.-I¥ 1» u@ .. 1. R. 5.). c In?” I . ,fi-IU‘ a afieunwwu. L A III; .I IJ 413.. .. . u . 2.3.1. .. .. 34!...er , .. 1. 7:111. ....9h....1%.m.1, ‘ .. HH....L4£TI.UM.J...3Q. . L. Xv ful‘ (£1421:de Heir... fiufwio .. ; YE u. 1.: x . ; 3:. 144 .. . - fin....fl..nu.-...au: 1.: ‘ |K?.I:w...->ut’.~ URI I”... .II K I) (u. ol I..I l..l I r: i'fl! . .. b). 1"." . . 4.0 ,o 1.1“). .VHMI ..t.1 .H 11. 3.1%.: 4/ ., I; . , u w w r I III. 1.1:.» t. s$é¢fls§é§£ III II... 111.51% | 33. def." «. 1,. 3.3 .934.» III LII - ill llllllllllllllllllllllllllllllllllllllll 3 1293 01050 61 LIBRARY Mlcnlgan State Unlverslty This is to certify that the dissertation entitled kIKETICS CHARACTERIZATION OF CONETABOLI 21M; COMMUNITIES AA/D ADAPT/Snot] T 0 MoflépowTH SUBSTRATE presented by [WAuCz- KUAN’ CHAUQ has been accepted towards fulfillment Pk. D. of the requirements for degreein ENVIRONMENHL EJéINEERIx/é Datel'Ll/l7/?g / F FmQiQsl/oe/L PLACE II RETURN BOX to remove We checkout from your record. TO AVOID FINES return on or before date due. «,_DATE Due DATE DUE DATE DUE 'é. W _ , ‘49! r _—l| J l__J MSU IeAn AfflrmetIve Action/Equal Opportunlty Inetttwon Wm: KINETICS CHARACTERIZATION OF COMETABOLIZING COMMUNITIES AND ADAPTATION TO N ONGROWTH SUBSTRATE By Wang-kuan Chang 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 KINETICS CHARACTERIZATION OF COMETABOLIZING COMMUNITIES AND ADAPTATION TO NONGROWTH SUBSTRATE By Wang-kuan Chang Many compounds of environmental and toxicological significance are transformed by cometabolism. For routine engineering application of cometabolism, several issues must be addressed. First, a generally accepted kinetic model is needed to describe transformation of growth and nongrowth substrates. Second, methods are needed to evaluate model parameters. Finally, changes in the cometabolic activity of a community due to long term, repeated exposure to nongrowth substrates need to be investigated. An unstructured model for cometabolism is presented and verified experimentally in this research. The model includes the effects of cell growth, endogenous cell decay, product toxicity, and competitive inhibition with the assumption that cometabolic transformation rates are enhanced by reducing power obtained from oxidation of growth substrates. A theoretical transformation yield is used to quantify the enhancement resulting from oxidation. A systematic method for evaluating model parameters is described. The applicability of the model is evaluated by comparing experimental data for methanotrophic cometabolism of TCE with model predictions from independently measured model parameters. Propagation of errors is used to quantify errors in parameter estimates and in the final prediction. The model predicts TCE and methane transformation successfully for a wide range of concentrations of TCE (0.5 - 9 mg/L) and methane (0.05 - 6 mg/L). The model was also sucessfully applied for the simplified case of nongrowth substrate transformation of HCFCs and HFCs by resting cells. To describe adaptive changes within cometabolizing communities, the verified cometabolism model was analyzed in terms of the "fitness" concept. Fitness was quantified in term of measurable kinetic parameters. The selection gradient for each parameter was defined as the partial derivative of fitness with respect to that parameter. The gradient indicates the selection acting on each fitness component, with other components held constant. It appears possible to use selection gradients as criteria for the stability of a community under a given perturbation. Finally, a methanotrophic mixed culture and a phenol-degrading culture were repeatedly exposed to different levels of TCE. Changes of community structure were monitored and the effects of TCE exposure were evaluated. Various phenotypic parameters were measured to monitor change within each community. Molecular methods of community analysis (ARDRA and DGGE) were used to monitor shifts in microbial community structure. The results indicate that the phenol-degrading community is very stable despite repeated exposure to TCE. Greater instability in a methanotrophic mixed culture is predicted and observed. A stable community is able to maintain its structure and performance under nongrowth substrate exposure. Based on the results of this study, a diverse community seems to have higher stability. To the Memory of My Father. Efiik iv ACKNOWLEDGMENTS I want to offer very special thanks to my advisor, Dr. Craig Criddle, without whom this work would never have happened. He is the one who finally convinced me to tackle it, and he did everything possible to help me to accomplish this research. Of the many people who helped during the research and writing of this work. I would especially like to thank Dr. Larry Fomey, Dr. Robert Hickey, and Dr. Thomas Voice, for their infinite patience and invaluable advice; I am grateful as well to Dr. Denise Searles, Hector Ayala, and Deborah Hogan, for their help in culture isolation and community analysis; and Dr. Chien-chun Shih and Dr. Xianda Zhao and Yanlyang Pan, for their friendship and for the right stuff at the right time. For sheer inspiration, energy, joy, and faith in the future, I am grateful to my wife, Shu- chu, who have added a whole lot of color to my life. TABLE OF CONTENTS LIST OF TABLE ................................................................................... ix LIST OF FIGURES ............................................................................... xi NOMENCLATURE ............................................................................... xiv CHAPTER 1 INTRODUCTION ................................................................................ 1 COMETABOLIC TRANSFORMATION ............................................. 1 ADAPTATION OF COMETABOLIZIN G COMMUNITIES ...................... 2 RESEARCH OBJECTIVES ............................................................ 5 OUTLINE OF THE THESIS ........................................................... 6 REFERENCES ........................................................................... 7 CHAPTER 2 BACKGROUND .................................................................................. 8 COMETABOLISM PRINCIPLES ..................................................... 8 General concept .................................................................. 8 Energy requirement ............................................................. 10 Substrate interaction ............................................................. 11 Product toxicity .................................................................. 12 Cometabolic community ........................................................ 12 QUANTIFICATION OF COMETABOLISM ........................................ 13 Cometabolism by resting cells ................................................. 13 Cometabolism by growing cells ............................................... 15 MODEL MICROBIAL SYSTEMS FOR COMETABOLISM ...................... 2O Methanotrophic communities .................................................. 21 Phenol-oxidizing bacteria ....................................................... 25 MODEL NONGROWTH SUBSTRATES ............................................ 29 HCFC/CFCs ..................................................................... 30 Trichloroethylene ................................................................ 3 l Methanotrophic transformation of trichloroethylene ........................ 32 MODEL REACTOR SYSTEMS FOR COMETABOLISM ......................... 35 REFERENCES ........................................................................... 38 CHAPTER 3 KINETICS OF COMETABOLISM BY RESTING CELLS ................................. 45 ABSTRACT .............................................................................. 45 INTRODUCTION ....................................................................... 46 Rationale for experimental system ............................................. 47 vi MATERIALS AND METHODS ....................................................... 48 Chemicals ........................................................................ 48 Analytical techniques ............................................................ 48 Measurement of Henry's constants ........................................... 48 Culture conditions ............................................................... 49 Batch biotransformation experiments ......................................... 50 Modeling transformation of HCFCs/HFCs .................................. 50 RESULTS ................................................................................. 52 Henry's Law constants ......................................................... 52 HCFCs/HFC transformation rates ............................................ 53 DISCUSSION ............................................................................ 61 REFERENCES ........................................................................... 64 CHAPTER 4 KINETICS OF COMETABOLISM BY GROWING CELLS ............................... 66 ABSTRACT .............................................................................. 66 INTRODUCTION ....................................................................... 66 Rationale for experimental system ............................................. 68 MATERIALS AND METHODS ....................................................... 69 Culture and culture conditions ................................................. 69 Analytical methods .............................................................. 70 Batch biotransformation experiments ......................................... 71 Model development ............................................................. 72 Model verification ............................................................... 74 RESULTS AND DISCUSSION ....................................................... 77 REFERENCES ........................................................................... 93 CHAPTER 5 THEORY OF ADAPTATION OF COMETABOLIZIN G COMMUNITIES .............. 96 INTRODUCTION ....................................................................... 96 MODEL DEVELOPMENT .............................................................. 97 Adaptive change of microbial communities .................................. 97 Fitness of cometabolizing cultures ............................................ 100 Selection gradients .............................................................. 102 Implications of theory for different model systems ......................... 104 MODEL ANALYSIS .................................................................... 106 RESULTS AND DISCUSSION ....................................................... 106 REFERENCES ........................................................................... 118 CHAPTER 6 CHANGE IN COMMUNITY STRUCTURE IN RESPONSE TO LONG TERM TCE EXPOSURE: METHANOTROPHIC MIXED CULTURE IN CI-IEMOSTAT ..... 119 INTRODUCTION ....................................................................... 119 MATERIALS AND METHODS ....................................................... 123 Culture conditions ............................................................... 123 Perturbation of TCE ............................................................. 124 Monitoring of consortia ......................................................... 128 Adaptation of cometabolizing community in chemostat ..................... 129 General procedures for biotransformation measurement ................... 131 Community Analysis ............................................................ 132 RESULTS ................................................................................. 133 vii DISCUSSION ............................................................................ 138 REFERENCES ........................................................................... 140 CHAPTER 7 CHANGE IN COMMUNITY STRUCTURE IN RESPONSE TO LONG TERM TCE EXPOSURE: PHENOL-DEGRADING COMMUNITY IN SEQUENCING BATCH REACTOR .............................................................................. 143 INTRODUCTION ....................................................................... 143 MATERIALS AND METHODS ....................................................... 146 Culture conditions ............................................................... 146 Perturbation of TCE ............................................................. 147 Monitoring of consortia ......................................................... 147 Adaptation of cometabolizing community in SBR ........................... 148 General procedures for biotransformation measurement ................... 151 Community Analysis ............................................................ 153 RESULTS ................................................................................. 154 DISCUSSION ............................................................................ 164 REFERENCES ........................................................................... 167 CHAPTER 8 ENGINEERING SIGNIFICANCE ............................................................ 169 CHAPTER 9 SUMMARY AND CONCLUSIONS ........................................................... 171 CONCLUSIONS ......................................................................... 171 FUTURE RESEARCH .................................................................. 173 viii LIST OF TABLES Table 2.1 Organic compounds subject to cometabolism and accumulated products ..... 9 Table 2.2 Microorganisms exhibiting the phenomenon of cometabolism ................. 10 Table 2.3 Summary of cometabolic transformation model by resting cells ............... 16 Table 2.4 Summary of cometabolic transformation model by growing cells ............. 19 Table 2.5 Examples of cometabolic transformation of trichloroethylene .................. 21 Table 2.6 Tentative classification Scheme for Methanotrophic Bacteria ................... 23 Table 2.7 Kinetic coefficients of methane utilization ........................................ 24 Table 2.8 Kinetic coefficients of TCE cometabolic transformation by methanotrophic communities ..................................................................... 25 Table 2.9 Kinetic coefficients of phenol utilization .......................................... 28 Table 2.10 Kinetic coefficients of TCE cometabolic transformation by phenol- oxidizing bacteria .................................................................................. 29 Table 2.11 Properties and purity of HCFCs and HFC evaluated in this work ............. 31 Table 2.12 General information and properties of trichloroethylene ........................ 32 Table 2.13 TCE transformation and product formation by methanotrophs ................. 34 Table 3.1 Measured values of Henry's constant vs. temperature .......................... 54 Table 3.2 Temperature regression for Henry's constant .................................... 55 Table 3.3 Salting-out coefficients .............................................................. 55 Table 3.4 Kinetic coefficients for HCFC/HFC transformation by methanotrophic mixed culture MM] ............................................................................... 60 Table 3.5 Kinetic coefficients for HCFC/HFC transformation by methanotrophic mixed culture MM] ............................................................................... 60 Table 4.1 Kinetic and stoichiometric parameters for methane utilization, growth, and TCE degradation for methanotrophic mixed culture MMl .............................. 83 ix Table 4.2 Stoichiometry of TCE cometabolism by methanotrophic mixed culture MMl for different initial TCE concentrations .................................................. 83 Table 4.3 Electron flow in methanotrophic mixed culture MMl for different initial TCE concentrations ............................................................................... 84 Table 4.4 The effect of methane concentration on observed transformation yield ....... 84 Table 5.1. The effects of fitness components on fitness ..................................... 104 Table 6.1 Operation conditions for chemostat ................................................ 126 Table 6.2 Endogeneous decay constant for methanotrophic mixed cultures during long-term TCE exposure ......................................................................... 135 Table 7.1 TCE feed concentration for long-term TCE ....................................... 150 Table 7.2 Summary of bench-scale SBR operating condition .............................. 151 Table 7.3 Theoretical transformation capacity for phenol—degrading reactor communities ........................................................................................ 158 Table 7.4 Endogeneous decay constant for phenol-degrading communities during long-term TCE exposure in SBRs ............................................................... 159 LIST OF FIGURES Figure 1.] Flow of reducing power in cometabolizing system ............................. 3 Figure 1.2 General scheme of cometabolic transformation ................................. 4 Figure 3.1 Biotransformation of HCFC-22 by methanotrophic mixed culture MM] ................................................................................................ 56 Figure 3.2 Biotransformation of HCFC-142b by methanotrophic mixed culture MM] ................................................................................................ 57 Figure 3.3 Biotransformation of HFC-134a by methanotrophic mixed culture MM] ................................................................................................ 5 8 Figure 3.4 Biotransformation of HCFC-123 by methanotrophic mixed culture MM] ................................................................................................ 59 Figure 4.] Approach for prediction of degradation of growth and nongrowth substrates by the proposed model ............................................................... 79 Figure 4.2 The sensitivity equations for parameters estimated nonlinearly from model ............................................................................................... 81 Figure 4.3 Biotransformation of TCE and methane by methanotrophic mixed culture ................................................................................................ 85 Figure 4.4 Biotransformation of TCE and methane by methanotrophic mixed culture ............................................................................................... 86 Figure 4.5 The effect of methane concentration on TCE and methane degradation rate .................................................................................................. 87 Figure 4.6 The observed transformation capacity as a function of the concentrations of growth and nongrowth substrates .......................................... 91 Figure 4.7 The observed transformation yield as a function of the concentration of growth and nongrowth substrates ............................................................ 92 Figure 5.1 Fitness space of cometabolizing community .................................... 102 Figure 5.2 Dimensionless selection gradient, (P/W)(8W/8P) as a function of methane and TCE concentrations for methanotrophic mixed culture when both substrates are present separately ................................................................. 110 xi Figure 5.3 Dimensionless selection gradient, (P/W)(8W/3P) as a function of phenol and TCE concentrations for phenol degrading culture when both substrates are present separately ............................................................................. 1 11 Figure 5.4 Dimensionless selection gradient, (P/W)(6W/8P) as a function of TCE concentration for methanotrophic mixed culture at a specified methane concentration ....................................................................................... 1 12 Figure 5.5 Dimensionless selection gradient, (P/W)(BW/8P) as a function of methane concentration for methanotrophic mixed culture at a specified TCE concentration ....................................................................................... 1 13 Figure 5.6 Sensitivity equation for qS as a function of TCE concentration for methanotrophic mixed culture at a specified methane concentration ......................... 114 Figure.5.7 Sensitivity equation for qs as a function of methane concentration for methanotrophic mixed culture at a specified TCE concentration ............................. 1 15 Figure 5.8 Sensitivity equation for qc as a function of TCE concentration for methanotrophic mixed culture at a specified methane concentration ......................... 116 Figure 5.9 Sensitivity equation for qc as a function'of methane concentration for methanotrophic mixed culture at a specified TCE concentration ............................. l 17 Figure 6.2 TCE feed concentration for long-term TCE exposure experiment in chemostate .......................................................................................... 126 Figure 6.3 Changes of Cell density of methanotrophic mixed cultures in respond to long-term TCE exposure ...................................................................... 134 Figure 6.4 Observed growth yield for methanotrophic mixed cultures during long- term TCE exposure ................................................................................ 135 Figure 6.5 Transformation capacity for methanotrophic mixed cultures during long-term TCE exposure ......................................................................... 135 Figure 6.6 Fitness of methanotrophic mixed cultures in respond to long-term TCE exposure in chemostats ........................................................................... 136 Figure 6.7 ARDRA fingerprints of control and TCE-exposed methanotrophic communities ........................................................................................ 137 Figure 7.1 Experimental setup for bench-scale sequencing batch reactors ................ 149 Figure 7.2 The operating mode of SBR in a cycle ........................................... 150 Figure 7.3 Changes in cell density of the phenol-degrading reactor communities. ...... 156 Figure 7.4 Second order rate coefficients of TCE transformation by phenol- degrading reactor communities. ................................................................. 157 Figure 7.5 Maximum specific rate of phenol utilization by phenol-degradin g reactor communities ............................................................................... 158 xii Figure 7.6 Long-term changes in observed yield for the phenol-degrading reactor communities ........................................................................................ 159 Figure 7.7 Fitness of phenol-degrading communities during period when the exposed reactor received an influent concentration 25 mg/L TCE ........................... 160 Figure 7.8 Change in the second order rate coefficient for TCE transformation in a typical SBR operating cycle ...................................................................... 161 Figure 7.9 TCE concentration in gas (Cg) and liquid (Cw) phase during the fill and react periods for a feed solution of 25 mg/L TCE ........................................ 162 Figure 7.10 DGGE fingerprints from control and TCE-exposed communities in sequencing batch reactors ........................................................................ 163 xiii NOMENCLATURE English symbols 996” a: 1:0 E a a '3»: fix. a. G is: 05 953333 “(139 (7; )obs (TQM first-order endogenous decay constant (d" ) concentration of nongrowth substrate (mg nongrowth substrate/L) initial concentration of nongrowth substrate (mg nongrowth substrate/L) final concentration of nongrowth substrate at t = 00 (mg nongrowth substrate/L) Selection gradient for fitness component, P Henry's constant of nongrowth substrate (-) Henry's constant of growth substrate (-) maximum specific rate of utilization of the nongrowth substrate (mg nongrowth substrate/mg cells-d) maximum specific rate of utilization of growth substrate (mg growth substrate/mg cells-d) = kc / Kc , second order rate coefficient of utilization of the nongrowth substrate (Umg cells-d) inhibition coefficient indicating the effect of nongrowth substrate concentration on growth substrate utilization rate (mg nongrowth substrate/L) inhibition coefficient indicating the effect of growth substrate concentration on nongrowth substrate utilization rate (mg growth substrate/L) half-saturation coefficient of nongrowth substrate (mg nongrowth substrate/L) half-saturation coefficient of growth substrate (mg growth substrate/L) mass of nongrowth substrate (mg) mass of growth substrate (mg) Malthusian parameter specific rate of utilization of nongrowth substrate (mg nongrowth substrate/mg cells-d) specific rate of utilization of growth substrate (mg growth substrate/mg cells-d) concentration of growth substrate (mg growth substrate/L) time (d) =- qclp , observed transformation capacity (mg nongrowth substrate/mg cells), theoretical transformation capacity in the absence of endogenous decay (mg nongrowth substrate/mg cells) = qc/q, , observed transformation yield (mg nongrowth substrate/mg growth substrate) xiv 7; theoretical transformation yield (mg nongrowth substrate/mg growth substrate) VG volume of gas phase (L) VL volume of liquid phase (L) W fitness X active organism concentration (mg cells/L) X 0 initial concentration of active organism (mg cells/L) Ym maximum yield or true growth yield (mg cells/mg growth substrate) Y observed yield (mg cells/mg growth substrate) Greek symbols [1 specific growth rate of organism (day'l) um maximum specific growth rate of organism (day'l) no specific growth rate of organism of the ancestor community (day-1) XV CHAPTER 1 INTRODUCTION The phenomenon of cometabolism was first reported by Leadbetter and Foster (1959). Since then, many microbial cometabolizing populations have been identified. Many compounds transformed by cometabolism are toxic and, therefore, of environmental concern because of deliberate or inadvertent release into waters and soils. Cometabolism may be a useful tool for removal of such contaminants, particularly those not readily catabolized, from natural environments and engineered systems. Quantitative understanding of cometabolic transformations will enable rational engineering design. COMETABOLIC TRANSFORMATION Under aerobic conditions, a number of organic compounds are transformed by cometabolism. These transformation are typically mediated by aerobic organisms possessing nonspecific oxygenase activity capable of oxidizing hydrocarbons as growth substrates and other compounds as nongrowth substrates. Transformation of growth substrate yields carbon or energy for the organisms; transformation of nongrowth substrate wastes the energy and reducing reserves of the cell. In the latter case, the transformation is termed cometabolic. This work focuses on monooxygenase activity. Monooxygenases (MO) cometabolically attack a broad range of compounds, including halogenated aliphatic compounds (RX): MO RX+02+2e-+2H+ —> RXO+H20 2 In a true sense, cometabolism is not metabolism (energy yielding) but fortuitous transformation of a compound by pathways that do not yield energy to the organisms. Enhancing these fortuitous side-reactions is a goal of engineered transformation using microorganisms. However, this goal may conflict with the primary objective of the microbe: the use of electrons for growth and respiration. The "dilemma" that microbes face in cometabolism is shown in Figure 1.1. In normal metabolism, the fraction of electrons used for energy generation (fe) plus the fraction of electrons used for synthesis (f8) will equal one. In cometabolism, a fraction of the electrons removed from the electron donor may also be used in cometabolic reactions (fco). Since the energy and the products of the transformation are unavailable for microbial use, fs-I-fe will decrease in the presence of a compound that is cometabolized. Thus, the successful removal of contaminants by cometabolism depends, at least in part, upon the electron donor requirements and the efficiency with which electrons can be directed to cometabolic transformation. ADAPTATION OF COMETABOLIZING COMMUNITIES Cometabolic transformation is a complex phenomenon, especially when both growth and nongrowth substrates are simultaneously present. Significant declines in methane conversion rates by methanotrophs following exposure to TCE are observed for both resting and formate-fed cells, suggesting toxic effects caused by TCE or its transformation products (Alvarez-Cohen and McCarty 1991a,b). The presence of toxic transformation products can be expected to have some impact on the development of microbial communities during long-term exposure to nongrowth substrate (Figure 1.2). Changes in populations are likely related to the level of exposure to nongrowth substrate, turnover of transformation products and utilization of growth substrate. | Growth substrate 1 e l f , Nongrowth substrate a CO S >| Cells 1 Electron acceptor Figure 1.1 Flow of reducing power in cometabolizing system The accumulation of stable nongrowth substrate breakdown products in pure cultures indicates that pure cultures are not able to mineralize nongrowth substrates. However, research shows that many mixed cultures and communities can achieve mineralization. Since pure cultures with oxygenases typically suffer from product toxicity, other populations capable of utilizing these toxic products may play an important role in detoxification. Under conditions of prolonged or repeated exposure to nongrowth substrate, selection pressures may favor shifts in the microbial community structure so as to enhance detoxification. Nongrowth Growth substrate \ substrate \ Enhanced \ \ transformation . . Cells . . Competltrve . Competitlve Inhibition NonSPeCIfic Inhlbltlon enzyme Finite transformation \\ capacity Useful products Useless and and/or biomass toxic products Figure 1.2 General scheme of cometabolic transformation. Dashed lines indicate the effects of substrates or products on transformation. 5 Different reactor environments may also have different effects on the adaptation of cometabolizing microbial communities. Communities in batch and plug flow reactors are exposed to high concentration of halogenated compound whereas continuous well-mixed systems are exposed to low concentrations continuously. Consequently, selective pressures in batch or plug flow reactors might be expected to favor cometabolizing communities with higher rate of transformation of halogenated compounds. RESEARCH OBJECTIVES The objectives of this research are to evaluate and verify a model for cometabolic transformation of nongrowth substrate and transformation of growth substrates and to use this model to characterize adaptive changes during long-term exposure to nongrowth substrate. The hypothesis are: l. Cometabolism can be quantified using a model based on saturation kinetics and incorporating terms for the loss of microbial biomass caused by endogenous decay, depletion of cofactors and product toxicity. 2. Under conditions of long term periodic exposure to nongrowth substrate, total biomass will be negatively affected by cometabolism, but can recover by means of changes in community structure. These changes can be characterized by phenotypic, morphological and genotypic parameters. 