".— ,1: . ‘44 a" .Jc' V St 5:12.: ’, Him ‘ Jung. ~:’ “ n S . . 1", ii i»- ” ‘IGB ”om- .'.. a: ... ’ “h ‘-I “7"?“ {I v [is if. .. .>-_ a“ \ V} ’, - Mum Eat-“33$? a. m‘ _, ‘;_‘4 «"1”: “LA- ,1, .' ..» 5; .‘.' 713% '3'.- "5'. «ac. .Q .n ‘9 ‘$ y 18“.in u u”. ' w . ‘fo‘li‘lfi. 1‘1",“ ~53! 33‘ 3'13 5 “"rf'w‘ : h“ “A 1‘1 ‘3‘ m.» .u ‘. u . .I‘L: THEE \'\J 565W?“ This is to certify that the dissertation entitled DEVELOPMENT AND CHARACTERIZATION OF A RAPID DECHLORINATING ENRICHMENT STIMULATED FROM AN AQUIFER CONTAMINATED WITH CHLORINATED ETHENES AND ETHANES presented by HAEKYUNG KIM has been accepted towards fulfillment of the requirements for the Ph.D degree in Civil and Environmental Engineering WProfé/sor's Signature May 25, 2004 Date MSU is an Affirmative Action/Equal Opportunity Institution ---u--o-.-o-----0-o-o-u-ncn—I—n-o-COO-O-O-C-I-O->-‘-‘ LIBRARV I Michigan State University PLACE IN RETURN BOX to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 6/01 c;/CIRC/DateDue.p65-p. 15 DEVELOPMENT AND CHARACTERIZATION OF A RAPID DECHLORINATING ENRICHMENT STIMULATED FROM AN AQUEER CONTAMINATED WITH CHLORINATED ETHENES AND ETHANES By Haekyung Kim 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 2004 ABSTRACT DEVELOPMENT AND CHARACTERIZATION OF A RAPID DECHLORINATING EN RICHMENT STINIULATED FROM AN AQUIFER CONTAMINATED WITH CHLORINATED ETHENES AND ETHANES By Haekyung Kim Biostimulation has become an accepted potential remedial alternative for treating contaminated aquifers. This study evaluates a novel approach to enrich for rapid dechlorinating microbes by pulse feedings in a batch system over an extended period. It was demonstrated that by varying the electron donor to acceptor ratio during pulse feedings of chlorinated compounds, two different types of activity could be developed. Cometabolic dechlorination was dominant under high lactate conditions, and dehalorespiration activity was enriched under low lactate conditions. The high lactate/PCB enrichments exhibited slow and incomplete dechlorination of PCB with cDCE as the major degradation product. The low lactate/PCB enrichments completely dechlorinated PCE to ethene, and further reduction of ethene to ethane was observed. Dehalococcoides 16S rRNA gene targeted real —time PCR confirmed that two orders of magnitude higher amounts of Dehalococcoides DNA was present in low lactate enrichments compared to high lactate enrichments. This demonstrated that pulse feedings with low-level electron donor and a chlorinated ethenes (at a 1:10 electron acceptor/donor ratio) favors stimulation of halorespirers. Using metabolic inhibitors revealed that the dechlorination observed in low lactate enrichments was not impacted by molybdate (inhibitor of sulfate-reducer), but was inhibited by high levels of sulfate, indicating sulfate reducers are not involved in the observed complete and rapid dechlorination. Low levels of BES (methanogen inhibitors) specifically inhibited methane productions, but the cultures retained dechlorination activity, indicating that the responsible microorganism(s) is (are) non-methanogenic. High levels of vancomycin (100 mg/L; acetogens inhibitors) successfiilly inhibited acetate production from lactate. The reduction in acetate production also inhibited dechlorination, flirther suggesting that acetate-utilizing dechlorinating microbes are responsible for the observed dechlorination. The development of methodology for the stimulation of dehalorespiring microorganisms presented in this dissertation represents a potential strategy for the design of successful biostirnulation systems for chlorinated solvents. It also has significant design implications for in—situ field systems. Copyright by HAEKYUNG KIM 2004 DEDICATION This dissertation is dedicated to my family and friends: my parents, NamYeun Chang and KiTae Kim; my sister and brother, SoAh Kim and ByungSoo Kim; and my biggest supporters, Y.A, TL, and AJN. No matter where I go in life, you all are in my heart and prayers, now and for always. I could not have done this without you. ACKNOWLEDGEMENTS I would like to express my appreciation to my advisor, Michael Dybas, for his encouragement and support. I am also grateful for my committee, Syed Hashsham, James Tiedje, and David Wiggert for their comments and guidance during this process. Also to, Davis Mckenzie, my mentor, for providing unconditional support and advise whenever I needed. Special thanks are extended to Lesile Dybas, for introducing me to the DNA extraction, Yan Pan and Xianda Zhao for answering every question I tossed at them. This work was fiinded by MDEQ Contract. Y403 86. TABLE OF CONTENTS LIST OF TABLES ........................................................................................................ x LIST OF FIGURES .................................................................................................... xiii LIST OF ABBREVIATIONS ..................................................................................... xvi CHAPTER 1 PROBLEM STATEMENT AND OBJECTIVES ............................................................ l 1.1 Hypothesis and research objective .................................................................... 2 1.2 Literature cited ................................................................................................. 4 CHAPTER 2 BACKGROUND AND THEORY ................................................................................... 5 2.1 Energetic Considerations ................................................................................. 5 2.2 Microbial Competition for Hydrogen ................................................................ 7 2.3 Design of Experiments ................................................................................... 11 2.4 Literature cited .............................................................................................. 13 CHAPTER 3 LITERATURE REVIEW ............................................................................................. 15 3.1 Chlorinated Solvents ...................................................................................... 15 3.2 Biotransformation of Chlorinated Ethenes ..................................................... 17 3.3 Mechanism of Chloroethenes Transformation ............................................... 17 3.4 Kinetics of Dechlorination ............................................................................. 20 3.5 Electron Donors for Reductive Dechlorination ............................................... 22 3.6 Methanogens vs. Sulfate Reducers ................................................................. 23 3.7 Abiotic Degradation of Chlorinated Compounds ............................................ 25 3.8 Literature cited .............................................................................................. 37 CHAPTER 4 MATERIALS AND METHODS ................................................................................. 43 4.1 Chemicals and Reagents ............................................................................... 43 4.2 Description of Site and Sediments ................................................................. 43 4.3 Microcosms .................................................................................................. 45 4.4 Analysis of Volatile Organics ........................................................................ 48 4.5 Analysis of Volatile Fatty Acids .................................................................... 49 vii 4.6 Anion Analysis ............................................................................................. 50 4.7 Hydrogen Analysis ........................................................................................ 