducts -l fluoride C5F1 1CF=CH—CH=O C5F1 1CF=C=O C 5F1 1CF=C=C=O C5F11CF\-C117H 0 Possible H-PFOSA products (-2 fluoride) C5F1 1CH=(|3-CH=O OH C5F1 1EH-CH2-CH 3 0 Molecular Weight 378 378 376 364 Molecular Weight 342 328 340 343 Molecular Weight 340 327 The fate of volatile fluorinated compounds in the environment has been studied internationally during the past twenty years. The chlorofluorocarbons (CFCs) have been implicated in the destruction of the ozone layer and global warming [13, 17]. 58 Replacements for the CFCs, hydrofluorocarbons (HFCs) and hydrochlorofluorocarbons (HCFCs), are expected to degrade before reaching the stratosphere and the ozone layer, but some of these can still contribute to global warming [26]. Although the identity of the volatile fluorinated products of H-PFOSA transformation is unknown, their production is important because it suggests an alternative route for the production of volatile fluorinated compounds. Until the identity of these compounds is determined, it is unclear what effect, if any, they might have on the environment or on global warming. Another finding of this research was the observation that only fluorinated compounds with one or more hydrogen substituents are transformed by strain D2. It is likely that the enzyme responsible for the transformation of these compounds must interact at or near the carbon-sulfur bond. In the case of molecules with only fluorine at the alpha-carbon, the strong electronegativity of fluorine (4.0) may preclude attachment of the enzyme to the molecule and thus limit biotransformation. The fluorocarbon chain is also more "rigid" due to fluorine atoms on the molecule [8]. This rigidness almost certainly interferes with molecule/enzyme interactions, protecting fluorocarbon molecules from biological attack. The results of this study clearly demonstrate that hydrogen-substituted fluorinated sulfonates are susceptible to biodegradation and defluorination. In addition, growth of bacteria may be supported by the sulfur byproduct of transformation under aerobic and sulfur-limiting conditions. By understanding the specific nature of the enzyme(s) or co- factor(s) involved in the transformation of these molecules, we may be able to gain insight into the mechanism of degradation of fluorinated sulfonates as well as non- fluorinated sulfonates. This insight may make it possible to construct fluorinated surfactants that are more readily biodegradable to harmless end products. REFERENCES Abe, T., and S. Nagase. 1982. Electrochemical fluorination (Simons process) as a route to perfluorinated organic compounds of industrial interest, pp. 19-44. In R. E. Banks (ed.), Preparation, properties, and industrial applications of organofluorine compounds. John Wiley & Sons, New York. Banitt, E. H., W. E. Coyne, K. T. McGurran, and J. E. Robertson. 1974. Monofluoromethanesulfonanilides. A new series of bronchodilators. J. Med. Chem. 17:116-120. Cartwright, D. 1994. Recent deve10pments in fluorine-containing agrochemicals, pp. 237-257. In R. E. Banks, B. E. Smart, and J. C. Tatlow (ed.), Organofluorine chemistry: Principles and commercial applications. Plenum Press, New York. . Elliot, A. J. 1994. Chlorofluorocarbons. DP. 145-157. In R. E. Banks, B. E. Smart, and J. C. Tatlow (ed.), Organofluorine chemistry: Principles and commercial applications. Plenum Press, New York. EPA. 1989. Pesticide Fact Sheet on Sulfluramid. EPA. 205 (540/FS-89-060). Fielding, H. C. 1979. Organofluorine surfactants and textile chemicals, pp. 214- 232. In R. E. Banks (ed.), Organofluorine chemicals and their industrial applications. Ellis Horwood Ltd., Chichester. Filler, R. 1993. Fluoromedicinal chemistry - An overview of recent developments, pp. 1-23. In R. Filler, Y. Kobayashi, and L. M. Yagupolskii (ed.), Organofluorine compounds in medicinal chemistry and biomedical applications. Elsevier, Amsterdam. Guo, W., T. A. Brown, and B. M. Fung. 1991. Micelles and aggregates of fluorinated surfactants. J. Phys. Chem. 95: 1829-1836. Higgins, T. P., M. Davey, J. Trickett, D. P. Kelly, and J. C. Murrell. 1996. . Metabolism of methanesulfonic acid involves a multicomponent monooxygenase enzyme. Microbiol. 142:251-260. 59 10. ll. 12. 13. 14. 15. 16. 17. 18. 19. 20. 60 Kelly, D. P., S. C. Baker, J. Trickett, M. Davey, and J. C. Murrell. 1994. Methanesulphonate utilization by a novel methylotrophic bacterium involves an unusual monooxygenase. Microbiol. 140: 1419-1426. Kissa, E. 1994. Fluorinated surfactants: synthesis, properties, and applications. Marcel Dekker, Inc, New York. Laue, H., J. A. Field, and A. M. Cook. 1996. Bacterial desulfonation of the ethanesulfonate metabolite of the chloroacetanilide herbicide metazachlor. Environ. Sci. Technol. 30: 1 129-1 132. Molina, M. J ., and F. S. Rowland. 1974. Stratospheric sink for chlorofluoromethanes: chlorine atom-catalysed destruction of ozone. Nature 249:810—812. Moore, G. G. I. 1974. Sulfonamides with antiinflarnmatory activity, pp. 159- 176. In R. A. Scherrer, and M. W. Whitehouse (ed.), Antiinflammatory agents: Chemistry and pharmacology. Academic Press, New York. Moore, G. G. I. 1979. Fluoroalkanesulfonyl Chlorides. J. Org. Chem. 14:1708- 1711. Rao, N. S., and B. E. Baker. 1994. Textile finishes and fluorosurfactants, pp. 321-336. In R. E. Banks, B. E. Smart, and J. C. Tatlow (ed.), Organofluorine chemistry: Principles and commercial applications. Plenum Press, New York. Rowland, F. S., and M. J. Molina. 1975. Chlorofluoromethanes in the environment. Rev. Geophys. Space Phys. 13:1-35. Seitz, A. P., E. R. Leadbetter, and W. Godchaux III. 1993. Utilization of sulfonates as sole sulfur source by soil bacteria including Comamonas acidovorans. Arch. Microbiol. 159:440-444. Stang, P. J ., and M. R. White. 1983. Triflic acid and its derivatives. Aldrichimica Acta 16:15-22. Thompson, A. S., N. J. P. Owens, and J. C. Murrell. 1995. Isolation and characterization of methanesulfonic acid-degrading bacteria from the marine environment. Appl. Environ. Microbiol. 61 :2388-2393. 21. 22. 23. 24. 25. 26. 61 Thysee, G. J. E., and T. H. Wanders. 1974. Initial steps in the degradation of n-alkane-l -sulphonates by Pseudomonas. Antonie Leeuwenhoek J. Microbiol. 40:25-37. Trepka, R. D., et al. 1974. Synthesis and herbicidal activity of fluorinated N- phenylalkanesulfonamides. J. Agr. Food Chem. 22:11] 1-1 1 19. Uria-Nickelsen, M. R., E. R. Leadbetter, and W. Godchaux III. 1993. Sulphonate utilization by enteric bacteria. J. Gen. Microbiol. 139:203-208. Uria-Nickelsen, M. R., E. R. Leadbetter, and W. Godchaux Ill. 1994. Comparative aspects of utilization of sulfonate and other sulfur sources by Escherichia coli K12. Arch. Microbiol. 161:434—438. Vander Meer, R. K., C. S. Lofgren, and D. F. Williams. 1985. Fluoroaliphatic sulfones: a new class of delayed-action insecticides for control of Solenopsis invicta. J. Econ. Ent. 78:1190-1197. Wallington, T. J ., et al. 1994. The environmental impact of CFC replacements- HFC's and HCFCs. Environ. Sci. Technol. 28:320A-326A. CHAPTER 4 PHYSIOLOGY OF DIFLUOROMETHANE SULFONATE TRANSFORMATION BY PSEUDOMONAS SP. STRAIN D2 ABSTRACT Difluoromethane sulfonate (DFMS; CHFZSO3Na), a fluorinated sulfonate, was chosen as a model compound to investigate the physiology of transformation of fluorinated sulfonates by Pseudomonas sp. strain D2. Previous research has established that a structural or molecular limitation to transformation is observed when the fluorinated sulfonates are completely fluorinated and do not possess hydrogen substituents. The present study investigates further the limitations on biotransformation of these compounds. In whole cell experiments, oxygen was required for growth and transformation of DFMS. It is also shown that DFMS cannot be utilized as a source of carbon and energy, but can be used as a sole source of sulfur under sulfur-limiting conditions with an added carbon source. It appears that actively metabolizing cells are required for the transformation of DFMS. This conclusion is based on the complete absence of transformation activity when glucose or ammonium were removed from the medium. It is hypothesized that transformation of DFMS and other fluorinated sulfonates is linked to a sulfur-scavenging system. Inhibition studies suggest that the sulfur-containing byproduct of DFMS transformation is assimilated through existing sulfur assimilation pathways. Non-competitive inhibition kinetics were observed with Ki values of 3-4 “M for sulfate, sulfite, methane sulfonate, cystine, and methionine. 62 63 INTRODUCTION Although fluorinated compounds are used more frequently today, the environmental fate of many of these compounds is largely unknown. Reductive defluorination has been observed for tri-, di-, and monofluoroacetate .[22]. However, Oremland et al. [15] were not able to replicate this work during further studies and concluded that trifluoroacetic acid was "generally refractory to microbial degradation." Hydrolytic defluorination of monofluoroacetate by bacteria has been reported [3, 4, 5, 6, 10, 13, l4, 19, 23]. More often, the transformation of fluorinated organics appears to require attack on ring structures and functional moieties such as sulfonic, carboxylic, and amine groups. Defluorination of and growth on sulfur-containing organofluorine compounds under aerobic and sulfur-limiting conditions by Pseudomonas sp. strain D2 (chapter 3) and other bacteria (chapter 5) has been demonstrated. Among the compounds degraded were difluoromethane sulfonate (DFMS; CHFZSO3N a), 1H,1H,2H,2H-perfluorooctane sulfonic acid (II-PFOSA; C6F13C2H4803H), and 2,2,2-trifluoroethane sulfonic acid (TES; CF3CH2803H). Trifluoromethane sulfonate (TFMS; CF3SO3Na) and perfluorooctane sulfonate (PFOSA; C3F17SO3K) were not degraded. This work suggested that the transformation of these compounds is restricted to molecules containing hydrogen. In this report other limitations to the transformation of sulfur- containing organofluorine compounds are examined using DFMS as a model compound. It is hypothesized that the transformation of DFMS is linked to sulfur metabolism or sulfur assimilation. Sulfur inhibition studies are used to evaluate whether utilization of DFMS as a sulfur source shares some intermediates or characteristics of sulfur assimilation pathways. 64 MATERIALS AND METHODS Media and chemicals. A defined mineral medium containing (in grams per liter): glucose, 2.0; Kzl-IPO4, 3.5; KH2P04, 2.0; NH4C1, 1.0; MgC12°6H20, 0.5; 0.15M CaC12-2H20 stock, 1.0 ml/liter; trace elements stock I and II, 1.0 mllliter. Trace elements stock I contained (in grams per liter): FeCl3, 1.36; CoC12°6H20, 0.2; MnC1204H20, 0.122; ZnC12, 0.07; NazMoO4-2H20, 0.036; NiClzo6H20, 0.12; B3(OH)3, 0.062; CuC12°2H20, 0.017 . Concentrated H2804 was added at 2.5 ml/L to the trace elements solution 1. Trace elements solution 11 contained (in grams per liter): NaZSeO3°5H20, 0.006; NaWO4-2H20, 0.033; Na2M004-2H20, 0.024. Sodium difluoromethane sulfonate (DFMS) was provided by 3M, St Paul, MN. All other carbon sources and chemicals were obtained from Sigma Chemical. Fluoride and DFMS analysis. Fluoride and DFMS were measured on a Dionex ion chromatography model 2000i/sp fitted with an IonPac AS4A ion exchange column and a Dionex IonPac AG4A guard column. This system utilizes an anion micromembrane suppressor with a Dionex Conductivity Detector-II (CDM). The eluant was a carbonate buffer (1.8 mM Na2C03 and 1.7 mM NaHCO3) with a flow rate of 2 ml/min. Fluoride concentration was used as an indirect method of detecting biotransformation of the fluorinated sulfonates. Fluoride was also measured using an ion selective electrode (Orion 96-09 BN). A total ionic strength adjuster (TISA) was used to stabilize fluoride measurements in samples. TISA was prepared by adding 57 ml of glacial acetic acid and 58 g reagent grade sodium chloride into 500 ml of deionized water. This mixture was cooled on ice, adjusted to a pH of 5.0-5.5 with 5M sodium hydroxide, and diluted with water to a final volume of one liter. One part TISA to one part sample was used to measure fluoride. 