3. Exposure to nongrowth substrate forces selective changes in the community and enhanced nongrowth substrate degradation as secondary populations capable of detoxifying reaction products become more prevalent. 4. Simple cometabolizing communities derived from growth upon a single rate-limiting substrate can be treated as a single population in terms of fitness or growth. Both genotypic and phenotypic adaptation within populations will contribute to phenotypic changes of the whole community. For the characterization of the community, the phenotypic properties eValuated will be apparent values for the whole community. 6 5. Adaptive changes of community structure can be described in term of a fitness parameter defined as the ratio of specific growth rate during nongrowth substrate exposure to specific growth rate before exposure to nongrowth substrate. The value of fitness at any instant will be the total result of phenotypic adaptation of the community. The sensitivity of change in each phenotypic property to fitness can also be evaluated as a measure of community stability. OUTLINE OF THE THESIS This work was conducted in three phases. The first phase, in chapters 3 and 4, entails presentation and verification of a model for cometabolism kinetics. In chapter 3, the model is verified for resting cells using transformation of HCFC/HFCs by a methanotrophic mixed culture as a model system. In chapter 4, the model is extended to growing cells and verified experimentally for the methanotrophic mixed culture with trichloroethylene as the nongrowth substrate. In second phase of this work (chapter 5), the model verified in chapters 3 and 4 is analyzed in terms of " fitness " concept to describe adaptive changes within cometabolizing communities, Fitness is quantified in term of measurable kinetic parameters. The final phase in chapters 6 and 7 focuses on long-term adaptations of cometabolic model systems in response to long-term repeated exposure to nongrowth substrate. A methanotrophic mixed culture in a chemostat and a phenol-degrading community in a sequencing batch reactor were chosen as model systems representing extreme cases for adaptation. The engineering significance of this work is described in chapter 8. Finally, the dissertation concludes with a summary of the most important results and possible future studies. REFERENCES 1.Alvarez-Cohen, L., and P. L. McCarty. 19913. A cometabolic biotransformation model for halogenated aliphatic compounds exhibiting product toxicity. Environ. Sci. Technol. 25:1381-1387. 2.Alvarez-Cohen, L., and P. L. McCarty. 1991b. Effects of toxicity, aeration, and reductant supply on trichloroethylene transformation by a mixed methanotrophic culture. Appl. Environ. Microbiol. 57:228 - 235. 3.Leadbetter, E. R., and J. W. Foster. 1959. Oxidation products formed from gaseous alkanes by the bacterium Pseudomonas methanica. Arch. Biochem. Biophys. 82:491-492. CHAPTER 2 BACKGROUND COMETABOLISM PRINCIPLES General concept Many compounds of environmental and toxicological significance are transformed by cometabolism. In this study, cometabolism is defined as transformation of a nongrowth substrate that depends upon the previous or concurrent utilization of a growth or nongrowth substrate (Criddle 1993; Horvath 1972). A growth substrate is defined as an electron donor that supports growth. An energy substrate is defined here as an electron donor that provides reducing power and energy for the transforming population, but does not by itself support growth. The term "co-oxidation" was first used to described cometabolism because the original observations all involved oxidations. Subsequently, reductive transformations were discovered that did not facilitate growth of the transforming organisms but still depended upon the concurrent or previous utilization of a growth or energy substrate. These "co- reduction" reactions led to the use of the broader term cometabolism. It now appears that certain cometabolic transformations are also hydrolytic. Thus, in addition to the well known examples of co-oxidation, we now recognize the potential for "co-reductions" and "co-hydrolyses". All of the known co-oxidations occur only under obligate aerobic conditions, while most co-reductions occur under anaerobic conditions 9 Cometabolism is important for many transformations, including some polynuclear aromatic hydrocarbons, halogenated aliphatic and aromatic hydrocarbons, and pesticides. Many of these compounds express toxic properties and are environmental concern. This phenomenon has been observed so frequently that it appears to represent a very important type of microbial metabolism (Table 2.1). Many microbial species exhibit the phenomenon of cometabolism. Table 2.2 lists those microorganisms which have been clearly shown to possess transformation by cometabolism. Table 2.] Organic compounds subject to cometabolism and accumulated products -Compound Product Ethane Acetic acid Propane Propionic acid, acetone Butane Butanoic acid, methyl ethyl ketone m-Chlorobenzoate 4-Chlorocatechol, 3-chlorocatechol o-Fluorobenzoate 3-F1uorocatechol, fluoroacetate 2-Fluoro-4-nitrobenzoate 2-Fluoroprotocatechuic acid 4—Chlorocatechol 2-Hydroxy-4-chloro-muconic semialdehyde 3, 5-Dichlorocatechol 2-Hydroxy-3,5-dichloro—muconic semialdehyde 3—Methylcatechol 2-Hydroxy-3~methyl-muconic semialdehyde o—Xylene o-Toluic acid p-Xylene p-Toluic acid, 2, 3-dihydroxytoluic acid Pyrrolidone Glutamic acid Cicerone Cinerolone naphthalene salicyclic acid n-Butylbenzene Phenylacetic acid Ethylbenzene Phenylacetic acid n-Propylbenzene Cynnamic acid p-Isopropyltoluene p-Isopropylbenzoate n-Butyl-cyclohexane Cyclohexaneacetic acid 2, 3, 6-Trichlorobenzoate 3, 5-Dichlorocatechol 2, 4, 5-Trichlorophenoxy acetic acid 3, 5-Dichlorocatechol p-p'-Dichlorodiphenyl methane p-Chlorophenylacetate l, 1-Diphenyl—2, 2, 2-trichloroethane 2-Phenyl-3, 3, 3-trichloropropionic acid 1, 1, l-Trichloroethene l, 2—dichloroethene Trichloroethylene Trichloroacetate, 2,2,2-trichloro-ethanol, dichloroacetate Source: (1) Horvath 1972; (2) Dalton and Stirling 1982 10 Table 2.2 Microorganisms exhibiting the phenomenon of cometabolism ficroorganism Acetobacterium woodii Achromobacter sp. Acinetobacter sp. Arthrobacter sp. Aspergillus niger Azotobacter chroococcum Azotobacter vinelandii Bacillus megaterium Bacillus sp. Brevibacterium sp. Clostridium sp. F lavobacterium sp. Hydrogenomonas sp. Methylomonas sp. Microbacterium sp. M icrococcus certficans Micrococcus sp. Nitrosomonas europaea Nocardia erythropolis Nocardia sp. Pseudomonas sp. P. fluorescens P. methanica P. putida Rhodococcus sp. Streptomyces aureofaciens Trichodemra viride Vibrio sp. _Xgnth0monas sp. _ Source: (1) Horvath 1972; (2) Dalton and Stirling 1982 ; (3) Criddle 1993 Energy requirement Many aerobic cometabolic reactions are catalyzed by non-specific oxygenase enzymes that use 02 as the electron acceptor and NADH as the reducing energy source to oxidize both growth and nongrowth substrates (Colby et al. 1977; Fox et al. 1990; Nelson et al. 1987; Wackett et al. 1989). These enzymes are the methane monooxygenases (MMO) of methanotrophs (Fox et al. 1990; Oldenhuis et al. 1989; Tsien et a1. 1989), ammonia monooxygenases of nitrifiers (Arciero et al. 1989; Hyman et a1. 1988; Rasche et a1. 1990; 11 Vannelli et al. 1990), propane monooxygenases (Arciero et al. 1989; Hyman et al. 1988; Vannelli et a1. 1990; Wackett et a]. 1989), certain toluene mono- and dioxygenases, and certain phenol monooxygenases (Winter et al. 1989; Zylstra et al. 1989). After the initial oxidation step, growth substrates are further degraded to regenerate reducing energy (NADH), which promotes more substrate oxidation. However, the oxidation of nongrowth substrate in the absence of growth substrate can cause the depletion of NADH in cells since NADH is not regenerated. Thus, energy or reducing power must be present to transform the nongrowth substrate. Transformation can not be sustained if growth substrate is not supplied continuously or intermittently. Substrate interaction In the metabolism of multiple substrates, competitive inhibition is frequently reported. A number of substrate interactions have been observed during hydrocarbon degradation by cometabolism involving monooxygenases and dioxygenases. Saéz and Rittmann (1991, 1993) reported that batch experiments on the simultaneous utilization of phenol and 4- chlorophenol by Pseudomonas putida PpG4 demonstrated 4-chlorophenol inhibited the oxidation of growth substrate and the cometabolic degradation of 4—chlorophenol was proportional to the rate of phenol oxidation. Competitive inhibition between phenol and trichloroethylene (TCE) was also observed for the degradation by Pseudomonas cepacia G4(Folsom et al. 1990). Studies on TCE degradation by methanotrophs (Alvarez-Cohen and McCarty 1991a; Anderson and McCarty 1994; Broholm et al. 1992; Chang and Alvarez-Cohen 1995a; Saéz and Rittmann 1993) indicated that competitive inhibition was generally present between methane (growth substrate) and TCE (nongrowth substrate). Chang et a]. (1993) also revealed competitive inhibition and cometabolic patterns using paired substrates (benzene, toluene, and p-xylene). Some research also concluded that the degradation of nongrowth substrates is enhanced in the presence of growth or energy 12 substrate (Chang and Alvarez-Cohen 1995a; Chang et al. 1993; Criddle 1993; Saéz and Rittmann 1993). Product toxicity Several examples indicate that cometabolism by pure cultures does not typically result in the mineralization of nongrowth substrates (Table 2.1). Horvath (1971) reported that cometabolism of 2, 3, 6-trichlorobenzoate resulted in accumulation of 3, 5-dichlorocatechol and development of a toxic environment for the cells. Several researchers have shown that stable and toxic TCE breakdown products accumulate in pure cultures of methanotrophs. This suggests that methanotrophic bacteria in isolation are not be able to effectively mineralize TCE (Henry and Grbié -Galié 1990; Little et al. 1988; Oldenhuis et al. 1989). Significant declines in methane conversion rates following exposure to TCE were observed for both resting and formate-fed cells, suggesting toxic effects caused by TCE or its transformation products (Alvarez-Cohen and McCarty 1991a; Alvarez—Cohen and McCarty 1991b). Cometabolic community The products of cometabolic transformation accumulate in pure cultures, but, in a mixed culture, they are typically used by other microorganisms. As a result, cometabolic transformations are key initiatory reactions in pathways that ultimately result in the complete degradation of many environmental pollutants. Some research shows that methanotrophic mixed cultures have advantages for complete degradation of TCE. Since methanotrophs suffer from product toxicity in TCE transformation, heterotrophs in mixed culture may play an important role in detoxification. Heterotrophic bacteria in the methanotrophic mixed cultures apparently can degrade most of the water-soluble breakdown products from l4C-TCE, decreasing levels of water-soluble radiolabel and l3 increasing production of 14C02 (Little et al. 1988). Futherrnore, Uchiyama and co- workers reported that a heterotrophic bacterium in a methanotrophic mixed culture, Xanthobacter autotrophicus, can can oxidize dichloroacetic and glyoxylic acid completely and can reduce trichloroacetic acid levels (Uchiyama et al. 1992). These results indicate that heterotrophic bacteria in microbial communities play an important role in complete degradation of nongrowth substrates. Another good example is the initial cometabolic transformation of PNAs, which is typically followed by a series of degradation steps leading to C02. The initial step is not mediated by the same organisms as the mineralization steps. QUANTIFICATION OF COMETABOLISM Cometabolism by resting cells Nongrowth substrates can be transformed by resting cells in the absence of growth substrates. Under these conditions, cells utilize nongrowth substrate in the absence of growth substrates. Transformations of growth substrate by cells can generally be described using a saturation kinetic expression: S K+S S CI. =k,( ) (1) Transformations of nongrowth substrate by resting cells can also be described using saturation kinetics: =k 4‘ ‘(K +C C ) (2) If the substrate concentration is low (S << K,, C << K, ), the specific transformation rate is directly proportional to the substrate concentration: 14 %=ks (» q, = k' C (4) At high substrate concentration (S >> Ks, C>> K), the specific rate of substrate transformation is independent of substrate concentration( qs = ks , qc = kc). In some cases, a growth substrate inhibits its own transformation at high concentration. To describe this situation, Haldane kinetics is often used: k S 4. = 4'37 (5) Ks + S + —- K As discussed previously, loss of cometabolic transformation activity can occur as a result of endogenous decay and product toxicity. To evaluate the loss of cell activity during the transformation of nongrowth substrate, first order decay of biomass is usually used: — = -bX (6) To account for the loss of transformation activity in resting cells caused by a depletion of reducing power (in the absence of growth or energy substrate) and by product toxicity, the concept of "transformation capacity" is applied. Transformation capacity was first defined as (Alvarez-Cohen and McCarty 1991a): dC (72),,” = Ti} (7) 15 Alvarez-Cohen and McCarty assumed that biomass transformation capacity was equal to the mass of nongrowth substrate ultimately degraded divided by the initial biomass used. Criddle (1993) defined a “theoretical” biomass transformation capacity by correcting for losses caused by endogenous decay. The later definition represents a theoretical maximum value in the absence of external reducing power. Most models proposed to describe transformations of nongrowth substrate under resting cell condition are a combination of Eq. (2), (6), and (7). Criddle (1993) proposed a model that unified the earlier models. This model assumed that nongrowth substrate is degraded with saturation kinetics and endogenous decay and that loss of cometabolic activity can be attributed to product toxicity are incorporated into cell decay term. A surrunary of these models is shown in Table 2.3. Cometabolism by growing cells Earlier cometabolic models presented in the literature focus on resting cells, yet frequently, both the rates and extent of cometabolism are enhanced in the presence growth substrate and energy substrates. Competitive inhibition between growth and nongrowth substrates (Folsom et al. 1990; Strand et al. 1990) or between multiple nongrowth substrates (Alvarez-Cohen and McCarty 1991c) is also observed. When there is competitive inhibition between the growth substrate and the nongrowth substrate, (12),,” and (KC )0,” replace K, and KC respectively in Eq. (1) and Eq. (2), where: C (Ks )obs _ Ks(1+ E) (8) S (Kclob, — Kc(1+ F) (9) 13 Transformations of growth and nongrowth substrate by cells under the conditions of competitive inhibition can generally be described as follow: 16 Table 2.3 Summary of cometabolic transformation model by resting cells Differential equations for substrate Model utilization rate Integrated form Ref. dC k or dX _ c _ —k X0 _b, 1a -'Z=K:+C and E—-bx Kcln(E;)+C—C0——;—(1-e ) 1’2 _£_ chXO e—bt dt K + C dC c__kx0e —br “3 dt —kc CX andd dt bX C- Coe b 23 c<< K, 80 ”g = chXoe'b' —i—C;kx and—--bX Sarneasmodel3where (T) -k—‘ 1° dr dt _ C 0’” _ b 4 dC - C>> Kc SO -E = kCCXOe b' _d_C__ chX and— = (T )0,” ,= _L Kc ln CXo 2a d—t— —KC + C dxl kc (( C0 “Ebb: — 0) FCO 5 kCC(X -——(C —C) X dc 0 (a)... 0 ) +...n n{—"}> so —— = — F dt Kc + C 1 where F: X0 -——)—(C0 — C) c obs dC ' —k' F': -_ _ F ‘ 2b dt = k CX and: = (T )obs C = C0 20 k. F} 4 dC X0 - ————e_ ( C<< Kc SO -7 = ch(X0 _ 1; (C0 _ C» (Tc)obs where F' = X0 — C0 (7;)obs dC d S kc 2c mat—=1: X an —=(T)0,,, ameasmodel 3where (T )obs—Z- 4 dC I so -— = k x —— C - C>> Kc d: c( 0 cm )obs ( 0 C3) dX 3 E-‘bX and 'd_C=(T)obS C=C0_(T~C)0bsxo(1_e-bl) 6 so —i(-:-- —bT Xoe "" d_t _d_C_ k CX and 4 d, = K +C No integrated form 4 (_1x_ = —bX-1—(——k cx ) d: T K + C Reference: (1) Galli and McCarty 1989; (2) Schmidt et al. 1985; (3) Criddle et al. 1990; (4) Criddle 1993; (5) Alvarez-Cohen and McCarty I991a; (6) Saéz and Rittmann 199]. 17 S K,(1+—C—)+s K. lC 6]. =k.( ) (10) c s KC(I+——)+C K. (S q. =k..( ) (11) To describe the observed enhancement in the rate of cometabolism in the presence of growth substrate, modification of Eq (10) or Eq (11) were proposed. Chang and Alvarez- Cohen (1995a) proposed a general model that includes reducing energy explicitly as a limiting reactant during cometabolism. They applied the model to describe degradation of TCE by methanotrophs. The effects of reducing power were separated from toxicity effects and quantified by supplying the cells with formate. For methanotrophs, forrnate provides energy as NADH, but does not support growth. Thus, the experimental approach of Chang and Alvarez-Cohen is somewhat specific to organisms for which energy substrates can be identified that do not support growth. By contrast, Criddle (1993) introduced a term, the growth substrate transformation capacity, or theoretical transformation yield, to quantify the enhancement of cometabolism resulting from oxidation of the growth substrate: 4. = (T,.q. + k.)( CS ) (12) K I+— +C .( K”) (S Use of Eq. (12) does not require explicit quantification of reducing power and may be applied to growth and energy substrates. Eq. (12) is related to the Luedeking-Piret (LP) model that is commonly used in fitting product formation data from many different fermentation (Bailey and Ollis 1986). In the LP model, qp = all + B, where qp = specific 18 rate of product formation, a = the growth-associated product yield, and ,8 = nongrowth- associated product formation rate. The LP model can be modified to apply to cometabolism. Consequently, qc =(ap+fi)/ Yp”, where Yp” =the mass of product formed by cometabolism of a unit mass of nongrowth substrate. The same result is derived with Eq. (12) when C >> Kc. Thus, the LP model is obtained as a limiting case of Eq. (12). The most widely used model of cell growth and decay is the Monod expression as modified by Herbert et a] (1956): Y kS =Y —-b=—”"‘——b 13 # qu K +S ( ) S For cells that are growing, decaying, and simultaneously carrying out cometabolic transformation, Eq. (13) is an inadequate description of growth. A modification proposed by Criddle (1993) and used by Anderson and McCarty (1994) and by Chang and Alvarez- Cohen (1995a) is: ‘1. =Y — ——‘ 14 u ’7qu T ( ) C Eq(l4) indicates that an increase in the rate of cometabolism causes a decrease in the specific growth rate of a cometabolizing population. While this appears to be true for pure cultures, it may not be true for mixed cultures where organisms capable of using the products of cometabolism for growth are frequently present. Eq(10), (12), and (14) provide the framework for a complete kinetic description of cometabolism by growing or resting cells and were used throughout this research. A summary of these models is shown in Table 2.4. 19 Table 2.4. Summary of cometabolic transformation model by growing cells 'l | Model Differential equations for substrate utilization rates and growth rate Ref. l ; -__1_£_ k,s (g) I 1 ‘15 th K,+s ___l_ig ‘K. 1 I X ‘1‘ (1+—S—+—C—) l K: KC 1 q __i£_ k,s q __i£€_ ch 2 X ‘1‘ K,(1+—)+s X ‘1‘ K,(1+-§—)+C 2 1 (IX =——=r —b u Lit q. as 1 dC ____._. + 3 (1+5: qc X d1 or], lab 3 q _ 1 dS_ K, 3__——" n2 2 X dt K,(1+Z—2-)+S+§— K2 Kh 1 (1X =——=Y -b u x it 4. q =_l£S__ kss q —.__1__‘.1£- ch 4 X ‘1’ K,(1+£)+s c X ‘1‘ Kc(1+i)+c 4 KC [(5 1 =X - (C -C) 0 (Tc)obs 0 5 q 1dS_ k,s q_ ldC_ ch s‘—_—- c-—_—"' X ‘1’ K,(1+—IS—)+S X ‘1' Kc(l+i)+C 5 c s | 1 (IX q l =——=Y — —- c ” x d: 4’ (a)... l 1 d5 R KS 1 dC R k C ' qs=———=( )( 5 ) qC=———=( )( C ) | 6 th KR-i-R K,(1+—C—)+s th KR+R Kc(1+-k§-)+C 6 ; 1 dX' q l =__=y _b.._£ ‘1 X dt 4 Tc ; 1 (15 as 1 dC ch 7 qs=———= qc=-——=(7}q.+kc)( ) th KAN-[EMS X dt Kc(1+ )+C 7 fr is l -i__ _ _4_c I Renferece: (1 trandt a] 1990; (2) Broholm et a1 1992; (3) Saéz and Rittmann 1993; (4) Alvme_hen and McCarty 1991a; (5)Anderson and McCarty 1994; (6) Chang and Alvarez-Cohen 1995a; (7) Criddle 1993. 20 MODEL MICROBIAL SYSTEMS FOR COMETABOLISM Several aerobic bacteria with non-specific oxygenase activity are capable of oxidizing halogenated hydrocarbons. Microorganisms possessing this ability include toluene- oxidizing bacteria (Nelson et al. 1987; Wackett and Gibson 1988), methane-oxidizing bacteria (Little et al. 1988; Oldenhuis et al. 1989; Tsien et al. 1989), ammonia-oxidizing bacteria (Arciero et a]. 1989; Hyman et al. 1988; Rasche et al. 1990; Vannelli et al. 1990), and propane-oxidizing bacteria (Wackett et al. 1989). The enzymes which have been implicated in catalyzing halocarbon oxidations are toluene mono- and dioxygenase (Winter et al. 1989; Zylstra et al. 1989), methane monooxygenase (Fox et al. 1990; Oldenhuis et al. 1989; Tsien et a]. 1989), ammonia monooxygenase (Arciero et a]. 1989; Hyman et al. 1988; Rasche et al. 1990; Vannelli et al. 1990), and propane monooxygenase (Wackett et al. 1989), respectively. A variety of non-specific oxygenases that attack TCE in aerobic environments are listed in Table 2.5. As discussed previously, a diverse range of cometabolizing activities are present in the environment. Non-specific oxygenase activities within communities are the focus of this work. Such activity can be found in many hydrocarbon-degrading communities, the most well studied of which is the methanotrophs. Accordingly, a simple well-defined methanotrophic community was selected as a model cometabolizing community for this work. The second community selected was a phenol-degrading enrichment. This community was chosen because of its high growth rates, its high transformation capacity for TCE, and ease of handling of phenol in laboratory studies. Additional discussion of both methanotrophic and phenol-degrading communities are provided in the following sections. 21 Table 2.5 Examples of cometabolic transformation of trichloroethylene i Microorganisms Growth substrate References l Pseudomonas cepacia strain phenol, toluene, o-cresol 1, 2 . G4 phenol 3, 4 Pseudomonas putida Fl toluene 5, 6 Strain 46-1 toluene 7, 8 Methylosinus trichosparium methane 9, 10 OB3b methanol 11 forrnate 12 IIMethylocystis sp. strain M methane 13 14, 15 Mycabacterium vaccae propane 16 J 0B5 Nitrosomonas europaea ammonia 17, 18, 19 Xanthobacter strain Py 2 propylene 20, 21 Genetically engineered toluene 21 Escherichia coli References: (1) Ne son et a1. 1986; (2) Nelson et a . 1 87; (3) F0 som et a . 1 9 ; ( ) Folsom &Chapman 1991; (5) Nelson et a1. 1988; (6) Wackett & Gibson 1988; (7) Fox et al. 1990; (8) Little et al. 1988; (9) Oldenhuis et al. 1989; (10) Oldenhuis et al. 1991; (11) Tsien et al. 1989; (12) Newman & Wackett 1991; (13) Nakajima et al. 1992; (14) Uchiyama et al. 1989; (15) Uchiyama et al. 1992; (16) Wackett et al. 1989; (17) Arciero et al. 1989; (18) Hyman et al. 1988; (19) Rasche et al. 1991; (20) Ensign et al. 1992; (21) Zylstra et al. 1989. Methanotrophic communities Methanotrophs are classified into two major groups depending on their internal cell structure and carbon assimilation pathway. Type I organisms assimilate one-carbon compounds via a unique pathway, the ribulose monophosphate cycle, whereas Type II organisms assimilate C-l intermediates via the serine pathway. The requirement for 02 as a reactant in the initial oxidation of methane explains why all methanotrophs are obligate aerobes, whereas some organisms using methanol as electron donor can grow anaerobically (with nitrate or sulfate as electron acceptor). Both groups of methanotrophs contain extensive internal membrane systems, which appear to be related their methane- 22 oxidizing ability. Type I methanotrophs are characterized by internal membranes arranged as bundles of disk-shaped vesicles distributed throughout the organism whereas Type II methanotrophs possess paired membranes running along the periphery of the cell. Type I methanotrophs are also characterized by a lack a complete tricarboxylic acid cycle (the enzyme a-ketoglutarate dehydrogenase is absent), whereas Type II methanotrophs possess a complete cycle. In addition, most Type II methanotrophs can fix molecular nitrogen whereas Type I organisms do not. The classification and characteristics of methanotrophic bacteria are listed in Table 2.6. The methylotrophs are capable of growth on a variety of organic compounds; however, they cannot use methane as carbon and energy sources. All methanotrophs can grow on methane, many are also able to utilize methanol and formaldehyde, and a few can use a wider range of organic compounds. Most methanotrophs are obligate methylophiles, which means that they are incapable of growth on compounds that contain carbon-carbon bonds. The responsible enzyme of methanotrophic bacteria, monooxygenase, catalyzes the incorporation of one oxygen atom from molecular oxygen into methane to produce methanol. The lack of substrate specificity of the monooxygenase enzyme enables it ability to oxidize a broad range of compounds, including halogenated aliphatic compounds. Monooxygenases can hydroxylate many alkanes and aromatic compounds and form epoxides from alkenes (Semprini et al. 1992). The epoxides are unstable and hydrolyze to acids. Since some products of these reactions are not further metabolized by methanotrophs, a community of microorganisms is necessary for mineralization. 23 Table 2.6 Tentative classification Scheme for Methanotrophic Bacteria* Characteristic Groupl GroupX Group H Morphology Straight rod Coccus Straight, curved or pear-shaped rod Membrane arrangement Bundles of vesicular disks + + - Paired peripheral membranes - - + Motility i - i: Resting stage Azotobacter-type Azotobacter-type Lipid cyst or terminal cyst cyst exospore Rosette - - + (most strains) Major carbon Rump Rump Serine assimilation pathway Autotrophic C02 - + - fixation Complete TCA cycle - - + Nitrogenase - + + Isocitrate dehydrogenase NAD’ and NAD(P)+ specific + ' - NAD l specific NAD(P)‘ specific - + - - - + Glucose-6—dehydrogenase +( NADP + — specific) +(NADP l - specific) "a” gigggefigwwc +(NADP ’ - specific) +(NADP; - specific) ' Predominant fatty acid 16** 16 18 carbon-chain length Growth at 45°C Variable + _ Mol% G+C of DNA 50-54 62.5 61.7-63.1 * Green 1992. Kinetic coefficients for methane utilization and TCE transformation reported in literature are summarized in Table 2.7 - 2.8. Although units have been standardized for purposes of comparison, differences in experimental protocol and methods of reporting sometimes made an accurate and complete comparison impossible. Clearly, broadly accepted unstructured models of cometabolism and standardized experimental protocols are needed to enable fair comparison of organisms from different sources and assist in the design and operation of engineering systems. 24 Table 2.7 Kinetic coefficients of methane utilization ks Ks k. 3 Re 3. . mg cells/ mg CH4] mg CH4/L [lmg cell- day-1 mg CH4 mg cell- day , day " Methanotrophic 0.34 (obs.) 1.13 0.67 Strand et 7 mixed culture 0.5] (true) J al. 1990 f Methanotrophic 0,221 1,728: 0.2 0.12 Broholm et 7 1, mixed culture (Ki=12 a1. 1992 . mg/L for 1 TCE) _ T Methanotrophic 0.33 0.94 1.07 0.18 Chang . j mixed culture &Alvarez- 1 Cohen ' ; 1995a . Methanotrophic 0.35 Alvarez- , mixed culture Cohen I &McCarty , 1991a ; Mefiylosinus 8.37 1.47 7.49 Oldenhuis ' trichosporium et al. 1991 . OB3b : Methanotrophic 0.14 Henry & mixed culture Grbié -Galié ; 1991 l Methylomonas 0.26-0.73 0.16-1.63 Henry & ’ Grbic’: ~Galic’: I 1990 l on protein content. 25 Table 2.8 Kinetic coefficients of TCE cometabolic transformation by methanotrophic communities Culture! growth substrate kc 6 (TV )0,” mg TCE/ mg TCE/ L leg mg TCE/ mg mg TCE/ mg cell—day cell-day growth mg cell substrate Methanotrophic mixed 0.009 culture/ methane Methanotrophic mixed (1011a culture/ methane Methanotrophic mixed 1.03 3.84 0.05 culture/ methane (- formate) (- formate) (- formate) 4.17 6.95 0.1 (+fonnate) (+fonnate) (+fonnate) l' Methanotrophic mixed 0.017 0.05 culture/ methane (- formate) (- formate) 0.034 0.1 (+fonnate) (+fonnate) Methanotrophic mixed 0.84 0.69 0.043 culture/ methane (- formate) (- formate) (-formate) 4.8 7.9 0.061 (+fonnate) (+fonnate) (+formate) Methanotrophic mixed 5.] 7.3 0.013 0.036 culture/ methane (+fonnate) (+fonnate) Methanotrophic mixed 0.84 1.5 0.042 culture/ methane Methylosinus trichosporium 41.6 18.1 OB3b/ methane (+fonnateL (+fonnate) Methylosinus trichosporium 54.9 19.] 2.88 OBBbl methane Methylomonas sp. MM2/ 0.046-0.29 0.51-1.35 0.003- methane 0.86 Methanotrophic mixed 0.61 culture/ methane sMMO from Methylosinus 64 4.83 m'chosporium OB3b/methane a. biomass based on protein content. b. References: (1) Strand et al. 1990; (2) Broholm et al. 1992; (3) Chang and Alvarez-Cohen 1995a; (4) Chang and Alvarez-Cohen 1995b; (5) Alvarez-Cohen and McCarty I991a; (6) Alvarez-Cohen and McCarty 1991b; (7) Alvarez-Cohen and McCarty 1991c; (8) Brusseau et al. 1990; (9) Oldenhuis et a1. 1991; (10) Henry and Grbié -Galié 1990; (11) Fox et a1. 1990. Phenol-oxidizing bacteria Another group of microorganism that exhibit cometabolic oxygenase activity are bacteria that degrade aromatic compounds. aromatic-oxidizing microorganisms usually grow faster than methanotrophs by one or two orders of magnitude. The faster growth kinetics and 26 relative ease of addition of aromatic compounds give aromatic-oxidizing microorganisms certain pratical advantages in reactor systems. Some Pseudomonas species produce aromatic oxygenases that can degrade halogenated alkenes, including TCE (Nelson et al. 1987; Wackett and Gibson 1988). Other research established that Pseudomonas putida PpG4 would utilize phenol as growth substrate and cometabolize 4—chlorophenol (Saéz and Rittmann 1991; Saéz and Rittmann 1993). Previous results (Gottschalk 1986) supported that the initial step of 4-chlorophenol degradation, a monooxygenase-mediated attack on 4- chlorophenol, requiring Oz and NADPH as cosubstrates. Phenol oxidation supplies the electrons needed to regenerate the NAPDH cosubstrate. Phenol and toluene are typical inducing agents for this activity. In aromatic oxygenase systems, TCE is degraded to formate, carbon monoxide, and glyoxylic acid in pure culture(Wackett and Householder 1989; Winter et al. 1989). P. cepacia G4, degraded TCE to C02, Cl' and unidentified, nonvolatile products (Nelson et al. 1987; Nelson et al. 1986). Microorganisms with phenol- or toluene-degrading ability include bacteria, such as Pseudomonas (Beltrame et al. 1980; Yang and Humphrey 1975), Nocardia (Rizzuti and Augueliaro 1982), and Bacillus (Buswell 1975); yeast, such as trichosporon (Gaal and Neujahr 1979); and multicellular fungi, such as Fusarium (Anselmo et a1. 