50 4.8 Hydrogen Sulfide Analysis ............................................................................ 51 4.9 Data Reduction ............................................................................................ 51 4.10 DNA Analysis .............................................................................................. 53 4.11 Literature cited ............................................................................................. 56 CHAPTER 5 EVALUATION OF EFFECT OF ELECTRON DONORS ON DECHLORINATION ................................................................................................... 58 5.1 Abstract ........................................................................................................ 58 5.2 Introduction .................................................................................................. 59 5.3 Materials and Methods ................................................................................... 62 5.4 Results ........................................................................................................... 65 5.5 Conclusions and Discussions ......................................................................... 78 5.6 Literature cited .............................................................................................. 82 CHAPTER 6 STIMULATION OF HIGH-RATE COMPLETE DECHLORINATION ACTIVITY BY PULSE FEEDINGS OF CHLOROETHENES ..................................................................................................... 85 6.1 Abstract ........................................................................................................ 85 6.2 Introduction .................................................................................................. 86 6.3 Materials and Methods ................................................................................... 88 6.4 Results ............................................................................................................ 92 6.5 Conclusions and Discussions ....................................................................... 116 6.6 Literature cited ............................................................................................ 120 CHAPTER 7 CHARACTERIZATION OF DECHLORINATING COMMUNITY BY METABOLIC INHIBITORS ..................................................... 122 7.1 Abstract ...................................................................................................... 122 7.2 Introduction ................................................................................................ 123 7.3 Materials and Methods ................................................................................. 125 7.4 Results .......................................................................................................... 128 7.5 Conclusions and Discussions ....................................................................... 145 7.6 Literature cited ............................................................................................ 149 CHAPTER 8 ENGINEERING APPLICATIONS ............................................................................ 151 8.1 Enrichment of Rapid and Complete Dechlorinating Activity ............................................................................... 151 viii 8.2 Field Application ............................................................................................ 154 8.3 Literature cited ................................................................................................. 160 CHAPTER 9 CONCLUSIONS AND FUTURE RESEARCH ........................................................... 161 9.1 Conclusions ..................................................................................................... 161 9.2 Future Research .............................................................................................. 162 APPENDIX A — ELECTRON DONOR AMOUNT CALCULATIONS .............................................................................. l 64 APPENDIX B — EFFECT OF ADDED SULFATE ON THE POWDER X-RAY DIFFRACTION PATTERN ................................. 168 APPENDIX C — DESIGN OF EXPERIMENT ............................................................. 169 APPENDIX D -— ADDITIONS AND CONSUMPTION FOR THE STIMULATIONS OF ENRICHMENTS FOR EXPERIMENT II ................................................................................ 172 APPENDIX E — T-RFLP PATTERNS FROM CHAPTER 6 ...................................... 174 APPENDIX F — T-RFLP PATTERNS FROM CHAPTER 7 ....................................... 176 LIST OF TABLES Table 2.1. Reduction potentials of redox couples found in anaerobic dechlorinating environments ........................................................ 6 Table 2.2. Maximum free energies (AG°’) for electron donors commonly used in dechlorination of chlorinated compounds ................................................................................................... 7 Table 2.3. Values of half-velocity constants K8 and maximum specific utilization rate k with respect to hydrogen, for hydrogenotrophic PCE dechlorinators and methanogens ................................ 9 Table 2.4. Threshold H2 concentrations and energetics of H2 utilizing Reactions ...................................................................................... 10 Table 2.5. Sign table for 2k factorial design ................................................................... 12 Table 3.1. Properties of common chlorinated organic compounds ................................ 16 Table 3.2. Rate coefficients for various mixed and pure cultures in dechlorination ............................................................................................. 21 Table 3.3. Summary of research efforts regarding the degradation of chlorinated ethenes and ethanes in biotic and abiotic conditions. ................................................................................................ 33 Table 4.1. Typical contaminants found from the study area ......................................... 44 Table 4.2. Typical chemical composition of G plume site ........................................... 45 Table 4.3. Basal medium for microcosm ...................................................................... 46 Table 5.1. Effect of substrates on transformation of Chloroethenes afier 240 days of incubation in sulfate-reducing conditions .................................................................................................... 69 Table 5.2. Reduction of TCE depending on presence of reductant after 180 days of incubation at 15°C ............................................................ 72 Table 5.3. The byproducts and mass recoveries for c-DCE and Table 5.4. Table 6.1. Table 6.2. Table 6.3. Table 6.4. Table 6.5. Table 6.6. Table 6.7. Table 6.8. Table 7.1. Table 7.2. Table 7.3. Table 7 .4. TCE transformation after 180 days ............................................................... 74 Estimated transformation rate and half-time of TCE and c-DCE biodegradation in Plume G .............................................................. 77 Experiment arrangements for respective conditions. Microcosms were exposed to pulse feedings of PCB over 1.5 years for Experiment I and pulse feedings of TCE, c-DCE, and TCA over 1 year for Experiment 11. ................................ 90 Effects of prolonged exposure of high versus low lactate on enhancing transformation activity (after 12 month of incubation). ................................................................................ 93 Electron balance for a single, PCE degrading microcosm (15 umol PCE /bottle, 12.5 mM (2500 umol) lactate; data obtained at day 10) . ..................................................... 