65 Growth conditions for Pseudomonas sp. strain D2. Plate colonies of strain D2 were inoculated into 5 ml of nutrient broth (Difco, Detroit, MI) and incubated at 30° C for 24 hours. A 1% inoculum prepared in this manner was introduced into the defined mineral medium (pH of 6.9-7.0) with DFMS as the sole sulfur source. Cells were typically grown aerobically at 30°C and were shaken on a rotary shaker at 160 rpm. To test whether strain D2 could grow under denitrifying conditions, growth medium was amended with 10 mM sodium nitrate, degassed with a 98% N2 and 2% H2 gas mixture, and capped with teflon-faced butyl rubber septa. Culture manipulations were performed in a Coy anaerobic glove box (Coy Laboratories, Ann Arbor, MI). Glucose or acetate provided the carbon and energy source for these incubations. To test whether DFMS could simultaneously serve as carbon source and sulfur source, medium (without glucose) was amended with 5 or 10 mM DFMS. To test whether DFMS could serve as carbon source, medium (without glucose) was amended with 50 uM sulfate and 5 or 10 mM DFMS. A 1% inoculum of nutrient broth grown cells was used for these experiments. Growth conditions and preparation of cells for inhibition experiments. Because fluoride concentration correlated with growth of strain D2, both fluoride and optical density measurements were used to determine stationary phase. Stationary phase cells were harvested by centrifugation (15 minutes at 12,100 X g in a Beckman SS-34 rotor at 4°C), washed in medium without a sulfur source, and resuspended to one tenth the original culture volume in the same medium to a cell density of approximately 1000- 1500 mg protein/L. One milliliter of the resulting 10X concentrated cell suspension was added to 4 ml of medium. Specific substrate utilization curves were determined by varying the concentration of DFMS as well as the concentration of the tested sulfur sources. Fluoride was measured as an indirect measurement of transformation over a 20 minute time period and total cell protein was determined using the modified Lowry 66 method, with bovine serum albumin as the standard [12]. Specific rates of substrate utilization were computed as the mass degraded over a given time period divided by the total mass of protein used in the assay. Modeling of Kinetic parameters. Difluoromethane sulfonate transformation rate coefficients were determined using a model for specific substrate utilization rate: U = (k*S) / (Ks + S) = -(dS/dt)/X where U is the specific substrate utilization coefficient (umoles DFMS/mg protein/hr), k is maximum Specific substrate utilization coefficient (pmoles DFMS/mg protein/hr), S is the rate-limiting substrate concentration (pM), Ks is the half-velocity coefficient (uM), t is time in hours (hr), and X is the total protein concentration (mg protein/L). Values for k and Ks were estimated using a nonlinear curve fit obtained from Systat version 5.2.1, based on initial DFMS degradation rates at a fixed cell protein concentration. Two models were used to determine the nature of the inhibition by other sulfur sources. The first model was the competitive inhibition model: U = (k * S) / (Ks*(l + I/Ki) + S) = -(dS/dt)/X where I is the inhibitor concentration (uM) and Ki is the inhibition constant (uM). The second model used was the non-competitive inhibition model: U = (k*S) / ((I/Ki + l)*(Ks + S)) = -(dS/dt)/X The inhibition rate coefficients were determined using a nonlinear fit of the initial DFMS degradation rates at a fixed cell protein concentration. Rates of transformation were determined at concentrations of DFMS varying from 1.3 uM to 65 pM and inhibitor concentrations from 2 uM to 200 uM. 67 Sonicated crude cell extract preparation. Cultures of strain D2 were grown in IL batches aerobically in 2L Erlenmeyer flasks for approximately 36 hours from a 1% inoculum. Pseudomonas strain D2 was screened for defluorination activity by measuring transformation rates prior to preparation of cell extracts. Actively transforming cultures were transferred to 250 m1 centrifuge tubes and centrifuged at 12,100 X g for 15 minutes at 4°C. Cells were then washed twice and resuspended in 50 ml of buffered growth medium without DFMS. Thirty milliliters of this concentrated cell suspension was sonicated on ice, for 10 minutes at 1 second bursts at stage 5 (50% time on and 50% time off). The effective sonication time was 5 minutes. The sonicate was centrifuged at 12,100 X g for one hour at 4°C and divided into a supernatant and cell pellet. Whole cells, whole sonicate, supernatant, and cell pellet (resuspended in 30 ml medium) were used to assay DFMS transformation. DFMS was added to these samples at 32 pM and 20 11M B-NADH was added to a subset of these samples. French press crude cell extract preparation. Cells grown and harvested as described above were resuspended in 5 ml 20 mM Tris-buffer (pH 7.0) with 1 mM EDTA, placed on ice, and supplemented with the following protease inhibitors: 1 til/ml of leupeptin solution and Sal/ml aprotinin. The cells were then passed three times through a chilled French pressure cell at 1000-1200 psi. The cell extract was diluted to 20 ml. Three reaction mixtures were used to evaluate cell free activity. The first consisted of growth medium, as described above, amended with glucose and 1 mM DFMS. The second was a mixture containing 50 pM ascorbic acid, 50 11M FeC12 , and 10 mM imidazole buffer (pH 6.75) amended with 1 mM DFMS. The third was 20 mM Tris-buffer (pH 7 .0) amended with 1 mM DFMS. Three different cofactors, ATP, B-NADH, and or- ketoglutarate were chosen to evaluate biotransformation requirements. A portion of the cell extract was centrifuged at 27,200 X g for 20 minutes and 0.5 ml of the supernatant was added to 1.5 ml of reaction mixture with one of the three cofactors. Controls 68 consisted of samples with no cofactor, samples with no cell extract, and cell extract that was not centrifuged. The 2 ml samples were placed into 12 ml screw-capped tubes and shaken for 20 minutes to one hour at 21 °C. RESULTS Although Pseudomonas sp. strain D2 can grow under denitrifying conditions with sulfate as the sole sulfur source, it was not capable of growth under denitrifying conditions when the sole sulfur source was difluoromethane sulfonate (Figure 4.1). 0.8 1 I ' I I I v I L glucose & aerobic V N 5. 0.6 /F o r: so . © acetate & aerobic e 0.4 -- - a l O _ .1 % fll 'E‘, 0.2 0 glucose & anaerobic acetate & anaeI'ObiC I 4 0.0 Eb ‘ ‘ TB ‘ d O l 2 3 4 5 Time (days) Figure 4.1. Growth of Pseudomonas sp. strain D2 with glucose and acetate under aerobic conditions. No growth was observed under anaerobic conditions. Error bars represent the standard deviation of triplicate samples. 69 As shown in Figure 4.2, Pseudomonas sp. strain D2 was not capable of using DFMS as a source of both carbon and sulfur nor was it capable of utilizing DFMS as a source of carbon when sulfur was provided in the form of sulfate. 0.8 ' I ' l . r . g 0'6 ‘ DFMS o (5 or 10 mM) as only a l sulfur source + glucose .E‘ 0 4 E ' -carbon - sulfur + DFMS 8 or '3 -carbon + SOuM sulfate + DFMS ‘33. 02 . _---. -. .---..“ l m.-- “.-----“ o l , 0.0 O l 2 3 4 Time (days) Figure 4.2. Growth of Pseudomonas sp. strain D2 with DFMS as sole source of sulfur. DFMS cannot be used as carbon source. Error bars represent the standard deviation of triplicate samples. As shown in Figure 4.3, transformation did not occur when cells that had been grown to stationary phase with DFMS as sole sulfur source were incubated without a source of carbon and energy (no glucose). In addition, transformation of DFMS did not occur when ammonium, the nitrogen source, was removed from the medium. 7O 21 ' I ' I r I 15 . Control (+g1ucose,+ammonium)—7/ , 7; 12 WWW-— :8 ° / 1 u, Ammonrum (-) GIUCIOSC H . 45g ._ . _- O l l l 0.0 0.5 1.0 1.5 2.0 Time (hours) Figure 4.3. Effects of removal of glucose or ammonium, from growth medium, on the transformation of DFMS in stationary phase cells. Error bars represent the standard deviation of triplicate samples. To determine whether transformation of DFMS was liked to sulfur metabolism or sulfur assimilation, sulfur inhibition studies were initiated to further elucidate the physiology of DFMS utilization. Sulfate, sulfite, methane sulfonate (a structural analogue of DFMS), cystine, and methionine were used as inhibitors of transformation. N on-competitive inhibition kinetics were observed. Figure 4.4 illustrates non-competitive inhibition of DFMS transformation for varying concentrations of sulfate. Figures 4.5, 4.6, and 4.7 show the effects of sulfite, methane sulfonate, and cystine respectively on DFMS transformation. Methionine had a similar effect on the transformation of DFMS. 71 0.025 . . . . . , . __1' 0 “M g 0.020 ? - '9 Ki = 3.24 i 0.92 E. g 0.015 _ a E El 2 uM o 0.010 - A 5 0.005 10 “M - 0.000 ' * L 40 60 Concentration (uM) Figure 4.4. Non-competitive inhibition of DFMS transformation rates by sodium sulfate. Error bars represent the stande deviation of triplicate samples. 80 72 0.025 . . . . . , ‘1’? O “M 4 3 0.020 - B 8 . g Ki = 3.59 i 0.20 g 0.015 ‘ 272 E Q 0.010 ‘ In 5 [1M .9 D E I 10 uM 0.005 . . - V (A D 15 uM 0.000 ' ‘ ' ‘ 40 60 80 Concentration (uM) Figure 4.5. Non-competitive inhibition of DFMS transformation rates by sodium sulfite. Error bars represent the standard deviation of triplicate samples. 73 0.025 ' I ' ' ' 1 OuM 3 0.020 ~ 0 ' ' ‘ .3 8 E. g 0.015 ~ ,, Ki=2.87tO.56 ‘ 275 [j ZuM E _ D U 4uM _ Q 0.010 I n I. .9 ’ . 3 0.005 if " D _A_ A 20 “M 4 A 0.000 ’ ‘ ‘ ‘ ‘ ‘ 0 2O 40 60 80 Concentration (uM) Figure 4.6. N on-competitive inhibition of DFMS transformation rates by sodium methane sulfonate. Error bars represent the standard deviation of triplicate samples. 74 0.025 ' I I I f I I E T 0 “M 3 0.020 ? a s 8 9' Ki=3.64:0.61 g 0.015 - :75 E a 0.010 CI 5 uM ~ .2 a 0.005 10 “M - E; 15 uM 0.000 ‘ L ' ‘ 4O 6O 8O Concentration (uM) Figure 4.7. Non-competitive inhibition of DFMS transformation rates by cystine. Error bars represent the standard deviation of triplicate samples. DISCUSSION We have previously shown that Pseudomonas sp. strain D2 is capable of growth with difluoromethane sulfonate (DFMS), 2,2,2-trifluoroethane sulfonic acid (TES), and lH,lH,2H,2H-perfluorooctane sulfonic acid (ll-PFOSA) as sole sulfur source (Chapter 3). DFMS was completely defluorinated while TES and H-PFOSA were only partially defluorinated. Molecules that were completely fluorinated and did not contain hydrogen (trifluoromethane sulfonate; TFMS and perfluorooctane sulfonic acid; PFOSA) were not utilized as a sulfur source and were not defluorinated. These results suggest that hydrogen-substitution is required for transformation. 75 Most reports on the utilization of sulfonates by bacteria have been restricted to aerobic conditions [7, 9, 11, 16, 17, 18, 20, 2]]. However, Chien et a1 [1] have shown that a fermenting Clostridium isolate can grow with taurine and isethionate as sulfur source, and a fermenting Klebsiella isolate can grow with cysteate as a sulfur source. Pseudomonas sp. strain D2 was not able to grow on or transform DFMS under anaerobic conditions even though it was capable of growth under denitrifying conditions with sulfate as the sulfur source. Evidently, whole cells of strain D2 require molecular oxygen to grow with fluorinated sulfonates as the sole sulfur source. Additional evidence is needed to establish whether molecular oxygen is required for the transformation. These results show that DFMS could not be used as a carbon and energy source. In most of these experiments, glucose served as the source of carbon and energy. When glucose was removed from the medium, cells did not grow or transform DFMS. When pregrown cells were used, removal of glucose prevented transformation (Figure 4.3). This suggests that energy from glucose metabolism or some metabolic product of glucose utilization may be required for transformation. In all experiments, ammonium was the source of nitrogen. Removal of ammonium from the medium also prevented transformation (Figure 4.3). It is unclear what role ammonium has on the transformation activity of pregrown cells. Perhaps it is related to enzymatic activation, energy production, or some transport mechanism. Presumably, the transformation of DFMS, as well as the other fluorinated sulfonates, is related to sulfur scavenging. A plausible hypothesis is that the first step in transformation of DFMS is desulfonation yielding sulfite (or another sulfur byproduct) and a fluorinated intermediate that spontaneously decomposes releasing fluoride. The sulfur byproduct is then assimilated into biomass through normal sulfur assimilation pathways. To evaluate this hypothesis, the effects of other sulfur sources on the transformation of DFMS were 76 investigated. Sulfate, sulfite, methane sulfonate, cystine, and methionine were used as possible inhibitors of DFMS transformation. As is shown in Figures 4.4 through 4.7, all of these sulfur sources were inhibitory to transformation (methionine not shown). The data obtained from these experiments did not fit the competitive inhibition model. However, the non—competitive model fit well, with an average correlation coefficient of 0.98 t 0.013. Of interest is the similarity in Ki values (3-4 uM) for these different sulfur sources. This suggests some commonality in the mechanism of inhibition. One possibility is the inhibition of sulfur transport mechanism(s). This inhibition could affect binding proteins in the periplasm or cell membrane complexes that are used for sulfur uptake. Perhaps the high electronegativity of the carbon-fluorine bond interferes with proper binding to transport proteins or other sulfur sources compete for binding proteins in the periplasm but not within the cell. The overall effect of competition outside of the cell combined with abscence of competition on the inside could result in non-competitive pattern of inhibition. It is important to note that the Michaelis-Menten model used to model non-competitive inhibition is based on the assumption of one enzyme interacting with one substrate and a inhibitor molecule. In the case of whole cells experiments, a one step reaction is unlikely to explain the observed non-competitive inhibition. In any case, these results strongly suggest that the transformation of DFMS is related to sulfur- scavenging activities of the cell and that the sulfur byproduct(s) enter via existing sulfur assimilation pathways. In an attempt to clarify the mechanism of transformation of DFMS, crude cell extract experiments were conducted. These experiments investigated the hypothesis that an NADH-dependent oxygenase was responsible for the initial attack on the molecule. This is similar to what has been shown by others [7, 8, 9, 18]. In addition, the possibility of an energy requirement was evaluated and ATP was used as a cofactor for these experiments. Finally, a-ketoglutarate was used to determine if a dioxygenase similar to 77 that encoded by the tfdA gene of Alcaligenes eutrophus (responsible for the first step in 2,4-dichlorophenoxyacetic acid degradation) was present [2]. No transformation activity was observed in cell extracts, even with added cofactors. Perhaps additional cofactors are required or critical membrane or cellular components were destroyed during lysis of cells. Because inhibition was observed with various forms of sulfur found in intact cells (such as cysteine) the negative results might be partially explained by sulfur release during cell lysis. This work establishes several limitations on the biotransformation of sulfur-containing organofluorine compounds and establishes the importance of sulfur assimilation processes in the transformation of DFMS. Structural or molecular limitations evidently prevent transformation when the molecule is completely fluorinated and the presence of hydrogen facilitates transformation under aerobic and sulfur-limiting conditions. REFERENCES Chien, C. C., E. R. Leadbetter, and W. Godchaux III. 1995. Sulfonate-sulfur can be assimilated for ferrnentative growth. FEMS Microbiol. Lett. 129:189-194. Fukumori, F., and R. P. Hausinger. 