1985). Although many of these cultures (Pseudomonas strain G4, Pseudomonas putida F1, Pseudomonas putida BS, Pseudomonas putida PpFl) can transform TCE (Nelson et al. 1987; Nelson et al. 1986; Nelson et al. 1988; Wackett and Gibson 1988), the ability to degrade aromatic compounds does not always correlate with the ability to degrade TCE. The MMO system of the methanotrophs appears to be somewhat more consistent in its ability to degrade TCE although the rates of oxidation vary substantially. Phenol-oxidizing microorganisms have demonstrated effective transformation of cis- and trans- dichloroethylene and trichloroethylene in laboratory and in-situ field studies 27 (Hopkins et al. 1993a). The phenol-oxidizing microorganisms appear to have a much higher capacity to degrade trichloroethylene than the methanotrophs. Trichloroethylene degradation of 90 percent has been achieved with 99.8 percent removal of injected phenol. Separate laboratory studies suggest that the addition of noncompetitive external reducing power may significantly increase the transfomation potential. Trichloroethylene transformation capacities were enhanced by the addition of aliphatic compounds, the greatest enhancement being with formate or lactate (Hopkins et al. 1993b). Kinetic coefficients for phenol utilization and TCE transformation reported in the literature are summarized in Table 2.9 - 2.10. 28 Table 2.9 Kinetic coefficients of phenol utilization 2 ture Y Us] a. k. K. K.- Re 8- l mg ce 1/hr m henol/ m henol/ m henol/ l "18 Pbenol 11.35.11-111 L g p L g p l l Mixed culture NA 0. 13]- 5-266 142-] 199 1 l ; 0.363 l LMixed culture 0.7-0.9 0.66 16.5 634.4 2 l P putida Sp 0.55 0.1 19 5.27 377 3 g P. putida 0.53-1.84 0.5-1.23 8-20 4 i Mixed culture 0.326 19.2 229 5 1 Mixed culture 0.545 0.260 24.5 173 6 (non filament) Mixed culture 0.616 0.223 5.8 934 6 (filaments) Mixed culture 0.2] 630.41 2 Mixed culture 0.45 0.1 17 245 7 Mixed culture 0.07 2 Mixed culture 0.0] 1- 8 0.030 P. fluorescens 0.08 9 NocarJia 0.37 9 P. cepacia G4 2.6 0.8 43 10 Mixed culture 0.55 1 1 Mixed culture 0.051- 12 0.135 =——————_—_—— Refences: (1) D'Adamo et al. 1984; (2) Auteinrieth et al. 199]; (3) Kotturi et a1. 1991; (4) SoKol 1988; (5) Szetela and Winnicki 198]; (6) Pawlowsky and Howell 1973; (7) Beltrame et al. 1980; (8) Tischler and Eckenfelder 1969; (9) Rizzuti and Augueliaro 1982; (10) Folsom et al. 1990; (1]) Chang and Alvarez-Cohn 1995b; (12) Shih et al. 1996. 29 Table 2.10. Kinetic coefficients of TCE cometabolic transformation by phenol-oxidizing bacteria ‘ Culegrtu/owth subsate — L/mg mg cell-day 031le growth mg cell . substrate ! Mixed culture! phenol 0.1 1 0.24 1 . (-phenol) (-phenol) 0.01 0.03 , (+phenol) (+phenol) ' Mixed culture! phenol 0.017 0.031 2 (-phenol) (-phenol) 0.019 0.034 (+phenol) (+phenol) Mixed culture/ phenol 0.12-0.20 3 Mixed culture/ phenol 0.10 0.35 0.0026- 4 0.11 Pseudomonas cepacia G4/ 0.74 0.40 5 phenol Pseudomonas putida Fll 0.0162 6 toluene E a. biomass based on protein content. b. References: (1) Hopkins et al. 199321; (2) Chang and Alvarez-Cohen 1995b; (3) Coyle et al. 1993; (4) Shih et a1. 1996; (5) Folsom et al. 1990; (6) Wackett and Gibson 1988. MODEL NONGROWTH SUBSTRATES Halogenated hydrocarbons containing one or two carbon atoms constitute a significant fraction of the hazardous substances from industrial , domestic, and agricultural sources. These compounds tend to be mobile and persistent in soil and groundwater. Some have the potential for ozone depletion. For this work, the model compounds studied as nongrowth substrates were selected hydrochlorofluorocarbons (HCFCs), a hydrofluorocarbons(HFC), and trichloroethylene(TCE). The ban on CFCs has promoted the widespread use of HCFCs and HFCs. The presence of hydrogen makes HCFCs and HFCs more susceptible to tropospheric oxidation than the CFCs, and thus less likely to migrate into the stratosphere. To date, there is relatively little information on the fate of HCFCs or HFCs in aquatic environments. Trichloroethylene is a commonly detected 30 groundwater contaminant and is classified as a priority pollutant by the U. S. Environmental Protection Agency. A detailed discussion of these model compounds and their known properties for biodegradation is provided in the following sections. HCFC/CFCS Chlorofluorocarbons (CFCs) are widely used refrigerants and aerosols in industry and domestic life. Over the past decade, they have been implicated as agents of depletion of stratospheric ozone and as contributors to global warming (Molina and Rowland 1974)(Molina & Roland 1974; Rowland & Molina 1975). As a result, worldwide production of CFCs will be banned under the terms of Montreal Protocol. Nevertheless, CFCs will continue to be released into the environment due to past production and continued use. In aerobic aquatic environments, CFCs are recalcitrant, but they are transformed anaerobically (Denovan and Strand 1992; Lesage et a1. 1992; Lovely and Woodward 1992; Semprini et al. 1992). The ban on CFCs has inspired a major research effort to assess two classes of CFC substitutes - the hydrochlorofluorocarbons (HCFCs) and the hydrofluorocarbons (HFCs). HCFCs and HFCs are one- and two-carbon aliphatics, similar in structure and physical properties to the CFCs, but containing one or more hydrogen atoms. The presence of hydrogen makes HCFCs and HFCs more susceptible to tropospheric oxidation than the CFCs, and thus less likely to migrate into the stratosphere. To date, there is relatively little information on the fate of HCFCs or HFCs in aquatic environments. Lessage et al. (1992) reported transformation of HCFC-123a to HCFC-133 and HCFC-l33b under methanogenic conditions. DeFlaun et al. (DeFlaun et al. 1992) reported aerobic transformation of three HCFCs and one HFC by Methylosinus trichosporium OB3b. The properties and purity of chemicals used as model nongrowth substrates in this work are summarized in Table 2.1 1. 31 Table 2.1] Properties and purity of HCFCs and HFC evaluated in this work. Solubility in Boiling DEnsity, water,wt%, point,°C, g/cm3, Purity, Compound Chemical name @25°C @760mm Hg @25°C % HCFC-22 chlorodifluoro- 0.30 -40.8 1.194 99.9788 methane HCFC-142b l-chloro-1,1— 0.14 -9.2 1.108 99.9609 difiuoroethane HCFC-123 1,1-dichloro- 0.21* 27.9 1.48* ** 2,2,2- trifluoroethane HFC-134a 1,2,2,2- ** -26.2 1.206 99.8483 tetrafluoro- ethane *@ 21.13C ; **Data unavailable from manufacturers. Trichloroethylene Trichloroethylene (T CE) has been widely used in industry for many years as a popular dry cleaning solvent, an excellent degreasing agent, an extraction agent in decaffeinating coffee, and in several other ways. Because of improper handling, inadequate disposal techniques, or accidental spillage, it is commonly found in soil and groundwater near industrial sites (Barbash and Roberts 1986; Verschueren 1983). The presence of TCE and other low- molecular-weight chlorinated aliphatic hydrocarbons in groundwater threatens drinking water supplies (Roberts et al. 1982) and endangers human health because of the toxicity and suspected or demonstrated carcinogenicity of these chemicals (Miller and Guengerich 1983). In 1976, TCE was included on the EPA list of hazardous substances. It has become the subject of extensive governmental regulation. Moreover, TCE is partially degraded anaerobically to vinyl chloride, which is more toxic than TCE and is a known carcinogen (Parsons et al. 1984; Vogel and McCarty 1985). Therefore, TCE is the most frequently reported contaminant at hazardous waste sites on the National Priority List of the US. Environmental Protection Agency . 32 Trichloroethylene (C12C=CHC1) is a synthetic, chlorinated organic chemical that fulfills all requirements for the degreasing solvent. It has high solvency for oils, greases, waxes, tars, resins, lubricants, and coolants generally found in the metal-processing industry. TCE is only slightly soluble in water (about 1100 ppm at 77°F) and forms an azeotrope with water, resulting in a mixture with a lower boiling point and vapor density. It is considered to be a highly volatile compound and favors environmental partitioning to the air rather than water. TCE is destroyed by photooxidation in the atmosphere, with a half-life of about one day. Table 2.12 General information and properties of trichloroethylene1 :roperty Value ChemicaT Abstracts Service (CAS) number 79-01-6 Chemical formula (:2H(313 Molecular weight 131.40 Physical state Colorless liquid Boiling point 86.7°C Melting point -73°C Density 1.4 g/mL at 25°C Vapor pressure 77 mm Hg at 25°C Water solubility 1 g/L at 20°C Henry's constant (dimensionless)2 0.392 at 25°C Log octanol/ water partition coefficient 2.29 Odor threshold 0.5 mg/L in water; 2.5-900 mg/m3 in air Air concentration conversion factor 5.46 mg / m3=1 ppm References: (1)1Ware 1988; (2) Gossett 1987. Methanotrophic transformation of trichloroethylene In 1985, Wilson and co-workers reported on the possibility of aerobic oxidation of TCE by soil microorganisms that were provided natural gas as a primary source of energy (Wilson and Wilson 1985). Since them, the ability of methane-utilizing bacteria to cometabolize TCE has been reported and confirmed by several researchers (Fliermans et al. 1988; Fogel et a]. 1986; Little et a1. 1988). It is generally believed that the enzyme methane monooxygenase (MMO) oxidizes TCE to epoxides, which spontaneously hydrolyzes to 33 glyoxylate and dichloroacetate under acidic conditions, and carbon monoxide and formate under basic conditions (Henry and Grbié -Galié 1986; Henschler et al. 1979; Little et a1. 1988; Miller and Guengerich 1982). Glyoxylate and dichloroacetate are oxidized to carbon dioxide by heterotrophic bacteria. Formate and carbon monoxide are oxidized to carbon dioxide by methanotrophs. Later, researchers reported that besides products resulting from epoxide hydrolysis, intramolecular halide or hydride migration can occur yielding 2,2,2-trichloroacetaldehyde (chloral hydrate) (Fox et al. 1990). Chloral hydrate can be reduced further to trichloroethanol or oxidized to trichloroacetic acid (Newman and Wackett 1991). Chloral is toxic and may be responsible for the product toxity observed during TCE transformation by methanotrophs. Trichloroacetate degraded slowly in one methanotrophic mixed culture (Uchiyama et al. 1989). Formation of significant levels of Trichloroacetate and epoxide degradation products by Methylocystis sp. strain M indicates that chlorine migration and epoxide formation proceed in parallel (Nakajima et a1. 1992). Other research also shows that these pathways proceed in parallel in Methylocystis trichosporium OB3B (Fox et al. 1990; Newman and Wackett 1991; Oldenhuis et al. 1989). Taken together, these reports suggest that both pathways occur simultaneously in type II methanotrophs. On the other hand, reports published to date suggest that type I methanotrophs transform exclusively by the epoxide pathway (Henry and Grbic’: -Galié 1986; Little et al. 1988). Methanotrophs and the products produced via transformation pathways are summarized in Table 2.13. 34 Table 2.13 TCE transformation and product formation by methanotrophs ehaotrop V _ I Tranmsforatron pth Product reported Referne l 1 Type 1. Methylomonas _ FA“, GA‘, DCAA‘ Henry & Grbié -Galié l . sp. strain MM2 1986 l } Type 1, strain 46-] 1# GA, DCAA Little et al. 1988 l l Type II, Methylosinus l&2# Chloral, TCetOH* Oldenhuis et al. 1989 l ‘ trichosporium OB3b l . | Soluble MMO from 1&2 FA, GA, CO, DCAA, Fox et al. 1990 l i Methylosinus Chloral l ; trichosporium OB3b l l Methylosinus 1&2 FA, GA, CO, DCAA, Newman and Wackett l ; trichosporium OB3b Chloral, TCAA“, TCetOH 1991 l l ; -Methylococcus _ Chloral Newman and Wackett ; 1 capsulatus, 1991 i l -Methylosinus, sporium, l -Methylosporovibrio I . methanica 812 I . l I Type II, Methylocystis 1&2 FA, GA. DCAA, Chloral, Nakajima er al. 1992 l 2 sp. strain M 1CAA,'ICetOH 1 i l i Type II, Methylocystis 1&2 'ICAA Uchiyama et al. 1992 1 sp. strain M + l Xanthobacter l autotrophicus DA4 l l l 1 *FA: formic acid GA: glyoxylic acid DCAA: dichloroacetic acid TCetOH: 2,2,2-trichloroethanol TCAA: trichloroacetic acid #Path 1: epoxidation Path 2: chloride migration Recent studies indicate that TCE transformation capacity is not only a function of the availability of reducing power, but also of the specific cometabolized compound and the toxicity of its transformation products (Alvarez-Cohen and McCarty 1991a; Henry and Grbic’: -Galié 1991a; Wackett and Householder 1989). Formate addition resulted in increased initial specific TCE transformation rates and elevated transformation capacity. Significant declines in methane conversion rates following exposure to TCE were observed for both resting and formate-fed cells, suggesting toxic effects caused by TCE or its transfomation products (Alvarez-Cohen and McCarty 1991a; Alvarez-Cohen and McCarty 35 1991b). Not many researchers have reported on the toxicity of the specific products. Oldenhuis and co-workers suggested that TCE epoxide can be expected to bind covalently to proteins and nucleic acids. Other possible reactive metabolites that might bind irreversibly are chloral, dichloroacetyl chloride, and formyl chloride (Oldenhuis et al. 1991). Organisms or communities capable of degrading a large amounts of TCE should possess detoxification systems or populations that degrade these compounds. The presence of toxic transformation products can be expected to have some impacts on the development of microbial communities during long-term TCE exposure. Changes in the populations are likely related to the level of TCE exposure, turnover of transformation products and utilization of growth substrate. Lackey et al. (1994) used total-recycle expanded-bed bioreactors to evaluate the degradation potential of TCE by a microbial consortium. Ester-linked phospholipid fatty acid profiles (PLFAME) were used to monitor the change of TCE-affected community during short-term perturbation. The results showed that a propane-utilizing bacterial biomaker increased as TCE was degraded and propane consumed. However, the relationship between community structure and extent of TCE exposure was not clear for these short-term exposures. MODEL REACTOR SYSTEMS FOR COMETABOLISM In order to study the development of microbial communities, a culture or community of microorganisms must be grown under defined conditions. Reactor configurations that have been evaluated include: completely-stirred tank reactor (Coyle et al. 1993; Landa et al. 1994), fed-batch reactor (Strand et al. 1990), fixed-bed reactors (Strand et a1. 1991; Strandberg et al. 1989), expanded-bed reactor(Lackey et al. 1994; Phelps et al. 1990) and multi-stage systems(Alvarez—Cohen and McCarty 199ld; Folsom and Chapman 1991). This work focuses on dispersed growth systems. 36 Three basic modes of dispersed growth culture are widely used. Simple batch systems are used to study substrate-sufficient growth with maximum specific growth rate. Continuous culture in chemostats permits full control over specific growth rate with a given environment or conversely, the environment may be varied with the specific growth rate held constant. A unique feature of chemostats is that a time-independent steady state can be attained which enables one to determine the relationship between microbial behavior (genetic and phenotypic expression) and the environmental conditions. The last basic mode is fed-batch systems. In such systems, the culture is provided with a substrate feed which permits substrate-limited growth with a decreasing specific growth rate. The three basic reactor systems described in last section affect the process parameters differently. Characteristically, batch systems show the four phases of growth (lag, logarithmic, stationary, and decline). In a chemostat culture at steady state, all the environmental factors are constant. In a variable volume fed-batch system in the "quasi steady state" all the environmental factors are virtually constant except the growth-limiting substrate; in constant volume fed-batch system with substrate-limiting growth, all nutrient concentrations vary throughout the cycle. Several model systems have been considered for the evaluation of cometabolism of TCE. The cases here are much different from the traditional growth study involving only growth substrates. In most cases, both growth and nongrowth substrates were injected into model systems. Thus, the interactions between both substrates are important. The questions discussed in these research can be divided into three categories: verification of kinetic models, evaluation of reactor operations and monitoring of development of microbial communitines. 37 The presence of toxic transformation products can be expected to have some impacts on the microbial community structure during long-term TCE exposure. Changes in the populations are likely related to the level of TCE exposure, turnover of transformation product and utilization of growth substrate. These factors will have different effects in simple mixed cultures compared to complex communities and in batch reactors compared to continuous reactors. In batch reactors, for example, the microbial community is exposed to a range of growth and nongrowth substrate concentrations. This may select for a more diverse community with "specialist" organisms that occupy a variety of niches created by substrate concentration gradients. In contrast, a chemostat favors selection of specific populations at a fixed specific growth rate. This can be expected to result in a less diverse culture. Since nongrowth substrates are not mineralized by cometabolizing species and since heterotrophs are known to play important roles in detoxification, more diverse cultures should have advantages for cometabolism. As discussed previously, batch system and chemostat represent two extremes for cometabolizing nongrowth substrate. One purpose of this work is to assess the effect of TCE exposure on microbial communities. Therefore, both systems were chosen as model systems for this study. Compared to conventional batch reactors, sequencing batch reactors (SBRs) have several advantages for cometabolic transformations. An SBR can alternate between periods of growth on growth substrates and periods of cometabolism of nongrowth substrate, eliminating the possibility of competitive inhibition for enzymes between growth and nongrowth substrates. SBRs also have a great ability to periodically change environmental condition, selecting or enriching specific microbial populations. Thus, SBRs offer a better model environment to study the dynamic changes of microbial communities than conventional batch system. 38 REFERENCES 1.A1varez-Cohen, L., and P. L. McCarty. 1991a. A cometabolic biotransformation model for halogenated aliphatic compounds exhibiting product toxicity. Environ. Sci. Techno]. 25:1381-1387. 2.Alvarez-Cohen, L., and P. L. McCarty. 1991b. Effects of toxicity, aeration, and reductant supply on trichloroethylene transformation by a mixed methanotrophic culture. App]. Environ. Microbiol. 57:228 - 235. 3.Alvarez-Cohen, L., and P. L. McCarty. 1991c. Product toxicity and cometabolic competitive inhibition modeling of chloroform and trichloroethylene transformation by methanotrophic resting cells. Appl. Environ. Microbiol. 57:1031 - 1037. 4.Alvarez-Cohen, L., and P. L. McCarty. 1991d. Two-stage dispersed-growth treatment of halogenated aliphatic compounds by cometabolism. Environ. Sci. Technol. 25: 1387- 1393. 5.Anderson, J. E., and P. L. McCarty. 1994. Model for treatment of trichloroethylene by methanotrophic biofilms. J. Environ. Eng. 120:379—400. 6.Anselmo, A. M., M. Mateus, J. M. S. Cabral, and J. M. Novais. 1985. Degradation of phenol by immobilized cells of F usarium flocciferum. Biotechnol. Letters. 7:889-894. 7.Arciero, D., T. Vannelli, M. Logan, and A. B. Hooper. 1989. Degradation of trichloroethylene by the ammonia-oxidizing bacterium Nitrosomonas europaea. Biochem. Biophys. Res. Commun. 159:640-643. 8.Auteinrieth, R. L., J. S. Bonner, A. Akgerman, and M. Okagun. 1991. Biodegradation of phenolic wastes. Journal of Hazardous Materials. 28:29-53. 9.Bailey, J. E., and D. F. Ollis. 1986. Biochemical engineering fundamentals, 2nd ed. McGraw-Hill, New York. 10.Beltrame, P., P. L. Beltrame, P. Camiti, and D. Pitea. 1980. Kinetics of phenol degradation by activated sludge in a continuous-stirred reactors. J. Water Pollut. Control Fed. 52:126-133. 11.Broholm, K., T. H. Christensen, and J. B. K. 1992. Modelling TCE degradation by a mixed culture of methane-oxidizing bacteria. Wat. Res. 26: 1 177-1 185. 12.Brusseau, G. A., H.-c. Tsien, R. S. Hanson, and L. P. Wackett. 1990. Optimization of trichloroethylene oxidation by methanotrophs and the use of a colorimetric assay to detect soluble methane monooxygenase activity. Biodegradation. 1:19-29. 13.Buswell, J. A. 1975. Metabolism of phenol and cresols by Bacillus stearothermophilus. J. Bacteriol. 124: 1077-1083. 14.Chang, H.-l., and L. Alvarez-Cohen. 1995a. Model for the cometabolic biodegradation of chlorinated organics. Environ. Sci. Technol. 29:2357-2367. 39 15.Chang, H.-L., and L. Alvarez-Cohen. 1995b. Transformation capacities of chlorinated organics by mixed cultures enriched on methane, propane, toluene, or phenol. Biotechnol. Bioeng. 45:440-449. 16.Chang, M.-k., T. C. Voice, and C. S. Criddle. 1993. Kinetics of competitive inhibition and cometabolism in the biodegradation of benzene, toluene, and p-xylene by two Pseudomonas isolates. Biotechnol. Bioeng. 41: 1057-1065. l7.Co]by, J ., D. I. Stirling, and H. Dalton. 1977. The soluble methane mono-oxygenase of Methylococcus capsulatus (bath). Its ability to oxygenate n-alkanes, n-alkenes, ethers, and alicylic, aromatic and heterocyclic compounds. Biochem. J. 1652395-402. 18.Coyle, C. G., G. F. Parkin, and D. T. Gibson. 1993. Aerobic, phenol-induced TCE degradation in completely mixed, continuous-culture reactors. Biodegradation. 4:59-69. l9.Criddle, C. S. 1993. The kinetics of cometabolism. Biotechnol. Bioeng. 41:1048- 1056. 20.Criddle, C. S., J. T. DeWitt, and P. L. McCarty. 1990. Reductive dehalogenation of carbon tetrachloride by Escherichia coli k-12. Appl. Environ. Microbiol. 56:3247-3254. 21.D'Adamo, P. D., A. F. Rozich, and A. F. J. Gaudy. 1984. Analysis of growth data with inhibitory substrate. Biotechnol. Bioeng. 26:397. 22.Dalton, H., and D. I. Stirling. 1982. Co-metabolism. Phil. Trans. R. Soc. Lond. B. 297 :48 1 -496. 23.DeFlaun, M. F., B. D. Ensley, and R. J. Steffan. 1992. Biological oxidation of hydrochlorofluorocarbons(HCFCs) by a methanotrophic bacterium. Biotechnology. 10:1576-1578. 24.Denovan, B. A., and S. E. Strand. 1992. Biological degradation of Chlorofluorocarbons in anaerobic environments. Chemosphere. 24:935-940. 25.Ensign, S. A., M. R. Hyman, and D. J. Arp. 1992. Cometabolic degradation of chlorinated alkenes by alkene monooxygenase in a propylene-grown Xanthobacter strain. Appl. Environ. Microbiol. 58:3038-3046. 26.F1iermans, C. B., T. J. Phelps, D. Ringelberg, A. T. Mikel], and D. C. White. 1988. Mineralization of trichloroethylene by heterotrophic enrichment cultures. Appl. Environ. Microbiol. 54: 1709-1714. 27.Fogel, M. M., A. R. Taddeo, and S. Fogel. 1986. Biodegradation of chlorinated ethene by a methane-utilizing mixed culture. Appl. Environ. Microbiol. 51:720-724. 28.Fo]som, B. R., and P. J. Chapman. 1991. Performance characterization of a mode] bioreactor for the biodegradation of trichloroethylene by Pseudomonas cepacia G4. Appl. Environ. Microbiol. 57: 1602- 1608. 29.Fo]som, B. R., P. J. Chapman, and P. H. Pritchard. 1990. Phenol and trichloroethylene degradation by Pseudomonas cepacia G4: kinetics and interactions between substrates. App]. Environ. Microbiol. 56: 1279-1285. 40 30.Fox, B. G., J. G. Bomeman, L. P. Wackett, and J. D. Lipscomb. 1990. Haloalkene oxidation by the soluble methane monooxygenase from Methylosinus trichosporium OB 3b: mechanistic and environmental implication. Biochemistry. 29:6419-6427. 31.Gaal, A., and H. Y. Neujahr. 1979. Metabolism of phenol and resorcino] in Trichosporon cutaneum. J. Bacteriol. 137213-21. 32.Galli, R., and P. L. McCarty. 1989. Biotransformation of 1,1,1-trichloroethane, trichloromethane, and tetrachloromethane by a Clostridium sp. Appl. Environ. Microbiol. 55:837-844. 33.Gossett, J. M. 1987. Measurement of Henry's law constants for C1 and C2 chlorinated hydrocarbons. Environ. Sci. Technol. 21:202-208. 34.Gottscha]k, G. 1986. Bacterial Metabolism, 2nd ed. Springer-Verlag, New York. 35.Green, P. N. 1992. Taxonomy of methylotrophic bacteria, p. 23-84. In J. C. Murrell and H. Dalton (ed.), Methane and methanol utilizers. Plenum Press, New York. 36.Henry, S. M., and D. Grbié -Ga]ié. 1986. Aerobic degradation of trichloroethylene (TCE) by methylotrophs isolated from a contaminated aquifer. Presented at the Abstr. 86th Annu. Meet. Am. Soc. Microbiol. abstr. Q-64z294. 37.Henry, S. M., and D. Grbié -Galié. 1990. Effect of mineral media on trichloroethylene oxidation by aquifer methanotrophs. Microbial Ecology. 20:151-169. 38.Henry, S. M., and D. Grbié -Galié. 199la. Influence of endogenous and exogenous electron donors and trichloroethylene oxidation toxicity on trichloroethylene oxidation by methanotrophic cultures from a groudwater aquifer. Appl. Environ. Microbiol. 57:236 - 244. 39.Henschler, D., W. R. Hoos, H. Fetz, E. Dallmeier, and M. Metzler. 1979. Reactions of trichloroethylene epoxide in aqueous systems. Biochem. Pharmacol. 28:543-548. 40.Herbet, D., R. E. Elsworth, and R. C. Telling. 1956. The continuous culture of bacteria: a theoretical and experimental study. J. Gen. Microbiol. 14:601-607. 41.Hopkins, G. D., J. Munakata, L. Semprini, and P. L. McCarty. 1993b. Trichloroethylene concentration effects on pilot field-scale in-situ groundwater bioremediation by phenol-oxidizing microorganisms. Environ. Sci. Technol. 27:2542- 2547. 42.Hopkins, G. D., L. Semprini, and P. L. McCarty. 1993a. Microcosm and in-situ field studies of enhanced biotransformation of trichloroethylene by phenol-utilizing microorganisms. Appl. Environ. microbiol. 59:2277-2285. 43.Horvath, R. S. 1971. Cometabolism of the herbicide, 2,3,6-trichlorobenzoate. J. Agr. Food Chem. 19:291-293. 44.Horvath, R. S. 1972. Microbial co-metabolism and the degradation of organic compounds in nature. Bacteriol. Rev. 36: 146-155. 41 45.Hyman, M. R., I. B. Murton, and D. J. Arp. 1988. Interaction of ammonia monooxygenase from Nitrosomonas europea with alkanes, alkenes, and alkynes. Appl. Environ. Microbiol. 54:3187 - 3190. 46.Kotturi, G., C. W. Robinson, and W. E. Inniss. 1991. Phenol degradation by a psychrotrophic strain of Pseudomonas putida. Appl. Microbiol. Biotechnol. 34:539-543. 47.Lackey, L. W., T. J. Phelps, V. Korde, S. Nold, D. Ringelberg, P. R. Bienkowski, and D. C. White. 1994. Feasibility testing for the on-site bioremediation of organic wastes by native microbial consortia. International Biodeterioration & Biodegradation. 33:41-59. 48.Landa, A. S., E. M. Sipkema, J. Weijma, A. A. C. M. Beenackers, J. Dolfing, and D. B. Janssen. 1994. Cometabolic degradation of trichloroethylene by pseudomonas cepacia G4 in a chemostat with toluene as the primary substrate. Appl. Environ. Microbiol. 60:3368-3374. 49.Lesage, S., S. Brown, and K. R. Hosler. 1992. Degradation of Chlorofluorocarbon- 113 under anaerobic conditions. Chemosphere. 24: 1225-1243. 50.Little, C. D., A. V. Palumbo, S. E. Herbes, M. E. Lidstrom, R. L. Tyndall, and P. J. Gilmer. 1988. Trichloroethylene biodegradation by a methane-oxidizing bacterium. Appl. Environ. Microbiol. 542951 - 956. 51.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. 52.Miller, R. E., and F. P. Guengerich. 1983. Metabolism of trichloroethylene in isolated hepatocytes, microsomers, and reconstituted enzyme systems containing cytochrome p- 450. Cancer Res. 43:1145-1152. 53.Miller, R. E., and F. P. Guengerich. 1982. Oxidation of trichloroethylene by liver microsomal cytochrome p-450: evidence for chlorine migration in a transition state not involving trichloroethylene oxide. Biochemistry. 2 1 : 1090- 1097. 54.Molina, M. J ., and F. S. Rowland. 1974. Stratospheric sink for chlorofluoromethanes: chlorine atomcatalysed destruction of ozone. Nature. 249:810-812. 55.Nakajima, T., H. Uchiyama, O. Yagi, and T. Nakahara. 1992. Novel metabolite of trichloroethylene in a methanotrophic bacterium, Methylocystis sp. M, and hypothetical degradation pathway. Bioscience, Biotechnology, and Biochemistry. 56:486-489. 56.Nelson, M. J. K., S. O. Montgomery, W. R. Mahaffey, and P. H. Pritchard. 1987. Biodegradation of trichloroethylene and involvement of an aromatic biodegradative pathway. Appl. Environ. Microbiol. 53:949—954. 57.Nelson, M. J. K., S. O. Montgomery, E. J. O'Neill, and P. H. Pritchard. 1986. Aerobic metabolism of trichloroethylene by a bacteria] isolate. Appl. Environ. Microbiol. 52:383-384. 58.Nelson, M. J. K., S. O. Montgomery, and P. H. Pritchard. 1988. Trichloroethylene metabolism by microorganisms that degrade aromatic compounds. Appl. Environ. Microbiol. 54:604-606. 42 59.Newman, L. M., and L. P. Wackett. 1991. Fate of 2,2,2-trichloroacetaldehyde (chloral hydrate) produced during trichloroethylene oxidation by methanotrophs. App]. Environ. Microbiol. 57:2399 - 2402. 60.0ldenhuis, R., J. Y. Oedzes, J. J. van der Waarde, and D. B. Janssen. 1991. Kinetics of chlorinated hydrocarbon degradation by Methylosinus trichosporium OB3b and toxicity of trichloroethylene. Appl. Environ. Microbiol. 57:7-14. 61.0ldenhuis, R., R. L. J. M. Vink, D. B. Janssen, and B. Witholt. 1989. Degradation of chlorinated aliphatic hydrocarbons by methylosinus trichosporium OB3b Expressing soluble methane monooxygenate. Appl. Environ. Microbiol. 55:2819-2826. 62.Parsons, F., P. R. Wood, and J. DeMarco. 1984. Transformation of tetrachloroethene and trichloroethene in microcosms and groundwater. J. Am. Water Works Assoc. 76:56- 59. 63.Pawlowsky, U., and J. A. Howell. 1973. Mixed culture biooxidation of phenol. 1. Determination of kinetic parameters. Biotechnol. Bioeng. 15:889-896. 64.Phelps, T. J., J. J. Niedzielski, R. M. Schram, S. E. Herbes, and D. C. White. 1990. Biodegradation of trichloroethylene in continuous-recycle expanded-bed bioreactors. Appl. Environ. Microbiol. 56: 1702 - 1709. 65.Rasche, M. E., M. R. Hyman, and D. J. Arp. 1990. Biodegradation of halogenated hydrocarbon fumigants by nitrifying bacteria. Appl. Environ. Microbiol. 56:2568 - 2571. 66.Rasche, M. E., M. R. Hyman, and D. J. Arp. 1991. Factors limiting aliphatic chlorocarbon degradation by Nitrosomonas europaea: cometabolic inactivation of ammonia monooxygenase and substrate specificity. Appl. Environ. Microbiol. 57:2986-2994. 67.Rizzuti, L., and V. Augueliaro. 1982. Kinetic parameter estimation in monoculture and monosubstrate biological reactors. Can. J. Chem. ENG. 60:608. 68.Roberts, P. V., J. E. Schreinger, and G. C. Hopkins. 1982. Field-study of organic- water quality changes during groundwater recharge in the Palo-Alto Baylands. Wat. Res. 16:1025-1035. 69.Saéz, P. B., and B. E. Rittmann. 199]. Biodegradation kinetics of 4-chlorophenol, an inhibitory co-metabolite. Research J. WPCF. 63:838-847. 70.Saéz, P. B., and B. E. Rittmann. 1993. Biodegradation kinetics of a mixture containing a primary substrate (phenol) and an inhibitory co-metabolite (4-chlorophenol). Biodegradation. 4:3-21 . 71.Semprini, L., D. Grbi’c-Gali’c, P. L. McCarty, and P. V. Roberts. 1992. Methodologies for evaluating in-situ bioremediation of chlorinated solvents EPA/600/R- 92/042. U. S. Environmental Protection Agency. 72.Shih, C.-C., M. Davey, J. Zhou, J. M. Tiedje, and C. S. Criddle. 1996. Effects of phenol feeding pattern on microbial community structure and cometabolism of trichloroethylene. Appl. Environ. Microbiol. 62:2953-2960. 73.SoKol, W. 1988. Uptake rate of phenol by Pseudomonas putida grown in unsteady State. Biotechnol. Bioeng. 32: 1097-1 103. 43 74.Strand, S. E., M. D. Bjelland, and H. D. Stensel. 1990. Kinetics of chlorinated hydrocarbon degradation by suspended cultures of methane-oxidizing bacteria. Research J. WPCF. 62:124-129. 