95 Electron balance for a single, PCE degrading microcosm (15.2 umol PCE /bottle, 0.5 mM, lactate; data obtained at day 10) ............................................................................. 99 Summary of dechlorination rate observed by low lactate/PCB, low lactate/T CE, and low lactate/c-DCE enrichments (/day). ................................................................................... 105 Effects of prolonged exposure of TCA on stimulation of transformation activity . ............................................................................ 107 Mass balance determined for PCE transformation to ethane after 60 days .................................................................................. 109 Quantitative estimation on total DNA and Dehalococcoides populations in pre-enrichment, Low/PCB, High/PCB, c-DCE, TCE, and TCA enriched bottles using PicoGreen and RTm-PCR ..................................................... l 15 The list of inhibitors and their concentrations used in this chapter ................................................................................................ 127 Effect of inhibitors on reductive dechlorination of PCB after 18 days .............................................................................................. 138 Influences of inhibitors on electron balance after 7 days ............................ 139 Quantitative Estimation of Total DNA and xi Dehalococcoides populations in the BES, molybdate, sulfate, and vancomycin impacted microcosms using PicoGreen and RTm-PCR ......................................................................... 144 Table 8.1. Enhancement of first order rate k (1/day) on disappearance of parent products by pulse feedings over time ........................................................................................................... 152 Table 8.2. Comparisons of maximum rate between the pulse fed and once-fed enrichments in enhancing dechlorination activity observed fi'om each condition. .................................................... 154 Table 8.3. Levels of Dehalococcoides DNA in the biostirnuation and bioaugmentation test sites ................................................................... 157 xii LIST OF FIGURES Figure 3.1. Pathway of microbial PCE transformation under anaerobic conditions ................................................................................... 18 Figure 4.1. Location of study area ................................................................................ 47 Figure 5.]. Powder X-ray diffraction pattern by experimentally-prepared FeS .................................................................... 64 Figure 5.2. Differences in (a) slow vs (b) rapid T CE transformation due to electron donors during 240 days. ............................. 66 Figure 5.3. The differences in dechlorination of TCE after 240 days expressed as percentage removal of abiotic soil control ......................................................................................................... 67 Figure 5.4. The differences in dechlorination of c-DCE after 240 days expressed as percent removal of abiotic soil control. ...................................................................................................... 68 Figure 5.5. Anion degradation in CMB fed microcosm ................................................ 70 Figure 5.6. Effect of sulfate on transformation of TCE in CMB fed microcosm. ................................................................................................. 71 Figure 5.7. Powder x-ray diffraction pattern by iron sulfide produced in microcosm. ............................................................................. 75 Figure 5.8. Possible pathways for the reductive dechlorination of chlorinated ethenes by iron sulfide ............................................................ 81 Figure 6.1. The overview and sequence of the experiments performed during enrichments .................................................................. 91 Figure 6.2. PCB additions and consumption by high lactate/PCB treatment over 1.5 years ........................................................................... 92 Figure 6.3. Effects of prolonged exposure of high lactate (12.5 leI) on enhancing transformation activity (after 12 month of incubation). ............................................................................. 93 xiii Figure 6.4. Figure 6.5. Figure 6.6. Figure 6.7. Figure 6.8. Figure 6.9. Figure 6.10. Figure 7.1. Figure 7 .2. Figure 7.3. Figure 7.4. Figure 7.5. PCB additions and consumption by low lactate/PCB treatment during 450 days .......................................... Transformation of PCB and intermediate productions from lactate/PCB enrichments received high dosage of PCB (80 umol). ...................................................... Conversion of chlorinated ethenes to ethene by lactate/PCB enrichment. This experiment was done with the microcosms showing highest activity. Approximately 15 umol of chlorinated ethenes and 0.5 mM of lactate were used: (a) PCE; (b) TCE; (c) c- DCE; (d) t-DCE; (e) 1,1-DCE; and (f) VC. ............... Conversion of chlorinated ethanes by low lactate/TCA enrichment. Individual serum bottles have received 15 umol of chlorinated ethanes and lactate of 0.5 mM: (a) TCA and (b) 1,1-DCA. .......... Temperature effects on PCE transformation on low lactate/PCB enrichment. Individual serum bottles have received 6 pmol of PCB and lactate of 0.5 mM. Final PCE transformation byproducts by low lactate/PCB enrichments Temperature effects on PCE transformation on low lactate/PCB enrichment .......... T-RFLP profiles generated using pre-enrichment, low lactate/PCB, and high lactate/PCB enrichments using HhaI. ........................................................................ Chemical Structures of inhibitors used in this study. The PCE transformation by control. ........................... The effects of PCB transformation by 0.5 mM 5 mM, and 50 mM of BES. Microcosms were amended at 0.5 mM, 5 mM, and 50 mM of BES at zero time only The effects of PCB transformation by 0.5 mM 2 mM, and 6 mM of molybdate. Microcosms were amended at 0.5 mM, 2 mM, and 6 mM of molybdate at zero time only .................................................................... The effects of PCB transformation by 0.6 mM and 2 mM of sulfate. Microcosms were amended with xiv .............................. 97 ............................ 101 ............................ 102 ............................ 106 ........................... 108 ............................ 110 ............................ 112 . ............................ 124 ............................ 129 ............................ 130 ............................ 133 sulfate at time zero only ......................................................................... 134 Figure 7.6. Sulfate reduction and effect of molybdate on residual sulfate in microcosms. The reduction of high sulfate (2 mM) is shown on the bigger scale. ................................................... 135 Figure 7.7. The effects of PCB transformation by 25 mg/L and 100 mg/L of vancomycin. Microcosms were amended with vancomycin at time zero only ......................................................... 137 Figure 7.8. T-RFLP profiles generated by using dechlorinating enrichments treated with inhibitors. Each fragment size indicates percent fragment area: (a) BES; (b) molybdate; (c) sulfate; and (d) vancomycin. .......................................... 141 Figure 7.9. Model based on DiStefano et al (2), for carbon and electron flow in a lactate-PCB anaerobic enrichment .............................. 148 Figure 8.1. A Schematic of the relationship between fermentators and Dehalococcoides ............................................................................. 