1993. Purification and characterization of 2,4-dichlorophenoxyacetic acid/a—ketoglutarate dioxygenase. J. Biol. Chem. 268:24311-24317. Goldman, P. 1965. The enzymatic cleavage of the carbon-fluorine bond in fluoroacetate. J. Biol. Chem. 240:3434—3438. Goldman, P. 1971. Enzymology of carbon-halogen bonds, pp. 147-165. In (ed.), Degradation of synthetic organic molecules in the biosphere. National Academy of Sciences, Washington DC. Goldman, P., and G. W. A. Milne. 1966. Carbon-fluorine bond cleavage; H. Studies on the mechanism of the defluorination of fluoroacetate. J. Biol. Chem. 241:5557-5559. Goldman, P., G. W. A. Milne, and D. B. Keister. 1968. Carbon-halogen bond cleavage; III. Studies on bacterial halidohydrolases. J. Biol. Chem. 243:428-434. Higgins, T. P., M. Davey, J. Trickett, D. P. Kelly, and J. C. Murrell. 1996. Metabolism of methanesulfonic acid involves a multicomponent monooxygenase enzyme. Microbiol. 142:251-260. Junker, F., T. Leisinger, and C. A.M. 1994. 3-Sulphocatechol 2,3-dioxygenase and other dioxygenases (EC 1.13.112 and EC 1.14.12.-) in the degradative pathways of 2-aminobenzesulphonic, benzesulphonic and 4-toluenesu1phonic acids in Alcaligenes sp. strain O-l. Microbiol. 140:1713-1722. Kelly, D. P., S. C. Baker, J. Trickett, M. Davey, and J. C. Murrell. 1994. Methanesulphonate utilization by a novel methylotrophic bacterium involves an unusual monooxygenase. Microbiol. 140:1419-1426. 78 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 79 Kelly, M. 1965. Isolation of bacteria able to metabolize fluoroacetate or fluoroacetamide. Nature (London) 208: 809-810. Laue, H., J. A. Field, and A. M. Cook. 1996. Bacterial desulfonation of the ethanesulfonate metabolite of the chloroacetanilide herbicide metazachlor. Environ. Sci. Technol. 30:1129-1132. Markwell, M. A., S. M. Hass, N. E. Tolbert, and L. L. Bieber. 1981. Protein determination in membrane lipoprotein samples: manual and automated procedures. Methods Enzymology 72:296-301. Meyer, J. J. M., N. Grobbelaar, and P. L. Steyn. 1990. Fluoroacetate- metabolizing Pseudomonad isolated from Dichapetalum cymosum. Appl. Environ. Microbiol. 56:2152-2155. Meyer, J. J. M., and D. O'Hagan. 1992. Conversion of 3-f1uoropyruvate to fluoroacetate by cell-free extracts of Dichapetalum cymosum. Phytochemistry 3 1:2699-2701. Oremland, R. S., L. J. Matheson, J. R. Guidetti, J. K. Schaefer, and P. T. Visscher. 1995. Summary of research results on bacterial degradation of trifluoroacetate (TFA), November, 1994-May,1995. USGS. OF 95-0422. Seitz, A. P., E. R. Leadbetter, and W. Godchaux III. 1993. Utilization of sulfonates as sole sulfur source by soil bacteria including Comamonas acidovorans. Arch. Microbiol. 159:440-444. Thompson, A. S., N. J. P. Owens, and J. C. Murrell. 1995. Isolation and characterization of methanesulfonic acid-degrading bacteria from the marine environment. Appl. Environ. Microbiol. 61 :2388-2393. Thysee, G. J. E., and T. H. Wanders. 1974. Initial steps in the degradation of n-alkane-l-sulphonates by Pseudomonas. Antonie Leeuwenhoek J. Microbiol. 40:25-37. Tonomura, K., F. Futai, O. Tanabe, and T. Yamaoka. 1965. Defluorination of monofluoroacetate by bacteria; Part 1. Isolation of bacteria and their activity of defluorination. Agr. Biol. Chem. 29:124—128. 20. 21. 22. 23. 80 Uria-Nickelsen, M. R., E. R. Leadbetter, and W. Godchaux III. 1993. Sulphonate utilization by enteric bacteria. J. Gen. Microbiol. 139:203-208. Uria-Nickelsen, M. R., E. R. Leadbetter, and W. Godchaux III. 1994. Comparative aspects of utilization of sulfonate and other sulfur sources by Escherichia coli K12. Arch. Microbiol. 161:434-438. Visscher, P. T., C. W. Culbertson, and R. S. Oremland. 1994. Degradation of trifiuoroacetate in oxic and anoxic sediments. Nature 369:729-731. Walker, J. R. L., and B. C. Lien. 1981. Metabolism of fluoroacetate by a soil Pseudomonas sp. and Fusan'um solani. Soil Biol. Biochem. 13:231-235. CHAPTER 5 BIOTRANSFORMATION OF SULFUR-CONTAINING ORGANOFLUORINE COMPOUNDS BY DIVERSE ORGANISMS ABSTRACT Previous work (Chapters 3 and 4) established that difluoromethane sulfonate (DFMS, CHF2803Na), 2,2,2-trifluoroethane sulfonic acid (TES; CF3CH2$O3H), and 1H,1H,2H,2H-perfluorooctane sulfonic acid (H-PFOSA; C6F13C2H4$O3H) are degraded by a Pseudomonad under aerobic and sulfur-limiting conditions in minimal medium. The present study focuses on generalizing this observation to other bacteria and yeast. The phylogenetically related Pseudomonasfluorescens, Pseudomonas chloroaphis, and Pseudomonas stutzeri KC were tested. Bacillus subtilis, Escherichia coli, and Saccharomyces cerevisiae were also evaluated. The bacteria were capable of transforming sulfur-containing organofluorine compounds. However, E. coli was not able to degrade H-PFOSA. Yeast was unable to utilize any of the organofluorine compounds as a source of sulfur. Transformation was also evaluated in aquatic samples and soil. No transformation was observed in the aquatic samples, but transformation was detected in the soil samples. It is concluded that the ability to degrade fluorinated sulfonates is not unique to strain D2 and in fact is widely distributed in nature. However, the results also suggest that transformation in the environment requires aerobic and sulfur-limiting conditions. 81 82 INTRODUCTION Sulfur-containing organofluorine compounds are commonly used as catalysts, reagents, surfactants, and pesticides [1, 2, 3, 4, 5, 6, 7, 9, 10]. Several of these are perfluorinated compound such as perfluorooctane sulfonate (PFOSA; C3F17SO3Na), perfluorooctane sulfonamide (C3F17SOzNH2), and trifluoromethane sulfonic acid (TFMS; CF3SO3H). TFMS is one of the strongest organic acids known, has great thermal stability, does not release fluoride in the presence of strong nucleophiles and resists both oxidation and reduction [6]. These compounds generally have excellent chemical and thermal stability. Partially fluorinated analogues of the perfluorinated sulfonates include difluoromethane sulfonate (DFMS; CHF2803N a), 1H,lH,2H,2H-perfluorooctane sulfonic acid (H- PFOSA; C5F13C2H4803H), and 2,2,2-trifluoroethane sulfonic acid (TES; CF3CH2503'). A previous report (Chapter 3) describes experiments to assess biodegradation of TFMS, DFMS, TES, H-PFOSA, and PFOSA by Pseudomonas strain D2. Under aerobic and sulfur-limiting conditions, strain D2 utilized compounds with hydrogen substituents (DFMS, TES, and H-PFOSA) as sulfur sources. D2 completely defluorinated DFMS and partially defluorinated TES and H-PFOSA. The present study generalizes these earlier findings to various other bacteria, yeast, and complex environmental samples (activated sludge, soil, and water). MATERIALS AND METHODS Media and chemicals. A defined mineral medium containing (in grams per liter): glucose, 2.0; K2HPO4, 3.5; KH2P04, 2.0; NH4C1, 1.0; MgC12°6H20, 0.5; 0.15M CaC12°2H20 stock, 1.0 mllliter; trace elements stock I and II, 1.0 mllliter. Trace elements 83 stock I contained (in grams per liter): FeCl3, 1.36; CoC12-6H20, 0.2; MnC12°4H20, 0.122; ZnClz, 0.07; N azMoO4°2H20, 0.036; NiC12°6H20, 0.12; B3(OH)3, 0.062; CuC12°2H20, 0.017. Concentrated H2SO4 was added at 2.5 ml/L to the trace elements solution 1. Trace elements solution 11 contained (in grams per liter): N aZSeO3°5H20, 0.006; NaWO4-2H20, 0.033; Na2M004-2H20, 0.024. To this medium was added the appropriate organofluorine sulfonate as the sulfur source. Ferrnenting cultures of E. coli were prepared similar to that described above except that 5.5 g/L of Kzl-IPO4 and 3.0 g/L KH2P04 were used. Nitrate-respiring cultures were prepared as described above with the addition of 3.88 g/L NaNO3. Both the fermenting medium and nitrate-respiring media were degassed under vacuum. Glucose ( 1 g/L) was used as a carbon source for fermenting cultures and glycerol (0.5 g/L) was used as carbon source in aerobic and nitrate-respiring cultures of E. coli. For experiments with yeast, 10 m1/L of a vitamin stock and 5 ml/L of a amino acid stock were added. The vitamin stock contained (in grams per liter): folate, 0.021; pyridoxine (B6), 0.21; nicotinic acid, 2.1; riboflavin; 0.18; pantothenic acid, 0.53; p-arrrinobenzoic acid 0.025; and B12, 0.025. The amino acid stock solution contained 4.2 g/L glycine and 4.2 g/L of the 21 essential L-amino acids except methionine, cystine, and cysteine. Sodium difluoromethane sulfonate (DFMS) was provided by 3M, St Paul, MN. Potassium perfluorooctane sulfonate (PFOSA) and 1H,lH,2H,2H-perfluorooctane sulfonic acid (H-PFOSA) were obtained from ICN Pharmaceuticals Inc., Costa Mesa, CA. 2,2,2-Trifluoroethane sulfonyl chloride was obtained from Sigma Chemical (St. Louis, MO). Hydrolysis of 2,2,2-trifluoroethane sulfonyl chloride by addition to water and autoclaving yielded trifluoromethane sulfonate (TES). All other chemicals were obtained from Sigma Chemical. Experiments with diverse cell types. Strain D2 was identified as Pseudonwnas fluorescens previously (Chapter 3). Pseudomonasfluorescens (ATCC deposit no. 17400) and Pseudomonas chloroaphis (ATCC deposit no. 9446) were obtained from 84 American Type Culture Collection (ATCC), Rockville, MD. Bacillus subtilis (ATCC deposit no. 6051) and Escherichia coli (ATCC deposit no. 10798) were obtained from the culture collection of the Microbiology Department at Michigan State University. Pseudomonas stutzeri KC (ATCC deposit no. 55595) is routinely cultured in our laboratory. Plate colonies of bacteria were inoculated into 5 ml of nutrient broth (Difco, Detroit, MI) and incubated at 30° C for 24 hours. A 1% inoculum of bacteria was used in the above medium (pH of 6.9-7.0) with the specified organofluorine compound as the sole sulfur source. Cells were grown aerobically at 30 °C and were shaken on a rotary shaker at 160 rpm. Saccharomyces cerevisiae (American Ale, type no. 1056) was obtained from Wyeast Laboratories, Inc., Mt Hood, OR. Saccharomyces cerevisiae was grown in malt extract for 36 hours and then inoculated at 1% of medium. Yeast were grown aerobically 21 °C and were shaken on a rotary shaker at 160 rpm. E. coli was used to determine if degradation of DFMS would occur under anaerobic conditions. A 1% inoculum of nutrient broth grown cell of E. coli were added to fermenting medium and nitrate-respiring medium. Control cultures were prepared in medium that was amended with sulfate as the source of sulfur. Experiments with activated sludge, river water, and soil. To evaluate the fate of fluorinated sulfonates in the environmental samples, samples of activated sludge were obtained from the East Lansing Wastewater Treatment Facility, East Lansing, MI and river water was obtained from the Red Cedar River, on the campus of Michigan State University, East Lansing, MI. In addition three soils samples were also obtained from various locations on campus including a sample from a woodland swamp (natural area), a sample from a botanical garden (agricultural area), and a sandy soil from the banks of the Red Cedar River. The river and activated sludge samples were prepared by placing nine milliliters of the sample into sterile flasks (50 ml). One subset was amended with 1 ml sterile water (no added carbon source), another subset was amended with 1 ml of 20 85 g/L stock of glucose, and another was amended with 1 ml of 10 times concentrated growth medium with glucose (20 g/L). Triplicate samples were prepared with 32 pM DFMS as a source of sulfur and incubated at 21 °C. These samples were shaken on a rotary shaker at 160 rpm. Soil samples were prepared by adding 2 g of soil to sterile balch tubes (28 ml) and amending with 100 pl of 0.1 pg/pl of H-PFOSA stock solution (+ H-PFOSA) or 100 pl sterile water (- H-PFOSA). In addition subsets of these samples were amended with 200 pl of 10 mM glucose stock (+carbon) or 200 pl sterile water (-carbon). For all tested conditions and soil types, killed controls were prepared by autoclaving at 121°C for 30 minutes. Analysis of biotransformation. Fluoride was measured using an ion selective electrode (Orion 96-09 BN). A total ionic strength adjuster (TISA) was used to