75.Strand, S. E., J. V. Wodrich, and H. D. Stensel. 1991. Biodegradation of chlorinated solvents in a sparged, methanotrophic biofilm reactor. Research J. WPCF. 63:859-867. 76.Strandberg, G. W., T. L. Donaldson, and L. L. Farr. 1989. Degradation of trichloroethylene and trans-1,2-dichloroethylene by a methanotrophic consortium in a fixed-film, packed-bed bioreactor. Environ. Sci. Technol. 23. 77.Szetela, R. W., and T. Z. Winnicki. 1981. A novel method for determining the parameters of microbial kinetics. Biotechnol. Bioeng. 23: 1485-1490. 78.Tischler, L. F., and W. W. J. Eckenfelder. 1969. Linear substrate removal in the activated sludge process. Presented at the Advances in Water Pollution Research, Oxford, England. 22361-374. 79.Tsien, H.-c., G. A. Brusseau, R. S. Hanson, and L. P. Wackett. 1989. Biodegradation of trichloroethylene by Methylosinus trichosporium OB3b. Appl. Environ. Microbiol. 55:3155-3161. 80.Uchiyama, H., T. Nakajima, O. Yagi, and T. Nakahara. 1992. Role of heterotrophic bacteria in complete mineralization of trichloroethylene by Methylocystis sp. strain M. Appl. Environ. Microbiol. 58:3067-3071. 81.Uchiyama, H., T. Nakajima, O. Yagi, and T. Tabuchi. 1989. Aerobic degradation of trichloroethylene by a new type II methane-utilizing bacterium, strain M. Agric. Biol. Chem. 53:2903 - 2907. 82.Vannel]i, T., M. Logan, D. A. Arciero, and A. B. Hooper. 1990. Degradation of halogenated aliphatic compounds by the ammonia-oxidizing bacterium Nitrosomonas europaea. Appl. Environ. Microbiol. 56:] 169-1 171. 83.Voge], T. M., and P. L. McCarty. 1985. Biotransformation of tetrachloroethylene to trichloroethylene, dichloroethylene, vinyl chloride, and carbon dioxide under methanogenic conditions. Appl. Environ. Microbiol. 49: 1080-1083. 84.Wackett, L. P., G. A. Brusseau, S. R. Householder, and R. S. Hanson. 1989. Survey of microbial oxygenases: trichloroethylene degradation by propane-oxidizing bacteria. Appl. Environ. Microbiol. 55:2960 - 2964. 85.Wackett, L. P., and D. T. Gibson. 1988. Degradation of trichloroethylene by toluene dioxygenase in whole-cell studies with Pseudomonas putida F1. Appl. Environ. Microbiol. 54:1703-1708. 86.Wackett, L. P., and S. R. Householder. 1989. Toxicity of trichloroethylene to Pseudomonas putida F1 is mediated by toluene dioxygenase. Appl. Environ. Microbiol. 55:2723-2725. 87.Ware, G. W. (ed.). 1988. Reviews of environmental contamination and toxicology, vol. 106. Springer-Verlag, Inc., New York. 44 88.Wi]son, J. T., and B. H. Wilson. 1985. Biotransformation of trichloroethylene in soil. Appl. Environ. Microbiol. 49:242-243. 89.Winter, R. B., K.-M. Yen, and B. D. Ensley. 1989. Efficient degradation of trichloroethylene by a recombinant Escherichia coli. Bio/Technology. 7:282-285. 90.Yang, R. D., and A. E. Humphrey. 1975. Dynamic and steady state studies of phenol biodegradation in pure and mixed culture. Biotechnol. Bioeng.: 1599-1615. 91.Zylstra, G. J ., L. P.Wackett, and D. T. Gibson. 1989. Trichloroethylene degradation by Escherichia coli containing the cloned Pseudomonas putida F1 toluene dioxygenase genes. Appl. Environ. Microbiol. 55:3162-3166. CHAPTER 3 KINETICS OF COMETABOLISM BY RESTING CELLS" ABSTRACT This research investigated the potential for methanotrophic biotransformation of three HCFCs -- chlorodifluoromethane (HCFC-22); 1-chloro-1,1-difiuoroethane (HCFC-142b); and 1,1-dichloro-2,2,2-trif]uoroethane (HCFC-123); and one HFC -- 1,2,2,2- tetrafluoroethane (HFC-134a). All of these compounds were biotransformed to differing degrees by methanotrophic mixed culture MM 1. Rates of transformation were obtained by monitoring disappearance of the target compounds from the headspace in batch experiments. Henry's constants were determined over a range of conditions to enable estimation of the intrinsic rates of transformation. Intrinsic rates of transformation were obtained by combining a second order rate expression with an expression describing loss of transformation activity due to either endogenous decay or product toxicity. For HCFC- 123 and HFC-134a, the independently measured endogenous decay rate for mixed culture MM] (0.594/day) was sufficient to account for the observed loss of transformation activity with time. However, the endogenous decay rate did not account for the loss of transformation activity for HCFC-22 and HCFC-l42b. A model based on product toxicity provided a reasonable representation of the loss of transformation activity for these compounds. The order of reactivity was HCFC-22 > HCFC-l42b > HFC-134a > HCFC- 123, with second order rate coefficients of 0.014, 0.0096, 0.00091, and 0.00054 L/mg- day, respectively. Transformation capacities for HCFC-22 and HCFC-142b were 2.47 * This study was published in W, 6: 1-9, 1995. 45 46 and 1.1] ug substrate/mg biomass, respectively. INTRODUCTION Cometabolic transfomation of nongrowth substrates under resting cell conditions is a specific case for cometabolism. Under this condition, reducing power or energy reserve is consumed for nongrowth substrate transformation with little or no energy return to microorganisms in the community. Several researchers have chosen resting cell conditions to study cometabolic transformation because interaction between growth and nongrowth substrates is not present. Kinetic models for cometabolic transformation by resting cells have been developed by several researchers (Alvarez-Cohen and McCarty 1991a; Criddle et a]. 1990; Galli and McCarty 1989; Saéz and Rittmann 1991; Schmidt et al. 1985). Criddle (1993) proposed a mode] that unified the earlier models. This model assumed that nongrowth substrate is degraded with saturation kinetics and endogenous decay and that loss of cometabolic activity due to product toxicity can be incorporated into the cell decay term. This chapter focuses on the methanotrophic biotransformation of three HCFCs (HCFC-22: chlorodifiuoromethane, HCFC-l42b: 1-chloro-1,l-difiuoroethane, and HCFC-123: 1,1- dichloro-2,2,2-trifluoroethane) and one HFC (HFC-134a: 1,2,2,2-tetrafluoroethane) in a defined methanotrophic mixed culture. The unified model was used to predict the cometabolic transformation of HCFCs/HFC. The headspace method was used to monitor disappearance of target compounds. To obtain intrinsic kinetic data using this method, Henry's constants are needed. Because these constants were not available for the compounds studied, Henry's constants were first determined. Disappearance of the target compounds was then monitored in methanotrophic mixed cultures. Henry's constants were measured over a range of ionic strengths to enable use of these measurements in environments beyond those of the present study, such as seawater. Knowledge of 47 methanotrophic transformations should assist in selecting environmentally acceptable HCFCs and HFCs, modeling environmental fate, developing treatment technologies for fugitive manufacturing emissions, and remediating future wastewater and groundwater contamination. Rationale for experimental work Chlorofluorocarbons (CFCs) are widely used refrigerants and aerosols in industry and domestic life. Over the past decade, they have been implicated as agents of depletion of stratospheric ozone and as contributors to global warming (Molina and Rowland 1974; Rowland and Molina 1975). As a result, worldwide production of CFCs will be banned under the terms of Montreal Protocol. Nevertheless, CFCs will continue to be released into the environment due to past production and continued use. In aerobic aquatic environments, CFCs are recalcitrant, but they are transformed anaerobically (Denovan and Strand 1992; Lesage et al. 1992; Lesage et al. 1990; Lovely and Woodward 1992; Semprini et al. 1992). The ban on CFCs has inspired a major research effort to assess two classes of CFC substitutes - the hydrochlorofluorocarbons (HCFCs) and the hydrofluorocarbons (HFCs). HCFCs and HFCs are one- and two-carbon aliphatics, similar in structure and physical properties to the CFCs, but containing one or more hydrogen atoms. The presence of hydrogen makes HCFCs and HFCs more susceptible to tropospheric oxidation than the CFCs, and thus less likely to migrate into the stratosphere. To date, there is relatively little information on the fate of HCFCs or HFCs in aquatic environments. Lesage et al. (1992) reported transformation of HCFC-123a to HCFC-133 and HCFC-133b under methanogenic conditions. DeFlaun et a]. (1992) reported aerobic transformation of three HCFCs and one HFC by Methylasinus trichosporium OB3b. 48 MATERIALS AND METHODS Chemicals HCFC-22, HCFC-142b and HFC-134a were obtained from Asahi Glass Co., LTD. (Yokohama, Japan). HCFC-l23 was obtained from Allied-Signal, Inc. (Morristown, NJ , USA). All chemicals used in media preparation were ACS grade, and all water used was 18 megaohm resistance or greater. Analytical techniques The study compounds were analyzed by withdrawing 0.5 mL of headspace from the test bottles using a Precision gas-tight syringe and injecting the sample onto a Perkin Elmer 8500 Gas Chromatograph (GC) equipped with a squalene packed column and a flame ionization detector. The GC was operated isothermally at 90°C with helium as carrier. Concentrations were obtained from an external standard calibration curve bracketing the concentration range of interest. Measurement of Henry's constants The modified EPICS procedure (Gossett 1987) was used to determine Henry's constants for each of the target compounds. Pure compounds were dissolved in methanol as stock solutions. To examine possible cosolvent interferences, five serum bottles with same amounts of HCFC-l34 but with different methanol content (0 to 5%) were prepared. The result showed no significant cosolvent effect for methanol levels below 2%. Therefore, all subsequent measurements were conducted under this condition. For each compound, Henry’s constant was measured in six 158.8 ml serum bottles: three containing 100 milliliters of distilled water, and three containing 25 milliliters. Both sets of bottles were sealed with Teflon/rubber septa and aluminum crimp caps. HCFCs/HFC solutions were injected into each bottle using a 0.5 mL gas-tight syringe. The bottles were then incubated 49 in an inverted position for 24 hrs at the desired temperature (6, 12, 22, 30 and 40°C, all i0.2°C) on a temperature-controlled shaker, and headspace samples were analyzed by gas chromatography. To assess the effects of ionic strength on the Henry's constant, six serum bottles were filled with 100 mL solution, each with different concentration of KC] (0, 0.2, 0.4, 0.6, 0.8, 1.0 M). These bottles were then analyzed by headspace gas chromatography. Culture conditions Mixed culture MM] , a methanotrophic enrichment obtained from aquifer material at Moffett Field, California, was used for these experiments (Henry and Grbié -Galié 1991a). This culture is a stable consortium consisting of one methanotroph and three or four heterotrophs containing predominantly Grain-negative pleomorphic coccobacilli and prosthecates as well as some Grain-negative bacilli and cocci. The methanotroph in the mixed culture expresses soluble MMO similar to that of Methylosinus trichosporium OB3b under similar growth conditions (Henry and Grbié -Ga]ié 1991a). Mixed culture MM] was grown in Whittenbury Mineral Medium containing (per liter of deionized water): 1.0 g of MgSO4-7H20, 1.0 g of KNO3, 200 mg of CaC12-2H20, 3.8 mg of FeEDTA, 0.5 mg of NazMoO4-2H20, 0.5 mg of FeSO4-7HzO, 0.4 mg of ZnSO4-7HzO, 0.02 mg of MnC12-4H20 0.05 mg of CoC12-6H20 0.01 mg of NiC12-6H20, 0.015 mg of H3BO3, 0.25 mg of EDTA, 260 mg of KH2PO4, and 330 mg of NazHPO4. One liter of culture was grown at room temperature (~21°C) in a continuously stirred 4-liter bottle supplied 30% methane in air at 68 mL/min. Growth curves were monitored and as stationary phase approached, approximately 10 mL of culture was transferred to a 1 liter of fresh Whittenbury Medium. Cells were harvested in mid-log growth phase for biotransformation experiments. 50 Batch biotransformation experiments HCFCs/HFC degradation studies were performed using 158.8 mL serum bottles sealed with Telfon/rubber septa and aluminum crimp caps. These bottles were incubated with 100 mL of of Whittenbury Mineral Media plus culture. An appropriate amount (measured as dry weight) of mixed culture MM] was added to each test bottle. Some bottles were autoclaved after cell addition (autoclaved cell controls) and others were filled with 100 mL pure water (water controls). HCFCs or HFC solutions (dissolved in water) were added to each bottle using Precision gas tight syringes, then vigorously shaken upside-down on a rotary shaker (250 rpm). Headspace samples were periodically analyzed by GC as described previously. Modeling transformation of HCFCs/HFCs To quantify the cometabolic transformation of HCFCs and the HFC studied, a second order rate expression was combined with an expression describing loss of activity due to endogenous decay (b ) and product toxicity (qc ITC): qc :chL (1) =dX/dt:_b_gc_ (2) X 1; where: qc = specific rate of transformation (mg substrate/mg cell-d) k; = second order rate coefficient (1ng cell-d) CL = liquid phase concentration of the substrate (mg/L) [1 = specific growth (or decay) rate (d'l) X = active organism concentration (mg/L) b = endogenous decay coefficient (d'l) 51 Tc = theoretical or true biomass transformation capacity (mg substrate/mg cell) The endogenous decay term b includes loss of activity caused by cell death and by depletion of reducing power required for monooxygenase activity. A more extensive discussion of these processes and of Eq. (1) and (2) is provided by Criddle (1993). For batch transformation of a volatile cometabolic substrate, a mass balance at equilibrium gives: dM —— C = XV 3 dt qc L ( ) Me = CL(VL + HcVG) (4) where: Mc = mass of substrate (mg) HC = Henry's constant (-) VL = liquid volume (L) VG = gas volume (L) Batch cometabolic transformation can be described by Eq. (1) - (4). These equations can be solved simultaneously using a Runge-Kutta algorithm. Two simplifying cases should be noted. The first occurs when product toxicity is absent or insignificant, I) >> qc /Tc and p = -b. For this case, Eq. (1) - (4) can be combined and integrated to give the mass of substrate Mc as a function of time: —(e-'~-z>} MC = Mcoe{ b (5) where: M60 = initial mass of substrate (mg) 52 X0 = initial active organism concentration (mg/L) A=nkn+m%) The second simplifying case is also obtained when product toxicity is the dominant factor causing loss of transformation activity. For this case, qc ITC >> b, and It = qc /TC and Eq. (1) - (4) can be combined and integrated to give: (—k',AF:) M=Mw F: _ m) X __,;_0_e(-k,AF:) 0 T. where: Cw: initial concentration of substrate in the aqueous phase (mg/L) F=Xo_CLo/7; Disappearance of the target compounds was modeled with both Eq. (5) and (6). Kinetic parameters were estimated by nonlinear regression using Systat 5.1 (Systat, Inc.). For all modeling with Eq. (5), an endogeous decay rate b of 0.594/day was assumed. This value was independently obtained by Clowater (1992) for loss of trichloroethylene (TCE) transformation activity in aerated batch cultures of mixed culture MM]. Cultures of MM] were aerated in the absence of methane and periodically assayed to determine the TCE transformation rate. The endogenous decay coefficient b was then computed as the slope taken from a plot of the logarithm of specific TCE transformation rate vs. aeration time. RESULTS Henry's Law constants Measured Henry's Law constants are provided in Table 3.], along with coefficients of variation. With two exceptions, all coefficients of variation were less than 6%. The effects 53 of temperature on Henry's constant followed the van't Hoff relationship (Gossett 1987). Results from a linear regression of In H vs. Tl( H in m3-atm/mole; T in K) are provided in Table 3.2. Salting-out coefficients are listed in Table 3.3. Henry's constants were relatively insensitive to salinity. For the most sensitive compound studied (HCFC-22), the ionic strength must exceed 0.35 M to cause a greater than 10% increase in the apparent Henry's constant. HCFCs/HFC transformation rates Figures 3.1-3.4 illustrate the methanotrophic transformation of the target compounds. All four compounds were degraded to different degrees over the concentration range studied (900-3000 [1 g/L). Mode] fits obtained using equations 5 and 6 are also illustrated in Figures 3.1-3.4. Estimates for the kinetic parameters used to describe the transformation of each compound are summarized in Table 3.4 and 3.5. 54 Table 3.] Measured values of Henry's constant vs. temperature. compound temperature Hc fiH , W— ‘C H m3.atm/mol % HCFC-22 6 0.622 0.0142 8.00 12 1.277 0.0298 5.36 22 1.679 0.0406 5.57 30 2.358 0.0586 3.87 40 3.535 0.0907 1.99 HCFC-142b 6 1.390 0.0318 4.33 12 1.749 0.0409 8.24 22 2.432 0.0588 4.75 30 3.213 0.0798 2.24 40 3.926 0.101 2.83 HCFC-123 6 0.571 0.0131 3.26 12 0.825 0.0193 2.96 22 1.057 0.0256 4.26 30 1.463 0.0364 5.02 40 1.979 0.0508 1.19 HFC-134a 6 1.190 0.0272 5.92 12 1.528 0.0357 2.47 22 2.067 0.0500 0.99 30 2.199 0.0546 3.86 *Percent coefficient of variation = 100(SD/mean). Triplicate measurements were performed for each compound and temperature. 55 Table 3.2 Temperature regression for Henry's constant*. H=exp(A?B-/Tl) Compound A B r2 HCFC-22 1 1.66 4387 0.956 HCFC-142b 7.363 3011 0.995 HCFC-123 7.805 3373 0.990 HFC-134a 5.714 2588 0.979 *Computed using values given in Table 3.1. Units of Henry's constant are m3-atm/mole; T is in degrees Kelvin. Table 3.3 Salting-out coefficients (22°C)*. log Y=kI Compound k, L/mole r2 HCFC-22 0.1 18 0.996 HCFC-142b 0.0838 0.960 HCFC-123 0.0860 0.997 HFC-134a 0.0761 0.972 *Based upon measurements from 0 to 1.0 M KC1 solution. Salting-out coefficients were determined by plotting log10 (activity coefficient) vs. ionic strength: 10g10'Y = kIwhere: y = activity coefficient (-), k = salting-out coefficient (L/mole), I=ionic strength (M). 56 0 m a f f I .2 Q to 8 0 0 WC 3 I AC a A Live .5 ———-Eq.5 . Eq.6 0 r: J ; * O ‘5 o _____________ N 400 - ‘2' 0 u. 200 1- O I 0 4 l l 1 l 0 25 50 75 100 125 150 Time (hrs) Figure 3.1 Biotransformation of HCFC-22 by methanotrophic mixed culture MM]. Fitting parameters are summarized in Tables 3.4 and 3.5. Error bars give standard deviations for three samples. WC = water control (no cells), AC = autoclaved control, LIVE = 275 mg/L MM] (dry weight). 57 1000 * a 900 a 3 800 3- “A 700 _l 5 T» 600 ‘5 3 c 500 o o o a a ,2 400 - N a. 0 we I L. E 200 - __‘__E:°5 o a I 100 .. Eq.6 O J i L J J 0 25 50 75 100 125 150 Time (hrs) Figure 3.2 Biotransformation of HCFC-142b by methanotrophic mixed culture MMl: model fit based on equation 5 and 6, respectively. Fitting parameters are summarized in Tables 3.4 and 3.5. Error bars give standard deviations for three samples. WC = water control (no cells), AC = autoclaved control, LIVE = 275 mg/L MM]. 58 3000 o o a .c a 2500 I n 3 g 2000 —* u- A “ _l ,5 B, 1500 - _ a o v 5 o 1000 - . we 3 I A0 a 500 _ A Live :3 —---Eq.5 ti Eq.6 o I I J A l 0 25 50 75 100 125 150 Time (hrs) Figure 3.3 Biotransformation of HFC-134a by methanotrophic mixed culture MM]. Fitting parameters are summarized in Tables 3.4 and 3.5. Error bars give standard deviations for three samples. WC = water control (no cells), AC = autoclaved control, LIVE = 275 mg/L MM]. 59 o 0 cu .: a a a o o :1 u- “ 3 2000 r \ 5 m 5 31500 L c 8 1000 - ° WC 3 I AC '7 A Live E 500 - ___-Eq,5 o Eq.6 I o l 1 I 0 25 50 75 100 Time (hrs) Figure 3.4. Biotransformation of HCFC-123 by methanotrophic mixed culture MM]. Fitting parameters are summarized in Tables 3.4 and 3.5. Error bars give standard deviations for three samples. WC = water control (no cells), AC = autoclaved control, LIVE = 275 mg/L MM]. 60 Table 3.4 Kinetic coefficients for HCFC/HFC transformation by methanotrophic mixed culture MM]. Best fit for the parameters of equation 5: comparison with TCE. b k ' Correlation Compound (l/day) (Umg-day)a coefficient 2 r HCFC-123 0.594b 0.00054 0.888 10.00014 HFC-134a 0,594b 0.0009] 0.915 10.00002 HCFC-142b 0.594b 0.0030 0.580 10.0002 HCFC-22 0,594b 0.0043 0.61 1 10.0001 TCE 0.594b 1.4 10.23c 0.998b a Determined by triplicate samples at 95% confidence interval. b Independently determined by Clowater (1992). c Clowater (1992) Table 3.5 Kinetic coefficients for HCFC/HFC transformation by methanotrophic mixed culture MM]. Best fit for the parameters of equation 6: comparison with TCE. T k ' Correlation Compound (11 g Substrate/mg (Umg-day)b coefficient cell)a r2 HCFC-123 1 .166 0.00090 0.934 10.264 10.00011 HFC-134a 1.636 0.0011 0.937 10.05] 10.0001 HCFC-142b 1.113 0.0096 0.981 10.068 10.0016 HCFC-22 2.468 0.014 0.984 10.050 10.002 TCE 47 10C 1.33 10.24c 0.997c a b Determined by triplicate samples at 95% confidence interval. C Clowater (1992). 61 DISCUSSION Different degrees of transformation were obtained for the compounds studied. DeFlaun et al. (1992) reported that HCFC-123, HCFC-142b, HFC-134a were not degraded by the pure culture Methylosinus trichosporium OB3b. In the present study, HCFC-123 and HFC-134a degraded slowly, while HCFC-142b was transformed at a somewhat higher rate by mixed culture MM]. A possible explanation for the difference between this work and the DeFlaun study is that the MM] methanotroph may possess enzymes with greater reactivity toward HCFCs and HFCs. It is also possible that the heterotrophs present in mixed culture MM] facilitated transformation. Uchiyama (1992) found that TCE was mineralized to a greater extent by a mixed culture containing heterotrophs. Pathways of transformation for the compounds evaluated in this study are not known. Presumably, in each case, oxygen is inserted at the carbon-hydrogen bond yielding an alcohol intermediate. Halogenated alcohols undergo further hydrolysis and elimination in aqueous systems giving rise to a variety of products. For HFC-134a, one of the possible products is trifluoroacetic acid. Trifluoroacetic acid is thought to be stable in aqueous environments, and it is an expected oxidation product in the troposphere. Analysis of the culture medium by ion chromatography after completion of HFC-134a transformation revealed a peak at a run time corresponding to that of trifluoroacetic acid. Additional analysis is needed to confirm this tentative identification. For two of the target compounds - HCFC-123 and HFC-134a, use of Eq. (5) and the independently measured endogenous decay rate of 0.594 day'l provided a reasonable fit to the data. It should be noted, however, that the model based on product toxicity (Eq. (6)) also fit the data well, indicating that conclusions about the mechanism for loss of transformation activity cannot be based on model fit alone. 62 As shown in Table 3.4, Eq. (5) provided a poor fit for HCFC-22 and HCFC-142b indicating that, for these compounds, another mechanism (besides endogenous decay) apparently contributes to the loss of transformation activity with time. As shown in Figures 3.1 and 3.2, Eq. (6) provided a good fit to these data. Thus, product toxicity may explain the loss of transformation activity for these compounds. To assist in the interpretation of data, the transformation of the targeted fluorocarbons was compared with trichloroethylene (TCE). TCE is a useful bench mark for comparison because many researchers have evaluated methanotrophic transformation of TCE, and there is an extensive dataset on its transformation kinetics. As indicated by Table 3.5, rates of transformation for all of the fluorinated compounds studied were considerably slower than rates of transformation for TCE. Second order rate coefficients were 100 to 1000 times smaller for mixed culture MM]. Transformation capacities for HCFC-142b and HCFC-22 were ten to twenty times smaller than the values reported for TCE. The rapid loss of activity for HCFC-22 seems reasonable inasmuch as this compound is structurally similar to chloroform, a compound previously known to exhibit product toxicity in methanotrophic mixed cultures (Alvarez-Cohen and McCarty 1991c). For chloroform, the toxic byproduct is believed to be carbonyl chloride (phosgene). An analogous carbonyl may be formed from HCFC-22. Alvarez-Cohen and McCarty (19910) reported a chloroform transformation capacity of 6.5 ug/mg cell, a value somewhat higher than that observed for HCFC-22. The results of this work suggest that methanotrophic transformation is not likely to be a significant sink for the removal of HFCs and HCFCs globally. As indicated in Table 3.5, the fastest second order rate observed in this study was 0.014 L/mg cell-day for HCFC-22. Typical microbial densities in the soil and marine environment are on the order of 106 - 107 organisms per cm3 of soil or water. Assuming that active methanotrophs constitute about 63 1% of these communities, first order rates of 10'10 to 10'9 3'1 would be possible for HCFC-22. In order to have a significant affect (>2-3%) on global lifetime estimates, first order rates exceeding 103 8'1 are required (Rodriguez et al. 1991). REFERENCES 1.A1varez-Cohen, L., and P. L. McCarty. 1991a. A cometabolic biotransformation model for halogenated aliphatic compounds exhibiting product toxicity. Environ. Sci. Technol. 25:1381-1387. 2.Alvarez-Cohen, L., and P. L. McCarty. 1991c. Product toxicity and cometabolic competitive inhibition modeling of chloroform and trichloroethylene transformation by methanotrophic resting cells. Appl. Environ. Microbiol. 57:1031 - 1037. 3.Clowater, L. M. 1992. M. S. thesis. Michigan Stata University. 4.Criddle, C. S. 1993. The kinetics of cometabolism. Biotechnol. Bioeng. 41 : 1048-1056. 5.Criddle, C. S., J. T. DeWitt, and P. L. McCarty. 1990. Reductive dehalogenation of carbon tetrachloride by Escherichia coli k-12. Appl. Environ. Microbiol. 56:3247-3254. 6.DeFlaun, M. F., B. D. Ensley, and R. J. Steffan. 1992. Biological oxidation of hydrochlorofluorocarbons(HCFCs) by a methanotrophic bacterium. Biotechnology. 10:1576—1578. 7.Denovan, B. A., and S. E. Strand. 1992. Biological degradation of Chlorofluorocarbons in anaerobic environments. Chemosphere. 24:935-940. 8.Galli, R., and P. L. McCarty. 1989. Biotransformation of 1,1,1-trichloroethane, trichloromethane, and tetrachloromethane by a Clostridium sp. Appl. Environ. Microbiol. 55:837-844. 9.Gossett, J. M. 1987. Measurement of Henry's law constants for C1 and C2 chlorinated hydrocarbons. Environ. Sci. Technol. 21:202-208. 10.Henry, S. M., and D. Grbié -Galié. 1991a. Influence of endogenous and exogenous electron donors and trichloroethylene oxidation toxicity on trichloroethylene oxidation by methanotrophic cultures from a groudwater aquifer. Appl. Environ. Microbiol. 57:236 - 244. 11.Lesage, S., S. Brown, and K. R. Hosler. 1992. Degradation of Chlorofluorocarbon- 113 under anaerobic conditions. Chemosphere. 24:1225-1243. 12.Lesage, S., R. E. Jackson, M. W. Priddle, and P. G. Riemann. 1990. Occurence and fate of organic solvent residues in anoxic groundwater at the Gloucester Landfill, Canada. Environ. Sci. Technol. 24:559-566. 13.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. 14.Molina, M. J., and F. S. Rowland. 1974. Stratospheric sink for chlorofluoromethanes: chlorine atomcatalysed destruction of ozone. Nature. 2492810-812. 15.Rodriguez, J. M., M. K. Ko, N. D. Sze, and C. W. Heisy. 1991. Modeling lifetimes and sinks for HCFCs/HFCs. Presented at the AFEAS Workshop on Aqueous Loss Processes for Halocarbons, Washington D. C. . 65 16.Rowland, F. S., and M. J. Molina. 1975. Chlorofluoromethanes in the environment. Rev. Geophys. Space Phys. 1321-35. 17.Saéz, P. B., and B. E. Rittmann. 1991. Biodegradation kinetics of 4-chlorophenol, an inhibitory co-metabolite. Research J. WPCF. 63:838-847. 18.Schmidt, S. K., S. Simkins, and M. Alexander. 1985. Model for the kinetics of biodegradation of organic compounds not supporting growth. Appl. Environ. Microbiol. 50:323-331. 19.Semprini, L., G. D. Hopkins, P. L. McCarty, and P. V. Robert. 1992. In-situ transformation of carbon tetrachloride and other halogenated compounds resulting from biotransformation under anoxic conditions. Environ. Sci. Technol. 26:2454-2461. 20.Uchiyama, H., T. Nakajima, O. Yagi, and T. Nakahara. 1992. Role of heterotrophic bacteria in complete mineralization of trichloroethylene by Methylocystis sp. strain M. Appl. Environ. Microbiol. 58:3067-3071. CHAPTER 4 KINETICS OF COMETABOLISM BY GROWING CELLS“ ABSTRACT An unstructured model for cometabolism is presented and verified experimentally for a defined methanotrophic mixed culture. The model includes the effects of cell growth, endogenous cell decay, product toxicity, and competitive inhibition with the assumption that cometabolic transformation rates are enhanced by reducing power obtained from oxidation of growth substrates. A theoretical transformation yield is used to quantify the enhancement resulting from oxidation. A systematic method for evaluating model parameters is described. The applicability of the model is evaluated by comparing experimental data for methanotrophic cometabolism of TCE with model predictions from independently measured model parameters. Propogation of errors is used to quantify errors in parameter estimates and in the final prediction. The model successfully predicts TCE and methane transformation successfully for a wide range of concentrations of TCE (0.5 - 9 mg/L) and methane (0.05 - 6 mg/L). INTRODUCTION Many compounds of environmental and toxicological significance are transformed by cometabolism. In this study, cometabolism is defined as transformation of a nongrowth substrate by cells that are growing in the presence of growth substrate or by resting cells in the absence of growth substrate (Criddle 1993; Horvath 1972). With oxidative * This study was submitted to bigtghnglggy and bioengineering, Jun. 1996. 66 67 cometabolism, the growth substrate may compete with nongrowth substrate for positions at the enzyme active site, hindering transformation of the nongrowth substrate. However, in the absence of the growth substrate, the ability to sustain cometabolic transformation is eventually exhausted. Loss of transformation capacity may also result from damage to cellular material caused by toxic transformation products. Recent studies indicate that cell inactivation can be quantified and incorporated into kinetic models by introducing a parameter termed biomass transformation capacity. This concept has been applied to quantify degradation kinetics for serveral oxgenase- expressing cultures with a range of chlorinated compounds (Alvarez-Cohen and McCarty 1991a; Chang and Alvarez-Cohen 1995b; Chang et al. 1993; Chang and Criddle 1995; Hopkins et al. 1993b). Alvarez-Cohen and McCarty (1991a) assumed that biomass transformation capacity was equal to the mass of nongrowth substrate ultimately degraded divided by the initial biomass used. Criddle (1993) defined a “theoretical” biomass transformation capacity by subtracting the effect of endogenous decay. The later definition represents a theoretical maximum value in the absence of external reducing power. Studies indicate that transformation capacity is not only a function of the availability of reducing power, but also of the specific cometabolized compound and the toxicity of its transformation products (Alvarez-Cohen and McCarty 1991c; Henry and Grbié -Gali<’: 1991a; Wackett and Householder 1989). Several models have been proposed to describe the cometabolic transformations of nongrowth substrate in the absence of growth substrate, many of which have been reviewed by Criddle (1993). Saéz and Rittmann (1991; 1993), for example, linked biomass decay with transformation of nongrowth substrate. Models have also been proposed to describe cometabolic degradation in the presence of growth substrates. Broholm et al. (1992) and Strand et al. (1990) modeled the interaction between growth 68 and nongrowth substrates by competitive inhibition, neglecting product toxicity and reducing power effects. Anderson and McCarty (1994) proposed a biofilm model that incorporated both product toxicity and competitive inhibition, but did not account for reducing power limitations. Recently, Chang and Alvarez-Cohen (1995a) proposed a general model that includes reducing energy explicitly as a limiting reactant during cometabolism. They applied the model to describe degradation of TCE by methanotrophs. The effects of reducing power were separated from toxicity effects and quantified by supplying the cells with formate. For methanotrophs, formate provides energy as NADH, but does not support growth. Thus, the experimental approach of Chang and Alvarez-Cohen is somewhat specific to organisms for which energy substrates can be identified that do not support growth. In this paper, a general model is presented that combines the effects of cell growth, endogenous cell decay, product toxicity, and competitive inhibition with the assumption that cometabolic degradation rates are enhanced by reducing power obtained from oxidation of growth substrates. The proposed model does not require use of energy substrates, such as formate. A theoretical transformation yield is used to quantify the enhancement of cometabolism resulting from oxidation of the growth substrate (Chang et a]. 