153 Figure 8.2 Layout of the tracer injection system ....................................................... 158 Figure 8.3. Comparison on the observed reduction in chlorinated compounds in the biostimuation and bioaugmentation test sites ................................................................................................. 159 ATP BES CA c-DCE t-DCE 1, l-DCE 1, l, l-TCA 1 , l-DCA 1 ,2-DCA DNA EPA PCE PCR rRNA TCE TOC T-RFLP VC VOC LIST OF ABBREVIATIONS Adenosine Triphosphate 2-Bromoethanesulfonic acid Chloroethane cis-l, 2 Dichloroethene trans-1,2 Dichloroethene Dichloroethene 1, 1, l-Trichloroethane 1 , l-Dichloroethane 1 ,2-Dichloroethane Deoxyribonucleic acid Environmental Protection Agency Perchloroethene Polymerase chain reaction Ribosomal Ribonucleic acid Trichloroethene Total Organic Carbon Terminal-Restriction Fragment Length Polymorphism Vinyl Chloride Volatile Organic Compound Chapter One: PROBLEM STATEMENT AND OBJECTIVES Due to past agricultural practices, accidental chemical spills, and inappropriate disposal of industrial and commercial wastes, groundwater contamination is a significant national and international problem. In the United States, for example, groundwater contamination can be found in 1200 Superfund sites and an estimated 7000 Department of Defense sites (EPA, 1993). The most ubiquitous class of organic compounds polluting groundwater is the chlorinated solvents (EPA, 1998). Traditional approaches to water containing these contaminants have relied on groundwater extraction, followed by a physical/chemical process (e. g., air stripping and carbon adsorption). This approach has some major disadvantages such as high cost, inefficiencies in removing contaminants sorbed to the aquifer material, and requirement of pumping, treating, and ultimately disposal of large volumes of water. Accordingly, there has been a great deal of interest in in-situ alternative treatment strategies, many of which incorporate biological remediation of contaminants. The three basic bioremediation approaches are natural attenuation, biostirnulation, and bioaugmentation. Natural attenuation refers to unenhanced “natural” processes, which may occur at contaminated sites if sufficient nutrients and degrading microbes are present. This can include both abiotic and biotic mechanisms. For chlorinated organics, natural attenuation may be insufficient, and electron donor addition is often needed to accelerate reductive dechlorination or stimulate cometabolic activities. Biostimulation refers to enhancing the metabolic activity of indigenous microflora to transform target compounds and usually requires amendments (electron donors and/or acceptors) to enhance the microbial populations already present. In some cases, indigenous bacteria may be unable to mediate the desired transformation (Mayotte et al. 1996; Dybas et al. 1998; Ellis et al. 1999) and specific microorganisms with the desired traits need to be provided (bioaugmentation). Although bioaugmentation offers the promise of increased control over transformation of a specific compound, competition between indigenous microflora and the introduced organisms usually presents a challenge. Factors such as the nutritional requirements of dehalogenating microbes and electron donor competition between dechlorinating and non-dechlorinating microorganisms have presented challenges to implementation of this promising technology in the field. The research presented in this thesis provides insight into the issue of selective biostimulation: that is, how to enrich microbial populations that are capable of complete and rapid degradation of contaminant. 1.1. Hypothesis and objectives The hypotheses of this research are: l. Palmitate, a slow releasing carbon source, will enhance the rate of dechlorination compared to simple, readily available carbon sources. 2. Relative abundance of electron donor with respect to PCE will influence the level of dehalogenation populations such as Dehalococcoides ethenogenes, will impact rates, and will determine the extent of transformation. A lower abundance of electron donor with respect to PCE will produce a higher yield of Dehalococcoides, which results in higher degradation rates and complete transformation. The specific objectives were: 1. Evaluate the effect of various types of nutrient on stimulation of dechlorination. 2. Evaluate the effect of various chlorinated compounds and 2 different levels of lactate on stimulation of dechlorinating enrichments. 3. Characterize the mixed dechlorinating community by the inhibitors and hydrogen effects on dechlorination and growth. 4. Characterize the mixed dechlorinating community by using the molecular techniques such as RTm-PCR and T-RF LP. 5. Evaluate the abiotic degradation of chlorinated compounds. The format of this thesis is as follows: theory supporting the aforementioned hypotheses are outlined in Chapter 2; Chapter 3 contains review of related researches; materials and methods are illustrated in Chapter 4; results from the screening of the various electron donors and byproduct analysis presented in Chapter 5; development strategies and complete dechlorination by enriched sediments are presented in Chapter 6; and effects of metabolic inhibitors on dechlorinating communities in Chapter 7; engineering applications in Chapter 8; conclusions and future research are summarized in Chapter 9. 1.2. Literature cited US. Environmental Protection Agency. 1993. Cleaning Up the Nation’s Waste Sites: Markets and Technology Trends. Office of Solid Waste and Emergency Response, EPA 542-R—92-012. US. Environmental Protection Agency. 1998. National Water Quality. Inventory Report to Congress. Mayotte, T., M. Dybas, and C. S. Criddle. 1996. Bench-scale evaluation of bioaugmentation to remediate carbon tetrachloride-contaminated aquifer materials. Ground Water 34 (2), 358-367. Dybas, M. 1., Barcelona, M., Bezborodnikov, 8., Davies, S., Fomey, L., Heuer, H., Kawka, O., Mayotte, T., Sepulveda-Torres, L., Smalla, K., Sneathen, M., Tiedje, J ., Voice, T., Wiggert, D. C., Witt, M. E., and C. S. Criddle. 1998. Pilot-Scale Evaluation of Bioaugmentation for In-Situ Remediation of a Carbon Tetrachloride-Contaminated Aquifer. Environ. Sci. Technol. 32(22): 3598-3611. Ellis, D.E., E.J. Lutz, R.J. Buchanan, M.D. Lee, C.L., Bartlett, M. Harkness, and K. DeWeerd. 1999. Bioaugmentation for accelerated in situ anaerobic bioremediation. In Platform Abstracts of the Fifth International Symposium on In-Situ and On-Site Bioremediation, Session D2, San Diego, California, April 19-22. Columbus, Ohio: Battelle Chapter Two: BACKGROUND AND THEORY 2.1 Energetic Considerations Microorganisms that are capable of utilizing the most energy rich electron acceptor/donor pair (the half reactions exhibiting the most positive electron EH°) will be dominant in a given environment. These organisms then in turn determine the redox conditions of the environment. The maximum free energy (AG°') microorganisms may gain is commonly presented by reducing-oxidation reactions. Table 2.1 contains standard reducing potentials and free energies (AG°') of redox couples and chlorinated compounds that are commonly found in contaminated groundwater at 25 °C. Thermodynamic consideration indicates that reductive dechlorination is exergonic and chlorinated compounds can be used as electron acceptors. Table 2.2 contains maximum free energies of selected organic acids and alcohols commonly used as electron donors in dechlorinating experiments. McCarty (1969) proposed stoichiometric methodology to evaluate the energy requirement for cell synthesis from various carbon and nitrogen sources. AG /k’"+AGc+AGn/k A=—— P (2.1) kAG f where: A = electron equivalent of electron donor to energy/ electron equivalent of cells synthesized : AGp represents the energy required to reduce carbon source to pyruvate : k is cell efficiency of energy transfer : AGn represents energy required to convert N source to NH] : AGr presents energy available energy per electron equivalent of substrate converted for energy : AGc presents energy of conversion of one electron equivalent of intermediate to one electron equivalent of cells. Table 2.1 Reduction Potentials of redox couples found in anaerobic dechlorinating environments. Eo’+ AG°’ Half-Reaction (V) (kJ /mol e) Fe3+ + e' = Fe2+ 0.77 -114 $0424 9H+ + 8e° = HS' + 4 H20 022 -146.7 C020,) + SPF + 8e' = CH4 (9 + 2 H20 025 -131.2 C2C14 + H*+ 2e‘ = C2HCI3 + Cl' 0.58 -191.0 C2HC13 + H+ + 2e‘ = C2H2C12+ Cl' 0.54 -183.0 C2H2C12 + H+ + 2e‘ = C2H3CI + Cl' 0.37 -151.1 C2H3Cl + I—F+ 2e‘ = C2H4 + Cl’ 0.49 -173.7 C2H4+ 2H+ + 2e' = C2H6 0.1 -98.9 C2H3Cl3 + H+ + 2e‘ = C2H4C12 + Cl' 0.57 -189.1 C2H4C12 + H+ + 2e' = C2H5Cl + Cl' 0.51 -177.5 I Redox-potentials were obtained from Criddle et al. (1991), Dolfing and Janssen (1994), and Vogel et al. (1987). In heterotrophic reactions, the value of A is by definition feis so that the fraction of organic substrate utilized for energy is A =_ 2.2 A+1 ( ) fe where: fs is fraction of electron donor synthesized and fe is fraction of electron donors converted to energy. Table 2.2. Maximum free energies (AG°’) for electron donors commonly used in dechlorination of chlorinated compoundsa AG°’ kcal . per mole Half-Reactions electrons 1/92CH3(CH2)14COO‘ + 31/92 H20 —+ 15/92 C02+1/92 HCO3‘+ IF+ e' -6657 Palmitate 1/12 C3H503' + 1/3 H209 1/6 C02 + 1/12 HCOg' + I-F + e' -7.873 Lactate 1/14 C3H502'+ 5/14 H20 —> 1/7 C02 + 1/14 HC03'+ W + e‘ -6.664 Propionate 1/6 CH40 +1/6 H20 —-> 1/6 C02 + W+e‘ -8.965 Methanol aEnergetics and Bacterial Growth. Perry L. McCarty. Fifth Rudolf Research Conference, Rutgers. The State University, New Jersey, 1969. July 2. 2.2 Microbial Competition for Hydrogen The ability of certain microorganisms to out compete others for the same grth nutrient is dependent upon both the kinetics and energetics of the competing reactions. When microbial strains compete for the same limiting nutrient, resource-based competition theory predicts that only one group of organisms will survive and all other will die out (Hansen et a1, 1980). Experimental results have shown that two species were observed to coexist only if either (1) each was limited by a different resource and met the theoretical criteria for coexistence, or (2) the species were limited by the same resource and did not differ significantly in their resource requirements. The rate of bacterial growth (rg) and the rate of substrate consumption (rsc) can be used to determine whether one organism has a kinetic advantage over another. The rate of bacterial grth is described mathematically by Monod’s equations. XS eta-:3- <2.” Where; um (hr'l) is the organism’s maximum specific growth rate; X represents the concentration of biomass (mg cells L'l) ;S is the concentration of the growth-limiting substrate (mgL'l) ; Ks is the half velocity constant which is the substrate concentration at half the maximum specific growth rate. The rate of substrate consumption, rsc is related to the rate of bacterial growth by r. = — (2.4) The organism’s maximum yield coefficient (Y) represents the maximum mass of cells that can be produced per mass of substrate consumed. The Y and mm define the maximum specific substrate utilization rate k (mg substrate consumed mg cells"1 h'l ) according to the relationship k = umax/Y . Combining equation 2.3 into equation 2.4 yield the following expression for rSC r50 = kXS (2.5) K, + S Examination of the Monod’s expression reveals that the rates of bacterial growth and substrate consumption are dependent upon values of Ks, Y, and k. Half-velocity rate constants represent the affinity of an organism for a substrate, and lower values of K8 are associated with higher affinities. Many organisms like acetogens, methanogens, and sulfidogens use hydrogen produced from anaerobic metabolism of organic matters (Brock et al. 1999). Since hydrogen is also needed for the hydrogenotrophic organisms (such as hydrogen-utilizing PCE dechlorinators) to survive, the microbial competition for hydrogen can play an important role in the transformation of the chlorinated compounds. Values of half- velocity constants with respect to hydrogen, Ks (Hg), for hydrogenotrophic dechlorinators and methanogens and measurements of the maximum specific utilization rate reported are summarized are listed in Table 2.3. For hydrogenotrophic dechlorinators, half-velocity Table 2.3. Values of half-velocity constants K8 and maximum specific utilization rate k with respect to hydrogen, for hydrogenotrophic PCE dechlorinators and methanogens. K5 (H2) k Methanogens 2,500 —l3,000 nMa loo-2,600Pmol-mgprotein 'lhr" c 960 i 180 M 1339111011113me 100 i 50 an 20 umol-mgprotein 'lhr'” 9-21 nM° ’ Robinson and Tiedje, 1984 " Smatlak et al. 1996 ° Ballapragada et al. 1997 constants were found to be one-tenth that of the methanogenic organisms (Smatlak et a1. 1996). The use of fermentable substrates that maintain low levels hydrogen was suggested to offer a competitive advantage for hydrogen utilizing dechlorinators. The energetics of competing reactions can also be used to determine whether the competition will emerge between organisms utilizing the same electron donor. Cord- Ruwisch et al. (1988) observed that the hydrogen threshold is inversely correlated with changes in Gibbs free energy (AG°'). As can be seen in Table 2.4, since dechlorination has lower free energy compared to other processes, dechlorinators are expected to grow at lower hydrogen partial pressure. Supporting observations had been made in many studies (Smatlak et al. 1996; Ballapragada et al. 1997). Table 2.4. Threshold H2 Concentrations and Energetics of H2 Utilizing Reactions (adopted from Loffler et al. 1999). Biological . Threshold H2 AG°’ Equation process (nM)' (KJ/mol H2) Dehalogenation PCE + 4H2 —) ethene +I—I+ + Cl’ < 0.3 -130 to -187 Methanogenesis C02 + 4H2 —> CH4 + 2H20 5 — 95 -33.9 Acetogenesis 2CO2 + 4H2 —) CH3COO' + H+ + 2H20 336- 3640 -26.1 Sulfate reduction 8042' + 4H2 +I-I+ —> HS' + H2O 1-15 -3 8.0 ’ In aqueous phase concentration 10 2.3 Design of Experiments Significant factors Many factors contribute to transformation of chemicals and it can be difficult to separate the contribution of any one factor to the overall degradation. Reported data on PCB and TCE transformation often does not include consideration of any role of abiotic mechanisms, assuming transformation occurred only biologically with microorganisms. To attempt to determine which factors are most significant, a statistical design approach (factorial design) is adopted which allows separation of significant factors from insignificant factors and thus the effects of interactions between each factor can be evaluated (Box et al. 1978). Design Matrix The main objective of factorial design is to see the effects of two or more independent variables (Wadsworth, 1990). It is usually more efficient to manipulate these variables in one experiment than run a separate experiment for each variable. Also it is possible to test for interactions among variables using factorial design. 2" factorial designs need k parameters and helps in sorting out parameters of impact. The method uses a nonlinear regression model and assuming there are only 2 parameters, x1, x2, gives: y=q0+q1*XI+QZ*X2+(112*X1*X2 (2.4) 11 Here, y is the system performance metric; qo is the average performance; q,- is the effect on the performance of parameter x,; q12 is the effect of the interaction of x1 and x2; The value of x,- is defined as:-1 if x, assumes the minimum value in the experiment, otherwise, +1. Expenment ........... .x.l. x2 xlxé y . 