stabilize fluoride measurements in samples. TISA was prepared by adding 57 ml of glacial acetic acid and 58 g reagent grade sodium chloride into 500 ml of deionized water. This mixture was placed on ice for cooling, and the pH was adjusted to between 5055 with 5M sodium hydroxide. This mixture was then diluted with 500 ml of water to a final volume of one liter. One part TISA to one part sample was used to measure fluoride. Mass spectrometry, GCIECD, and GC/AED analysis. GC/MS data were obtained using a Perkin-Elmer GC with a Finnigan Ion Trap Mass Spectrometer. A DB624 capillary column (30 m x 0.25 mm x 1.4 pm) was used for separation of volatile byproducts (J & W Scientific, Inc., Folsom, CA). Operating conditions were as follows: flow of 30 cm/sec linear velocity, initial temperature 40° C for 4 minutes followed by a 10° C/minute ramp to 200° C, injector temperature of 250° C, transfer line temperature of 225° C. Samples were taken by injecting a solid phase rrricroextraction (SPME) fiber assembly with a 100 pM polydimethylsiloxane coating (Supelco Inc., Bellefonte, PA) in through the septum of the sample. This assembly was allowed to equilibrate for 30 86 minutes within the sample headspace before injecting onto the GC/MS. Identical conditions were used for GC/ECD analysis with a Hewlett-Packard 5890 series 11 GC/ECD and a DB624 capillary column. A Hewlett-Packard 5890 series 11 GC with operating conditions identical to those above was used with a Hewlett-Packard 5921 A atomic emission detector (AED) for elemental analysis of volatile products of H- PFOSA. Emission wavelengths and plasma gases used for the various elements were previously detailed (Chapter 3). Sulfate and fluoride measurement. Sulfate was measured on a Dionex ion chromatograph model 2000i/sp fitted with an IonPac AS4A ion exchange column and a Dionex IonPac AG4A guard column. This system utilizes an anion micromembrane suppressor with a Dionex Conductivity Detector-II (CDM). The eluant was a carbonate buffer (1.8 mM N azCO3 and 1.7 mM N aHCO3) with a flow rate of 2 ml/min. Fluoride was measured using an ion selective electrode (Orion 96-09 BN). RESULTS Biotransformation of fluorinated sulfonates by diverse cell types. In previous studies, Pseudomonas sp. strain D2 completely defluorinated DFMS and partially defluorinated TES and H-PFOSA. In the present study, the phylogenetically related Gram-negative and oxidase-positive bacteria, P. fluorescens , P. chloroaphis, and P. stutzeri KC were all capable of complete defluorination of DFMS (Figure 5.1, P. chloroaphis and P. stutzeri KC are not shown) with stoichiometric yield of fluoride. These organisms showed partial defluorination of TES and H-PFOSA (Figure 5.2 and Figure 5.3, P. chloroaphis and P. stutzeri KC are not shown). None of the tested Pseudomonad strains were capable of growth on or transformation of the completely fluorinated sulfonates TFMS and PFOSA. 87 Bacillus subtilis was chosen as a representative of the Gram-positive bacteria. This organism was capable of defluorination of DFMS as shown in Figure 5.1. However, it was not capable of degrading as much DFMS as the Pseudomonas strains in this medium. Defluorination was also observed with TES (Figure 5.2) and H-PFOSA (Figure 5.3), but again a smaller amount was degraded, as compared to the Pseudomonas strains. B. subtilis was not capable of growth on or transformation of TFMS or PFOSA. Escherichia coli K-12 was selected as a representative Gram-negative, oxidase-negative, facultative anaerobe. As shown in Figure 5 .1, when grown aerobically, E. coli K-12 was capable of defluorinating DFMS, but like B. subtilis, it did not degrade as much DFMS as the Pseudomonas strains. E. coli K-12 also showed a similar partial defluorination of TES like that of the Pseudomonas strains and B. subtilis. However, E. coli K-12 was not capable of growth on or transformation of H-PFOSA. In addition, K-l2 was not capable of growth on or transformation of TFMS or PFOSA, when these compounds were added as the sole source of sulfur. Furthermore, E. coli was not capable of growth or transformation of DFMS under anaerobic conditions. Both fermenting and nitrate- respiring conditions were evaluated with DFMS as the sole source of sulfur. When sulfate was added to these anaerobic cultures, growth was observed yet no transformation of DFMS occurred. Yeast were not capable of degrading any of the tested fluorinated sulfonates even when supplemented with amino acids and vitamins. W m 2: ZWA Control P. DZ P. fluorescens B. subtilus E. coli S. cerevisiae Organism Figure 5.1. Defluorination of DFMS (32 pM) by various bacteria and yeast under aerobic and sulfur-limiting conditions. Error bars represent the standard deviation of triplicate samples. ‘: EEEE Control P. DZ P. fluorescens B. subtilus E. coli S. cerevisiae Organism Figure 5.2. Defluorination of 2,2,2-trifluoroethane sulfonic acid (32 pM) by various bacteria and yeast under aerobic and sulfur-limiting conditions. Error bars represent the standard deviation of triplicate samples. .m. E E, Emm‘ Control P. DZ P. fluorescens B. subtilus E. coli S. cerevisiae Organism Figure 5.3. Defluorination of 1H,lH,2H,2H-perfluorooctane sulfonic acid (32 pM) by various bacteria and yeast under aerobic and sulfur-limiting conditions. Error bars represent the standard deviation of triplicate samples. 91 Biotransformations in aqueous samples. Two different aquatic samples were used to determine the fate of fluorinated sulfonates in natural microbial communities: activated sludge from a wastewater treatment facility and river water. To insure that these samples were not carbon limited 2 g/L glucose was added to all but the no carbon control. It was shown previously (Chapter 4) that ammonium is also required for the transformation of DFMS, so complete growth medium was added to the carbon plus ammonium replicates. No transformation was observed in any of the activated sludge and river samples. Sulfate concentration in these samples was 1.1 mM for the activated sludge and 0.90 mM for the river water. Biotransformations in soils. A volatile transformation product was detected by gas chromatography with an electron capture detector (Figure 5.4) and GC/MS ion trap (Figures 5.6 and 5.7) in all soil samples with and without added carbon. This peak was not observed in the autoclaved controls nor in the samples without H-PFOSA. This peak was identical to a peak found as a transformation product of H-PFOSA by Pseudomonas sp. strain D2 (Figures 5.5, 5.8 and 5.9). Atomic emission analysis of these peaks indicates that they are fluorinated and do not contain sulfur. 1.2e61 1.0e6j l ti0e5€ t10e5€ 4.0e5‘: 2.0e5‘: 0- 92 Peak“ 0 L... M 6 a 10 12 1h 16 Time (min.) Figure 5.4. GC/ECD chromatogram of the natural area incubation with glucose and H- PFOSA. Peak :1 ] W ”W 15 Time (min.) Figure 5.5. GC/ECD chromatogram of strain D2 incubated with glucose and H- PFOSA. 93 Scan Range: 1 - 984 Int = 968 1002 = 10144 mg - Peak #1 TOI- - l b ,.I.,.I.,.l.,.]..., 200 400 800 800 1000 3:21 8:41 10:01 13:21 Figure 5.6. GC/MS chromatogram of the natural area incubation with glucose and H- PFOSA. fluerageol‘l 110805 l‘lims258510589 1002:1533 m; . i l- 5’9’4 77 . 11KB. ‘ 51 we 13: ' . 93' 119 .. 219 255 J.,“. I. .I 141 1". 1 1-. I 2?? '|'IT ‘1']‘l‘l‘l'l'l']‘l'l'l'l'l‘l'l'l‘lrl'l'l'l'l so we 150 200 250 Figure 5.7. GC/MS spectrum number 604 of the natural area incubation with glucose and H-PFOSA. 94 8cm Range! 1 - 1200 Int = 548 1002 = 218844 100; 4 10 1 1 1'01- Peak #1 8 . 67 9 1 2 3 5 . 1._._,~1.L1 M I A'fij—‘WJ . I , I ,— I I 1 3M 51': 900 5:01 10:01 15:01 Figure 5.8. GC/MS chromatogram of strain D2 incubated with glucose and H-PFOSA. average oftfltom Hims3821101m=23526 1M; 1 sari mg! 1 1 77 119 . 277 1.1913. 1 1" 2117.2332151 fl'l' rtzri]'I'I'I'I'VITI'I'I']'I'I'T'T'I'I'I'I’I‘] 50 m 150 200 250 300 Figure 5.9. GC/MS spectrum number 604 of strain D2 incubated with glucose and H- PFOSA. 95 DISCUSSION As is shown in Figures 5.1 through 5.3, Pseudomonads related to strain D2 showed transformation patterns that were similar to those of strain D2 with respect to DFMS, TES, and H-PFOSA. B. subtilis was not as effective at transforming all of the fluorinated sulfonate present. A minimal medium was used for all of these experiments, so that the inability of B. subtilis to degrade all of the sulfonate present may be associated with depletion of some essential nutrient or carbon source. This limited transformation may also be the result of production of a toxic byproduct. A similar result was observed with E. coli when DFMS was the sole sulfur source but not when TES was the sole sulfur source. Interestingly, E. coli was not capable of growth on or transformation of H-PFOSA. E. coli may lack the necessary transport mechanism to transport H-PFOSA across the cell membrane or a different mechanism(s) of transformation may be present in E. coli. Uria-Nickelsen et al. [8] demonstrated growth of S. cerevisiae on various sulfonates such as taurine, isothionate, and cysteate. However, S. cerevisiae was not capable of growth on any of the fluorinated sulfonates tested. The electronegativity of the carbon-fluorine bond may inhibit the uptake of organofluorine sulfur compound or it may inhibit the transformation of these compounds. No transformation was observed in the river water or activated sludge samples. This is most likely due to high background sulfate concentrations (0.9-1.1 mM sulfate) that inhibits transformation of DFMS. However, in soil incubations with H-PFOSA a volatile peak was observed. This product had a mass spectrum identical to one of the peaks generated by incubations of Pseudomonas sp. strain D2 with H-PFOSA. The mass/charge ratio for many of the fragments from this product are typical of fluorinated compounds (i.e. 69, 100, 119, 131, 169, and 219). Figures 5.4 and 5.6 illustrate that this product had the same run time and response to GC/ECD. In addition to these results, 96 atomic emission analysis of this peak indicates that it is fluorinated and does not contain sulfur. These observations support the conclusion that H-PFOSA is transformed in soils with the production of a volatile fluorinated compound that is identical to one of the products of H-PFOSA degradation by Pseudomonas sp. strain D2. Further characterization of this product may provide clues to the mechanism(s) of transformation. The results of this work demonstrate that fluorinated sulfonates can be degraded by a wide range of bacteria as well as within complex communities of microorganisms. In addition, this data suggest that if the environmental restrictions of sulfur limitation and aerobic conditions are satisfied then degradation of these compounds is possible. Future work is needed to identify the mechanism(s) of transformation and to explain why transformation is apparently limited to bacteria, with differences in the extent of transformation and types of sulfonates susceptible to transformation. 10. REFERENCES Kissa, E. 1994. Fluorinated surfactants in blood. J. Fluor. Chem. 66:5-6. Kissa, E. 1994. Fluorinated surfactants: synthesis, properties, and applications. Marcel Dekker, Inc, New York. Langlois, B. R. 1990. Difluoromethanesulfonic acid. Part II. A two-step route to the free acid from monohydrated sodium difluoromethane sulfonate. J. Fluor. Chem. 48:293-305. Moore, G. G. I. 1979. Fluoroalkanesulfonyl Chlorides. J. Org. Chem. 14:1708- 1711. Moore, G. G. 1., and J. K. Harrington. 1975. Antiinflammatory fluoroalkanesulfonanilides. 3. Other fluoroalkanesulfonamido diaryl systems. J. Med. Chem. 18:386-391. Stang, P. J., and M. R. White. 1983. Triflic acid and its derivatives. Aldrichimica Acta 16:15-22. Trepka, R. D., et al. 1974. Synthesis and herbicidal activity of fluorinated N - phenylalkanesulfonamides. J. Agr. Food Chem. 22:11] 1-1 1 19. Uria-Nickelsen, M. R., E. R. Leadbetter, and W. Godchaux III. 1993. Sulphonate-sulfur assimilation by yeasts resembles that of bacteria. FEMS Microbiol. Lett. 114:73-78. Vander Meer, R. K., C. S. Lofgren, and D. F. Williams. 1985.F1uoroaliphatic sulfones: a new class of delayed-action insecticides for control of Solenopsis invicta. J. Econ. Ent. 78:1190—1197. Vander Meer, R. K., C. S. Lofgren, and D. F. Williams. 1986. Control of Solenopsis invicta with delayed-action fluorinated toxicants. Pestic. Sci. 17:449- 455. 97 CHAPTER 6 BIODEGRADATION OF MONOFLUOROACETATE UNDER DENITRIFYING CONDITIONS ABSTRACT This research investigated the potential for biodegradation of monofluoroacetate (MFA), a fluorinated pesticide, under anaerobic conditions. Enrichment cultures were evaluated for MFA degradation under denitrifying and sulfate-reducing conditions. Defluorination was detected under both conditions. Although no isolates were obtained from the sulfate-reducing enrichments, a bacterium, designated strain M7, was isolated from the denitrifying enrichment. The isolate was capable of growth on MFA and defluorination under denitrifying conditions. Phylogenetic analysis of the 16S rRNA sequence for the isolate indicated a 96.0 to 97.5% match for the Bradyrhizobium genus. When incubated with MFA, crude cell extracts of aerobically or anaerobically grown strain M7 released glycolate and fluoride indicating hydrolytic defluorination. Strain M7 was not capable of defluorinating molecules with more than one fluorine substituent, but was capable of dechlorinating monochloroacetate, dichloroacetate, and trichloroacetate. Other bacteria with the ability to utilize MFA under aerobic conditions were evaluated under denitrifying conditions. Two of five isolates were capable of growth on MFA under denitrifying conditions. These findings demonstrate that defluorination of MFA is not limited to aerobic conditions and is likely widespread in the biosphere. 