1993; Criddle 1993). A systematic method for evaluating model parameters is developed. The applicability of the model is evaluated by comparing experimental data for methanotrophic cometabolism of TCE with model predictions from independently measured model parameters. Rationale for experimental system The experimental system selected for investigation in this work was methanotrophic transformation of trichloroethylene. In 1985, Wilson and co-workers reported aerobic oxidation of TCE by soil microorganisms provided natural gas as a primary source of 69 energy (Wilson and Wilson 1985). Since then, the ability of methane-utilizing bacteria to cometabolize TCE and other chlorinated organic solvents has been firmly established by many researchers (Fliermans et a]. 1988; Fogel et al. 1986; Little et al. 1988). The enzyme methane monooxygenase (MMO) oxidizes TCE to an epoxide, which spontaneously degrades to intermediates that can be further metabolized, including glyoxylic acid, dichloroacetic acid, carbon monoxide, and formate (Little et al. 1988; Uchiyama et al. 1992) . Some researchers have reported that, in addition to products resulting from epoxide hydrolysis, intramolecular halide or hydride migration can occur, yielding 2,2,2-trichloroacetaldehyde (chloral hydrate). Chloral hydrate can be reduced to trichloroethanol and oxidized to trichloroacetic acid. All these products are potentially toxic and may cause cellular inactivation (Fox et al. 1990; Newman and Wackett 1991). For TCE transformation, formate addition resulted in increased initial specific transformation rates and elevated transformation capacity. Significant declines in methane conversion rates were observed following exposure to TCE for both resting and formate-fed cells, suggesting toxic effects by TCE or its transformation products (Alvarez-Cohen and McCarty 1991b; Alvarez-Cohen and McCarty 1991c). Oldenhuis and co-workers suggested that TCE epoxide can bind covalently to proteins and nucleic acids. Other possible reactive metabolites that might bind irreversibly are chloral, dichloroacetyl chloride, and formyl chloride (Oldenhuis et al. 1991; Oldenhuis et al. 1989). MATERIALS AND METHODS Culture and culture conditions The methanotrophic culture used for these experiments was a mixed culture originally derived from aquifer material at Moffett Field, California (courtesy S. M. Henry). This culture is a stable consortium consisting of one methanotroph, one hyphomicrobium, and ' several heterotrophs containing Gram-negative thin and fat rods as well as some gram- 70 positive rods and cocci. The methanotroph in the mixed culture expresses soluble MMO similar to that of Methylosinus trichosporium OB3b under similar growth conditions (Henry and Grbic’ -Galié 1991a). Mixed culture MM] was grown in Whittenbury Mineral Medium containing (per liter of deionized water): 1.0 g of MgSO4-7H20, 1.0 g of KNO3, 276 mg of CaSO4-2H20, 3.8 mg of FeEDTA, 0.5 mg of NazMoO4~2H20, 0.5 mg of FeSO4-7H20, 0.4 mg of ZnSO4-7H20, 0.02 mg of MnC12-4H20 0.05 mg of CoC12-6H20 0.01 mg of NiC12-6H20, 0.015 mg of H3BO3, 0.25 mg of EDTA, 260 mg of KH2PO4, and 330 mg of NazHPO4. One liter of culture was grown at room temperature (~21°C) in a continuously stirred 2-liter bottle supplied 30% methane in air at 68 mL/min. Growth curves were monitored and as stationary phase approached, approximately 10 mL of culture was transferred to 1 liter of fresh Whittenbury Medium. Cells were harvested in mid-log growth phase for biotransformation experiments. Analytical methods A TCE-saturated water solution was used as the spike solution in all experiments. The spike solution was prepared by adding excess TCE (991% pure ACS reagent, Aldrich Chemicals Co., Milwankee, WI) to a 250 ml glass bottle capped with TFE-lined Mininert valve. The bottle was vigorously shaken and allowed to settle at least 24 hrs. The upper layer of the solution was transferred to another bottle and capped with a Mininert valve. The spike solution was stored in a refrigerator until needed. One hour before use, it was shaken again and allowed to settle. TCE was analyzed by withdrawing 0.1 ml of headspace from the test bottles using a 0.5 ml Pressure-Lok Series A-2 gas syringe and injecting the samples onto a Hewlett Packard 5890 gas Chromatograph (GC) equipped with a column (DB624, 30m x 0.53mm ID.) and a flame ionization detector. The GC was operated isothermally at 90°C with helium as ' carrier. The temperature at the injection port and detector was 250°C. 71 Methane and oxygen were analyzed by withdrawing 0.1 ml of headspace from the test bottles using a 0.5 ml Pressure-Lok Series A-2 gas syringe and injecting the samples onto a Hewlett Packard 5890 series H gas chromatograph equipped with a column (6 ft x 1/8 in SS packed with 80/ 100 washed molecular sieve 13X) and a thermal conductivity detector. The GC was operated isothermally at 50°C with helium as carrier. The temperature of the injection port and detector were 50°C and 90°C respectively. Cell biomass was determined on a dry weight basis using 0.2 pm filters (Gelman Sciences Inc., Ann Arbor, MI). The filters were prepared by first soaking them in mineral media for 10 minutes, rinsing on a vacuum filter with deionized water, drying overnight in a 103°C oven, and cooling in a desiccator until needed. The filters were weighed, and once a known amount of culture was filtered through them, they were rinsed, dried, cooled and reweighed. Batch biotransformation experiments TCE degradation studies were performed using 25 ml glass vials capped with teflon-lined Mininert valves. These vials were incubated with 5 mL of of Whittenbury Mineral Media plus culture. An appropriate amount (measured as dry weight) of mixed culture MMl was added to each test vial. TCE solutions (dissolved in water) were added to each bottle using Precision gas tight syringes. Methane were withdrawn from Scotty II cyclinders (99.0% CH4, Alltech Associate, Inc., Deerfield, IL) at fixed exit pressure and injected into batch vials. After adding substrates, the vials were vigorously shaken upside-down on a rotary shaker (250 rpm). Headspace samples were periodically analyzed by GC as described previously. 72 Model development A cometabolic model was evaluated that included terms for the loss of microbial biomass or enzyme activity caused by autooxidation (endogenous decay), proteolysis, depletion of cofactors (such as NADH), product toxicity, and suicide inactivation. A theoretical discussion of this model is provided elsewhere (Criddle 1993). For cometabolism in the presence of growth substrate, the following equations provide a complete mathematical description of the specific growth rate and the specific rates of utilization of the growth and the nongrowth substrates throughout the growth and decay periods. s =k q: 3(KS+S) (1) =(T +k)( C > (2) q‘ "q" ‘ KC+C fl=qu.-- 1'- (3) C When there is competitive inhibition between the growth substrate and the nongrowth substrate, (Ks)obs and (Kc)ob3 replace K, and Kc , respectively, in Eq. (1) and Eq. (2), where: C (K: )obs : Ks(l + 2:) (4) S (12),,” — Kc(1 +7) (5) (5 In the absence of growth substrate, the mode] simplifies to: C K+C q. = k.( ) (6) (7) 73 Two important stoichiometric parameters are the observed transformation capacity,(Tc)obs, and the observed transformation yield, (7;)053- (72),,“ is obtained by dividing -qc by u: I (72).)!» = b— qus +1. (8) q. T. For resting cells (qs =0), Eq. (8) simplifies to: (72),“ =—,,——'—, (9) _+_ q. T C For resting cells, the observed transformation capacity is determined by the theoretical transformation capacity, 7;, and by the ratio of the endogenous decay rate b to the specific rate of TCE transformation, qc. The observed transformation yield, (7;)0b3, is obtained by dividing qc by qs. In the absence of cometabolism, electrons from the growth substrate are used exclusively for cell synthesis and respiration so f, + fl = I, where f: = fraction of electrons for cell synthesis and fi = fraction of electrons for energy generation. In cometabolic reaction, however, electrons are consumed for growth, respiration and cometabolism. In this case, f, + fl + fw = I , where fa, = fraction of electrons used for cometabolism (Criddle 1993; Criddle et al. 1991). For oxygenase-mediated reactions, two moles of electrons are consumed for every mole of nongrowth substrate transformed, but this ratio will decrease if the byproducts of transformation are further oxidized by the cometabolizing community. 74 For batch transformation of a volatile cometabolic substrate in the presence of growth substrate, a mass balance for growth and nongrowth substrates at equilibrium gives: _nd =chVL (10) (It MC :C(VL+HCCVG) (11) dM _ s -_—, XV 12 dt qs L ( ) Ms =S(VL+HCSVG) (13) Batch cometabolic transformation in the presence of growth substrate can be described by 5 Eq. (1), (2), (3), (4), (5), (10), (11), (12) and (13). Once the parameters of the model (b, kc, Kc, 7;, k,. K3, Y, K. tc’ K“, 7;) are determined, these equations can be solved simultaneously using a Runge-Kutta numerical method. Simplified cases of the model (C << Kc in the absence of growth substrate) have been previously verified (Chang and Criddle 1995). Model verification The model evaluated in this work was verified using the procedure illustrated in Figure 4.1. Experimental data were compared with predictions from separate measurements of the kinetic parameters. Using the measured parameters, degradation rates for methane and TCE were predicted for specified initial conditions. The predictions were evaluated experimentally. Four independent series of experiments were conducted to measure the maximum specific rate of utilization of substrate (kc and IQ). the half-saturation . coefficient (K. and K,), growth yield (Y), endogenous decay constant (b) and the 75 theoretical transformation capacity (72) in the absence of endogenous decay. To measure the remaining parameters (KINKL, and 7;), an additional set of experiments was conducted over a range of concentrations of growth substrate with high initial TCE concentrations. Details of the experimental evaluation for each of these parameters is described in the following sections. Propogation of errors was used to quantify errors in parameter estimates and in the final prediction (Mandel 1984). All non-linear parameter estimates were obtained by nonlinear regression using Systat 5.2.1 (Systat, Inc., Evanston, IL). Sensitivity analysis (Robinson and Characklis 1984; Robinson 1985) were performed to evaluate the uniqueness of parameter estimates and the relative importance of parameters over the range of substrate concentration. Three equations derived from model, Eq. (1), (6), and (2) were used to estimate three sets of parameters (k, and K3 , kc and Kc , KL, and Ty). The derivatives of dependent variable with respect to each set of parameters (qu I dk, and dqs/sz, dqc ldkc and dqc lch,qu/dKis and dqc/d7; ) were evaluated for a range of substrate concentration. If sensitivity equations for each pair of parameters are not multiples of each other over a wide range of substrate concentration, a unique combination of parameters can be estimated from the data set. To determine the relative importance of each parameter on the specific rate of transformation of growth and nongrowth substrates, the derivatives of qc and qs with respect to related parameters were also evaluated over a range of substrate concentrations. -x_1_1__ 1.1 .0; '_ -. e 0. f u at” f h nu: wt sustrat and half-sat ratin ' t e w r K.- and kc were determined by adding a range of concentrations of TCE to batch cultures of resting cells that were fully induced for the desired cometabolic activity. The initial concentrations spanned a range that bracketed the concentration above which specific rate ' of transformation are maximum and the concentration corresponding to the half- 76 saturation coefficient. The initial slope of the resulting degradation curves for each initial concentration was determined. A nonlinear regression on Eq. (6) was used to estimate kc and Kc. .11 1 ucf at 0. 1112-10! ofrowth us ..t- 1. -a :01 cof' in of w s rv i K, and k3 were determined by adding a range of concentrations of growth substrate to batch cultures. The initial concentrations spanned a range that bracketed the concentration above which specific rate of transformation are maximum and the concentration corresponding to the half-saturation coefficient. The initial slope of the resulting degradation curves for each initial concentration was determined. A nonlinear regression on Eq. (1) was used to estimate k, and K. Y was determined by measuring the increase of dry weight of biomass with the consumption of growth substrate during a period of time. The value was obtained during the growth phase, before decay of cell biomass was significant. En n f't Subsamples were withdrawn from the decaying culture and spiked with high concentrations of TCE, so that Co >> K. The initial slope of the resulting degradation curve was proportional to the concentration of cometabolizing cells. The active fraction remaining at any time was computed by dividing the initial slopes for each subsample by the initial slope at the beginning of the decay period. A semilog plot of active fraction vs. time yields a straight line with slope of -b. I] .1 E . i Once the endogenous decay coefficient b and maximum specific rate of utilization kc 77 were quantified, 7; was determined by adding a high concentration of TCE to a batch culture of resting cells . For C0 >> K; , Eq. (6) simplifies to q; =kc. By combining this result with Eq. (9) and allowing time to become infinite, the actual or observed transformation capacity, (7; )ohsis given by dM VL+HVdC VL+HV Co— —C,, 1 (T,),,,_,= : =(——£—G>— =( C: 6x )=—-———— , VLdX VL dX VL X0 -b—+— kc T. 7; was calculated from the above relationship. [ICL‘ 1. ‘f'l.1.!_ 1"] ° ioO‘M'ioanfii frw _1, - rt 0. or: wh . trat "zain an inhibiti oeffi ien f n wrw 1 . t 1 e 01 ° OWL! substrate utilization To evaluate these three parameters, initial specific rates of utilization of growth substrate and nongrowth substrate were measured over a range of concentrations of growth substrate with high initial TCE concentration (10 mg/L). A nonlinear regression on Eq. (1), (2), (4) and (5) with previously determined values for k; , K; , k; and Kc was used to estimate K1 , K; and 7;. RESULTS AND DISCUSSION Cometabolism is a complex phenomenon, especially when both growth and nongrowth substrates are simultaneously present. For oxygenase-mediated reactions, nongrowth substrate competitively inhibits utilization of the growth substrate, yet utilization of growth substrate is needed for sustained transformation of nongrowth substrate. The model evaluated in this work attempted to capture this paradox. In this research, a systematic method was developed to predict simultaneous degradation 78 of growth and nongrowth substrates. Parameters for growth and nongrowth substrate degradation were first measured alone in the absence of competitive interactions. Thereafter, parameters indicating interaction between growth and nongrowth substrates (K. I; , K; and 7;) were measured in the presence of both substrates. The sensitivity equations for each pair of parameters are not multiples of one another for the wide range of substrate concentration (Figure 4.2). This implies that unique combination of parameters can be estimated from the data set. Sensitivity equations with respect to parameters in Eq. (1), (2), (4), and (5) were also evaluated. The results show much greater sensitivity to maximum specific rate of utilization of growth and nongrowth substrates than to the respective half-saturation coefficients. Of all parameters, K1: was the most sensitive parameter affecting the specific utilization rate of nongrowth substrate. K‘; was less sensitive. The sensitivity equation with respect to K is reached a maximum at the lower concentration of growth substrate. 79 Adequate amount of cells in vials (fully induced) Measure Measure ks , Ks& Y Measure kc & Kc Measure endogenous for a series of growth for a series of transfo- decay substrates in the nongrowth substrates rmation constant, absence of nongrowth in the absence of capacrty, b substrate growth substrate TC Measure Kic for a Measure Kis &Ty for series of growth a series of growth substrates and high substrates and high ‘9— initial cone. of initial conc. of nongrowth substrate nongrowth substrate l Solve model with Runge-Kutta numerical method 1 Prediction of degradation of growth & nongrowth substrates Figure 4.1 Approach for prediction of degradation of growth and nongrowth substrates by the proposed model. (a) sensitivity Parameter (b) sensitivity Parameter 80 0.15 dqs/dks .0 —L 0.05 -0.05 - “ § ‘- Q ‘- ~- ‘- ‘-- ‘-- - ‘-- -_- -0.1 -oo15 l l l l 0 0.5 1 1.5 2 2.5 Methane concentration (mg/L) 0.8 0.7 - 0-5 ” dqc/dkc 0.5 0.4 0.3 0.2 0.1 dQC/ch O ................................................ .. 0.1 l J l l TCE concentration (mg/L) 81 (c v 9° 01 10 mg/L TCE sensitivity N I _L 01 I Parameter dQC/dTy 0 __----"' 'l' 0 0.5 1 1.5 2 2.5 3 Methane concentration (mg/L) Figure 4.2 The sensitivity equations for parameters estimated nonlinearly from model, concentrations shown here are the ranges for each parameter determination. The half-saturation coefficients (K, and K;) are often assumed to be equal to the inhibition coefficients for the respective substrates (KL; and KC) (Alvarez-Cohen and McCarty 1991c; Anderson and McCarty 1994; Broholm et a]. 1992; Chang and Alvarez- Cohen 1995a). Our data indicate that they were significantly different under the present experimental conditions (Table 4.1). This may be caused by the fact that the measurements were performed with whole cells rather than purified enzymes. Factors other than competition for the active site of the enzyme may influence the interactions of growth and nongrowth substrates, e.g., reductant supply or substrate transport to the enzyme (Landa et a]. 1994). Thus, the assumption that both values are the same and that 82 the substrates are substitutable appears to be inappropriate in this case. A similar observation was reported for toluene and TCE by Landa et a]. (18). Independently measured kinetic parameters for TCE degradation are summarized in Table 4.1. Using these parameters, the mode] was solved numerically to predict methane and TCE degradation under various conditions. Figure 4.3 illustrates model predictions and experimental data for batch transformation of growth substrate in the presence of a high initial TCE concentration. To confirm the consistency of predictions, further independent batch degradation experiments were conducted with lower initial TCE concentrations (Figure 4.4). Methane degraded more slowly in all cases when TCE was present, indicating that methane utilization was strongly inhibited by TCE transformation. The results show that the model can predict methane and TCE degradation with reasonable accuracy. Addition of a limited level of methane enhanced TCE degradation over a specific range of methane concentrations (Figure 4.5). However, competitive inhibition between growth and nongrowth substrates also played an important role. Thus, further increases in methane concentration eventually decreased TCE degradation. This result is in agreement with the observations of Chang and Alvarez-Cohen (Chang and Alvarez-Cohen 1995b). The effect of methane is less significant for lower initial TCE concentration. Also, maximum degradation rates for TCE are achieved at lower methane concentrations for lower initial TCE concentrations. This phenomenon can be explained by competitive inhibition between growth and nongrowth substrates. When TCE concentration is low, methane has more chance to occupy active sites in the methane monooxygenase, and TCE transformation is inhibited. 83 Table 4.] Kinetic and stoichiometric parameters for methane utilization, growth, and TCE degradation for methanotrophic mixed culture MM]. Parameters maximum specific rate of utilization of methane, k; half-saturation coefficient of methane, K3 maximum specific rate of utilization of TCE, k; half-saturation coefficient of TCE, Kc theoretical transformation capacity in the absence of endogenous decay, 7; first-order endogenous decay constant, b observed yield, Y inhibition coefficient indicating the effect of methane on TCE utilization rate, Ki; inhibition coefficient indicating the effect of TCE on methane utilization rate, Ki; growth substrate transformation capacity, T .V Value1 3.77 (+/- 0.83) mg/mg cell-day 6.85 (+/- 1.86) mg/L 0.152 (+/- 0.018) mg/mg cell-day 1.94 (+/- 0.46) mg/L 0.0602 (+/- 0.0005) mg TCE Img cell 0.549 (+/- 0.044) /day 0.426 (+/- 0.023) mg cell/mg methane 0.119 (+/- 0.052) mg/L 10.8 (+/- 1.45) mg/L 4.01 (+/- 1.20) mg TCE/mg methane 1. Values represent the 95% confidence interval for triplicate data. Table 4.2 Stoichiometry of TCE cometabolism by methanotrophic mixed culture MM] for different initial TCE concentrations. Initial TCE Initial CH4 ATCE/ACH4 AOzlACH4 AX/ACH41’2 concentration concentraton (mole TCE/ (mole 02/ (mole cells/ (mg/L) (mg/L) mole CH4) mole CH4) mole CH4) 0 6.42 0.88 0.059 0.98 6.19 0.003 1.01 0.054 3.98 6.88 0.232 2.09 0.008 10.2 6.73 0.563 2.35 0 17.2 7.03 0.759 2.34 -0.021 1. The formula of cells was assumed C 5H7OzN. 2. Yields were measured after two-day incubation. 84 Table 4.3 Electron flow in methanotrophic mixed culture MM] for different initial TCE concentrations. Initial TCE Initial CH4 Fraction of Fraction of Fraction of concentration concentraton electrons for electron for electron for (mg/L) (mg/L) cometabolism, energy, fe synthesis, fs 2’3 fco 1 0 6.42 0 0.44 0.832 0.98 6.19 0.008 0.50 0.760 3.98 6.88 0.058 1.05 0.1 19 10.2 6.73 0.141 1.18 0 17.2 7.03 0.190 1.17 0 1. Assumes two moles of electrons required per mole of TCE transformation. 2. The formula of cells was assumed C5H702N. 3. Yields were measured after two-day incubation. Table 4.4 The effect of methane concentration on observed transformation yield when both substrates are simultaneously present: comparison of measured and predicted values. Initial TCE Methane concentration concentration Observed transformation yield (mg TCE/mg CH4) (mg/L) (mg/L) Predictedlv2 Measured2 4.] 0.8 0.88+/-0.42 0.94+/-0.22 1.6 0.52+/-0.25 0.60+/-0.13 3.4 0.28+/-0.13 0.22+/-0.02 4.4 0.22+/-0. 10 0. 14+/-0.03 6.4 0. 15+/-0.07 0.12+/-0.03 8.5 0.6 1 .99+/-0.89 1.84+/-0.36 1 .2 1 .34+/-0.62 0.74+/-0.09 2.9 0.73+/-0.35 0.52+/-0.08 3.4 0.65+/-0.30 0.50+/-0.06 5.8 0.42+/-0.19 0.38+/-0.05 1. Errors calculated from the law of propagation of error. 2. Values represent the 95% confidence interval for triplicate data. 85 9.0 8.0 A _l 3, 7.0 - E I: TCE,2 v 6.0 " q- I c 50 _ 1 TCE,1 E o . I: e 4.0 - on g 3.0 1 0 Methane,1 g 2.0 g g, 0 1.0 Methane.2 0.0 ‘ ___.__._._ ‘ 0.0 0.5 1.0 1.5 2.0 Time (days) Figure 4.3 Biotransformation of TCE and methane by methanotrophic mixed culture: comparison with model predictions (for different initial methane concentration). Error bars indicate the 95% confidence interval for triplicate samples. 86 9.0 8.0 A d m 7.0 " E "' 6° ' TCE1 : _ I n O 5.0 "3 h 4.0 P on 5 3.0 - 2 O 2.0 " O Methane,1,2 1.0 .————111 4 .. TCE.2 0.0 I J l 0.0 0.5 1.0 1.5 2.0 Time (days) Figure 4.4 Biotransformation of TCE and methane by methanotrophic mixed culture: comparison with model prediction (for different initial TCE concentration). Error bars indicate the 95% confidence interval for triplicate samples. 87 (a) 1.8 3 1.5 A g $1.4 GU 1: ' 1.2 '6; E 8 1 m 08 52 U\ 0.6 U'lu 50 0|- 0.4 a. 0.2 ._ 0 Methane concentration (mg/L) Figure 4.5 The effect of methane concentration on TCE and methane degradation rate when both substrates are present at same time. (a) Initial concentration of TCE is 8 mg/L and (b) Initial concentration of TCE is 4 mg/L. Error bars indicate 95% confidence interval. Dashed lines indicate error range for the prediction. 88 (b) 2.5 f f . 4mg/LTCE o A 01>. gm =3 “ 0: Em o m Em 0E \ Um $3 a a' Methane concentration (mg/L) Figure 4.5 The effect of methane concentration on TCE and methane degradation rate when both substrates are present at same time. (a) Initial concentration of TCE is 8 mg/L and (b) Initial concentration of TCE is 4 mg/L. Error bars indicate 95% confidence interval. Dashed lines indicate error range for the prediction. 89 The effect of nongrowth substrate on stoichiometry of cometabolic transformation was also evaluated. The results (Table 4.2) show that an increase in TCE concentration increases the observed transformation yield. It appears that methane is more efficiently used for transformation of TCE at high concentrations of TCE. This result is confirmed by the results in Table 4.4 and Figure 4.7. For higher ratios of TCE to methane, the cultures exhibited net decay after two days of incubation. Therefore, some minimum level of methane was needed to sustain transformation of TCE. The amount of methane needed to sustain transformation of TCE was related to incubation time in a batch system. Microorganisms can not sustain transformation of TCE if TCE concentration is too high with respect to methane. The stoichiometric ratio for oxygen to methane consumed increases with increasing TCE concentration. Electron flow calculations (Table 4.3) indicate that almost all electrons supplied by methane go to energy generation and TCE transformation with none left for cell synthesis at high TCE concentrations. Under these conditions, utilization of growth substrate is inhibited by transformation of nongrowth substrate, and cell growth can not occur. In Table III the sum of f; , f; and fa, exceeded 1. This indicates some error in the electron balance assumptions- perhaps due to errors in the assumed biomass formula or, more likely, the assumption that TCE transformation byproducts are not further utilized. Calculation of fco assumed two moles of electrons were required per mole of TCE transformed. In a mixed culture, this assumption is likely to be incorrect. Inactivation of methane monooxygenase can be caused by the availability of reducing power, endogeneous decay, and transformation of TCE. When there are no external energy sources, a theoretical transformation capacity (7;) can be computed by correcting for cell inactivation caused by endogenous decay and depletion of reducing power. This is a finite value that is independent of the presence of growth and nongrowth substrate and represents a maximum value that cells can theoretically attain. For this culture, 7; 90 appeared to be a characteristic value under the specified experimental conditions. Observed transfomation capacity values predicted using 7; and b are shown in Figure 4.6. (7;) is much lower than 7; when growth substrate concentration is low. This is obs caused by the fact that endogeneous decay becomes more significant at low concentration of growth substrate, when toxicity and use of reducing power become less significant. (T;)0,,s approaches the theoretical value when the concentration of growth substrate is sufficiently high. Table IV shows that observed transformation yield decreased with increasing methane concentration. The observed transformation yield is very small compared with the theoretical value (7;) measured by the mode] when higher levels of methane are present. Apparently, inhibition of growth substrate plays an important role. Transformation of TCE was seriously inhibited when high levels of methane were present. The effect of inhibition of methane offset its enhancement effects. On the other hand, the observed transformation yield was higher than the theoretical value when trace methane was present. Under this condition, enhanced transformation of nongrowth substrate by growth substrate was dominant and a higher observed transformation yield was attained. Therefore, the theoretical transformation yield represents a theoretical value that the cultures could have and is independent of the effect of growth and nongrowth substrates. The observed transformation yield was a function of concentration of growth and nongrowth substrates, as shown in Figure 4.7: the lower the concentration of TCE, the less significant is the effect of methane on observed transformation yield. In summary, the model presented here can be used to predict transformation of growth and nongrowth substrate accurately. A systematic method was developed to measure parameters to describe simultaneous degradation of growth and nongrowth substrates. In previous work, a simplified form of the present model (in the absence of growth 91 substrate) was verified for I-IFC/HCFC degradation by the same methanotrophic mixed culture (Chang and Criddle 1995). These results suggest that the proposed model can be applied to other cometabolic transformations for a range of concentrations and substrate types. ~‘ TCE concentration (m 9! L) \ \ \ \‘ Observed transformation capacity (mg TCE/mg cells) Methane concentration ( m 9 I L) Figure 4.6 The observed transformation capacity as a function of the concentrations of growth and nongrowth substrates. Prediction is based on parameters listed in Table I. 92 MW“? 1"“ 1 i = 4.5 1 5 4 1 1 3'5 Observed 3 transformation yield, qclqs 0 (mg TCE/mg methane) TCE concentration (In 9/L) Methane concentration (m glL) Figure 4.7 The observed transfomation yield as a function of the concentration of growth and nongrowth substrates. Prediction is based on parameters listed in Table I. 93 REFERENCES 1.A1varez-Cohen, L., and P. L. McCarty. 1991a. A cometabolic biotransformation mode] for halogenated aliphatic compounds exhibiting product toxicity. Environ. Sci. Technol. 25:1381-1387. 2.Alvarez-Cohen, L., and P. L. McCarty. 1991b. Effects of toxicity, aeration, and reductant supply on trichloroethylene transformation by a mixed methanotrophic culture. Appl. Environ. Microbiol. 