0-1 -1 1.y.0. ............... 1........... . . 1_1’_'1 yfll 2 -1W1._1“ yz 3 1 1 1. y3 Table 2.5. Sign table for 2k factorial designs In this example, 22 experiments with 4 different combination of x1,x2 are needed. Then, by substituting x, and y into the above performance equation, 4 equations with 4 unknowns qo, q1, q2, q12 are obtained. Then, we can solve for qo, q1, q2, q12 unknowns. Analysis of Variance (AN OVA) In order to further verify the degradation of Chloroethenes from these preliminary results, collected data will be analyzed using (1) Qalysis gf ygriance (ANOVA), which tests the difference between 2 or more means by examing the ratio of variability between two conditions and variability within each condition; and (2) factorial design, which identifies positive and negative factors on degradation of Chloroethenes. Values of the F 12 statistics and and p-value will be computed to determine the significance of main and interaction effects. 2.4. Literature cited McCarty, Perry L. 1969. Energetics and Bacterial Growth. The Fifth Rudolf Research Conference, Rutgers, New Brunswick, New Jersey Criddle, CS. and PL. McCarty. 1991. Electrolytic model system for reductive dehalogenation in aqueous environments. Environ. Sci. T echnol. 25:973 -978 Dolfmg J. and DB. Janssen. 1994. Estimates of Gibbs free energies of formation of chlorinated aliphatic compounds. Biodegradation 5: 21-28 Vogel, T.M., C.S. Criddle, and PL. McCarty. 1987. Transformation of halogenated aliphatic compounds. Environ. Sci. T echnol. 21: 722-736. Hansen, S. R., and S. P. Hubbell. 1980. Single-nutrient microbial competition: qualitative agreement between experimental and theoretically forecast outcomes. Science. 207: 1491— 1493. Brock, T. D., Madigan, M. T., Martinko, J. M., and Parker, J. 1999. Biology of Microorganisms, Prentice-Hall, Inc, Englewood Cliffs. Smatlak, C. R., Gossett, J. M., and Zinder, S. H. 1996. Comparative kinetics of hydrogen utilization for reductive dechlorination of tetrachloroethene and methanogenesis in an anaerobic enrichment culture. Environ. Sci. Techno]. 30 (9): 2850- 2858. Ballapragada, B. S., Stensel, H. D., Puhakka, J. A., and Ferguson, J. F. 1997. Effect of hydrogen on reductive dechlorination of chlorinated ethenes. Environ. Sci. Technol. 31(6): 1728-1734. Robinson, J .A., and Tiedje, J.M. 1984. Competition between sulfate-reducing and methanogenic bacterial for H2 under resting and growing conditions. Archives of Microbiology. 137 : 26-32. Cord-Ruwisch, R, H.-J. Seitz, and R. Conrad. 1988. The capacity of hydrogenotrophic anaerobic bacteria to compete for traces of hydrogen depends on the redox potential of the terminal electron acceptor. Arch. Microbiol. 149:3 50-357 . Loffler F. E., James M. Tiedje, and Robert A. Sanford. 1999. Appl. Environ. Microbiol. 65(9): 4049-4056 13 Distefano, T. D., Gossett, J. M., and Zinder, S. H. 1992. Hydrogen as an electron-donor for dechlorination of Tetrachloroethene by an anaerobic mixed culture. App]. Environ. Microbiol. 58 (11): 3622-3629. Monod, J. 1949. ”The growth of bacterial cultures.” Annual Review of Microbiology. Box, GE. 1978. Statistics for Experimenters: An introduction to design, data analysis, and model building. John Wiley & Sons. Harrison M. Wadswroth. 1990. Handbook of statistical methods for engineers and scientist. McCraw-Hill Publishing Company. 14 Chapter Three: LITERATURE REVIEW This chapter provides previous research efforts on biotransformation of chlorinated ethenes under anaerobic conditions. The review begins with a general overview of chlorinated ethenes properties. The remaining review is divided into two sections. The first contains discussions on mechanisms and transformation kinetics of chlorinated solvents. A summary of previous research efforts on biotransformation of chlorinated ethenes follows shortly after. 3.1. Chlorinated Solvents Due to the similarities in the electrical charges (or electronegativites) of oxygen and chlorine, chlorinated solvents are considered highly oxidized. Perchloroethylene (PCE), the most highly chlorinated ethylene molecule, is environmentally persistent partly because it only degrades anaerobically. PCE degradation proceeds through trichloroethylene (TCE), the TCE is then reduced to isomers of dichloroethylene. Although the cis-isomer (c-DCE) appears to be the predominantly produced dechlorination byproduct, trans-isomer (t-DCE) is also reported to be produced by D. ethenogenes 195 (Maymo-Gatell et al. 1999). A third isomer, 1, l-dichloroethylene (1,1- DCE) is chemically possible, however, there is little evidence that it is biologically produced from TCE dechlorination. The next step of reductive dechlorination is the production of vinyl chloride (VC) from the dichloroethylene isomers. Removal of the final chlorine produces the innocuous ethylene molecule making this product the end point of the reductive dechlorination process, which satisfies the treatment purpose. Note 15 that because of the double bond in the ethylene molecule can be further reduced to ethane, which contains only a single bond between the carbons with three hydrogens on each carbon. All of the chlorinated ethylene are listed as priority pollutants under the Safe Drinking Water Act Amendments of 1986. Typical chemical and physical properties of these compounds are summarized in Table 3.1. Table 3.1. Properties of Common Chlorinated Organic Compounds Chemical Molecular Density Solubility in Henry’s Compound Formula Weight (m g/m’)’ water b Constant Hcc (g/mol) (mM at 25 °C) (dimensronless) Tricthlrblethane CHCl3 133.41 1350 8.5 0.703 Tetrachloroethene C2Cl4 165.83 1626 0.91 0.723 Trichloroethene C2HC13 13 1.39 1460 8 .4 03” c-Dichloroethene C2H2Cl2 96.94 1284 3 6 O. 167 Vinyl Chloride C2H3Cl 62.50 91 1 43 1.137 Ethene C2H4 28.04 2085 4.7 8.5b ‘ Data taken from Sawyer and McCarty (1994). bYaws, C. L. (1999) ° Gossett, J. M. (1987) at 24.8 °C 16 3.2. Biotransformation of Chlorinated Ethenes Reductive Dechlorination Reductive dechlorination or hydrogenolysis is a term to describe reactions where sequential electron and hydrogen additions replace chlorine atoms. Chlorinated ethenes are susceptible to biologically mediated reductive dechlorination reactions under anaerobic conditions. For example, PCE is sequentially reduced to TCE, c-DCE or t- DCE, VC, and then to ethene (Figure 3.1). The extent of reaction varies for different compounds and with differing environmental conditions. Generally more chlorinated compounds are more susceptible to rapid reductive dechlorination reactions while the more toxic intermediates such as DCE, VC are less readily degraded (Haston & McCarty, 1999; Vogel et al. 1987). These less chlorinated intermediates are susceptible to oxidation under aerobic and iron reducing conditions but PCE is only removed by anaerobically (Freedman and Gossett, 1989). 3.3. Mechanism of Chloroethenes Transformation: Halorespiration vs. Cometabolism Most of these transformations were explained as cometabolism, which are fortuitous transformation of chemicals by enzymes or cofactors produced by organisms for other purposes. During the cometabolic process, microorganisms that are responsible for catalyzing such reactions do not obtain nutritional or energetic benefits. Researchers showed that the majority of chlorinated compounds could undergo cometabolic transformation to less chlorinated products and in some instances complete dechlorinations to benign products in environments. Reductive dechlorination has been reported to occur cometabolically under various redox environments including 17 denitrifying, sulfate reducing, fermentating, acetogenic, and methanogenic conditions (Fetzner & Lingens, 1994; Mohn & Tiedje, 1992). C1 C1 PCE Cl L C1 C1 C1 >:.