98 99 INTRODUCTION Monofluoroacetate (MFA) is one of the most toxic substances known, based on a lethal dose (LD50) of 0.7-2.1 mg/kg for man [1]. Its toxicity is due to "lethal synthesis" of fluorocitrate which inhibits the aconitase enzyme of the Kreb's cycle [22] although recent investigations implicate fluorocitrate as a "suicide" substrate instead of a competitive inhibitor [3]. MFA is produced naturally by plants in the genus Dichapetalum, as well as Palicourea marcgravii, Acacia georginae, Gastrolobium grandiflorum, and Oxylobium species [11, 12, 23]. The West African plant, Dichapetalum toxicarium, produces m-fluorooleic acid, w-fluoropalmitic acid, and possibly co-fluorocaprate and m-fluoromyristate [24]. Certain fungi are known to produce fluorinated organics: Streptomyces clavus and Streptomyces cattleya produce the fluorine containing antibiotics, nucleocidin and 4-fluorothreonine, respectively [12, 24, 25]. Streptomyces cattleya is also capable of producing MFA [25]. Given that certain plants can produce MFA, it is not surprising that several microorganisms can metabolize MFA. While many investigators have reported aerobic defluorination of MFA by bacteria (mainly Pseudomonads ) and fungi [6, 7, 8, 9, 15, 18, 19, 26, 29], few reports of MFA biodegradation under anaerobic environments are available. Gregg et al. [10] demonstrated defluorination of MFA by genetically modified rumen bacteria Butyrivibrio fibrisolvens. The engineered strain contained a plasmid encoding MFA defluorination genes derived from Moraxella sp. strain B. The goal of this research was to protect cattle from MFA poisoning by accidental ingestion of MFA-containing plants. Evidently, there is only one other report of anaerobic MFA degradation. Visscher et al. [28] reported reductive dehalogenation of MFA to acetate under methanogenic conditions, but this transformation was not reproducible in subsequent investigations [21]. 100 The present report documents isolation and characterization of Bradyrhizobium sp. strain M7, a facultative aerobe that is capable of growth on and defluorination of MFA under denitrifying conditions. The defluorination reaction is shown to proceed by a hydrolytic pathway like that utilized by aerobic organisms. The observation of MFA degradation is then generalized to other bacterial species. MATERIALS AND METHODS Chemicals. Sodium monofluoroacetate, 95 %, was obtained from Aldrich Chemical (Milwaukee, WI). Sodium difluoromethane sulfonate, sodium difluoromalonate, and sodium tetrafluorosuccinate were provided for by 3M, St Paul, MN. All other carbon sources and chemicals were obtained from Sigma Chemical (St. Louis, MO). Nutrient agar and nutrient broth were purchased from Difco (Detroit, MI) Denitrifying and sulfate-reducing enrichments. Primary effluent from the East Lansing Wastewater Treatment Facility, East Lansing, Michigan was used as a source of organisms for enrichment culture. Denitrifying enrichments were prepared by amending this effluent with 2.35 mM sodium nitrate as electron acceptor and 5 mM sodium monofluoroacetate as a carbon source. Sulfate-reducing enrichments were initiated with 2.80 mM sodium sulfate as an electron acceptor. Enrichments were sealed with teflon lined stoppers, and incubated at 20° C for approximately four weeks. Subsequent enrichments and subculturing were performed in defined mineral medium containing (in grams per liter 18 Mohm water): K2HPO4, 0.7; KHZPO4, 0.4; NaNO3, 0.4; NH4SO4, 0.2; MgSO4-7H20, 0.1; 0.15M CaC12-2H20 stock, 0.2 mllliter; trace elements stock, 0.2 mllliter; yeast extract, 0.05. Trace elements stock contained (in grams per 500 ml of 18 Mohm water): FeSO4, 0.68; NazMoO4-2H20, 0.12; 101 CuSO4°5H20, 0.125; ZnSO4-7H20, 0.29; Co(NO3)2°6H20, 0.145; NiSO4°6H20, 0.11; NaSeO3, 0.018; B3(OH)3, 0.031; NH4VO3, 0.06; and MnSO4°H20, 0.505. Concentrated H2804 was added at lml/L to the trace metals solution. After initial subculturing, yeast extract was removed from the medium. Bacterial strains and growth conditions. Strain M7 was grown to stationary phase aerobically in medium containing 10 mM MFA, unless otherwise noted. Cells were typically grown at 30 °C and shaken at 150 rpm on a rotary shaker. For MFA degradation experiments, cells were harvested by centrifugation (15 minutes at 12,100 X g in a Beckman SS-34 rotor at 4°C), washed in buffered medium then resuspended in fresh medium. For anaerobic incubations, bottles with cells and medium were degassed with nitrogen, amended with MFA, and capped with teflon stoppers. Protein was assayed using the modified Lowry method, with bovine serum albumin as the standard [17]. Five isolates capable of aerobic growth on MFA were provided by M. Emptage (DuPont Central Research and Development, Wilmington, DE). All of these isolates were Gram-negative rods. Strain WSZ6b-12-Z and strain CFR1-16-BO have not been identified, strain WS3-12-Z has been identified as a Pseudomonad by Vitek and Biolog, strain N Zl4-5a has been identified asAncylobacter aquaticus by Biolog, and strain DWI-9-G has been identified as Agrobacterium tumefaciens by MIDI (M. Emptage, personal communication). These isolates were grown in the same medium used for strain M7. Chemical analysis. MFA, nitrate, nitrite, and fluoride were measured on a Dionex ion chromatography model 2000i/sp fitted with an IonPac AS4A ion exchange column and a Dionex IonPac AG4A guard column. This system utilizes an anion micromembrane suppressor with a Dionex Conductivity Detector-II (CDM). The eluant was a carbonate buffer (1.8 mM NazCO3 and 1.7 mM NaHCO3) with a flow rate of 2 ml/min. Fluoride 102 was also measured using an ion selective electrode (Orion 96-09 BN). In addition, ion exclusion chromatography was used to determine fluoride and MFA using a Dionex IonPac ICE-AS] column. The eluant for ion exclusion chromatography was 1.0 mM octanesulfonic acid / 2% 2-propanol at a flow of 0.8 ml/min. Nitrous oxide was detected using a thermal conductivity detector (TCD). A Hewlett-Packard 5890 GC fitted with a molecular sieve column (13x, 80/100 mesh, Alltech Associates, Inc., Deerfield, IL) was operated isotherrnally at 150 C. Headspace samples (100 pl) were injected with a 1 ml gas-tight syringe. Metabolite analysis. Strain M7 was grown in 500 ml batches in IL Erlenmeyer flasks for 5 days from a 1% inoculum. Actively transforming cultures were centrifuged at 12,100 X g for 15 rrrinutes in a Beckman SS-34 rotor at 4°C. Cells were then washed twice and resuspended in 8 ml of buffered growth medium without MFA. Six milliliters of this concentrated cell suspension was placed on ice and was disrupted by a 550 Sonic Dismembrator (Fisher Scientific, Pittsburgh, PA) for 10 minutes at 1 second bursts at stage 5 (50% time on and 50% time off). The effective sonication time was 5 minutes. The sonicate was centrifuged for 14,000 X g for 30 minutes at 4°C and divided into a supernatant and cell pellet. The supernatant (final protein concentration of 35 pg/ml) from this crude cell extract was then incubated with 1 mM MFA in growth medium and shaken at 150 rpm at 30 °C. One fraction was degassed with nitrogen, and the other fraction remained aerobic. As a control for glycolate accumulation, whole cells were also incubated with 1 mM MFA aerobically and anaerobically. Samples were withdrawn periodically from both fractions and assayed for glycolate. Glycolate was measured by ion exclusion chromatography as described above and by the colorimetric method of Calkins [2]. 103 Kinetic characterization. Growth kinetic parameters were determined using van Uden's modified version of Monod's microbial growth model which includes endogenous decay of microorganisms : p = “max * (S / (Ks + S )) -b = (dX/dt)/X where p is specific growth rate (d'l), llmax is the maximum specific growth rate (d'1 ), S is the rate-limiting substrate concentration (mg/L), Ks is the half-velocity coefficient (mg/L), b is the endogenous decay rate (d'1 ), t is time in days (d), and X is the concentration of microorganisms (mg/L) [27]. The model used for substrate consumption was: U = (k * S) / (Ks + S) = -(dS/dt)/X where U is the specific substrate utilization coefficient (d‘l) and k is maximum specific substrate utilization coefficient (d'l ). Values for k and Ks were estimated using a nonlinear curve fit obtained from Systat version 5.2.1 based on initial MFA degradation rates at a fixed cell protein concentration. Yield (Y; mg protein/mg MFA) was determined by dividing the change in total cell protein by the change in MFA concentration. Maximum specific growth rate was calculated by taking the slope of the equation: 111(X/Xo) = 11 max *1 where X0 is the initial cell concentration (mg/L). 104 The endogenous decay rate was calculated by taking the slope of the equation: ln(X) = b*t. Molecular characterization. To establish the phylogenetic relationship for M7, 16S rRNA gene-based molecular analyses were performed. The total genomic DNA from strain M7 was isolated using the SDS-based lysis method [30]. The 16S rRNA gene was amplified as described previously [30]. The PCR amplified products were purified and directly used as the templates for automated fluorescent sequencing [31]. The DNA sequence was determined using 10 primers from both directions [30]. Sequences were assembled using assembling programs in the Genetic Computer Group (GCG) software package [4], and a preliminary analysis was done by searching the current databases (GenBank release 91.0 and EMBL release 44.0) using the program FASTA. Sequences were then aligned manually to the 16S rDNA sequences of the species, which showed high similarity scores in the outputs of FASTA, in the previously aligned l6S rDNA sequence database, RDP (Ribosomal Database Project) [16] using the GDE multiple sequence editor program from RDP. Initial phylogenetic screening was constructed using the DNA distance program, Neighbor-Joining, in the PHYLIP package [5] based on all 16S rDNA sequences from alpha subdivision. Based on the initial phylogenetic results, appropriate subsets of 16S rDNA sequences were selected and subjected to final phylogenetic analyses through maximum parsimony method with a bootstrap procedure of 500 replicates, DNA distance matrix method, Neighbor-Joining, and maximum-likelihood method. The SEQBOOT program was used to obtain confidence levels. 105 RESULTS Evaluation of initial enrichments and isolation of denitrifying bacteria. Both the denitrifying and sulfate-reducing enrichments demonstrated defluorination of MFA (Figure 6.1). The sulfate-reducing enrichments turned black as evidence of sulfate reduction with removal of sulfate as measured by ion chromatography. Nitrate and nitrite were completely depleted in the denitrifying enrichments. No isolates were obtained from the sulfate-reducing enrichments. However, an isolate designated strain M7 was isolated aerobically from the denitrifying enrichment by streaking on 1.5 % (wt/vol) nutrient agar plates. Colonies were small (approximately 1 mm) and appeared only after a lengthy incubation period. Colonies were opaque specks and did not spread. Figure 6.2 illustrates defluorination of MFA, with the corresponding reduction of nitrate and accumulation of fluoride ion. A minor accumulation of nitrite during log growth phase was observed followed by subsequent utilization. Nitrous oxide was detected in the headspace of anaerobically grown cultures. Growth kinetic data for this strain M7 under both aerobic and anaerobic conditions are provided in Table 6.1. 106 6 1 .. . :3“ E E I sulfate control E / t; B sulfate +MFA E / :3 1 II]! nitrate control ; E / g D nitrate + MFA E / 11.. / __ / 2 E E / r = r E a e r a E O W // f — n é O :3 34: Am D. u x< w v 0‘ C Figure 6.1. Defluorination of MFA by denitrifying and sulfate-reducing enrichments. Error bars represent the standard deviation of triplicate sammes. Concentration (mM) Concentration (mM) 0.6 0.2 107 fluoride Time (days) 10 12 I r 8.2 (_— log(CFU/ml) - 8.1 1- 8.0 ~ 7.9 L I l l 7.8 4 6 8 10 12 Figure 6.2. Growth of Bradyrhizobium sp. strain M7 on MFA under denitrifying conditions: (a) consumption of MFA and production of fluoride, (b) consumption of nitrate and growth of cells. Error bars represent the standard deviation of triplicate samples. 