57:228 - 235. 3.Alvarez-Cohen, L., and P. L. McCarty. 1991c. Product toxicity and cometabolic competitive inhibition modeling of chloroform and trichloroethylene transformation by methanotrophic resting cells. Appl. Environ. Microbiol. 57:1031 - 1037. 4.Anderson, J. E., and P. L. McCarty. 1994. Model for treatment of trichloroethylene by methanotrophic biofilms. J. Environ. Eng. 120:379-400. 5.Broholm, K., T. H. Christensen, and J. B. K. 1992. Modelling TCE degradation by a mixed culture of methane-oxidizing bacteria. Wat. Res. 26:1177-1185. 6.Chang, H.-]., and L. Alvarez-Cohen. 1995a. Model for the cometabolic biodegradation of chlorinated organics. Environ. Sci. Technol. 29:2357-2367. 7.Chang, H.-L., and L. Alvarez-Cohen. 1995b. Transformation capacities of chlorinated organics by mixed cultures enriched on methane, propane, toluene, or phenol. Biotechnol. Bioeng. 45:440-449. 8.Chang, M.—k., T. C. Voice, and C. S. Criddle. 1993. Kinetics of competitive inhibition and cometabolism in the biodegradation of benzene, toluene, and p-xylene by two Pseudomonas isolates. Biotechnol. Bioeng. 41:1057-1065. 9.Chang, W.-k., and C. S. Criddle. 1995. Biotransformation of HCFC-22, HCFC-142b, HCFC-123, HFC-134a by methanotrophic mixed culture MM]. Biodegradation. 6: 1-9. 10.Criddle, C. S. 1993. The kinetics of cometabolism. Biotechnol. Bioeng. 41 : 1048-1056. 11.Criddle, C. S., L. A. Alvarez, and P. L. McCarty. 1991. Microbial processes in porous media, p. 639-691. In J. Bear and M. Y. Corapcioglu (ed.), Transport processes in porous media. Kluwer Academic Publishers, Netherlands. 12.Fliermans, C. B., T. J. Phelps, D. Ringelberg, A. T. Mikel], and D. C. White. 1988. Mineralization of trichloroethylene by heterotrophic enrichment cultures. Appl. Environ. Microbiol. 54: 1709-1714. 13.Fogel, M. M., A. R. Taddeo, and S. Fogel. 1986. Biodegradation of chlorinated ethene by a methane-utilizing mixed culture. Appl. Environ. Microbiol. 51:720-724. 14.Fox, B. G., J. G. Bomeman, L. P. Wackett, and J. D. Lipscomb. 1990. Haloalkene oxidation by the soluble methane monooxygenase from Methylosinus trichosporium OB3b: mechanistic and environmental implication. Biochemistry. 29:6419-6427. 15.Henry, S. M., and D. Grbf: -Galié. 1991a. Influence of endogenous and exogenous electron donors and trichloroethylene oxidation toxicity on trichloroethylene oxidation by 94 methanotrophic cultures from a groudwater aquifer. Appl. Environ. Microbiol. 57:236 - 244. 16.Hopkins, G. D., J. Munakata, L. Semprini, and P. L. McCarty. 1993b. Trichloroethylene concentration effects on pilot field-scale in-situ groundwater bioremediation by phenol-oxidizing microorganisms. Environ. Sci. Technol. 27 :2542- 2547. l7.Horvath, R. S. 1972. Microbial co-metabolism and the degradation of organic compounds in nature. Bacteriol. Rev. 36:146-155. 18.Landa, A. S., E. M. Sipkema, J. Weijma, A. A. C. M. Beenackers, J. Dolfing, and D. B. J anssen. 1994. Cometabolic degradation of trichloroethylene by pseudomonas cepacia G4 in a chemostat with toluene as the primary substrate. Appl. Environ. Microbiol. 60:3368-3374. 19.Little, C. D., A. V. Palumbo, S. E. Herbes, M. E. Lidstrom, R. L. Tyndall, and P. J. Gilmer. 1988. Trichloroethylene biodegradation by a methane-oxidizing bacterium. Appl. Environ. Microbiol. 54:95] - 956. 20.Mandel, J. 1984. The Statistical Analysis of Experimental Data. Dover Publication, Inc., New York. 21.Newman, L. M., and L. P. Wackett. 1991. Fate of 2,2,2-trichloroacetaldehyde (chloral hydrate) produced during trichloroethylene oxidation by methanotrophs. Appl. Environ. Microbiol. 57:2399 - 2402. 22.0]denhuis, R., J. Y. Oedzes, J. J. van der Waarde, and D. B. Janssen. 1991. Kinetics of chlorinated hydrocarbon degradation by Methylosinus trichosporium OB3b and toxicity of trichloroethylene. Appl. Environ. Microbiol. 5727-14. 23.0ldenhuis, R., R. L. J. M. Vink, D. B. Janssen, and B. Witholt. 1989. Degradation of chlorinated aliphatic hydrocarbons by methylosinus trichosporium OB3b Expressing soluble methane monooxygenate. Appl. Environ. Microbiol. 55:2819-2826. 24.Robinson, J., and W. G. Characklis. 1984. Simultaneous estimation of Vmax, Km, and the rate of endogenous substrate production (R) from substrate depletion data. Microbial Ecology. 10:165-178. 25.Robinson, J. A. 1985. Determining microbial kinetic parameters using nonlinear regression analysis, p. 61-114. In M. K. C. (ed.), Advances in Microbial Ecology, vol. 8. Plenum Press, New York. 26.Saéz, P. B., and B. E. Rittmann. 1991. Biodegradation kinetics of 4-chloropheno], an inhibitory co-metabolite. Research J. WPCF. 63:838-847. 27.Saéz, P. B., and B. E. Rittmann. 1993. Biodegradation kinetics of a mixture containing a primary substrate (phenol) and an inhibitory co-metabolite (4-chlorophenol). Biodegradation. 4:3-21 . 28.Strand, S. E., M. D. Bjelland, and H. D. Stensel. 1990. Kinetics of chlorinated hydrocarbon degradation by suspended cultures of methane-oxidizing bacteria. Research J. WPCF. 62: 124-129. 95 29.Uchiyama, H., T. Nakajima, O. Yagi, and T. Nakahara. 1992. Role of heterotrophic bacteria in complete mineralization of trichloroethylene by Methylocystis sp. strain M. Appl. Environ. Microbiol. 58:3067-3071. 30.Wackett, L. P., and S. R. Householder. 1989. Toxicity of trichloroethylene to Pseudomonas putida F] is mediated by toluene dioxygenase. Appl. Environ. Microbiol. 55:2723-2725. 31.Wilson, J. T., and B. H. Wilson. 1985. Biotransformation of trichloroethylene in soil. Appl. Environ. Microbiol. 49:242-243. CHAPTER 5 THEORY OF ADAPTATION OF COMETABOLIZING COMMUNITIES INTRODUCTION Selection in a continuous culture on a limiting carbon source can lead to the development of stable microbial communities. The relative dominance of populations within these communities can be altered by environmental perturbations, such as exposure to nongrowth substrates. Selection among microbial populations usually occurs because of competition between different species or competition between an ancestor and its derived genotypes. As a result of such processes, a new microbial community can develop that is better adapted to its new environment. For the microbial communities considered in this work, commensal interactions are likely. For such interactions, one species produces compounds which serve as an energy or carbon source for a second species. Such commensal realtionships are often strung together in a chain so that over time a succession of commensal pairs appears. For example, Wilkinson et al . (1974) studied the nature of commensal interactions for a microbial community utilizing methane. The community consisted of a methane-utilizing Pseudomonas sp., a methanol utilizing Hyphomicrobium sp., an Acinetobacter sp., and a F lavobacter sp. The authors concluded that the Hyphomicrobium sp. served to remove the small amounts of methanol which are produced during methane utilization by the Pseudomonas sp. and which inhibited its growth. Acinetobacter and F lavobacter sp. removed the complex products of growth or cell lysis. 96 97 The commensal relationship has a twist under cometabolic conditions. The species that are responsible for utilization of growth substrates also transform nongrowth substrates simutaneously. Cometabolizing pure cultures do not mineralize nongrowth substrates and accumulation of transformation products is typical (Henry and Grbié -Gali<’: 1990; Little et al. 1988; Oldenhuis et al. 1989). Not surprisingly, mixed cultures or communities have advantages for mineralization of nongrowth substrates (Uchiyama et a1. 1992). Since pure cultures with oxygenase activity typically suffer from product toxicity, the presence of heterotrophs capable of degrading such products may facilitate detoxification, enhancing the growth rate of the community. Thus, selection in cometabolizing environments may favor community structures with higher capabilities for detoxication. In this chapter, the fitness concept (Lenski et a]. 1991) is adapted for evaluation of cometabolism and the changes that occur in a cometabolizing community. The major concept for the model is that changes in a microbial community can be described by a "fitness" parameter which is quantified by changes in the kinetic parameters of the community. The selection gradient for each parameter is defined by the partial derivative of fitness with respect to that parameter. The gradient therefore reflects the direct selection acting on each fitness component, with the other components held constant. The experimental systems selected for investigation in this work is TCE transformation by a methanotrophic mixed culture and by a phenol-degrading community. MODEL DEVELOPMENT Adaptive change of microbial communities In a competitive situation, we are interested in discovering whether a species enjoys a natural advantage. Microorganisms (r strategist) with the fastest growth rate should come to dominate when resources are temporarily abundant, since by virtue of their rapid growth, they will be able to utilize more of the limiting factor than the slower-growing organism. K 98 strategists, which reproduce more slowly than r strategist, tend to be successful in resource-limited situations (Andrews and Hall 1986). However, considerations should also be given to application of this concept for mixed cultures, consortia, and communities where interactions other than simple competition are important. For simple communities fed a single substrate, one or a few species are predominant. Intermediates from oxidation of the growth substrate or products of decay of the dominant species sustain other species within the community. Since the whole community is interconnected by amensual or mutualistic relationships, we assume that the whole community can be viewed as a single population with a single phenotype. Under this assumption, phenotypic changes during exposure to nongrowth substrate will be expressed as changes in the phenotype of the whole community. The Malthusian parameter for a specific species over any time period t is given by (Lenski et al. 1991). m=ln(N/N0)/t (1) Where No and N are initial and final cell number during a growth period, respectively. We assume that the above definition can be applied to a whole community derived from a single growth substrate: m = ln(X/X0)/t (2) Where X0 and X are initial and final cell density, respectively. The fitness, W of a derived community relative to the ancestor community is expressed as the ratio of their respective Malthusian parameters : 99 W:_ (3) For any short period, the specific growth rate, 11 is a constant . Fitness at any instant is then expressed as: (4) For a continuous culture, the specific growth rate, 11 is the actual Malthusian parameter. Generally, we can measure the concentrations of growth and nongrowth substrate, and kinetic parameters to obtain 11 at any instant. The specific growth rate prior to TCE exposure is 1.10. The selection gradient for any trait P is the partial derivative of fitness with respect to that trait, 8W I 3P. The gradient therefore reflects the direct selection acting on each fitness component, with the other components held constant. To facilitate comparison among selection gradients for the several fitness components, gradients are normalized to create dimensionless quantities that reflect the proportional sensitivity of fitness to each component (Vasi et a]. 1994): P [W GP - (WK?) (5) Component fitness (fitness contributed by the change of a trait) can be calculated from : aWMP AW = n (aPn n (6) Therefore, the change of fitness for a derived community can be expressed as: 100 AW = 2 —AP (7) n=1 The above equation is assumed to express the development of the community. In principle, changes in community structure can be quantified in this way. Fitness of cometabolizing cultures For the case where growth substrate utilization and cometabolic transformation occur in separate stages, growth phase change is described by: as YkS = m 2 ms 8 fl KS+S KS+S U and W=-'—"—=-‘—‘— (4) mo ”0 Where W=1 for the ancestor community and W=oo for the theoretical maximum fitness in the community (um z oo). During the death phase, cell decay occurs as a result of endogenous decay and transformation of nongrowth substrate (Chapter 3): q =—b——° 9 fl 7; () and kC = c 10 q, K+C ( ) C In order to have a consistent trend in fitness over periods of net growth and net decay, a 101 modified definition of fitness and selection gradient is proposed: Wzfi; 00 01111-111 Where W=-1 for the ancestor community and W=0 for the maximum fitness of the community (b = 0 and qC I7; = 0) For cometabolism in the presence of growth substrate, the following equations provide a complete mathematical description of the specific growth rate and the specific rates of utilization of growth and nongrowth substrates throughout the growth and the decay periods (Chapter 4). S q, = k,( C ) (13) le+——)+S 1C C q. = (T,.q, + k.)( S ) (14) K;(1+ —) + C Kis u=na— -$- 03 C and the definition of fitness given by Eq. (1 1) can be used for all cases. a W = —— Incl (11) Where W=1 or - 1 for the ancestor community and W=oo for the theoretical maximum fitness in the community (11,m 2 co, b=0 and q; / 7; = 0). 102 Fitness in a cometabolizing community can begin with 1 or - 1 for the ancestor community depending upon whether there is sufficient growth substrate to support growth. Subsequently, fitness of the community can evolve throughout the whole ‘fitness space’. negative positive growth growth in in ancestor ancestor community community -oo -1 0 1 oo - 1 1 1 > ————> increasing fitness Figure 5.1 Fitness space of cometabolizing community Selection gradients The selection gradient is a normalized parameter defined so as to indicate the relative effects of different fitness components on fitness and to enable comparison of selection gradients for several fitness components. The definition of selection gradient has been shown in Eq. (5). Selection gradients with respect to um and K, can be derived for growth phase using Eq. (8): 0.1-:11) ._. £1 1%: 16 0,, (Wxaflm) I < > 103 KS 8W —Ks G - — = 17 "5 (W)(8Ks_) KS+S ( ) The change of fitness for the derived community can be expressed as: 8W 8W AW: A —AK 18 an," ”771+ 8K5 S ( ) Selection gradients for um and K8 are zero when there is no growth substrate. Selection gradients for b, qc and T; can then be derived as follows: G, =(|_v%|)(%):_b +711)” (19) G=(1115—|’(aq)= bf—IITT— (20) (7:01am 1231.77; (21) The change of fitness for the derived community can be expressed as: AW=a—W-Ab+a—W—AT+flAq (22) 3b 312. (94. " Selection within a community is more complex when both growth and nongrowth substrates are present at the same time. However, referring to Eq. (13), (14), and (15), the change of fitness for the derived community can be expressed as follows: 104 AW=-a—W—AY +2—W—Ab+flA7; +2—W-A7; +9341: +flAk ay 36 are ar, 31 31 +8W AK 8W 8W 8W 3K, 8K 8K "‘ (3K; "‘ (23) The effect of fitness components on fitness can be deduced from Eq. (16), (17), (19), (20), and (21). Fitness can increase if maximum specific growth rate, 1.1m and theoretical biomass transformation capacity, 7; increase. Alternatively, fitness decreases if half- saturation coefficient of growth substrate, Ks, endogenous decay constant, b and specific rate of utilization of nongrowth substrate, qc increase. These predictions are summarized in Table 5.1. Table 5.1. The effects of fitness components on fitness Increase of fitness components Effects on fitness,W maximum specific growth rate, pm Increase half-saturation coefficient of growth substrate, K s Decrease endogenous decay constant, b Decrease specific rate of utilization of nongrowth substrate, q; Decrease Theoretical biomass transformation capacity, 7; Increase Implications of theory for different model systems Several model systems have been considered for the evaluation of cometabolism. In most cases, both growth and nongrowth substrates are introduced into the model systems. Thus, the interactions between both substrates are important. The presence of toxic transformation products can be expected to have some impacts on the microbial community structure during long-term TCE exposure. These factors will have different effects in simple mixed 105 cultures compared to complex communities and in batch reactors compared to continuous reactors. In sequencing batch reactors (SBRs), for example, the microbial community is exposed to a range of growth and nongrowth substrate concentrations. This may select for a more diverse community with "specialist" organisms that occupy variety of niches created by substrate concentration gradients. In contrast, a chemostat favors selection of specific populations at a fixed specific growth rate. This can be expected to result in a less diverse culture. Since nongrowth substrates are not mineralized by cometabolizing species and since heterotrophs are known to play important roles in detoxification, more diverse cultures should have advantages for cometabolism. In this work, two model communities are evaluated, representing different extremes in the continuum of community. A methanotrophic mixed culture in a chemostat is used to represent one extreme in the continuum. The phenol-degrading SBR community represents a more complex case. The simple methanotroph community is essentially a consortium, with very limited diversity. Methane utilizers, methanol utilizers, and associated heterotrophs are bound together in a commensal relationship, with methane as the sole source of carbon and energy. By contrast, more complex interactions are likely in the batch-fed phenol-degrading community. As discussed previously, the selection gradient is a normalized parameter that indicates the effects of different parameters on fitness. This definition may be useful in predicting adaptation of specific systems. For example, the selection gradient with respect to K, could be -1 and O for chemostat and SBR respectively because substrate concentration is low for a chemostat (S << K,) and periodically high for a SBR (S >> K). It may also be possible to use selection gradients (especially more sensitive ones) as criteria for the stability of a community under different environments or different communities in the same environment. Magnitude of selection gradients for different organisms and different 106 environments can be compared. Lower values of selection gradients indicate a more stable community for a specified environment. MODEL ANALYSIS The experimental systems selected for investigation in this work were TCE transformation by a methanotrophic mixed culture and a phenol-degrading community. The model for cometabolism given by Eq.(13), (14), and (15) was verified previously (chapters 3 and 4). Kinetic parameters for the methanotrophs were listed in Table 4-1. The parameters for phenol-degrading community are provided in chapter 7. As discussed previously, the selection gradient is an important indicator of the relative importance of parameters (fitness components) on fitness. These parameters were evaluated over the ranges of growth substrate (methane and phenol) and nongrowth substrate (TCE) concentration. Selection when both substrates are present at the same time is compared to selection when both substrates are supplied at different times for the methanotrophic mixed culture. The phenol- degrading community was also evaluated when both substrates were present separately. For this case, the phenol-degrading community and the methanotroph culture can be compared. To determine the relative importance of each parameter on the specific rate of transformation of growth and nongrowth substrates, the derivatives of q; and q; with respect to related parameters were also evaluated over a range of substrate concentrations. RESULTS AND DISCUSSION Selection gradients for the methanotrophic mixed culture when growth and nongrowth substrates are supplied separately is shown in Figure 5.2. Apparently, there is no selection with respect to the inhibition constants. The selection gradients of growth and decay phases (including nongrowth substrate transformation) are evaluated separately. The effects of fitness components are in agreement with predictions in Table 1. Similar results were also 107 observed for the phenol-degrading community when growth and nongrowth substrates are not supplied simultaneously (Figure 5.3). Comparing Figures 5.2 and 5.3, both the methanotrophic mixed culture and the phenol- degrading community have the same selection gradient with respect to Y," and k; (or um). However, there are some differences for selection gradients with respect to 7; , b , kc, and K;. For example, a 100% increase in theoretical transformation capacity (7;) causes 50% increase of fitness at 2 mg/L TCE for phenol-degrading community. The same TCE exposure would cause 70% increase in fitness for the methanotroph culture. There are similar observations for other parameters. Thus, the phenol-degrading community is more stable than methanotrophic mixed culture based on these criteria. This implies that it may be possible to use selection gradients (especially more sensitive ones) as criteria for the stability of a community under a given perturbation. Figure 5.4 and 5.5 show selection gradients for each fitness component as functions of T CE and methane concentration when both substrates are present at the same time. Over a wide range of concentrations, K; and K; are the most sensitive gradients with respect to TCE and methane concentration. The selection gradients are higher at higher substrate concentration for these two parameters. This implies that a smaller change in K; and K .1 is needed to attain the same level of change in fitness. As a result, the half-saturation coefficient of nongrowth substrate and inhibition coefficient have more significant effects on fitness at high substrate concentrations. Some other differences can also be observed. The selection gradient for k; shifts to negative values and the selection gradient for K; shifts to positive values when both substrates are present at the same time. This indicates that improvements in the transformation of TCE (such as an increase of K; and 7;) become more critical for adaptive changes. 108 Figure 5.6 to 5.9 also provide a sensitivity analysis for q; and q; as a function of TCE and methane concentrations, respectively. The results show that an increase in K; and a decrease in K; can increase qc. However, the effect on q; is less significant. Compared to fitness with respect to several parameters, q; and q; are not so sensitive to substrate concentration and change in parameters. Various model environments also have different effects on selection. The batch growth environment selects strongly for a higher maximum specific growth rate, with much weaker selection for populations that have higher affinity for substrate. The ratio between selection gradients with respect to um and K; represents a measure of fitness that can be attributed to these two parameters. Dividing Eq. (16) by (17), the proportional selection gradient is obtained: _Esfi Gm/stz K S 11 In continuous culture, the dilution rates through the reactor, D, are usually a fraction of the maximum specific growth rate, um. Assuming p. = umS/(KS +S) = 0.211,“, the proportional selection gradient for um and KS is -1.25. By comparison, the proportional selection gradient for um and KS in a batch reactor can differ by a factor of several hundred(since S>> Ks ). In the absence of growth substrate, there is no selection with respect to um and K5 in cultures conducting cometabolic transformations. However, batch cultures are exposed to a higher concentration of nongrowth substrate initially than continuous cultures. Therefore, the selection gradient with respect to qc is much higher in batch culture since q; is directly proportional to concentration of nongrowth substrate. However 7; can offset the effect of q;, so it is difficult to differentiate the selection advantages of batch and continuous 109 culture. The proportional selection gradient for qc and 7; is a constant factor (Ge. IGT;_ = —-I). This implies that there is counter selection with respect to qc and 7; over any range of nongrowth substrate concentration. In summary, this work extended the fitness concept to describe adaptation of microbial communities. Selection gradient is the key concept underlying a theory for the quantitative description of stability of communities. It appears possible to use selection gradients (especially more sensitive ones) as criteria for the stability of a community under a given perturbation. Communities are likely to exhibit adaptive changes in response to given perturbations (such as, long term nongrowth substrate exposure). A more stable community shows a smaller change in kinetic parameters as a result of a given perturbation. However, a less stable community has a high capacity for change in performance because it is characterized by larger selection gradients. Thus, fitness may be used to track the stability and adaptation of communities under perturbation. An unstable community characterized by high selection gradients and low fitness can be adapted to create a more stable community characterized by low selection gradients and high fitness. 110 (a) 1.5 ‘5 1 - Ym, ks .2 u 2 a: 0.5 - C .2 0 - fl 0 .2 ° 05 - m ' ,Ks -1 P 1 l 0 1 2 3 4 Methane concentration (mg/L) (b) Tc u C .2 'u . u Kc h a: C o b a: 0 .2 «1’5 kc 0 1 2 3 4 TCE concentration (mg/L) Figure 5.2 Dimensionless selection gradient, (P/W)(8W/8P) as a function of methane and TCE concentrations for methanotrophic mixed culture when both substrates are present separately. 1]] 1.4 - 1.2 ' 1 - Ym, ks (L8 - 0u4 - Selection gradient (L2 - o I I I 0 50 100 150 200 Phenol concentration (mg/L) (b) (16 To 054 - 0.2 L Kc -0.6 ~ kc Selection gradient TCE concentration (mg/L) Figure 5.3 Dimensionless selection gradient, (P/W)(BW/3P) as a function of phenol and TCE concentrations for phenol degrading culture when both substrates are present separately. 112 (a) 2 1 Methane: 1.5 - 0.2 mg/L fl 5 '5' Tc 8 Ks a) Kc Kis : o O Ym, KIC 33 b 2 kc a ks, Ty (I) -1.5 I I I I O 2 4 6 8 10 TCE concentration (mg/L) (b) 14 , Methane: 12 ' 2 mg/L ‘5 .2 10 ~ '0 a 3 L h a: 5 _ c o 4 ' . '3 KC,K|S 8 2 - Tc, Ks (g 0 :1 Ym, Kic, b ' kc, ks, Ty -2 -4 , 1 4 - 1 , O 2 4 6 8 10 TCE concentration (mg/L) Figure. 5.4 Dimensionless selection gradient, (P/W)(3W/8P) as a function of TCE concentration for methanotrophic mixed culture at a specified methane concentration. 113 (a) Kc Kis “ I: .2 'U a b O) t: O '3 0 2 o (I) 0 W Tc, Ks, Ym 5 . . ] Kic, kc, b , ks,Ty 0 2 4 6 Methane concentration (mg/L) (b) 9 Kc 3 Kis ‘5 7 c1 '5 6 2 U) 5 4 t: .2 3 ‘6 m 2 0 To U) 1 Ks O Ym,Kic,kc,b _1 ks, Ty O 2 4 6 Methane concentration (mg/L) Figure. 5.5 Dimensionless selection gradient, (P/W)(8W/3P) as a function of methane concentration for methanotrophic mixed culture at a specified TCE concentration. 114 m 0.15 Methane: g 0.1 b 0.2 /L 2:. «I g 0.05 . ks(dqs/dks) 0 /' Kic(dqs/dKic) > 0 .2: > E a Ks(dqs/sz) : 0 (I) -0.15- ' 1 ' . 0 2 4 6 8 10 TCE concentration (mg/L) I Methane: N: ks(dqs/dks) Kic(dqs/dKic) _L U" -L I Sensitivity equation 0 I o 0" I Ks(dqs/sz) I I I I 0 2 4 6 a 10 TCE concentration (mg/L) I A Figure. 5.6 Sensitivity equation for q; as a function of TCE concentration for methanotrophic mixed culture at a specified methane concentration. 115 (a) ks(dqs/dks) Kic(dqs/dKic) Sensitivity equation - 1 ‘ ' ‘ ‘ Ks(dqs/sz) O 1 2 3 4 5 Methane concentration (mg/L) A b) ks(dqs/dks) Kic(dqs/dKic) Sensitivity equation Ks(dqs/d Ks) 0 1 2 3 4 5 Methane concentration (mg/L) Figure. 5.7 Sensitivity equation for q; as a function of methane concentration for methanotrophic mixed culture at a specified TCE concentration. 116 (a) 0.8 Kic(dqc/dKic) 0 6 bMethane: : ’ 0.2 m/L ks(dqc/dks) .2 0.4 - fl " 02 g. ' _, --—---— ------ — -------------------- Ty(dqc/dTy) o O kC(dqc/dkc) o 2 Kis(dqc/dKis) 3‘ ' ' Kc(dqc/ch) 'S' -0.4 .3 "5 -0.6 E. -0.8 m _1 _ Ks(dqc/sz) _1.2 I I I A O 2 4 6 8 10 TCE concentration (mg/L) (b) 2 , ks(dqc/dks) ,_._ 1.5 - figs/1° Kic(dqc/dKic) O “:- 1 ' a g. 0.5 TY(dqc/dTY) m Kis(dqc/dKis) 0 kC(dqc/dkc) E. _0.5 Kc(dqc/ch) > E -1 a) C -1.5 I) m -2 Ks(dqc/sz) -2.5 I I I I O 2 4 6 8 10 TCE concentration (mg/L) Figure. 5.8 Sensitivity equation for q; as a function of TCE concentration for methanotrophic mixed culture at a specified methane concentration. 117 "”3 0.25 - 0.5 m/L ks(dqc/dks) Ty(dqc/dTy) Kis(dqc/dKis) Kic(dqc/dKic) kc(dqc/dkc) Kc(dqc/ch) Ks(dqc/sz) Sensitivity equation 0 1 2 3 4 5 A U' V Kic(dqc/dKic) ks(dqc/dks) A Ty(dqc/dTy) Kis(dqc/dKis) kc(dqc/dkc) Kc(dqc/ch) Ks(dqc/sz) Sensitivity equation -3 I I I I 0 1 2 3 4 5 Methane concentration (mg/L) Figure. 5.9 Sensitivity equation for q; as a function of methane concentration for methanotrophic mixed culture at a specified TCE concentration. 1 18 REFERENCES 1.Andrews, J. H., and R. F. Hall. 1986. r- and K selection and microbial ecology. Advances in Microbial Ecology. 9199-147. 2.Henry, S. M., and D. Grbié -Galic. 1990. Effect of mineral media on trichloroethylene oxidation by aquifer methanotrophs. Microbial Ecology. 20:151-169. 3.Lenski, R. E., M. R. Rose, S. C. Simpson, and S. C. Tadler. 1991. Long-term experimental evolution in Escherichia Cali. 1. Adaptation and divergence during 2,000 generations. The American Naturalist. 138:1315-1341. 4.Little, C. D., A. V. Palumbo, S. E. Herbes, M. E. Lidstrom, R. L. Tyndall, and P. J. Gilmer. 1988. Trichloroethylene biodegradation by a methane-oxidizing bacterium. Appl. Environ. Microbiol. 54:95] - 956. 5.0ldenhuis, R., R. L. J. M. Vink, D. B. Janssen, and B. Witholt. 1989. Degradation of chlorinated aliphatic hydrocarbons by methylosinus trichosporium OB3b Expressing soluble methane monooxygenate. Appl. Environ. Microbiol. 55:2819—2826. 6.Uchiyama, H., T. Nakajima, O. Yagi, and T. Nakahara. 1992. Role of heterotrophic bacteria in complete mineralization of trichloroethylene by Methylocystis sp. strain M. Appl. Environ. Microbiol. 58:3067-3071. 7.Vasi, F., M. Travisano, and R. E. Lenski. 1994. Long-term experimental evolution in Escherichia Cali. 11. Changes in life-history Traits During Adaptation to a Seasonal Environment. The American Naturalist. 144. 8.Wilkinson, T. G., H. H. Topiwala, and G. Hamer. 1974. Interaction in a mixed bacterial population growing on methane in continuous culture. Biotechnol. Bioeng. 16:41- 59. CHAPTER 6 CHANGE IN COMMUNITY STRUCTURE IN RESPONSE TO LONG TERM TCE EXPOSURE: METHANOTROPHIC MIXED CULTURE IN CHEMOSTAT‘ INTRODUCTION Trichloroethylene (TCE) is widely found in soil and groundwater near industrial sites. In 1985, Wilson and co-workers reported on the possibility of aerobic oxidation of TCE by soil microorganisms with natural gas as the primary energy source (Wilson and Wilson 1985). Since then, the ability of methane-utilizing bacteria to cometabolize TCE has been reported and confirmed by several researchers (Fliermans et al. 1988; Fogel et a]. 1986; Little et al. 1988). This work was concentrated on the kinetics (Alvarez-Cohen and McCarty 1991c; Anderson and McCarty 1994; Chang and Alvarez-Cohen 1995a; Criddle 1993; Folsom et a]. 1990; Strand et a]. 1990) and pathways of degradation (Fliermans et a]. 1988; Fogel et a]. 1986; Fox et al. 1990; Little et al. 1988; Nakajima et al. 1992; Newman and Wackett 1991; Oldenhuis et a]. 