< TCE Cl H A/ \ C1 C1 H Cl H H Cl H cis-DCE trans-DCE \ A/ H H H: :H Ethene H H H H M Ethane H H H H Figure 3.1 Pathway of microbial PCE transformation under anaerobic conditions. 18 However, recent research has indicated that in addition to cometabolic processes, anaerobic bacteria may degrade chlorinated ethenes by using them as an electron acceptor, which couples chlorinated ethenes directly to energy production (McCarty, 1997). In the absence of 02, many anaerobes use electron acceptors such as nitrate, ferric iron, sulfate, protons, and CO2. Reductive dechlorination has been shown to occur in all of the above mentioned alternative electron accepting environments, although it is more frequent in the environments with the lower redox potentials (F etzner & Lingens, 1994; Mohn & Tiedje, 1992; Vogel et a]. 1987). Due to similarity in redox potentials between chlorinated compounds to NO3-/NO2- redox couple, chlorinated compounds have the thermodynamic potential to serve as terminal electron acceptors in anaerobic microorganisms. Currently, only a few bacteria have been isolated that use chlorinated compounds as electron acceptors and conserve energy for growth. Holliger et al. (1993) have described Dehalobacter restrictus, a highly purified enrichment culture that is able to grow by the reduction of PCB to c-DCE using hydrogen as the electron donor. This bacterium was able to grow on a limited number of electron donors, especially; hydrogen and formate, and could use only PCB and TCE as electron acceptors with carbon dioxide and yeast extract acting as carbon sources. Five additional isolates, all capable of utilizing PCE as their terminal electron acceptors, have since been isolated: Dehalaspirillum multivorans, strain TEA, Desulfirromonas chloroethenica, Desulfitobacterium sp. strain PCB], and Dehalococcoides ethenogens strain 195. Dehalosporillium multivorans was able to grow on a range of electron donors including pyruvate, formate, hydrogen, lactate, ethanol and glycerol (Neumann et a], 1994). It was further demonstrated that only PCB and fiimarate l9 could act as electron acceptors for this organism. Strain MS-l has also been shown to dechlorinate PCE in pure culture, but PCE dependent growth has not been confirmed. Another organism capable of the incomplete dechlorination of PCB to c-DCE was the freshwater anaerobe TT4B, which used acetate or pyruvate as electron donor and only PCE or fumarate as an electron acceptor (Krumholz eta]. 1996). Of these isolates, only Dehalacoccodes ethenogenes strain 195 catalyzes the complete dechlorination of PCB to ethene. Dehalococcoides ethenogenes was extracted from anaerobic digester supernatant and grown on hydrogen and PCB using carbon dioxide and yeast extract as carbon sources (Maymo-Gatell et al. 1997). It is reported that the growth of strain 195 was resistant to ampicillin and vancomycin and its cell wall did not react with a peptidoglycan —specific lectin and its ultra structure resembled S- layers of Archaea (Maymo-Gatell et al. 1997). Based on its 16S ribosomal DNA sequence, strain 195 clustered phylogenetically with the eubacteria but did not fall in any of the previously described eubacterial branches (Maymo-Gatell et a]. 1997). The final step of VC transformation to ethene was found to occur only after PCE was completely degraded and cometabolic. No other known microorganisms capable of dechlorination beyond c-DCE have been reported so far. 3.4. Kinetics of Dechlorination As reductive dechlorination proceeds, successive steps are increasingly slow and the conversion of VC to ethene has been observed to be a rate—limiting step in complete degradation (Freedman & Gossett, 1989). Studies of the kinetics of chlorinated ethene biodegradation have shown that DCE and VC have higher half velocity coefficients and 20 lower maximum transformation rates than their parent compounds so that their rate of degradation will be slower, especially at low concentrations (Haston and McCarty, 1999). The kinetics of dechlorination has been determined in both mixed and pure cultures, although inconsistencies in the manner the dechlorination rates were determined have made comparisons between studies difficult. Table 3.2 lists rate coefficients for various cultures and electron donors used to maintain the dechlorination activity. Table 3.2. Rate coefficients for various mixed and pure cultures in dechlorination Electron Rate of dechlorination Reference Donor (5) Culture Dechlorination Methanasarcina PCE —) TCE Methanol 3.5 pmol-mgprotein ' Fathepure et 1 . hr sp. and acetate al. (1987) Sulfate-reducing TCE —) c-DCE Sodium 213 p.rnol-L"d‘l Pavlstathis & enrichment lactate and Zhuang acetate (1991) Enrichment PCE —) ethene ' Methanol 275 umol-L'Id'l Distefano et culture a]. (1991) Fixed-bed PCE -—> ethane Lactate 3,7 pmolL'Ihr'l DeBruin et al. column (1992) Anaerobic C-DCE —) Glucose 4.2 timolg'l VSSd'l Komatsu et al. sewage sludge — ethene, ethane (1994) Batch Aquifer solids PCE —-) TCE Mixture of 0,3 nMd'1 Gibson et al. from polluted TCE _> c-DCE lactate, (1994) sediment — Batch acetate and propionate Laboratory PCE —) VC Methanol 4.6 i 0.4 pmolg“ Tandoi er al. culture VSSd" (1994) Anaerobic TCE —> DCE Glucose 149 ter'1 Wild et al. Reactor treating (1995) dechloromethane 21 Dehalaspirillum PCE—> c-DCE multivarans Strain MS-l PCE —) c-DCE Enriched PCE ——) c-DCE contaminated soil sample Enrichment from c-DCE —) polluted sediment ethene - Batch Pyruvate 50 nmolmin'l (mg cell protein ‘1) Glucose, 0.5 umol'mgCell 'lhr'1 pyruvate, formate, lactate, acetate, yeast extract, amino acid Citrate, 0.4 umol-mg VSS'lhr'1 pyruvate, succinate, formate, and acetate with hydrogen H2 l6uMd" Scholz et a]. (1995) Shanna & McCarty (1996) Lee et al. (1997) Luijten et al. (1997) 3.5. Electron Donors for Reductive Dechlorination A wide variety of fatty acids, alcohols, sugars and other compounds have been successfully demonstrated to support reductive dechlorination, and current researches focus on finding a particular electron donor that best support mixed dechlorinating anaerobes. Researchers reported that hydrogen is a key electron donor and organic compounds are mainly precursors to supply the needed hydrogen via fermentation (Distefano et a]. 1992; Fennell et al. 1997). However, several studies have suggested the ability to enhance and sustain in situ dechlorination can be influenced by the selection of electron donor. Accordingly, many different types of electron donors were screened to stimulate and sustain dechlorination activity. Some researchers focused on hydrogen at 22 high partial pressures, where methanogens and other hydrogenotrophs may compete with dechlorinating microorganisms for hydrogen, resulting in incomplete dechlorination activity or its exclusion entirely (Ballapragada et a]. 1997; Yang & McCarty, 1998). Meanwhile other researcher reported that the difference in the hydrogen partial pressure fail to represent long-term dechlorination activity (F ennell et al. 1997). 3.6. Methanogens vs. Sulfate Reducers Anaerobic conditions encompass a range of redox potentials, depending on the terminal electron acceptors. Reductive dechlorination has been observed repeatedly in both natural and laboratory methanogenic environments (F athepure et a]. 1987; Fathepure & Boyd, 1988; DiStefano et a]. 1992), under nitrate reducing conditions (Bouwer & McCarty, 1983; Jafvert & Wolfe, 1987), and in sulfate reducing conditions (Egli et al. 1987; Bagley & Gossett, 1990). Many reports on dechlorination in sulfate reducing conditions followed (Kohring et a]. 