10g(CFUlm1) 108 Table 6.1. Kinetic parameters for Bradyrhizobium sp. strain M7 Kinetic parameter Denitrifying conditions Aerobic conditions k(d'1) 3.0 :t 0.30 4.9 1 0.40 Ks (mg MFA/L) 4.9 3: 0.20 26 :1: 7.0 b1 (d'l) 0.14 i 0.02 0.14 x 0.02 b2 (d'l) 0.013 :t 0.002 N/A Y (mg protein/mg MFA) 0.13 t 0.01 0.15 i 0.01 “max (d'l) 0.29 i 0.02 0.52 1 0.03 Physiological characterization . Strain M7 is a motile, Gram-negative rod (2.5-4 by 0.8 pM) that utilizes MFA as its sole source of carbon under aerobic and denitrifying conditions. Liquid cultures of strain M7 grow slowly, requiring 2-3 days to enter log phase with a 1% inoculum in nutrient broth. It is both catalase positive and oxidase positive. Optimal growth is observed at 30° C. Strain M7 was classified as a denitrifier based on a growth yield that was proportional to the amount of nitrate reduced, use of both nitrate and nitrite as electron acceptors for MFA degradation, and detection of N20 during anaerobic growth. Fatty acid profiles were performed by Microbial ID, Inc. (MIDI), Newark, Delaware. Although no match to the MIDI library was obtained there were some unique fatty acids present: 48% of the fatty acids were present as cis- 11-octadecenoic acid, 18.5% as octadecenoic acid, and 14.3% as cis-9-hexadecenoic acid. Phylogenetic analysis. Phylogenetic analyses established by bootstrap parsimony method showed that strain M7 is affiliated with the Rhizobium-Agrobacterium group of the alpha subdivision of the Proteobacteria, and very closely related to Bradyrhizobium species (Figure 6.3). Similar trees were also obtained by DNA distance matrix methods 109 and maximum likelihood methods. Strain M7 has 96.0 to 97.5% similarities to Bradyrhizobium species. Since the similarities of 16S rRNA genes from most Bradyrhizobium species range from 95.5 to 99.9%, it is proposed that strain M7 is a new unidentified species of the genus Bradyrhizobium. _E Rhodopseudomonas palustris strain ATH 2.1.6 75 F1. palustrist Bradyrhizobium sp. strain LMG 9980 8. sp. strain LMG 9966 —*34 B. sp. strain LMG 9514 8. sp. strain LMG 10689 * _ 100 '8. sp. strain M7| B. japonicum strain LMG 6138 B. japonicum strain USDA 136 ‘ 1100 1 Nitrobacter sp. strain LL N. winogradskyi strain W . hamburgensis strain X14 _EAgmbacterium tumefaciens strain ATCC 4720 100 A. rubi LMG 156 — Escherichia coli Figure 6.3 Phylogenetic tree showing the location of Bradyrhizobium sp. strain M7 in relation to other Bradyrhizobium species. 110 MFA degradation. When aerobically grown cells were converted to oxygen-free conditions, specific substrate utilization rates were reduced by a sixth (k, 0.83 :t 0.17 d' 1; Ks, 11 :i: 2 mg/L) as compared to cells that were incubated aerobically (see Figure 6.4). When anaerobically grown cells were converted to aerobic conditions however, there was no detectable inhibition of rates as compared to cells that were incubated anaerobically (data not shown). Cells that were grown anaerobically and then incubated anaerobically with MFA had a lower specific substrate utilization rate (R) and a lower half-velocity coefficient as compared to aerobically grown and aerobically incubated cells. 111 I 1 I ' r ' I ' I 5 -0 aerobic growth/aerobic incubation . 'r 7 4.0 I. T 3.0 anaerobic growth/anaerobic incubation T 4 2.0 aerobic growth/anaerobic incubation 1.0 i _I q Specific substrate utilization me (mg MFAimg proteiniday) 0. 0 1 l . l 1 I l l n l O 100 200 300 400 500 MFA (mg/L) Figure 6.4. Comparison of specific substrate utilization rates for aerobically and anaerobically grown Bradyrhizobium sp. strain M7. Error bars represent the standard deviation of triplicate samples. Identification of glycolate as a metabolite of MFA degradation. Several investigators have demonstrated that the first step in degradation of MFA is hydrolytic attack of the carbon-fluorine bond resulting in glycolic acid as an intermediate [6, 7, 26]. Aerobic and anaerobic cell extracts and whole cells of strain M7 were incubated with MFA to assay glycolate production. No accumulation of glycolate was observed for aerobic or anaerobic whole cells. However, supernatant from crude cell extracts from aerobically grown cells, incubated both anaerobically and aerobically, showed stoichiometric accumulation of glycolate and fluoride (see Table 6.2) 112 Table 6.2. Glycolate determination in cell free extract of strain M7 . Values were obtained by ion exclusion chromatography. Glycolate levels were confirmed by colorimetric assay. Glycolate (uM) MFA (uM) produced converted Anaerobic whole cells 0 1215 Anaerobic supernatant 276 282 Aerobic whole cells 0 1200 Aerobic supernatant 305 326 Degradation of other halogenated compounds. As shown in Table 6.3, strain M7 was unable to defluorinate compounds with more than one fluorine substituent. However, chlorinated analogues of MFA monochloroacetate, dichloroacetate, and trichloroacetate degraded both aerobically and anaerobically using cells that had been grown on MFA. Bradyrhizobium sp. strain M7 was not capable of growth on the chlorinated analogues under aerobic or anaerobic conditions. MFA degradation by other bacteria under denitrifying conditions. Five strains capable of MFA degradation aerobically were evaluated for MFA degradation under denitrifying conditions. Two of the five tested strains, strains DWl-9-G and WS3-12- Z, were capable of growth on MFA under denitrifying conditions. Strain WS3-l2-Z showed some accumulation of nitrite with MFA degradation. None of the other strains were capable of growth under denitrifying conditions with MFA as the sole carbon source. 113 Table 6.3. Substrate degradation and growth by Bradyrhizobium sp. strain M7 Denitrifying conditions Aerobic conditions substrate substrate Substrate degradation growth degradation growth monofluoroacetate + + + + difluoroacetate - - - - trifluoroacetate - - - - difluoromalonate - - - - tetrafluorosuccinate - - - - trifluoromethane sulfonate - - - - difluoromethane sulfonate - - - .. monochloroacetate + - + - dichloroacetate + - + - trichloroacetate + - + - glycolate + + + + glyoxylate + + + + oxalate + + + + DISCUSSION Although many bacteria are known to use MFA as a growth substrate under aerobic conditions [6, 7, 8, 9, 15, 18, 19, 26, 29], this is the first report of a naturally occurring bacterium using MFA as a growth substrate under anaerobic conditions. It is also the first report of a Bradyrhizobium species possessing this trait. The results with other 114 MFA-degrading bacteria demonstrate that the ability to degrade MFA under anaerobic conditions is not unique to strain M7. For example, strains DWI—9-G and WS3-l2-Z were also able to grow on MFA under denitrifying conditions. Although no isolate was obtained from the sulfate—reducing enrichments, defluorination of MFA was observed under this condition. Therefore, it may be possible to obtain a sulfate-reducing isolate that is capable of degrading MFA. Clearly, the ability to degrade MFA under anaerobic conditions is not restricted to the Bradyrhizobium and in fact several other bacteria can degrade MFA under anaerobic conditions. These findings suggest that this trait is widely distributed in nature. Defluorination of MFA, by strain M7 under denitrifying conditions appears to proceed by the same hydrolytic mechanism and pathway as aerobic MFA degradation although, it remains to be seen whether a single enzyme system is responsible for the transformation under both conditions. MFA was defluorinated hydrolytically with production of glycolate as an intermediate under both aerobic and anaerobic conditions. The results of this study suggest that the same or a similar halidohydrolase produced by aerobes capable of MFA degradation is also produced by strain M7 and other denitrifying bacteria. The pattern of hydrolytic defluorination, with the production of glycolate, and degradation of the chlorinated analogues , is similar to results obtained by others [6, 9, 29]. The existence of a transmittable plasmid could explain the broad distribution of MFA-degrading capabilities among different genera of bacteria. Kawasaki et a1.[l3] demonstrated that the halidohydrolase responsible for defluorination of MFA is plasmid encoded in Moraxella sp. strain B. Kawasaki et a1. and others have cloned this plasmid into Escherichia coli [14] and Buryrivibrio fibrisolvens. [10]. In addition, Meyer and van Rooyen [20] have transferred the plasmid responsible for MFA defluorination from Pseudomonas cepacia into Bacillus subtilis 115 and have shown that the genetically engineered B. subtilis was also capable of defluorinating MFA. As shown in Table 6.1, the maximum specific growth rate (u max ), the maximum specific substrate utilization rate (k), and the half-velocity coefficient constant (Ks) are different for aerobically grown cells as compared to the anaerobically grown cells. Another difference between aerobic and anaerobic degradation of MFA can be seen in Figure 6.4. When aerobically grown cells are incubated under anaerobic conditions, transformation of MFA is inhibited. However, the contrary is not true when anaerobic cells are incubated aerobically. The inhibition of aerobically grown cells incubated under anaerobic conditions is dramatic, with a one sixth reduction of maximum specific utilization rates. These findings are not readily explained. It is possible that some transport system that functions under aerobic conditions is affected under anaerobic conditions or perhaps different enzyme systems are involved under aerobic and anaerobic conditions. The results of this study demonstrate defluorination of MFA and growth on MFA as sole carbon and energy source under aerobic and anaerobic conditions by Bradyrhizobium sp. strain M7, a genus previously unknown to possess dehalogenating activity. In addition this work establishes defluorination of MFA by other MFA- degrading organisms and by a sulfate-reducing enrichment. The mechanism for degradation of MFA by strain M7 is hydrolytic attack yielding glycolate and fluoride under both aerobic and anaerobic conditions. As observed in aerobic systems the hydrolytic activity is not observed with more highly fluorinated analogues, but is observed with chlorinated analogues. 10. REFERENCES Atzert, S. P. 1971. A review of sodium monofluoroacetate (compound 1080) its properties, toxicology, and use in predator and rodent control. US. Fish and Wildlife Service, Special Science Report on Wildlife No. 146. Calkins, V. P. 1943. Microdetermination of glycolic and oxalic acids. Industrial and Engineering Chemistry-Analytical Ed. 15:762-763. Clarke, D. D. 1991. Fluoroacetate and fluorocitrate: Mechanism of action. Neurochemical Research 16: 1055-1058. Devereaux, J., P. Haeberli, and O. Smithies. 1984. A comprehensive set of sequence analysis programs for the VAX. Nucleic Acids Res. 12:387-395. Felsenstein, J. 1989. PHYLIP - Phylogeny inference package (Version 3.2). Cladistics 5: 164-166. Goldman, P. 1965. The enzymatic cleavage of the carbon-fluorine bond in fluoroacetate. J. Biol. Chem. 240:3434-3438. Goldman, P. 1971. Enzymology of carbon-halogen bonds, pp. 147-165. In (ed.), Degradation of synthetic organic molecules in the biosphere. National Academy of Sciences, Washington DC. Goldman, P., and G. W. A. Milne. 1966. Carbon-fluorine bond cleavage; 11. Studies on the mechanism of the defluorination of fluoroacetate. J. Biol. Chem. 241:5557-5559. Goldman, P., G. W. A. Milne, and D. B. Keister. 1968. Carbon-halogen bond cleavage; 1]]. Studies on bacterial halidohydrolases. J. Biol. Chem. 243:428-434. Gregg, K., et al. 1994. Detoxification of plant toxin fluoroacetate by a genetically modified rumen bacterium. Bio/Technology 12:1361-1365. 116 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 117 Hall, R. J. 1972. The distribution of organic fluorine in some toxic tropical plants. New Phytol. 71:855-871. Harper, D. B., and D. O'Hagen. 1994. The fluorinated natural products. Natural Product Reports 11: 123-133. Kawasaki, H., N. Tone, and K. Tonomura. 1981. Purification and properties of haloacetate halidohydrolase specified by plasmid form Moraxella sp. strain B. Agric. Biol. Chem. 45:35-42. Kawasaki, H., H. Yahara, and K. Tonomura. 1984. Cloning and expression in Eschericia coli of the haloacetate dehalogenase genes from Moraxella plasmid pUOl. Agric. Biol. Chem. 48:2627-2632. Kelly, M. 1965. Isolation of bacteria able to metabolize fluoroacetate or fluoroacetamide. Nature (London) 208:809-810. Larsen, N., et al. 1993. The ribosomal database project. Nucleic Acids Res. (suppl.) 21:3021-3023. Markwell, M. A., S. M. Hass, N. E. Tolbert, and L. L. Bieber. 1981. Protein determination in membrane lipoprotein samples: manual and automated procedures. Methods Enzymology 72:296-301. Meyer, J. J. M., N. Grobbelaar, and P. L. Steyn. 1990. Fluoroacetate- metabolizing Pseudomonad isolated from Dichapetalum cymosum. Appl. Environ. Microbiol. 56:2152-2155. Meyer, J. J. M., and D. O'Hagan. 1992. Conversion of 3-fluoropyruvate to fluoroacetate by cell-free extracts of Dichapetalum cymosum. Phytochemistry 31:2699-2701. Meyer, J. J. M., and S. W. van Rooyen. 1996. Genetically transformed Bacillus subtilis with defluorinating ability. S. Afr. J. Bot. 62:65-66. Oremland, R. S., L. J. Matheson, J. R. Guidetti, J. K. Schaefer, and P. T. Visscher. 1995. Summary of research results on bacterial degradation of trifluoroacetate (TFA), November, 1994-May,1995. USGS. OF 95-0422. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 118 Peters, R. 1972. Some metabolic aspects of fluoroacetate especially related to fluorocitrate, pp. 55-76. In (ed.), Carbon-fluorine compounds: Chemistry, biochemistry, and biological activities (A Ciba Foundation Symposium). Associated Scientific Publishers, Amsterdam. Peters, R., and R. J. Hall. 1960. Fluorine compounds in nature; the distribution of carbon-fluorine compounds in some species of Dichapetalum. Nature 187:573-575. Suida, J. F., and J. F. DeBernardis. 1973. Naturally occurring halogenated organic compounds. Lloydia 36:107-143. Tamura, T., M. Wada, N. Esaki, and K. Soda. 1995. Synthesis of fluoroacetate from fluoride, glycerol, and beta-hydroxypyruvate by Streptomyces cattleya. J. Bacteriol. 177:2265-9. Tonomura, K., F. Futai, 0. Tanabe, and T. Yamaoka. 