1989). To date, few researchers have investigated changes in community structure and performance in response to long-term TCE exposure. Wilkinson et al . (1974) evaluated a stable mixture of four bacterial species in continuous culture with methane as the sole carbon source. The community consisted of a methane- utilizing Pseudomonas sp., a methanol utilizing Hyphamicrobium sp., and, in addition, an * The genetic analyses described in this study performed with the assistance of Dr. Denise Searles 119 120 Acinetobacter sp. and a F lavobacter sp. The Pseudomonas was the only species that could utilize methane as a carbon and energy source, and it constituted the dominant member of the community (ca. 90% of the biomass). The authors concluded that Hyphomicrobium sp. removed small amounts of methanol produced during methane utilization by the Pseudomonas sp.. Methanol was believed to be inhibitory to the Pseudomonas sp.. The Acinetobacter and F lavobacter sp. apparently removed complex products generated during cell growth or lysis. Recent studies indicate that TCE transformation capacity is not only a function of the availability of reducing power, but also of the specific cometabolized compound and the toxicity of its transformation products (Alvarez-Cohen and McCarty 1991a; Henry and Grbic’: -Galié 1991a; Wackett and Householder 1989). Formate addition resulted in increased initial specific TCE transformation rates and elevated transformation capacity. Significant declines in methane conversion rates following exposure to TCE were observed for both resting and formate-fed cells, suggesting toxic effects caused by TCE or its transformation products (Alvarez-Cohen and McCarty 19913; Alvarez-Cohen and McCarty 1991b). Only a few researchers have examined on the toxicity of transformation products. Oldenhuis and co-workers suggested that TCE epoxide can be expected to bind covalently to proteins and nucleic acids. Other possible reactive metabolites that might bind irreversibly are chloral, dichloroacetyl chloride, and formyl chloride (Oldenhuis et a]. 1991). Organisms capable of degrading a large amounts of TCE should possess active detoxification systems for these compounds. The accumulation of stable TCE breakdown products in methanotroph pure cultures indicates that methanotrophic bacteria alone are not be able to mineralize TCE completely (Henry and Grbit’: -Galié 1990; Little et al. 1988; Oldenhuis et al. 1989).Some research has shown that methanotrophic mixed cultures are advantageous for mineralization of TCE. 121 Since methanotrophs suffer from product toxicity when transforming TCE, heterotrophs that degrade the toxic products may play an important role in detoxification. Heterotrophic bacteria in methanotrophic mixed cultures apparently can degrade most of the water-soluble T CE breakdown products, decreasing levels of water-soluble radiolabel and increasing production of l"’COz (Little et al. 1988). Futherrnore, Uchiyama and co-workers reported that a heterotrophic bacterium isolated from a methanotrophic mixed culture, Xanthobacter autotrophicus, can oxidize dichloroacetic and glyoxylic acid completely and can reduce trichloroacetic acid to lower levels. These results indicate that heterotrophic bacteria play an important role in TCE degradation (Uchiyama et al. 1992). The presence of toxic transformation products can be expected to have some impacts on the development of microbial communities during long-term TCE exposure. Changes in the populations are likely related to the level of TCE exposure, turnover of transformation products and utilization of growth substrate. Lackey et al . (1994) used total-recycle expanded-bed bioreactors to evaluate the degradation potential of TCE by a microbial consortium. Ester-linked phospholipid fatty acid profiles (PLFAME) were used to monitor the change of TCE-affected community during short-term perturbation. The results showed that a propane-utilizing bacteria] biomaker increased as TCE was degraded and propane consumed. However, the relationship between community structure and extent of TCE exposure was not clear for these short-term exposures. Several genetic analysis techniques have been developed for identification of methanotrophic bacteria. Tsuji et al . (1990) demonstrated that it is possible to distinguish and classify methanotrophic bacteria using 16S rRNA sequence analysis. Another report described the use of PFGE (pulsed-field gel electrophoresis) -restriction fragment length polymorphisms and the use of a cloned DNA fragment carrying the component B gene to detect soluble MMO genes from methanotrophs on Southern blots prepared from gels on 122 which large DNA restriction fragments were separated by PFGE. This technique, when combined with fluorescence-labeled oligodeoxynucleotide signature probes and Western blot analysis, enabled characterization of methanotrophs and detection of methanotrophs that synthesize soluble methane monooxygenase (MMO) (Tsien and Hanson 1992). More recently, PCR primers specific for four of the five structural genes in the soluble MMO gene clusters for several methanotrophs were used to amplify specific DNA sequences for direct detection of methanotrophs in natural environments (McDonald et al. 1995). Molecular biology techniques offer new opportunities for the analysis of the structure and species composition of microbial communities. Some approaches obtain information about microbial communities directly without the need for sequencing. Amplified Ribosomal DNA Restriction Analysis (ARDRA) or Restriction fragment length polymorphism provides a fingerprint of the microbial community under study(Martinez-Murcia et al. 1995; Massol-Deya et a1. 1995).If the rDNA fingerprints for individual bacteria in a community are sufficiently different, then one can examine the amplified products for a series of distinct patterns resulting from the different populations that make up the community. Another technique is based on the separation of PCR-amplified fragments of genes coding for 16S rRNA, all the same length, by denaturing gradient gel electrophoresis (DGGE). DGGE analysis of different microbial communities demonstrated the presence of up to 10 distinguishable bands in the separation pattern, which were likely derived from 10 different species constituting these populations. It is possible to identify constitutes which represent only 1% of the total community with this technique (Muyzer et al. 1993). These methods could be used for a quick assessment of genotypic changes over time or between different location reflecting different environmental conditions. In this study, a methanotrophic mixed culture was exposed to different levels of TCE for extended time periods. A chemostat was chosen as the model environment for the 123 methanotrophic mixed culture. Changes in the consortia were monitored and the effects of TCE exposure on community structure were evaluated. Various phenotypic parameters were also monitored. Community analyses were also conducted to monitor shifts in microbial community structure. Changes in the community were analyzed using the fitness theory presented in chapter 5. MATERIALS AND METHODS Culture conditions A methanotrophic enrichment obtained from aquifer material at Moffett Field, California, was used for these experiments. This culture is a stable consortium consisting of one methanotroph, one hyphomicrobium, and several heterotrophs containing Gram-negative thin and fat rods as well as some Gram-positive rods and cocci. The methanotroph in the mixed culture expresses soluble MMO similar to that of Methylosinus trichosporium OB3b under similar growth conditions (Henry and Grbié -Galié 1991a). Mixed culture MM] was grown in Whittenbury Mineral Medium containing (per liter of deionized water): 1.0 g of MgSO4-7HZO, 1.0 g of KNO3, 276 mg of CaSO4-2H20, 3.8 mg of FeEDTA, 0.5 mg of NazMoO4-2H20, 0.5 mg of FeSO4-7H20, 0.4 mg of ZnSO4-7H20, 0.02 mg of MnC12-4HzO, 0.05 mg of CoC12-6HzO, 0.01 mg of NiC12-6HzO, 0.015 mg of H3BO3, 0.25 mg of EDTA, 260 mg of KH2PO4, and 330 mg of NazHPO4. One liter of culture was grown at room temperature (~21°C) in a continuously stirred 2-liter bottle supplied 30% methane in air at 68 mL/min. Growth curves were monitored and as stationary phase approached, approximately 10 mL of culture was transferred to a 1 liter of fresh Whittenbury Medium. Cells were harvested in mid-log growth phase for inoculation of two chemostats. 1 24 Perturbation of TCE Possible feed modes for TCE peturbation are pulse input and step input. Pulse input should be able to adapt consortia for enhanced transformation of TCE. After exposure to TCE for a period of time, the consortia is allowed to recover and grow back. Alternatively, a step input can be used to evaluate the impact of toxicity of TCE transformation on the community. In this study, an input mode combining the two basic modes was used. TCE was injected in an impulse way but with step increases of the TCE level for more extended periods of time . Duplicate chemostats were seeded with the methanotrophic mixed culture taken from the batch reactor. The set-up of chemostats is shown in Figure 6.1. After the reactors had grown to steady state, one was injected with TCE solution. TCE exposure was continuous over a finite period, halted to allow recovery of the consortia, and then reintroduced for another period at a higher concentration with another recovery period. This pattern was repeated ten times (Figure 6.2). The operating conditions for the chemostats are summarized in Table 6.1. 2330 “02:2 oqubocmfio—Z a .3 souaccommcfih NOE 28 228825 .26 253m 228805 .258 28 59: MOB 23 . .So as,» .53.. 2.5 «Bk? .2 d 348 z_ «Bk? _ f _ .\ .52 023000 _ j L ~=<>Vm0 J a A O 1 O O «26 O O 38 1 2.32% o o o o 26 O _ _ O O O 0 {ii amt; /. ozsooo thee 22.5.28 mu... «Em: _ 302 9.0 A — _ _ _ .52 225:2 . _ .50 250285 _ _ _ m. "I .‘.'.' 500 *- . I m C I I a 400 - I I = 1: . I 300 ' II I I " I 3 I 0 20° ‘ n . lExposed 100 ' oControl O J I l I 0 100 200 300 400 500 Time (Days) Figure 6.3 Changes of Cell density of methanotrophic mixed cultures in respond to long- terrn TCE exposure Observed growth yield (mg cells/mg methane) 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 -0.1 135 Time (days) Figure 6.4 Observed growth yield for methanotrophic mixed cultures during long-term TCE exposure Table 6.2 Endogeneous decay constant for methanotrophic mixed cultures during long-term TCE exposure 150 1a 100 12 r——'l n n ['1 ll “n so 6 mg TCE/L mg/d P I I I ’ I I Exposed OControl 0 100 200 300 400 -—Time fidogenous decay r2 for regression (Days) constant, 1/day Exposed Control Exposed Control 25 0.881 0.804 0.824 0.878 54 0.657 0.921 0.999 0.953 94 0.631 0.988 0.718 0.811 115 0.438 0.786 0.954 0.899 124 0.62 0.758 0.586 0.663 136 0.261 0.616 0.843 0.857 164 0.654 0.459 0.777 0.807 193 0.549 0.225 0.945 0.955 136 _150 18 “n” -100 12 g, IExposed mgTCE/L mg/d V 025 _ OControl b . I- A a 2 f “-5 0.2 - IB 0 ° I g 5,015 - I a E I E ... \ "' ll] 20 0.1 ~ i i b l- O ‘3', 0.05 - g I h r- o - - . 0 100 200 300 400 Time (days) Figure 6.5 Transformation capacity for methanotrophic mixed cultures during long-term TCE exposure 1150 18 2 .———-. nnnnnfl nnlidgg° 1: . mgTCE/L mg/d 1.5 - I .- 3 1 ' ”’1' .. 0.5 - a 3 5 ° ' i: I -0.5 - 1 __ IExposed I OControl _1.5 l l l l L l l O 50 100 150 200 250 300 350 400 Time (Days) Figure 6.6 Fitness of methanotrophic mixed cultures in respond to long-term TCE exposure in chemostats 137 E _:_1 S Figure 6.7 ARDRA fingerprints of control and TCE-exposed methanotrophic communities. 138 DISCUSSION Changes in cell density for methanotrophic mixed culture exposed to periods of continuous loading of TCE are shown in Figure 6.3. Continuous exposure to TCE initially resulted in loss of capability to utilize methane and degrade TCE . Cell densities dropped drastically and increased slowly after TCE addition stopped. TCE concentration in aqueous phase within the exposed reactor changed from 0.01 to 0.1 mg/L approximately when cell densities dropped. This suggests that product toxicity impacted the microbial community. Greater TCE concentrations had a similar effect, and more time was needed for recovery of the biomass. For this culture the strong inhibition of TCE on methane utilization ( Kic: 10.8 mg/L) may be another factor in the decline of cell density. Clearly, simultaneous feeding of TCE to the community had a great impact on growth. Continuous exposure to TCE initially resulted in decreased rates of methane utilization and TCE transformation. Kinetic parameter measurements showed that maximum utilization rates of TCE decreased initially but returned as cell density increased. Growth yield and transformation capacity showed similar trends during TCE exposure. However, greater values of transformation capacity were observed between cell recovery and the next TCE exposure. This result suggests that the community adapted to TCE exposure by increasing transformation capacity. The relationship between the improvement in transformation capacity and diversity of community is still not clear from the data obtained. Specific growth rates of methanotrophic mixed culture in the chemostat were calculated from cell density data by mass balance of biomass. Fitness was calculated as the ratio of specific growth rate of the culture at any time to the respected value before the initiation of TCE exposure. The results indicate some decline in fitness just after initiation of TCE exposure (50 mg/L TCE in feeding), with subsequent recovery to the original value. After '150 mg/L TCE feeding was initiated, a significant drop in fitness was observed(from 1 to 139 -1). The fitness recovered to 1 after TCE feeding was halted. This result indicates that the microbial community was not sustainable when TCE was fed continuously. A possible explanation is the limited diversity of this methanotrophic mixed culture, perhaps contributing to the low tolerance to TCE exposure. Another explanation is the sensitivity of this particular culture to competitive inhibition of methane utilization by TCE. As discussed previously, the relatively high selection gradient for this culture indicates a greater instability for the culture (chapter 5). ARDRA results indicated changes in both the chemostat cultures from the original batch culture inoculum. However, there were no further changes in the control reactor and only minor changes could be seen in the TCE-exposed community, even after TCE feed concentration was raised to 150 mg/L (Figure 6.7). These results indicate that the community structure of the methanotrophic mixed culture was stable despite exposure to TCE. While transitory phenotypic changes occurred, the populations present within the community remained unchanged. Greater changes in community structure occurred when TCE feed concentration was subsequently reduced to 100 mg/L. Less diversity was observed at this exposure. The more diverse community again appeared when influent levels were restored to 150 mg/L. The results suggest that higher diversity corresponded to higher TCE exposure. 140 REFERENCES 1.A1varez-Cohen, L., and P. L. McCarty. 1991a. A cometabolic biotransformation model for halogenated aliphatic compounds exhibiting product toxicity. Environ. Sci. Technol. 25:1381-1387. 2.Alvarez—Cohen, L., and P. L. McCarty. 1991b. Effects of toxicity, aeration, and reductant supply on trichloroethylene transformation by a mixed methanotrophic culture. Appl. Environ. Microbiol. 57:228 - 235. 3.Alvarez-Cohen, L., and P. L. McCarty. 1991c. Product toxicity and cometabolic competitive inhibition modeling of chloroform and trichloroethylene transformation by methanotrophic resting cells. Appl. Environ. Microbiol. 57:1031 - 1037. 4.Anderson, J. E., and P. L. McCarty. 1994. Model for treatment of trichloroethylene by methanotrophic biofilms. J. Environ. Eng. 120:379-400. 5.Chang, H.-l., and L. Alvarez-Cohen. 1995a. Model for the cometabolic biodegradation of chlorinated organics. Environ. Sci. Technol. 29:2357-2367. 6.Criddle, C. S. 1993. The kinetics of cometabolism. Biotechnol. Bioeng. 41 : 1048-1056. 7.Fliermans, C. B., T. J. Phelps, D. Ringelberg, A. T. Mikell, and D. C. White. 1988. Mineralization of trichloroethylene by heterotrophic enrichment cultures. Appl. Environ. Microbiol. 54: 1709-1714. 8.Fogel, M. M., A. R. Taddeo, and S. Fogel. 1986. Biodegradation of chlorinated ethene by a methane-utilizing mixed culture. Appl. Environ. Microbiol. 51:720-724. 9.Folsom, B. R., P. J. Chapman, and P. H. Pritchard. 1990. Phenol and trichloroethylene degradation by Pseudomonas cepacia G4: kinetics and interactions between substrates. Appl. Environ. Microbiol. 56:1279-1285. 10.Fox, B. G., J. G. Bomeman, L. P. Wackett, and J. D. Lipscomb. 1990. Haloalkene oxidation by the soluble methane monooxygenase from Methylosinus trichosporium OB3b: mechanistic and environmental implication. Biochemistry. 29:6419-6427. 11.Henry, S. M., and D. Grbié -Galié. 1990. Effect of mineral media on trichloroethylene oxidation by aquifer methanotrophs. Microbial Ecology. 20:151-169. 12.Henry, S. M., and D. Grbié -Galié. 1991a. Influence of endogenous and exogenous electron donors and trichloroethylene oxidation toxicity on trichloroethylene oxidation by methanotrophic cultures from a groudwater aquifer. Appl. Environ. Microbiol. 57:236 - 244. 13.Lackey, L. W., T. J. Phelps, V. Korde, S. Nold, D. Ringelberg, P. R. Bienkowski, and D. C. White. 1994. Feasibility testing for the on-site bioremediation of organic wastes by native microbial consortia. International Biodeterioration & Biodegradation. 33:41-59. 14.Little, C. D., A. V. Palumbo, S. E. Herbes, M. E. Lidstrom, R. L. Tyndall, and P. J. Gilmer. 1988. Trichloroethylene biodegradation by a methane-oxidizing bacterium. Appl. Environ. Microbiol. 54:951 - 956. 141 15.Martinez-Murcia, A. J ., S. G. Acinas, and F. Rodriguez-Valera. 1995. Evaluation of prokaryotic diversity by restrictase digestion of 168 rRNA directly amplified from hypersaline environments. FEMS Microbiol. Ecology. 17:247-256. 16.Massol-Deya, A. A., D. A. Odelson, R. F. Hickey, and J. M. Tiedje. 1995. Bacterial community fingerprinting of amplified 168 and 16-23S ribosomal DNA gene sequences and restriction endonuclease analysis (ARDRA), p. 1-8. In A. D. L. Akkermans, J. D. V. Elsas, and F. J. (1. Bruijn (ed.), Molecular Microbial Ecology Manual, vol. 3.3.2. Kluwer Academic Publishers, Dordrecht, The Netherlands. 17.McDonald, I. R., E. M. Kenna, and J. C. Murrell. 1995. Detection of methanotrophic bacteria in environmental samples with the PCR. Appl. Environ. Microbiol. 61:116-121. 18.Muyzer, G., E. C. d. Waal, and A. G. Uitterlinden. 1993. Profiling of complex microbial populations by denaturing gradient gel electrophoresis analysis of polymerase chain reaction-amplified genes coding for 16S rRNA. Appl. Environ. Microbiol. 59:695- 700. 19.Nakajima, T., H. Uchiyama, O. Yagi, and T. Nakahara. 1992. Novel metabolite of trichloroethylene in a methanotrophic bacterium, Methylocystis sp. M, and hypothetical degradation pathway. Bioscience, Biotechnology, and Biochemistry. 56:486-489. 20.Newman, L. M., and L. P. Wackett. 1991. Fate of 2,2,2-trichloroacetaldehyde (chloral hydrate) produced during trichloroethylene oxidation by methanotrophs. Appl. Environ. Microbiol. 57:2399 - 2402. 21.0ldenhuis, R., J. Y. Oedzes, J. J. van der Waarde, and D. B. Janssen. 1991. Kinetics of chlorinated hydrocarbon degradation by Methylosinus trichosporium OB 3b and toxicity of trichloroethylene. Appl. Environ. Microbiol. 57:7-14. 22.0]denhuis, R., R. L. J. M. Vink, D. B. Janssen, and B. Witholt. 1989. Degradation of chlorinated aliphatic hydrocarbons by methylosinus trichosporium OB3b Expressing soluble methane monooxygenate. Appl. Environ. Microbiol. 55:2819-2826. 23.Strand, S. E., M. D. Bjelland, and H. D. Stensel. 1990. Kinetics of chlorinated hydrocarbon degradation by suspended cultures of methane-oxidizing bacteria. Research J. WPCF. 62:124-129. 24.Tsien, H.-c., and R. S. Hanson. 1992. Soluble methane monooxygenase component B gene probe for identification of methanotrophs that rapidly degrade trichloroethylene. Appl. Environ. Microbiol. 58:953-960. 25.Tsuji, K., H. C. Tsien, R. S. Hanson, S. R. DePalma, R. Scholtz, and S. LaRoche. 1990. 168 ribosomal RNA sequence analysis for determination of phylogenetic relationship among methylotrophs. J. Gen. Microbiol. 136:1-10. 26.Uchiyama, H., T. Nakajima, O. Yagi, and T. Nakahara. 1992. Role of heterotrophic bacteria in complete mineralization of trichloroethylene by Methylocystis sp. strain M. Appl. Environ. Microbiol. 58:3067-3071. 27.Vasi, F., M. Travisano, and R. E. Lenski. 1994. Long-term experimental evolution in Escherichia Cali. 11. Changes in life-history Traits During Adaptation to a Seasonal Environment. The American Naturalist. 144. 142 28.Wackett, L. P., and S. R. Householder. 1989. Toxicity of trichloroethylene to Pseudomonas putida F1 is mediated by toluene dioxygenase. Appl. Environ. Microbiol. 55:2723-2725. 29.Wilkinson, T. G., H. H. Topiwala, and G. Hamer. 1974. Interaction in a mixed bacterial population growing on methane in continuous culture. Biotechnol. Bioeng. 16:41- 30.Wilson, J. T., and B. H. Wilson. 1985. Biotransformation of trichloroethylene in soil. Appl. Environ. Microbiol. 49:242-243. CHAPTER 7 CHANGE IN COMMUNITY STRUCTURE IN RESPONSE TO LONG TERM TCE EXPOSURE: PHENOL-DEGRADING COMMUNITY IN SEQUENCING BATCH REACTOR‘ INTRODUCTION Trichloroethylene (T CE) is widely found in soil and groundwater. Considerable research has established that certain phenol- and toluene-degrading Pseudomonas species can rapidly cometabolize TCE (Folsom et al. 1990; Nelson et al. 1987; Wackett and Gibson 1988). To date, this research has concentrated on kinetics (Alvarez-Cohen and McCarty 1991c; Chang and Alvarez-Cohen 1995a; Criddle 1993; Folsom et al. 1990; Strand et a1. 1990) and pathways (Harker and Kim 1990; Nelson et al. 1988; Shields et al. 1989; Wackett and Gibson 1988) with relatively little research on adaptive changes in microbial community structure. TCE transformation capacity is a function of available reducing power and toxicity of transformation products (Alvarez-Cohen and McCarty 1991a; Henry and Grbié -Galié 1991a; Wackett and Householder 1989). Reducing power is required for monooxygenase activity and can be derived from either the growth substrate or supplemental electron donors, such as formate. For example, formate addition to a phenol-degrading community resulted in increased transformation yield and elevated transformation capacity (Hopkins et * The genetic analyses described in this study were performed with the assistance of Dr. Denise Searles 143 144 al. 1993b). Other reports indicate that the toxicity of certain transformation products can limit the extent of transformation. Oldenhuis and co-workers suggested that TCE epoxide can bind covalently to proteins and nucleic acids. Other possible reactive metabolites that might bind irreversibly are chloral, dichloroacetyl chloride, and formyl chloride (Oldenhuis et al. 1991). Communities and populations capable of degrading a large amounts of TCE should possess active detoxification systems for these compounds. Several reports indicate that TCE cometabolism by pure cultures does not result in the mineralization of TCE (Henry and Grbic’: -Galié 1990; Little et al. 1988; Oldenhuis et al. 1989). Other research suggests that microbial consortia or community have advantages for mineralization of TCE. Because the TCE-oxidizers suffer from product toxicity, associated heterotrophs may play an important role in detoxification (Little et al. 1988). Uchiyama and co-workers (1992) reported that a heterotrophic bacterium in a methanotrophic mixed culture, Xanthobacter autotrophicus, can oxidize dichloroacetic and glyoxylic acid completely and can reduce trichloroacetic acid to lower levels. These results indicate that heterotrophic bacteria play an important role in TCE detoxification. The presence of toxic transformation products can be expected to impact the development of microbial communities during long-term TCE exposure. Changes in the populations are likely related to the level of TCE exposure, turnover of transformation products and utilization of growth substrate. Lackey et al . (1994) used total-recycle expanded-bed bioreactors to evaluate the degradation potential of TCE by a microbial consortium. Ester- linked phospholipid fatty acid profiles (PLFAME) were used to monitor changes in a propane-fed community during short-term perturbation with TCE. A propane-utilizing bacterial biomaker increased as TCE was degraded and as propane was consumed. However, the relationship between community structure and extent of TCE exposure was not clear for these short-term exposures. 145 The effect of growth substrate feeding pattern on the structure and potential cometabolic activity of a phenol-degrading community was evaluated by Shih et al. (1996). The results indicated that the manner of growth substrate addition can have a pronounced effect on community structure and cometabolic activity. Communities enriched with continuous or protracted feeding intervals exhibited limited long-term capacity for TCE transformation. Communities enriched with short feeding intervals maintained higher TCE transformation rates. Pulse feeding also resulted in more stable and diverse communities. However, this research did not investigate community changes during long-term exposure to TCE. Adaptation of communities to changing environments may proceed in several ways. Genetic recombination is one mechanism that can lead to new genotypes in the absence of mutation. Genetic elements brought together may enable microorganisms to carry out new functions, and can result in adaptative changes. Selection among species is another mechanism of adaptation. The impact of the selection in evidenced by the characteristic of the community. In chapter 5, a theory of community "fitness" was presented. In this chapter, the cometabolism model verified previously (Chapter 5) is used to analyze adaptive changes in terms of the " fitness " concept. In this manner, fitness was quantified in terms of measurable kinetic parameters. Sequencing batch reactors (SBRs) have several advantages for cometabolic transformations. An SBR can alternate between periods of growth on growth substrates and periods of cometabolism of nongrowth substrate, eliminating the possibility of competitive inhibition for enzymes between growth and nongrowth substrates. SBRs also have a great ability to periodically change environmental condition, creating temporal concentration gradients, selecting or enriching specific microbial populations. Thus, SBRs offer a good model environment to study the dynamic changes of microbial communities. 146 In this study, sequencing batch reactors were chosen as model environments. A phenol- degrading community was exposed to increasing levels of TCE over time. Changes in the community were monitored and the effects of TCE exposure on community structure were evaluated. Various phenotypic parameters were measured. Community analyses were also conducted to monitor shifts in microbial community structure. Changes in the community were analyzed using the fitness theory presented in chapter 5. MATERIALS AND METHODS Culture conditions A stable phenol-degrading microbial community was obtained by seeding a chemostat with activated sludge from a municipal wastewater treatment plant (East Lansing, Michigan) and providing a phenol-containing medium for two months at a dilution rate of 0.1 day‘l. The enrichment was matained at 21.5 i 10°C. This culture was inoculated into a 2-liter stirred reactor supplied continuously with air. Two hundred milliliters of medium (2000 mg/L phenol) was provided as a single daily pulse immediately after removing the same volume of biomass from the reactor (Shih et al. 1996). This fed-batch system was operated for over 400 days, yielding a stable community with good TCE transformation properties. Microscopic examination revealed a community with distinctive floc structures of spherical and rod-shaped bacteria. Some fungi were also observed. The community was then used to inoculate duplicate SBRs. Phenol feed medium contained the following (per liter of deionized water): 2 g of phenol, 2.13 g of NazHPO4, 2.04 g of KHzPO4, 1 g of (NH4)2SO4, 0.067 g CaC12-2H20, 0.248 g of MgC12-6H20, 0.5 mg of FeSO4-7H20, 0.4 mg of ZnSO4-7H20, 0.02 mg of MnC12-4H20, 0.05 mg of CoC12-6H20, 0.01 mg of NiC12-6H20, 0.015 mg of H3BO3, 0.25 mg of EDTA. The pH of the medium was 6.8. 147 Perturbation of TCE The set-up of SBRs is shown in Figure 7.1. After both reactors had stabilized, one was injected with TCE over a one-hour interval during the filling period of cycle operation. Phenol was provided in a separate period for both reactors. Influent TCE levels were increased gradually from 0.5 to 25 mg/L over 4 months (Table 7.1). The operating mode of the SBR during a cycle is shown in Figure 7.2. The operating conditions for the SBRs are summarized in Table 7.2. Monitoring of consortia To monitor changes in the community, the following parameters were measured periodically (the methods for measuring these parameter are described in chapter 4) : (a). Theoretical transformation capacity in the absence of endogenous decay, Tc (b). Endogenous decay rate, b (c). Second order rate coeficient of transfomation of TCE in the absence of methane, k; ((1). Maximum specific rate of utilization of phenol in the absence of TCE, k, (e). Observed yield of cells, Y Specific growth rate of the community was obtained by multiplying observed yield of cells by maximum specific rate of utilization of phenol. Also, cell samples were taken periodically and the change in community structure were observed. Microscopic examination was also performed periodically. Genetic analysis was used to evaluate changes in community structure. The change in kinetic parameter values was compared to the results from morphological observation or genetic analysis. 