1965. Defluorination of monofluoroacetate by bacteria; Part 1. Isolation of bacteria and their activity of defluorination. Agr. Biol. Chem. 29:124-128. van Uden, N. 1967. Transport-limited growth in the chemostat and its competitive inhibition; a theoretical treatment. Archiv fur Mikrobiologie 58:145-154. Visscher, P. T., C. W. Culbertson, and R. S. Oremland. 1994. Degradation of trifluoroacetate in oxic and anoxic sediments. Nature 369:729-731. Walker, J. R. L., and B. C. Lien. 1981. Metabolism of fluoroacetate by a soil Pseudomonas sp. and Fusarium solani. Soil Biol. Biochem. 13:231-235. Zhou, J.-Z., M. R. Fries, J. C. Chee-Sanford, and J. M. Tiedje. 1995. Phylogenetic analyses of a new group of denitrifiers capable of anaerobic growth on toluene: Description of Azoarcus tolulyticus sp. nov. Int. J. Syst. Bacterol. 45:500-506. Zhou, J.-Z., and J. M. Tiedje. 1995. Gene transfer from a bacterium injected into an aquifer to an indigenous bacteria. Mol. Ecol. 4:613-638. CHAPTER 7 CONCLUSIONS AND FUTURE WORK RECOMMENDATIONS CONCLUSIONS 1. The ability to transform fluorinated sulfonates is restricted at a molecular or structural level. Completely fluorinated sulfonates were not degraded by the organisms evaluated in this work. However, fluorinated sulfonates containing hydrogen (DFMS, TES, and H-PFOSA) were defluorinated and used as sources of sulfur for growth. 2. The ability to transform fluorinated sulfonates was restricted by physiological conditions. No transformation of fluorinated sulfonates was observed under anaerobic conditions. Transformation was inhibited by other sulfur sources and was completely inhibited when other sulfur sources were present at levels sufficient to support growth. DFMS was not utilized as a carbon source. A carbon and nitrogen source were required for transformation by whole cells. 3. Transformation of fluorinated sulfonates under aerobic and sulfur-limiting conditions was linked to a sulfur-scavenging system. Based upon inhibition studies with other sulfur sources it is suggested that desulfonation of the molecule is followed by sulfur assimilation through existing sulfur assimilation pathways. 119 120 4. Although DFMS was completely defluorinated, with stoichiometric yield of fluoride, TES and H-PFOSA were only partial defluorinated. Transformation of H-PFOSA by Pseudomonas sp strain D2 generated several volatile and fluorinated products. None of these products contained sulfur. 5. In addition to strain D2, several other Pseudomonas species could utilize DFMS, TES and H-PFOSA as sulfur sources for growth. The gram positive species Bacillus subtilis was also capable of transformation of DFMS, TES, and H-PFOSA. Escherichia coli was capable of degrading DFMS and TES, but did not grow on or transform H- PFOSA. Saccharomyces cerevisiae was not able to utilize fluorinated sulfonates as a sulfur source. 6. Samples of river water and activated sludge showed no transformation of DFMS. However, soil samples incubated with H-PFOSA generated a volatile product that is most likely fluorinated. It appears that the ability to degrade partially fluorinated sulfonates is widely distributed among bacteria, but this ability may not be expressed if environmental conditions are not suitable. 7. The bacterium Bradyrhizobium sp. strain M7 was shown to have MFA.degrading abilities. Strain M7 used MFA as a carbon source under both aerobic and anaerobic conditions. 8. A few bacteria that use MFA as a carbon source under aerobic conditions can also use it as the carbon source under denitrifying conditions. In addition, defluorination was observed under sulfate-reducing conditions. These findings suggest that the ability to defluorinate MFA is most likely widespread throughout nature including some anaerobic environments. 121 9. The pathway for MFA degradation under denitrifying conditions was hydrolytic attack, with release of fluoride and production of glycolate. This pathway is similar under both aerobic and anaerobic conditions. FUTURE RESEARCH 1. The volatile byproducts of H-PFOSA transformation should be identified. 2. Improved analytical methods for direct detection and quantification of fluorinated sulfonates in environmental samples are needed. This should include improved methods for sample preparation, separation of mixtures, and detection. 3. The enzyme(s) or cofactors responsible for the transformation of the fluorinated sulfonates should be characterized. This could provide insight into the reasons for inhibition by other sulfur sources and will likely provide clues into the nature of the reaction mechanism ( i.e. oxidative, hydrolytic, nucleophilic, etc.). It will also help to explain why some organisms are capable of the transformation (bacteria) while others are not (yeast). 4. Further evaluation of the fate of these compounds in environmental samples is needed. In addition, the fate of fluorinated sulfonates in anaerobic environments should be evaluated. 5. Further work is needed to determine the fate and identity of the byproduct(s) of TES transformation. APPENDIX A APPENDIX A GC/MS DATA USING ELECTRON IMPACT WITH ION TRAP DETECTOR GC/MS (Ion Trap) analysis: GC: MS: Column: Operation: Injector: Transfer Line: Sample: Perkin Elmer Autosystem GC Perkin Elmer Ion Trap Detector DB624 capillary column (30 m x 0.25 mm x1.4 pm) from J & W Scientific (Folsom, CA) 40°C for 4 minutes, 10°C/minute ramp to 200°C 250°C 225°C Headspace for 15-30 minutes with a solid phase microextraction (SPME) fiber assembly with a 100 uM polydimethylsiloxane coating (Supelco, Inc., Bellefonte, PA) 122 123 Scanllangei 1-1200 lnt=548 lm=218844 100; 4 10 1'01- Peak #1 300 5201 Figure A.l. GC/MS chromatogram of strain D2 incubated with H-PFOSA and glucose. 124 Mange ofiflto I7 Hinus282110025 1002=23528 100% 00 mi 0881 +51 100 131 ‘ 77 118 . 277 119?. I 1'. 47.233275 I I|Tl' 'I'I'[’I'l'l'l'l'l'l'l'l'I'erT'l—FI'I'T‘I'I'I'] 50 100 150 200 250 300 Figure A.2. GC/MS spectrum for peak #1 of strain D2 incubated with glucose and H- PFOSA. Hass Intensity 2 Base 50 4.584 13 13 51 8.426 24.24 65 4,213 12.12 68 34,758 100.00 75 3.158 8.08 77 4,315 14.14 80 2,808 8.08 83 3.510 10.10 84 2,108 0.00 100 8,479 27.27 131 8,777 25.25 163 2,808 8.08 207 2,808 8.08 255 3,153 3.03 277 4.564 13.13 Table A.1. GC/MS listing for peak #1 of strain D2 incubated with glucose and H- PFOSA. 125 liver-age or: 71810722 Minus: 70710711 1002:4234 m; as _ 1 . 95 sup. . m. 55 77 131 118' . 1| .11 ..I 10231 295. 1'1 'I'I‘IJIITIFTT'TTI'I'I li'l'rl'l'r'l'l'lrftl't'v so m 150 200 250 see Figure A.3. GC/MS spectrum for peak #2 of strain D2 incubated with glucose and H- PFOSA. Hess Intensity 2 Base Mass Intensity 2 Base 50 454 4.71 100 1,565 16.23 51 1,565 16.23 118 808 8.42 54 404 4.18 127 1,816 16.75 55 1,212 12.57 131 2.020 20.84 56 707 7.33 181 353 3.66 57 5.76 285 303 3.14 64 858 8.80 305 303 3.14 65 3,283 34.03 67 505 5.24 00 8,648 100.00 70 353 3.66 74 404 4.18 75 1,010 10.47 77 1,111 11.52 83 1,212 12.57 85 4.182 43.46 Table A2. GC/MS listing for peak #2 of strain D2 incubated with glucose and H- PFOSA. 126 fluerage or: 738 to 742 Minus: 74710 751 100: = 1543 mg 75 _ , 9?. r m. ‘ 127 1" 157 1- 276 11.1.: I. 1-191 257 m9 l'l’ IJ'ITI'TTT r'r'r' so 100 159 200 250 an Figure A.4. GC/MS spectrum for peak #3 of strain D2 incubated with glucose and H- PFOSA. - Hess Intensity 2 Base ass Intensity 2 Base 50 180 5.53 112 120 3.68 51 587 17.87 118 256 7.83 57 256 7.83 124 165 5.07 65 150 4.61 127 602 18.43 67 120 3.68 131 873 26.73 68 3,268 100.00 137 135 4.15 71 120 3.68 157 361 11.06 75 2,801 85.71 168 210 6.45 76 150 4.61 181 451 13.82 77 813 24.88 218 105 3.23 83 286 8.76 257 120 3.68 85 225 6.81 276 185 5.88 100 487 15.21 281 225 6.81 107 331 10.14 108 165 5.07 111 75 2.30 Table A3. GC/MS listing for peak #3 of strain D2 incubated with glucose and H- PFOSA. 127 Huerage or: 753 to 753 11171115: 747 to 751 1333 = 43252 133; 75 5111’. mo. 1 127 157 137 278 327 “9 133 213 257 293 so 103 15o 233 253 Figure A.5. GC/MS spectrum for peak #4 of strain D2 incubated with glucose and H- PFOSA. 3' E! 5 _li_as_s Intensity ZBase 50 2,800 4.32 157 10.181 15.. 51 7,372 11.35 257 2,67 3.78 57 6,318 8.73 275 2,67 3.78 65 6,318 8.73 276 6.319 8.73 29,11! 44.86 327 8.075 12.43 70 2,106 3.24 75 64,852 100.00 76 3,510 5.41 77 13,341 20.54 1“ 6,670 10.27 107 6,318 8.73 108 3,158 4.86 113 2,106 3.24 124 3,862 5.85 127 11,235 17.30 131 2,608 4.32 Table A.4. GC/MS listing for peak #4 of strain D2 incubated with glucose and H- PFOSA. 128 Mrageol‘27661077011ims275010754 1332:3333 133; 31 ”4 3113. ‘ 75 133 “9 157 213 243 275233VMW so no 153 zoo 253 333 Figure A.6. GC/MS spectrum for peak #5 of strain D2 incubated with glucose and H- PFOSA. Mass Intensity 2 Base Mass Intensity 2 Base 50 144 2.55 101 144 2.55 51 372 6.57 107 186 3.28 56 124 2.18 118 310 5.47 58 248 4.38 131 578 10.22 00 186 3.28 181 331 5.84 61 5,672 100.00 207 144 2.55 62 165 2.82 00 372 6.57 65 227 4.01 00 3,208 56.57 70 124 2.18 75 000 15.33 77 268 4.74 78 805 6.83 83 310 5.47 100 486 8.76 Table A5. GC/MS listing for peak #5 of strain D2 incubated with glucose and H- PFOSA. 129 Mrageoft77613780111ms278410788 1333:2317 133; 33 135 121 1 ‘35 213 II [IlrlrltIrrvlr'I'r'UIW'TrIIIIUITTrIIIIIIIIIIII'lll 150 2” 250 3“ Figure A.7. GC/MS spectrum for peak #6 of strain D2 incubated with glucose and H- PFOSA. Hass Intensity 2 Base Mass Intensity 2 Base 50 4 7.78 81 5.84 51 532 13.64 81 3,142 80.52 53 354 8.08 82 1,78 46.10 55 633 16.23 83 3,802 100.00 62 152 3.80 84 30 7.78 63 177 4.55 100 22 5.84 65 28 8.44 105 456 11.68 66 481 12.34 106 202 5.18 67 887 22.73 107 202 5.18 68 177 4.55 121 532 13.64 68 1,485 38.31 131 532 13.64 70 126 3.25 136 228 5.84 77 1,875 48.05 78 760 18 48 78 2,230 57 14 8 557 14 28 Table A6. GC/MS listing for peak #6 of strain D2 incubated with glucose and H- PFOSA. 130 average of: 782 to 786 Minus: 800 to 804 1002 = 3588 100; sari BKCJ 59 77 91 245 1' 281 1 19131 195 218 327 50 133 150 200 250 30 Figure A.8. GC/MS spectrum for peak #7 of strain D2 incubated with glucose and H- PFOSA. Mass Intensitu 2 Base 0355 Intensity 2 Base 5 518 6.67 100 1,178 15.15 51 1,743 22.42 108 1,036 13.33 5 658 8.48 118 612 7.88 58 1,508 18.38 131 1,318 16.87 61 706 8.08 185 518 6.67 64 518 6.67 218 424 5.45 65 842 12.12 245 1,602 20.61 68 7,775 100.00 281 1,036 13.33 70 328 4.24 71 1,036 13 33 75 706 8 08 77 1,460 18 78 88 754 8 70 81 1,508 18 38 83 471 6 06 85 565 7 27 Table A7. GC/MS listing for peak #7 of strain D2 incubated with glucose and H- PFOSA. r r‘r 7" 131 averageoi‘202610830 Hims261800623 100$=1H 1332 57 _ i 1 3112; . 01181 . 231 t 77 109 139 . ' _ 13217518 237 233 265 333 336 50 100 150 2M 250 Figure A.9. GC/MS spectrum for peak #8 of strain D2 incubated with glucose and H- PFOSA. Hass Intensity 20352 11355 Intensity 20352 11355 Intensity 28352 50 672 3.48 1,185 6.20 308 672 3.48 51 3,52 1S.I I 672 3.48 308 448 2.33 56 7% 3.88 81 587 3.10 336 522 2.71 57 18,268 133.33 100 1.269 6.58 337 522 2.71 58 870 5.04 107 1,418 7.36 348 448 2.33 58 3,0I18.77 1I1,841 10.08 64 886 4.65 118 1,06 5.43 65 886 4.65 126 448 2.33 66 373 1.84 131 870 5.04 8,365 43.41 138 1.568 8.14 75 1,344 6.88 181 587 3.10 76 7% 3.88 185 821 4.26 77 1,782 8.30 207 522 2.71 78 1,06 5.43 252 522 2.71 87 587 3.10 255 522 2.71 88 672 3.48 257 746 3.88 Table A8. GC/MS listing for peak #8 of strain D2 incubated with glucose and H- PFOSA. 132 8057392011 334111333 "1711152710801 1332=5572 133; 153 _ , 59 117 337, - 3113. ‘ so 4 144 . 100 131 3‘7 1.. . .J l 175 251275 332 l 3i! '1‘ 'I'l'l'l T'l'l 'I'I'F'I'ITI'I‘I‘I'l‘l‘l'l'l‘l"I'l' so 133 153 233 253 333 Figure A.10. GC/MS spectrum for peak #9 of strain D2 incubated with glucose and H- PFOSA. Hass Intensity 2 Base 0355 Intensity 2 Base 50 784 7.74 275 463 4.52 51 827 8.03 317 1.182 11.61 57 463 4.52 331 528 5.16 68 8,080 78.71 88 3,377 32.80 83 662 6.45 88 528 5.16 100 883 8.68 116 1,125 10.87 117 6,480 63.23 118 1,7881 17.42 131 1,457 14.18 143 387 3.87 144 2,648 25.81 145 528 5.16 158 10,266 100.00 Table A9. GC/MS listing for.peak #9 of strain D2 incubated with glucose and H- PFOSA. 133 liver-age 01‘! 858 to $2 Minus: 87310 877 1002 = 60345 133255 - i ‘ i 68 ’ 8111’. 83 _ 3x9, 87 111 125 154 253 237 T'rTI'I'l'I'I’iWTWI'I'r'r‘r' 50 1M 150 233 250 333 Figure A.11. GC/MS spectrum for peak #10 of strain D2 incubated with glucose and H-PFOSA. Hass Intensity 2 Base Hass lntensrtg 2 Base 51 6,670 5.48 81 20,012 16.43 52 4,213 3.46 82 16.501 13.54 53 20,012 16.43 83 84,262 03.16 54 16,501 13.54 84 14.384 11.82 55 121,828 100.00 85 16,852 13.83 56 73.378 60.23 85 12,288 10.08 57 -87,071 71.47 86 11,837 8.80 58 8,830 8.07 87 70,568 57.83 65 4,213 3.46 88 6,318 5.18 66 3,862 3.17 110 4,564 3.75 67 37,215 30.55 111 12,880 10.66 68 15.788 12.87 125 4,815 4.03 68 833% 76.66 154 1,755 1.44 70 62,143 51.01 71 36,162 28.68 78 4,815 4.03 Table A.10. GC/MS listing for peak #10 of strain D2 incubated with glucose and H- PFOSA. 134 GC/MS DATA USING ELECTRON IMPACT WITH QUADRAPOLE DETECTOR GC/MS (Ouadrapole) analysis: GC: Column: Operation: Injector: Transfer Line: Sample: Hewlett-Packard 5995 Series II GC/MS DB624 capillary column (30 m x 0.25 mm xl.4 pm) from J & W Scientific (Folsom, CA) 40°C for 4 minutes, 10°C/minute ramp to 200°C 250°C 225°C Headspace for 15-30 minutes with a solid phase microextraction (SPME) fiber assembly with a 100 uM polydimethylsiloxane coating (Supelco, Inc., Bellefonte, PA) 135 soooof D 52333; 48333; 443337 43333; 36000: 323331 28333: 24333; 23333: ,1 16000: B 1233 1 0 A F I 33333 n\ 43331 l l-%-- 24C- 4 6 'I'I'T'I' ITWVI'I'I'I'I'I' . PH- 8 10 12 14 16 18 20 22 24 Time (minutes) Figure A.12. GC/MS chromatogram for strain D2 incubated with glucose and H- PFOSA. 100 136 69 95 § 125 ISO 175 103% =26528 275 300 325 341 350 Figure A.13. GC/MS spectrum for peak A of strain D2 incubated with glucose and H- PFOSA. 137 mil intensity _1T_1/_§ intensity 31 8.919 122 31.416 32 0.516 123 6.472 40 0.345 124 0.833 42 0.456 125 3.675 43 0.422 126 1.421 44 2.337 131 8.617 45 5.259 137 2.119 47 0.75 140 0.456 50 1.76 141 0.818 51 3.966 145 1.764 53 1.172 153 1.218 55 0.95 157 1.12 56 3.012 163 3.061 57 3.679 169 4.471 68 0.912 172 6.416 69 71.592 173 0.426 70 10.63 175 3.023 71 2.198 176 0.539 72 1.022 181 2.247 73 2.077 187 1.598 74 1.685 195 1.116 75 22.542 207 1.138 76 2.284 213 1.239 87 0.841 225 5.602 88 0.494 231 4.821 91 3.613 242 0.66 93 3.837 243 0.622 94 9.993 245 0.988 95 78.04 275 4.667 96 2.94 291 14.438 100 7.641 292 1.225 104 1.037 341 100 106 2.258 342 49.389 113 2.36 343 4.412 119 8.041 121 0.471 Table A.11. GC/MS listing for peak A of strain D2 incubated with glucose and H- PFOSA. 100 138 93 .. 33 .5 73 + 60 .. 50 3'- 43 1» 30 .1- 23 .L 13 ., 31 100% =62206 ‘59 95 J1 - .131 169 213 231 295 34,4 333 ‘4 A I l l .; UV'F'TI'I'YIIITT'YYUTjYIIIIITYTTWTIV"YII'II’V'YYT'YUUIUV'UTTTY'lIII Beggggggassaa N N (V) ("I 4") Figure A.14. GC/MS spectrum for peak 18 of strain D2 incubated with glucose and H- PFOSA. 138 100 31 90 ‘5 1005:62206 80 «- 60 4. 50 1- 40 .1. 30 1* 20 4c 69 '° " 131 344 . 169 213 231 295 363 l U A A I IVY TYVTTIVYUVYUTVVVVVV'VUUIUUUUUUUIUU UTTTUVVVU'V‘I 33-333335333 N N N t"! Figure A.14. GC/MS spectrum for peak B of strain D2 incubated with glucose and H- PFOSA. 139 M intensity M intensity 31 100 106 0.254 39 0.186 107 0.305 40 0.164 112 0.233 42 0.524 113 0.725 43 1.163 119 2.249 44 1.088 124 0.296 45 1.629 127 3.207 46 0.412 131 3.559 47 1.008 137 0.268 49 4.6004 145 0.416 50 0.593 157 0.296 51 2.677 169 0.897 57 0.995 207 0.252 64 2.514 213 0.362 65 6.884 219 0.35 67 0.858 231 0.436 68 0.262 295 1.27 69 14.741 296 1.174 71 0.217 305 0.453 73 0.95 314 1.101 74 0.354 315 0.497 75 2.762 344 2.516 76 0.379 345 0.701 77 2.99 363 0.701 80 0.24 81 0.256 83 0.212 88 0.166 93 1.534 95 12.926 96 0.423 100 2.262 105 0.198 Table A.12. GC/MS listing for peak B of strain D2 incubated with glucose and H- PFOSA. 140 100 90 ._ 75 1335:1935 80 .. 7o "31 69 60 ,_. so ~- 40 7r 127 51 157 137 276 30 ‘1- 20 «- 101- .11 25 53 75 33 125 153 17s 233 225 « so 275 3 25 Figure A.15. GC/MS spectrum for peak C of strain D2 incubated with glucose and H- PFOSA. 141 M intensity 31 64.91 34 1.525 36 10.956 40 6.718 42 6.305 44 14.522 45 9.612 49 14.884 51 29.147 55 11.473 57 13.54 58 7.907 59 8.217 61 6.615 69 65.426 73 6.822 74 100 75 32.558 100 6.357 103 7.183 107 21.499 124 12.765 127 35.711 131 10.853 157 28.217 207 8.269 276 38.76 Table A.13. GC/MS listing for peak C of strain D2 incubated with glucose and H- PFOSA. 142 100 ‘1- 93 -- 33 .. 73 ~- 60 «31 so .. 30 .1.- 20 .. 10 4b 51 69 1113‘ = 123862 276 127 157 137 343 169 237 213 257 296 32., +' ' 'l'JEA?‘Ul‘LI:U Ul'li I :AU‘IL' :I 0“. I IA. '1‘ ULTL‘ 1A UUUUUUUUU d‘ Figure A.16. GC/MS spectrum for peak D of strain D2 incubated with glucose and H- PFOSA. 143 m/z intensity m/z intensity m/z intensity m/z intensity 29 56.226 83 0.314 145 2.055 291 0.782 31 58.158 86 0.655 146 0.109 293 1.998 33 1.515 87 0.806 150 0.352 294 0.146 37 0.678 88 2.274 154 2.189 296 4.308 38 1.62 89 1.891 155 1.035 297 0.389 39 5.529 91 0.808 156 1.02 308 0.235 40 0.318 92 0.177 157 26.085 323 0.188 42 1.106 93 4.683 158 1.931 327 0.775 43 2.238 94 2.895 163 1.648 341 4.622 44 2.235 95 12.162 168 0.304 342 1.248 45 3.008 96 0.77 169 3.924 343 7.364 46 1.577 97 0.75 170 0.154 344 0.61 47 4.797 99 0.802 172 0.477 49 14.25 100 7.814 173 0.412 50 2.388 101 0.528 174 0.134 51 21.958 104 3.132 175 0.683 52 0.357 105 3.323 176 0.201 53 0.758 106 3.354 177 0.566 55 3.067 107 17.056 181 1.476 56 2.467 108 6.964 187 0.877 57 12.894 109 0.554 188 0.325 58 2.149 112 0.559 195 0.677 59 0.58 113 4.28 205 0.142 62 0.558 117 0.359 207 4.068 63 0.309 119 10.463 208 0.363 64 1.117 120 0.501 213 4.802 65 7.982 122 2.962 219 0.433 66 0.258 123 2.674 225 1.382 67 0.703 124 11.42 227 0.412 69 61.566 125 3.099 231 1.389 70 0.936 126 1.547 237 0.899 71 0.603 127 34.241 239 0.121 72 0.22 128 1.547 243 0.315 73 3.354 131 8.779 245 0.323 75 100 132 0.287 257 2.793 76 7.297 135 0.828 258 0.16 77 30.108 136 0.323 263 0.3 78 1.611 137 3.776 273 0.416 79 0.15 138 1.333 275 0.948 80 0.273 139 1.867 276 46.02 81 0.902 143 0.437 277 3.567 82 0.665 144 0.277 281 1.786 Table A.14. GC/MS listing for peak D of strain D2 incubated with glucose and H- PFOSA. 100 144 33 .. 73 4. 6O ,. so .1- 30 3- 20 ..- IO .1- 31 47 11 8 69 77 91 19 .8. 109 131 3 169 195 § 219 18 N 245 8 259 lCDi=Il3136 291 337 328 § § 350 . Figure A.17. GC/MS spectrum for peak E of strain D2 incubated with glucose and H- PFOSA. 145 M intensity M intensity g1/_z intensity 30 0.59 78 0.725 201 0.479 31 100 88 0.486 213 0.411 32 0.402 89 1.849 219 0.681 33 0.814 90 0.698 221 0.271 37 0.025 91 3.643 225 0.218 38 0.369 92 0.285 227 0.51 39 2.139 93 1.003 239 0.688 40 0.246 95 1.115 241 0.253 41 3.042 97 0.13 245 4.596 42 0.821 100 1.863 246 0.301 43 2.097 101 0.468 258 0.32 44 1.652 106 0.182 259 0.462 45 2.181 107 0.21 278 0.209 46 0.27 108 0.371 291 0.789 47 6.299 109 2.586 307 0.944 49 0.633 113 0.742 308 0.105 50 0.416 115 0.171 327 0.737 51 4.173 119 1.813 328 0.41 52 0.262 120 0.209 55 0.366 121 0.381 56 0.133 122 0.364 57 1.536 123 0.122 59 3.707 129 0.313 60 0.354 131 2.3 61 1.424 132 0.226 64 0.811 139 0.468 65 1.707 145 0.619 67 0.392 152 0.195 69 10.311 159 0.108 71 1.732 163 0.51 72 0.077 169 0.655 75 0.701 177 0.252 77 4.02 195 1.39 Table A.15. GC/MS listing for peak E of strain D2 incubated with glucose and H- PFOSA. 146 100 90 w 339 100% :26924 80 41- 73 3- 60 4»- 50 <1- 272 31 l 232 232 252 296 320 l. I I T11 null ! A l 'IYTrTYIIY T'IITW'VI’T § 3 6 E 3 6 350 Figure A.18. GC/MS spectrum for peak F of strain D2 incubated with glucose and H- PFOSA. 147 M intensity M intensity M intensity 31 4.513 97 0.416 273 1.72 32 1.549 100 5.59 291 0.917 39 1.192 101 5.649 292 1.252 40 4.598 102 1.601 293 0.813 41 12.073 103 2.002 296 3.87 42 31.411 106 0.94 300 2.722 43 4.895 113 0.579 311 21.579 44 1.445 117 0.379 312 1.612 45 0.42 118 0.602 320 3.759 50 1.753 119 26.326 338 37.695 51 1.319 120 1.909 339 100 52 3.283 121 1.645 340 8.799 53 2.429 122 1.801 341 0.602 54 1.419 131 6.47 57 0.758 137 0.773 60 0.977 141 2.254 66 3.124 149 0.825 68 5.868 150 0.813 69 26.122 152 0.984 70 36.272 153 1.553 71 1.606 169 1.772 72 4.3347 182 1.062 73 2.563 202 2.897 74 0.739 203 1.255 75 6.47 216 0.728 76 1.616 222 1.701 77 1.679 225 1.189 78 2.046 232 2.403 79 1.437 242 1.4 80 0.49 250 0.895 90 4.085 252 1.994 91 2.414 267 0.661 92 8.877 272 19.009 Table A.16. GC/MS listing for peak F of strain D2 incubated with glucose and H- PFOSA. 148 100 57 11131::12228 80 ‘1‘ 70 <1- 60 «p- 50 .. 30‘ 20< 10‘ x: a to g x: a :2 g g 250 275 300 325 Figure A.19. GC/MS spectrum for peak G of strain D2 incubated with glucose and H- PFOSA. 149 g; intensity mi; intensity 31 20.527 108 1.382 38 2.535 109 17.722 39 11.833 113 1.497 41 2.077 119 5.757 42 1.047 127 1.914 43 1.922 131 5.904 47 1.652 137 1.398 51 15.289 139 8.096 52 0.131 145 2.412 55 3.439 146 1.333 56 1.865 150 1.079 57 100 169 2.543 58 4.899 189 3.304 59 15.718 195 3.688 64 2.584 219 2.355 67 2.486 239 2.241 69 29.31 258 7.924 75 4.342 278 1.717 76 1.832 325 0.99 77 13.036 - 78 1.088 79 6.608 81 1.317 87 1.946 89 4.236 90 3.042 93 1.807 95 2.453 100 5.684 101 2.633 104 2.085 106 5.806 107 5.397 Table A.17. GC/MS listing for peak G of strain D2 incubated with glucose and H- PFOSA. 100 150 so .. 7o ‘- 60 .. so -- 4o ._ so 4- 2o ._ 240 100% = 56744 317 267 1 I. A 363 378 1 k I 'TTTWrT'T'TTTTT ' a fi V‘TVT rsé's A ‘VIT‘I II Figure A20. GC/MS spectrum for peak H of strain D2 incubated with glucose and H- PFOSA. 151 gill intensity ml; intensity 31 0.511 119 2.192 38 2.226 120 0.289 39 1.353 130 2.132 41 0.906 131 1.165 42 1.269 138 0.296 43 9.821 144 8.319 44 0.236 145 0.737 49 0.303 148 1.163 50 0.684 158 0.34 51 0.927 159 24.129 52 0.633 160 1.914 53 0.687 169 0.624 56 0.465 175 0.888 57 0.732 190 0.497 62 0.46 198 0.62 63 0.405 225 0.293 69 4.786 240 0.793 72 0.375 267 0.936 77 0.363 309 3.221 79 0.43 310 0.226 81 0.513 317 5.072 83 0.189 318 0.522 86 0.384 349 0.809 87 0.33 359 10.177 88 4.504 360 1.348 89 0.441 363 100 99 0.342 364 11.192 100 0.691 365 0.895 105 0.264 378 39.053 115 0.208 379 5.139 116 5.694 380 0.4 117 8.154 118 0.541 Table A.18. GC/MS listing for peak H of strain D2 incubated with glucose and H- PFOSA. 100 152 90 .. 80 - 70 4 103%:8210 I Y F T I I V j I TTTTT Figure A.21. GC/MS spectrum for peak I of strain D2 incubated with glucose and H- PFOSA. 153 g1/_z intensity ml; intensity 32 1.23 111 6.164 38 1.309 112 3.63 39 40.977 115 2.473 40 7.759 121 1.34 41 91.711 125 3.435 42 34.728 126 3.131 43 100 154 2.631 44 5.019 50 1.961 51 2.351 53 9.733 54 16.725 55 83.044 56 72.587 57 44.095 63 2.217 65 3.386 67 13.557 68 11.304 69 59.005 70 54.875 71 14.483 74 1.3852 77 1.376 81 5.652 82 11.974 83 33.973 84 23.881 85 4.726 95 1.571 96 4.032 97 16.773 98 6.042 Table A.19. GC/MS listing for peak I of strain D2 incubated with glucose and H- PFOSA. APPENDIX B APPENDIX B GC/MS DATA USING CHEMICAL IONIZATION WITH QUADRAPOLE DETECTOR GC/MS (Chemical Ionization) analysis: GC: Varian 3400 MS: Axtrel ELQ 400 (Methane as ionization agent) Column: Hewlett-Packard HP-FFAP (25 m x 0.32 mm x 0.52 pm) Operation: -20°C for 4 minutes, 25°C/minute ramp to 225°C, hold 10minutes Injector: 250°C Transfer Line: 225°C Sample: Headspace for 15-30 minutes with a solid phase microextraction (SPME) fiber assembly with a 100 uM polydimethylsiloxane coating (Supelco, Inc., Bellefonte, PA) 03 TIC. Sns: 1-2127. Max: 231922. C Max: 231922. RT: 0.0. 19.9 5 3 s ,_ a L 2E 31:1]? 17. at _ - I I Fl 1 l 1 ‘I ‘I I 1‘1 1 I 100 300 500 700 900 1100 1300 1500 1700 1900 Figure 3.1. GC/MS chromatogram for strain D2 incubated with glucose and H- PFOSA. 154 155 ID 3 M55: 255.0-380.0. Sns: 627-632. Max: 68165. ART: 5.9 343 A Ivvvv'vVvvlvvvafivvleYVV'vvvv vvvv vvvv VVVY Vvvv vvvr rwvr' ' T 960 270 290 900 300 310 390 330 940 QSIO 280 370 Figure 8.2. GC/MS spectrum for peak #1 strain D2 incubated with glucose and H- PFOSA. Q3 Mss: 240.0—675.0. ire: 992—999. B: 1.0 993.997.999. Max: 1638. ART: 9.3 3 5 j 327 I 393 . 295 l 431 300 - - '400 vvvvvvv - '560' - vvvvvvv sr'm' I - i - Figure 8.3. GC/MS spectrum for peak #2 strain D2 incubated with glucose and H- PFOSA. ID 3 M55: 274.0-436.0. Sns: 1001-1003. Max: 2735 TART: 9.4 319 359 341 114141 395 295 303 316 327 I.:Ij".VIIII"I‘_'III...1.4'11"rI u :II'JII‘ V'J'II'A' 393 VI: A vjll' 1!. A vi. A vvvv Figure 8.4. GC/MS spectrum for peak #3 strain D2 incubated with glucose and H- PFOSA. 156 3 M55: 26290-4730. Sns: 1031-1034. Max: 3156. ART: 9.7 2 1 Figure 8.5. GC/MS spectrum for peak #4 strain D2 incubated with glucose and H- PFOSA. JD 3 M85: 248.0-386.0. Sns: 1034-1039. Max: 52900. ART: 9.7 3?? « _ 217 age 309 AA vvvvr V‘vvv vvvw -- -.c: -... .372 -:L.- h--‘T‘v.f -311 -.fif ...3 3-3 aéo 270' aéo 290 360 310 aéo 350 330 350 360 310 350 Figure 8.6. GC/MS spectrum for peak #5 strain D2 incubated with glucose and H- PFOSA. Q3 Mss: 276.0—414.0. Sns: 1061-1065. Max: 2166. ART: 10.0 300 1 A 324 281 I 304 3“ I .gw'-. 1: 1 n I- -II-. ‘ "T 'T"' -1--- - rt fi;‘;.‘r-‘-~!‘ fiat‘..v. ..4fi --c- v- 290 350 310 aéo 330 340 330' 360 310 350 390 460 410 41.1 l 368 I .11- 1- vv—v firvw Figure B.7. GC/MS spectrum for peak #6 strain D2 incubated with glucose and H- PFOSA. 157 3 M33: 245.0-408.0. Sns: 1112-1116. Max: 352. ART: 10.4 3 7 291 279 ‘_ A I A 14 I A l A In 254 265 1.1.11111 - :1_1'1' 1111 1 _____ 20270 250 290 360 310 30 30 340 30 30 30 30 30 40 Figure 8.8. GC/MS spectrum for peak #7 strain D2 incubated with glucose and H- PFOSA. JO 3 M55: 352.0-465.0. Sns: 1353-1362. Max: 881. ART: 12.7 416 400 4441 A 1 A l L l A 38 80 339 428 II III III :1111i li‘- I I : : III! - -II I v '_I.' 36 o 390 400 410 420 450 430 830 ”Fm o _____ Figure 8.9. GC/MS spectrum for peak #8 strain D2 incubated with glucose and H- PFOSA. APPENDIX C APPENDIX C GC/ATOMIC EMISSION DETECTION - 3 12B. ‘51 " ‘1' 180‘ 80‘ 68‘ L0 1 m U') ‘ m- tn SS 2 ‘ m m L0. 20- 53 — v‘m 1 U” - Ln L—41 l L a? . . . . . . . . . . . . .9 , B S 10 15 Time (min.) Figure C.1. AED chromatogram for fluorine (at 690 nm) strain D2 incubated with glucose and H-PFOSA. 158 159 4 4 B S 10 15 20 Time (min.) Figure C.2. AED chromatogram for sulfur (at 181 nm) strain D2 incubated with glucose and H-PFOSA. 160 . E 48: EE 3%; 28‘: 7‘1 mi T V V V I V I I Y I an d B 5 10 15 28 Time (min.) Figure C.3. AED chromatogram for hydrogen (at 486 nm) strain D2 incubated with glucose and H-PFOSA. 161 n14 m . fl, 4%“ 1% 3B“ 28‘ ‘ v CD CD 1\ <1- m . . , m m c—o 10" ‘ v . 0". cs.) 4 N 8 1| 8. 1:) @— I r T Y 1 T ' I f I I U 13 15 28 Time (min.) Figure C.4. AED chromatogram for carbon (at 496 nm) strain D2 incubated with glucose and H-PFOSA. 162 147 I1.TWT 1A 8 S 18 15 28 Time (min.) 1 I 1 .4 Figure C.5. AED chromatogram for oxygen (at 777 nm) strain D2 incubated with glucose and H-PFOSA. TE UN V. IBRRRIES 1111111 ll um 815643343 MICHIGAN STR 11111111111111“ 31293