148 Adaptation of cometabolizing community in SBR In this work, phenol was the sole growth and energy substrate, so that tha actual substrate available to members of the community is phenol, intermediates generated by oxidation of phenol or products of cell decay. Under these conditions, it is assumed that the community can be treated as a single population (chapter 5). The fitness, W of such a community relative to ancestor community is expressed as the ratio of the respective specific growth rates: W=— (1) The phenol-degrading community in SBR utilized phenol according to zero order kinetic, thus qc = k3. and the specific growth rate of the cells is u = Ymks. Selection gradients with respect to Ym and k, can be derived as follows: _ Y... W _ _ k, aw _ G“ _(W)(8k )"1 (3) The change of fitness for the derived community can be expressed as: 8W 8W AW: AY + 31;, "' A al., k: (4) 149 Off Gas r121 —l —1 F —1 ‘—l 80 O Fl O O f) i I OC 0 0 1:1 I O O “ 001:1 Oo- Air Pump Water Bath Medium Feel TCE Feed Syringe Syringe Decant Medium Pump Pump 0 g D g l ._l Phenol Feed Phenol Feed! Syringe Syringe _ Recycle Timer pump :1 El q L Effluent Medium Tank Tank Figure 7.1. Experimental setup for bench—scale sequencing batch reactors 150 Table 7.1. TCE feed concentration for long-term TCE exposure experiment Start days (days) TCE feed concentration TCE loading per cycle (mg/L) (mg) 0 0 0 76 0.5 0.5 100 2.5 2.5 140 5 5 150 10 10 183 15 15 191 25 25 Full Read Settle Decant Recharge 0 1 4 5 6 65 12 l l I l l L l I l l l I T ”'1 2500 - TCE feed Pheno| Cell 2 f d wastin g 2000 - 9° 9 2 1500 - O > 7 b 1000 , O 4.1 o 3 500 P a: o 1 1 1 1 1 1 1 1 1 1 1 01 2 3 4 5 6 7 Time(hrs) Figure 7.2. The operating mode of SBR in a cycle 10 11 12 151 Table 7.2 Summary of bench-scale SBR operating condition Parameter Value Regtgr volume Total volume 2500 mL Liquid volume 1100 - 2200 mL Headspace volume 300 - 1400 mL Initial volume 1100 mL flow rate Influent flow rate 16.7 mL/min TCE concentration 0.5-25 mg/L Recharge flow rate 1.2 mL/min Phenol concentration 5000 mg/L Air flow 220-280 mL/min Minimum oxygen 2 mg/L concentration in reactor Magus. Fill time 1 hr Reaction time 3 hrs Settle time 1 hr Decant time 1 hr Recharge time 6 hr - phenol feed mode 0.5 hr - phenol react mode 5.5 hr Operating cycle time 12 hrs/cycle Sludge age, SRT 10 day General procedures for biotransformation measurement Bc' f 'n ' ts Biotransformation studies were performed using 20-ml glass vials sealed with teflon-coated butyl rubber stoppers and aluminum crimp caps. These vials were incubated with 5 mL of of Whittenbury Mineral Media plus culture. An appropriate amount (measured as dry 152 weight) of mixed culture MMl was added to each test vial. TCE solutions (dissolved in water) were added to each bottle using Precision gas tight syringes. Methane were withdrawn from Scotty H cyclinders (99.0% CH4, Alltech Associate, Inc., Deerfield, IL) at a fixed exit pressure and injected into batch vials. Phenol was added to vials from a 40 g/L stock solution. After adding substrates, the vials were vigorously shaken upside-down on a rotary shaker (250 rpm). Headspace samples of TCE and methane were periodically analyzed by GC. Cell solution was filtered with 0.2 um syringe filter. Phenol in the filtrate was analyzed by HPLC. Analytical methods A TCE-saturated water solution was used as the spike solution in all experiments. The spike solution was prepared by adding excess TCE (99+% pure ACS reagent, Aldrich Chemicals Co., Milwankee, WI) to a 250 ml glass bottle capped with TFE-lined Mininert valve. The bottle was vigorously shaken and allowed to settle at least 24 hrs. The upper layer of the solution was transferred to another bottle and capped with a Mininert valve. The spike solution was stored in a refrigerator until needed. One hour before use, it was shaken again and allowed to settle. TCE was analyzed by withdrawing 0.1 ml of headspace from the test bottles using a 0.5 ml Pressure-Lok Series A-2 gas syringe and injecting the samples onto a Hewlett Packard 5890 gas chromatograph (GC) equipped with a capillary column (DB624, 30m x 0.53mm I.D.), a flame ionization detector (FID) and a Electron capture detector (ECD). The GC was operated isothermally at 90°C with helium as carrier (12 mL/min). The injection port was set at 250°C. The temperature of FID and ECD were 250°C and 350°C, respectively. Phenol was analyzed by HPLC. Cell samples were collected by syringes and injected through 0.2 ttm NYLON syringe filters. Filtrate (2 ml) was collected for analysis. Water 153 HPLC (WISP 710B+ Model 510 pump) equipped with a column (Econosil C18, 10 micron, 250 mm, Alltech Cat No. 288138) and a UV detector (Lambda-Max Model 481 LC spectrophotometer) was operated isocratically (60% acetonitrile + 40% water) at a total flow rate 1 mL/min. The wavelength of UV detector was 235 nm and the injection amount of samples were 30 (IL. The limit of detection for phenol was approximately 1 mg/L. Cell biomass was determined on a dry weight basis using 0.2 pm filters (Gelman Sciences Inc., Ann Arbor, MI). The filters were prepared by first soaking them in mineral media for 10 minutes, rinsing on a vacuum filter with deionized water, drying overnight in a 103°C oven, and cooling in a desiccator until needed. The filters were weighed, and once a known amount of culture was filtered through them, they were rinsed, dried, cooled and reweighed. Community Analysis DN ' 'rcin Fifteen milliliter samples were collected from reactors at intervals, and cells were harvested by centrifugation and frozen. Pellets were later resuspended in a buffer consisting of 100 mM Tris, 100 mM EDTA (pH 8.0), 1.5 M NaCl, 1% CTAB. Two milliliters SDS (10%) was added and the cells were incubated at 65°C for 30 minutes in a rotary water bath (150 rpm). Samples were cooled to 37°C, then supplemented with 50 pl proteinase K (10 pig/[.11). The lysates were incubated at 30°C for 2 hours with shaking. DNA was purified by extraction with chloroform/isoamyl alcohol (24:1) and phenol/chloroform/isoamyl alcohol (25:24: 1) and precipitated with cold isopropanol. 154 El r sis Changes in the communities were monitored using DGGE (denaturing gradient gel electrophoresis). Thirty microliter PCR reactions contained approximately 10 ng template DNA, 1U Taq DNA polymerase, 10X buffer, 10% DMSO (v/v), 2mM Mg”, 200 mM dNTPs, and 1 11M of primers GM5F and 907R described by Muyzer, et al. (Muyzer et al. 1995). A CG clamp was added to the 5’ end of GMSF. Products were analyzed on an 8% polyacrylamide gel (37.5 :1 acrylamide/bisacrylamide) containing a 40 -60% gradient of urea and forrnamide. The separationwas achieved at 200 V and a temperature of 60°C. RESULTS For the TCE exposure experiment, TCE was injected over a one-hour interval during the fill period. Phenol was provided in a separate period, following the decant step. Influent TCE levels were increased gradually from 0.5 to 25 mg/L over 6 months (Table 7.1). Changes of cell density in response to long-term TCE exposure are shown in Figure. 7.3. To monitor changes in community structure and function, the following parameters were measured periodically: second order rate coefficient for TCE transformation, maximum specific rate of utilization of phenol, observed yield, biomass transformation capacity, and endogenous decay rate. The history of these parameters during long-term TCE exposure is shown in Figures 7.4 to 7.6 and Table 7.3 to 7.4. Figure 7.7 illustrates fitness of the two communities during the period when the TCE-exposed reactor was exposed to 25 mg/L of TCE feed solution. Typical changes in the apparent second order rate coefficient for TCE transformation during a SBR operating cycle are illustrated in Figure 7.8. A decline in TCE transformation was observed during the fill, react, settle and decant periods. However, recovery of activity was observed during the recharge period. From a mass balance on TCE, more than 95% of the 155 added TCE was removed by microbial degradation and only 3% was removed by air stripping. TCE concentration in gas and liquid phase within the reactor during the fill period are shown in Figure 7.9. Gas and liquid phase approach equilibrium except during the initial fill. One molecular technique - denaturing gradient gel electrophoresis (DGGE) - was used to detect changes in community structure. A small fragment (~400 bp) of the 16S gene is amplified and the products are resolved on an acrylamide gel containing a gradient of urea and forrnamide. The amplified PCR products from different species are of the same size and do not resolve on agarose or acrylamide. However, the different GC content of each fragment allows resolution on the gradient gel as fragments with higher GC content are transported further into the gel before denaturing. Most species will yield one band so a simplified community fingerprint is obtained. Changes in community structure can be detected through the appearance and/or disappearance of bands. DGGE fingerprints from control and TCE-exposed communities are shown in Figure 7.10. 156 TCE Feed concentration (mg/L) 0 10.5, 2.5 1 5 1 25 1800 f f 1 I A 1600 - . 0......” .1 B: 1400 -I Q I l ’I‘ II 5 1200 :EWfigéézo 9' III I I §1000 II °fi .3- .fi'. 2 800 - o '0 600 - = 400 - 8 oControl 20° ’ IExposed 0 l l l l 0 50 100 150 200 250 Time(Days) Figure 7.3 Changes in cell density of the phenol-degrading reactor communities. 157 TCE feed concentration (mg/L) O 10.5 | 2.5 l 5'15 1 25 0.8 l I l T I I one. o . "3 ’° 0’ ,I 5 i 00 o a: 0'5 ' C I I I I E b I ” I .- .‘I. _l 0.4 . I 0.3 - I 3: II' 0.2 - II. 0Control 0'1 ' IExposed O l l l l o 50 100 150 200 250 Time(Days) Figure 7.4 Second order rate coefficients of TCE transformation by phenol-degrading reactor communities. 158 TCE feed concentration (mg/L) 0 .0.512.5 .5-15I 25 06 1 I I J ”53 \ . O 3505 II I : u I a O .3 04 0 °. 0'. .2 r: ' 0’ . I . III :0 I .00 O i- 0: I 0' fit 0 I mg0.3 " .I ‘ a O I I I a c ’00. *II‘ 0 I 5002 P . I - a: J I I I l g“ I. .. I! 0.1 - OControl :1: IExposed E a I l l l o 0 50 100 150 200 250 Time (Days) Figure 7.5 Maximum specific rate of phenol utilization by phenol-degrading reactor communities. Table 7.3. Theoretical transformation capacity for phenol-degrading reactor communities. Time Theoretical transforrnationcapacity, mg fimg cells T,exposure/ (Days) T,control Exposed Control 68 0.366 i0.016 0.345 :i:0.045 1.06 96 0.333 21:0.023 0.324 21:0.043 1.03 174 0.342 $0.055 0.355 i0.044 0.96 189 0.222 $0.031 0.344 21:0.005 0.65 210 0.332 i0.017 0.34 i0.022 0.98 222 0.345 21:0.041 0.351 i0.055 0.98 159 Table 7.4. Endogeneous decay constant for phenol-degrading communities during long- term TCE exposure in SBRs. Time Endogenous decay constant, b,exposure/ r2 for regression (Days) 1/day b,control Exposed Control Exposed Control 68 0.398 0.349 1.14 0.875 0.771 96 0.177 0.202 0.88 0.735 0.668 174 0.369 0.318 1.16 0.948 0.685 189 0.267 0.248 1.08 0.744 0.791 210 0.152 0.367 0.41 0.704 0.762 222 0.314 0.718 0.44 0.998 0.923 g 1.4 3 1.2 - I 15 mg/L TCE initiated 3 o 1 _ J I g c I ' ' . ' v o 0.8 - . . II I C I I E 2 o . I o 0 0.6 r o O o g Q 0” . O. : :0 : o 'o 0.4 - I 0° m I 2 i .o OControl O 0 . . . 160 180 200 220 240 Time (Days) Figure 7.6 Long-term changes in observed yield for the phenol-degrading reactor communities. 160 1.4 -—Ii 1.2 - I p 1 P 3 0.8 - «T to 0.6 - 0 E. E 0.4 - 0.2 - ° 89°39“ l Control 0 ' 15mg/LTCE ____CE""°st°f' initiated °" '° -o.2 . - 160 180 200 220 240 Time (Days) Figure 7.7 Fitness of phenol-degrading communities during period when the exposed reactor received an influent concentration 25 mg/L TCE. Fitness calculations are based on )1 measurements (0 , 0) and fitness components (-—, ---), respectively. 161 Fill React Settle Decant Recharge l l l l l 0.9 I l I rJ 0Control l Exposed .0 m I (leg-d) O O-Il o 03 l-OI'I l—O—ll Apparent second order rate “ C Q ‘5 05 - C . O O OAT 0.3 l l l I L 8 10 12 14 16 18 20 Sampling time (hr) Figure 7.8 Change in the second order rate coefficient for TCE transformation in a typical SBR operating cycle. 162 0.7 OCw(mg/L) 3 0.6 _ O . IICg(mg/L)| \ m 0 E 0 v 0.5 - 0 c o O 2:- 0.4 - I I- E o 0 0.3 - o O c 8 L 0.2 I I I I.l.l I ' U I I- o.1 ~ I it ° ' t I 0 l l l l l l 0 10 20 30 40 50 60 70 Time (min) Figure 7.9 TCE concentration in gas (Cg) and liquid (Cw) phase during the fill and react periods for a feed solution of 25 mg/L TCE. 163 Figure 7.10 DGGE fingerprints from control and TCE-exposed communities in sequencing batch reactors. 164 DISCUSSION The SBR microbial community cometabolized TCE for extended periods without loss of transformation activity. No significant changes in phenotypic parameters occurred when the TCE concentration was less than 10 mg/L. With increased influent TCE concentrations, declines in biomass, observed yield, specific rates of phenol utilization and specific rates of TCE transformation were observed. Theoretical transformation capacity (defined as the mass of nongrowth substrate transformed per unit mass of cells in the absence of endogenous decay) decreased to 60% of the control value (from 0.344+l-0.005 to 0.222+l- 0.031). After one more month of perturbation, theoretical transformation capacity recovered to the original value. Of interest was the appearance of a persistent yellow color after the influent TCE concentration was raised to 15 mg/L. The yellow color may have been due to the accumulation of a-hydroxymuconic semialdehyde, an intermediate of phenol degradation resulting from the incomplete oxidation of phenol (Shih et al. 1996). The endogenous decay coefficient decreased slightly after prolonged TCE exposure. This suggests that the enzymes present in the TCE exposure culture became more resistant to endogenous decay. This may be because of adaptation to TCE or its transformation products. Fitness of phenol-degrading community in SBR was calculated from the ratio of specific rate of phenol utilization at any time to the value just before initiating the 25 mg/L TCE feed. The results show that a decline in fitness occurred just after the initiation of 25 mg/L TCE feed, with subsequently recovery to the original value. The changes in fitness during TCE exposure closely match those of the observed yield, specific rates of phenol utilization, specific rates of TCE transformation and theoretical transformation capacity. These observations suggest that microbial community can adapt to TCE exposure, and that this adaptation can be quantitatively described by fitness. To further demonstrate the fitness ' theory, selection gradients with respect to fitness components (Ym, ks ) and history of these 165 fitness components were used to calculated fitness (Eq.(l) to Eq. (4)). The results indicate fitness can be calculated accurately without knowledge of fitness itself (Figure 7.7). Microscopic examination revealed that the community has a distinctive floc structure of spherical and rod-shaped bacteria. This characteristic is similar to that of the inoculum from the fed-batch pulse reactor (Shih et al. 1996). During the settle period, the TCE-exposed reactor community had a finer floc structure than the control reactor. This difference appeared soon after TCE exposure and persisted after higher TCE exposures were implemented. A possible explanation for lower cell density in the TCE-exposed reactor may be loss of biomass in the decant period because of poor settling properties. The apparent second order rate coefficients declined during the fill, react, settle and decant periods and recovered during the recharge period. For specific growth rate, similar trends were observed. The apparent first order rate coefficient is equal to the true second order coefficient (k;) multiplied by the concentration of organisms (X) that can degrade TCE (X). Therefore, changes in the apparent second order rate coefficient during a operating cycle may correspond to the changes in the concentration of TCE degrading organisms (lower X ) or to loss of reducing power during TCE transformation (lower k). The organisms that were deactivated or killed during TCE transformation were reactivated by regrowth on the phenol. The fact that there was no significant difference in endogenous decay rate between the exposed and control reactor communities suggests that toxicity did not play an important role for this level of TCE exposure (25 mg/L in feed). Changes in the structure of the microbial communities were monitored using denaturing gradient gel electrophoresis (DGGE). Total community DNA was isolated from samples taken from the reactor and regions of the 168 rDNA genes were amplified by PCR using universal primers, and resolved by electrophoresis in an acrylamide gel with a urea and 166 formamide gradient. The control community showed no change over time. The TCE- exposed community was also largely unchanged, with the exception of the appearance of a distinct new band after TCE concentration was increased to 2.5 mg/L (Figure 7.10). This change suggests the gain of a population within a community, or perhaps a shift in dominance in the community structure. Previous research established that communities fed phenol over short intervals were more stable and diverse. This research establishes that such a community is capable of long-term TCE transformation. SBR reactor environments are well suited to cometabolism because they offer a wide variety niches for metabolism of phenol and cometabolism. In this work, the methanotrophic mixed culture chemostat experiment (Chapter 6) and the phenol- degrading SBR community represent two extreme cases for adaptation of cultures. Higher diversity within the SBR system offers a more stable community structure and higher potential for TCE transformation than the chemostat under the ranges tested (1-50 mg TCE/day and 6-18 mg TCE/day, respectively). As discussed previously, it may be possible to use selection gradients (especially more sensitive ones) as criteria for the stability of a community under a given perturbation. The selection gradients with respect to 7; , b , kc, and Kc for methanotrophic mixed culture are higher than those of phenol-degrading community (Chapter 5). Thus, the phenol-degrading community was expected to be more stable than methanotrophic mixed culture. This was confirmed by experimental results. Mass balance data show that over 95% of TCE was removed by microbial degradation and only 3% of TCE was stripped by aeration. TCE concentration in gas and liquid phase within the SBR approached equilibrium. From a process engineering point of view, these data verify that the current SBR design can treat 25 mg/L TCE with satisfactory results. 167 REFERENCES 1.A1varez-Cohen, L., and P. L. McCarty. 1991a. A cometabolic biotransformation model for halogenated aliphatic compounds exhibiting product toxicity. Environ. Sci. Techno]. 25:1381-1387. 2.Alvarez-Cohen, L., and P. L. McCarty. 1991c. Product toxicity and cometabolic competitive inhibition modeling of chloroform and trichloroethylene transformation by methanotrophic resting cells. Appl. Environ. Microbiol. 57:1031 - 1037. 3.Chang, H.-l., and L. Alvarez-Cohen. 1995a. Model for the cometabolic biodegradation of chlorinated organics. Environ. Sci. Technol. 29:2357-2367. 4.Criddle, C. S. 1993. The kinetics of cometabolism. Biotechnol. Bioeng. 41:1048-1056. 5.Folsom, B. R., P. J. Chapman, and P. H. Pritchard. 1990. Phenol and trichloroethylene degradation by Pseudomonas cepacia G4: kinetics and interactions between substrates. Appl. Environ. Microbiol. 56:1279-1285. 6.Harker, A. R., and Y. Kim. 1990. Trichloroethylene degradation by two independent aromatic-degrading pathways in Alcaligenes eutrophus JMP134. Appl. Environ. Microbiol. 56:] 179-1 181. 7.Henry, S. M., and D. Grbic’: -Galié. 1990. Effect of mineral media on trichloroethylene oxidation by aquifer methanotrophs. Microbial Ecology. 20: 151-169. 8.Henry, S. M., and D. Grbié -Galic’:. 1991a. Influence of endogenous and exogenous electron donors and trichloroethylene oxidation toxicity on trichloroethylene oxidation by methanotrophic cultures from a groudwater aquifer. Appl. Environ. Microbiol. 57:236 - 244. 9.Hopkins, G. D., J. Munakata, L. Semprini, and P. L. McCarty. 1993b. Trichloroethylene concentration effects on pilot field-scale in-situ groundwater bioremediation by phenol-oxidizing microorganisms. Environ. Sci. Technol. 27:2542- 2547. 10.Lackey, L. W., T. J. Phelps, V. Korde, S. Nold, D. Ringelberg, P. R. Bienkowski, and D. C. White. 1994. Feasibility testing for the on-site bioremediation of organic wastes by native microbial consortia. International Biodeterioration & Biodegradation. 33:41-59. 11.Little, C. D., A. V. Palumbo, S. E. Herbes, M. E. Lidstrom, R. L. Tyndall, and P. J. Gilmer. 1988. Trichloroethylene biodegradation by a methane-oxidizing bacterium. Appl. Environ. Microbiol. 54:951 - 956. 12.Muyzer, G., A. Teske, and C. O. Wirsen. 1995. Phylogenetic relationships of Thiamicrosopira species and their identification in deep-sea hydrothermal vent samples by denaturing gradient gel electrophoresis of 168 rRNA fragments. Arch. Microbiol. 13.Ne]son, M. J. K., S. 0. Montgomery, W. R. Mahaffey, and P. H. Pritchard. 1987. Biodegradation of trichloroethylene and involvement of an aromatic biodegradative pathway. Appl. Environ. Microbiol. 53:949-954. 168 14.Nelson, M. J. K., S. 0. Montgomery, and P. H. Pritchard. 1988. Trichloroethylene metabolism by microorganisms that degrade aromatic compounds. Appl. Environ. Microbiol. 54:604-606. 15.0]denhuis, R., J. Y. Oedzes, J. J. van der Waarde, and D. B. Janssen. 1991. Kinetics of chlorinated hydrocarbon degradation by Methylosinus trichosporium OB3b and toxicity of trichloroethylene. Appl. Environ. Microbiol. 57:7-14. 16.0]denhuis, R., R. L. J. M. Vink, D. B. Janssen, and B. Witholt. 1989. Degradation of chlorinated aliphatic hydrocarbons by methylosinus trichosporium OB3b Expressing soluble methane monooxygenate. Appl. Environ. Microbiol. 55 :28 19-2826. l7.Shields, M. S., S. 0. Montgomery, P. J. Chapman, S. M. Cuskey, and P. H. Pritchard. 1989. Novel pathway of toluene catabolism in the trichloroethylene-degrading bacterium G4. Appl. Environ. Microbiol. 55: 1624 - 1629. 18.Shih, C.-C., M. Davey, J. Zhou, J. M. Tiedje, and C. S. Criddle. 1996. Effects of phenol feeding pattern on microbial community structure and cometabolism of trichloroethylene. Appl. Environ. Microbiol. 62:2953-2960. 19.Strand, S. E., M. D. Bjelland, and H. D. Stensel. 1990. Kinetics of chlorinated hydrocarbon degradation by suspended cultures of methane—oxidizing bacteria. Research J. WPCF. 62:124-129. 20.Wackett, L. P., and D. T. Gibson. 1988. Degradation of trichloroethylene by toluene dioxygenase in whole-cell studies with Pseudomonas putida F1. Appl. Environ. Microbiol. 54:1703-1708. 21.Wackett, L. P., and S. R. Householder. 1989. Toxicity of trichloroethylene to Pseudomonas putida F1 is mediated by toluene dioxygenase. Appl. Environ. Microbiol. 55:2723-2725. CHAPTER 8 ENGINEERING SIGNIFICANCE Cometabolism is a complicated phenomenon. Several factors interact simultaneously. This presents a challenge for engineering application since rational design requires use of appropriate kinetic expressions. A generally accepted mode] is needed with model parameters that can be evaluated without difficulty. In this research, a model for the most general case was verified. The model also covers a simplied case in which growth and nongrowth substrate transformation occurs separately. Different microbial systems can therefore be evaluated under this same framework. Furthermore, the model can serve as the basis for design of treatment processes and for prediction of treatment efficiency. Cometabolic transformations often generate toxic products. The extent to which these products affect the microorganisms depends upon the type of organisms and the compounds transformed. Toxicity can affect microbial communities at both the population level and the community level. At the population level, specific populations may experience mutation and selection or changes in gene expression. At the community level, the community can be expected to undergo dynamic changes in composition. All of these factors contribute to changes in the phenotype of the community. By analyzing selection gradients, the sensitivity of different parameters on growth can be evaluated. This may suggest means whereby microbial populations or communities can be "designed" or improved. For example, improvements in transformation capacity in cultures that are sensitive to that parameter would enhance the sustainability of cometabolic transformation. 169 170 The patterns of addition of growth and nongrowth substrate can have a great impact on the performance of a system. In this work, two extreme cases were examined. A continuous system with both substrates fed simultaneously, and a SBR with both substrates fed at separate stages. When both substrates are provided simultaneously, growth and nongrowth substrates compete for the same enzymes, growth substrate is used inefficiently, and growth of cells declines. Futherrnore, product toxicity accelerates cell decay. All these factors contribute to the instability of continuous systems fed both substrates simultaneously. On the other hand, sequencing batch reactors have certain advantages. An SBR can alternate between periods of growth on growth substrates and periods of cometabolism of nongrowth substrate, eliminating the possibility of competitive inhibition for enzymes between growth and nongrowth substrates. Since microorganisms are reactivated after the transformation of nongrowth substrate, a steady and sustainable community can be maintained. Thus, SBRs offer a good treatment technology for cometabolic transformation of substance that do not support growth of microorganisms. Stability of microbial communities is an important factor for engineering application. Communities usually undergo dynamic changes in structures and performance in response to perturbation. In this research, nongrowth substrate exposure is the sole perturbation that the communities experience. A stable community should be able to maintain its structure and performance under nongrowth substrate exposure. Based on the results of this study, a diverse community seems to have better stability. It is feasible to use sensitive selection gradients as the criteria of stability of community. The parameter can be used as a standard for comparison between different communities in the same environment or the same communities in different environments. CHAPTER 9 SUMMARY AND CONCLUSIONS CONCLUSIONS 1. An unstructured model for cometabolism is presented and verified experimentally for a defined methanotrophic mixed culture. The model includes the effects of cell growth, endogenous cell decay, product toxicity, and competitive inhibition with the assumption that cometabolic transformation rates are enhanced by reducing power obtained from oxidation of growth substrates. A theoretical transformation yield is used to quantify the enhancement resulting from oxidation. A systematic method for evaluating model parameters is described. The applicability of the mode] is evaluated by comparing experimental data for methanotrophic cometabolism of TCE with model predictions from independently measured model parameters. Propagation of errors is used to quantify errors in parameter estimates and in the final prediction. The model successfully predicts TCE and methane transformation for a wide range of concentrations of TCE (0.5 - 9 mg/L) and methane (0.05 - 6 mg/L). 2. This research investigated the potential for methanotrophic biotransformation of three HCFCs -- chlorodifluoromethane (HCFC-22); 1-chloro—1,1-difluoroethane (HCFC-142b); and l,1-dichloro-2,2,2-trifluoroethane (HCFC-123); and one HFC -- 1,2,2,2- tetrafluoroethane (HFC-134a). All of these compounds were biotransformed to differing degrees by methanotrophic mixed culture MM]. Intrinsic rates of transformation were obtained by combining a second order rate expression with an expression describing loss of transformation activity due to either endogenous decay or product toxicity. For 171 172 HCFC-123 and HFC-134a, the independently measured endogenous decay rate for mixed culture MM] (0.594/day) was sufficient to account for the observed loss of transformation activity with time for the one case examined. However, the endogenous decay rate did not account for the loss of transformation activity for HCFC-22 and HCFC-142b. A model based on product toxicity provided a reasonable representation of the loss of transformation activity for all these compounds. The order of reactivity was HCFC-22 > HCFC-142b > HFC-134a > HCFC-123, with second order rate coefficients of 0.014, 0.0096, 0.00091, and 0.00054 L/mg-day, respectively. Transformation capacities for HCFC-22 and HCFC-142b were 2.47 and 1.1] ug substrate/mg biomass, respectively. 3. Theoretical transformation capacity is a function of organisms and target compounds. For two groups of compounds and two types of organisms studied here, the order of theoretical transformation capacity is phenol degrader/TCE > methanotrophic mixed culture/ TCE > methanotrophic mixed culture/ HCFC/HFC, with typical values of 0.35, 0.06, and 0.002 mg TCE or HCFC/mg cells. This indicates that product toxicity plays a much more important effect on methanotrophic mixed culture with HCFC/HFC transformation, but less important for phenol degrader with TCE transformation. 4. A fitness concept was developed for communities and combined with the cometabolism model to describe adaptation of cometabolizing communities. A major concept of the model is that gross phenotype changes within a microbial community can be quantitified by a " fitness" parameter which can be calculated using kinetic parameters that describe the community. The selection gradient for each parameter is defined by the partial derivative of fitness with respect to that parameter. The gradient, therefore, reflects the direct selection acting on each fitness component, with the other components held 173 constant. It appears possible to use selection gradients as criteria for the stability of a community under a given perturbation. 5. A methanotrophic mixed culture and a phenol degrading culture were exposed to different levels of TCE over extended periods of time. The changes of community were monitored and the effects of TCE exposure on community structure were evaluated. Various phenotypic parameters were measured. Genetic community analysis (ARDRA and DGGE) was also used to monitor shifts in microbial community structure. The results indicate that phenotypic and genetic changes occurred during TCE exposure. Both microbial communities adapted to TCE exposure with improvement in the observed transformation capacities and endogenous decay constants. 6. This research establishes that a phenol-degrading community is capable of long-term TCE transformation. SBR reactor environments are well suited to cometabolism because they offer a wide variety niches for metabolism of phenol and cometabolism. In this work, the methanotrophic mixed culture chemostat experiment and the phenol-degrading SBR community represent two extreme cases for adaptation of cultures. Based on an analysis of the selection gradients for these two communities, the phenol fed SBR community was expected to be more stable than methanotrophic chemostat mixed culture. This was confirmed by experimental results for the ranges tested. FUTURE RESEARCH l. Hyphomicrobia in methanotrophic mixed culture were thought to be possible indicators of MMO activity. However, in this research, no significant change in this population was detectable by image analysis. This may be because an insufficient number of images were taken. Further research is needed to determine the number of images required to accurately quantify a given morphotype. 174 2. SBRs provide flexible operation for cometabolism. Several operating parameters (length of periods, exposure level, growth substrate concentration and extent of aeration) can affect performance. Optimization of SBR systems will assist the engineering application of SBRs. 3. Genetic analyses (ARDRA and DGGE) were used in this research for characterization of community structure. Other methods of community analysis, such as analysis of fatty acid methyl esters (FAME), might also be employed for analysis of cometabolizing communities. 4. Key populations responsible for phenol and TCE transformations in the phenol- degrading community should be isolated and characterized. 5. Use of selection gradients to quantify community stability should be evaluated further. More extensive studies are needed for different communities within the same reactor type and for the same community within the different reactor types. 6. For the range of TCE exposures studied in this work, the phenol-degrading community was always able to adapt to TCE exposure. Further research should be conducted to determine the upper limit of exposure and to assess whether TCE could conceivably be used as a growth substrate. GR STRTE UNIV. HICHI N llllllllllllIll]llllllllllmllll ll] 3 293010 1 5064 LIBRRRIES “MINIMUM“ 112