1989; Stevens et al. 1988), but for most compounds, dechlorination occurred more readily when methanogenesis predominates (Bouwer & Wright, 1988; Gibson & Suflita, 1986). Therefore, early works on transformation of chloroethene were dominated in methanogenic mixed cultures or pure cultures of specific methanogens. However, many cases reported interactions/inhibitions in dechlorinations between microorganisms. Bacterial such as Methanosarcina sp. and Methanosarcina mazei could convert PCE to TCE with acetate as an electron donor and this conversion was found to be faster in mixed cultures and was inhibited when methanogensis was inhibited (Fathepure et al. 1987). 23 Freedman and Gossett (1989) reported a mixed culture capable of completely reducing PCE to ethene did experience inhibition when methanogenesis was inhibited, however, PCE degradation was only temporarily halted and did resume even with repeated additions of a methanogenesis inhibitor bromoethanesulfonic acid (BES). The same inhibitor was also found to prevent ethane production, but have no effect on other steps in reductive dechlorination of PCE to ethene in column studies (deBruin et al.1992). In other experiments with a methanol enrichment culture, high concentrations of PCB were reductively dechlorinated to ethene and methanogenic activity decreased as PCE loading and subsequent dechlorination increased (DiStefano et al. 1991). While these studies suggested that methanogens are still involved in partial PCE or TCE reductive dechlorination and may be necessary members of bacterial populations of complete reductive dechlorination, other works found that PCB and TCE could be degraded in the complete absence of methanogenesis. Bagley and Gossett (1990), reported that PCE degradation to TCE and c-DCE was enhanced in sulfate reducing cultures when methanogenesis was inhibited. There are still no published reports of complete degradation of PCB to ethene under sulfate reducing conditions in mixed cultures. But pure cultures of Desulfovibrio desulfirricans are reported to degrade PCE without any other productions of dichloroethene (Fathepure et a]. 1987). In addition, sulfidogenic bacterium DCB-l (Shelton & Tiedje, 1984; Stevens et al. 1988) has been shown to dechlorinate PCE to TCE, and Desulfobacterium autotrophicum reduces PCE quantitatively to TCE and DCE (Egli et al. 1987), indicating the existence of reductive dechlorination capabilities in sufidogenic and sulfate-reducing bacteria. The research efforts on biotransformation of chlorinated ethenes are summarized in Table 3 .3. 24 3.7. Abiotic degradation of chlorinated compounds The biological reductive dechlorination of PCB, TCE to ethene and ethane has been demonstrated at a number of sites. However, biological activity often leads to incomplete or slow degradation of a specific compound. This is generally due to a deficiency in an electron donor or acceptor, other nutrients required for the dechlorinating microbial population, and/or absence of appropriate dechlorinating microbial populations. FeS has been identified as a soil precipitate in sulfate-reducing environments where it is formed through the biologically mediated reduction of sulfate to sulfide and subsequent reaction of sulfide with available iron species (Rickard 1969; Freney 1979). Several forms of ferrous sulfide can be produced; amorphous iron sulfide, makinawite (F eo,995-1,o23S), greigite (Fe3S4), pyrrohotitie (F es”), and pyrite (F eS2). These soil minerals are reported to transform halogenated organic compounds. (Magnetite: Sivavec and Horney 1997; McCormick et al. 1998), pyrite (Kriegman-King & Reinhard, 1991 and 1994). Butler and Hayes (1999) have demonstrated that ferric sulfides and ferric disulfides such as mackinwite and pyrite can promote the abiotic dechlorination of Chloroethenes. The major byproduct of was acetylene after 120 days. Lee and Batchelor (2002) also reported that pyrite degraded TCE to 3.3% aDCE, 43% acetylene, 2.2% ethene, and 50 % residual TCE after 32 days. The research efforts on abiotic transformation of chlorinated ethenes are also summarized in Table 3.3. 25 Table 3.3. Summary of Research Efforts regarding the degradation of chlorinated ethenes and ethanes in biotic and abiotic conditions. (a) Electron donors and source of inoculums in methanogenic mixed and pure cultures. Source of Electron Donor (s) Summary Reference Inoculum and Dechlorination Methanogenic None, except Degradation of PCB and TCE in anaerobic Bouwer and mixed culture from methanol used in and aerobic cultures. Both compounds were McCarty laboratory PCB and TCE recalcitrant under aerobic conditions and (1981) digester solutions appreciable degradation under anaerobic conditions was not observed within 16 PCE —> DCE weeks. TCE —> DCE Same as above Acetate Under methanogenic conditions, PCB (100 Bouwer and < ug/L) was reductively dechlorinated to McCarty PCE —> TCE TCE. Addition of BES causes a reduction (1983) in acetate utilization, but did not affect the extent of PCB dechlorination The role of acetoclasts in PCE dechlorination remained uncertain Soil from a TCE Soybean meal With radiolabeled TCE, authors proved that Kleopfer er contaminated site it reduced to 1, 2-DCE isomers in sail a]. (1985) PCE —> DCE microcosms. Laboratory culture Acetate PCE dechlorinate to TCE, DCEs, and VC in Vogel and fixed film methanogenic columns. 24 % of McCarty PCE -—> DCE the initial PCE added was reported to be (1985) mineralized to CO2 Pure cultures of Methanol and Pure cultures of anaerobic bacteria capable F athepure, acetoclastic acetate 0f dechlorinating PCE were identified. Nengu, and methanogens DCB-l, a chlorobenzoate fed anaerobe, and Boyd PCE —> TCE two strains of Methanosarcina dechlorinated (1987) PCE. DCB-l stoichiometrically converted PCE to TCE. When DCB-l was mixed with the two Methanosarcina strains, higher rates of PCB dechlorination were achieved and TCE degraded. 26 Sewer sludge and acetoclastic methanogens in pure culture Purecultureof Methanosarcina sp. Strain DCM Laboratory anaerobic digester seeded with digested sludge from NY wastewater treatment plant Aquifer solids Laboratory culture used by Freedman and Gossett Methanol and acetate PCE —) TCE Methanol, acetate, methylamine, and trirnethylamine PCE —) TCE Methanol, hydrogen, formate, acetate, and glucose PCE —> ethene Toluene PCE -> DCE Methanol PCE —> DCE BES significantly inhibited PCE dechlorination and methanogensis. Acetoclastic methanogens, Methanosarcina sp. and Ms. mazei were shown to dechlorinate PCE in pure cultures. Relationships may exist between PCE dechlorination and methanogesis. Methanosarcina sp. Strain DCM as shown to dechlorinate PCE to TCE while growing on methanol, acetate, methylamine and trimethylamine. PCE dechlorination was observed only during mathanogenesis, and was found to be contingent upon methanol utilization. Authors suggest that a reduced electron carrier diverted electrons generated during methane synthesis to PCE, and that stimulating methanogenesis could enhance PCE dechlorination. First publication of the complete dechlorination of PCB to ethene in mixed cultures. Under methanogenic conditions, the reduction of VC to ethene was found to be the rate-limiting step in the dechlorination process. Hydrogen, formate, acetate, and glucose were found to sustain dechlorination, although methanol was the most effective. TCE dechlorination and methanogenesis were inhibited in cultures amended with BES. BES did not immediately inhibit PCE dechlorination, but resulted in the accumulation of TCE and 1,2-DCEs. Authors concluded that methanogens may have played a key role in PCE dechlorination, yet emphasized that complete dechlorination was only occurring in mixed cultures. Metabolism of toluene was found to create an initial source of reducing equivalents for the dechlorination of PCE. Either benzoate or acetate, both toluene metabolites, served as the final electron donor. PCE (55mg/L) was degraded to ethene ( DCE Rhine River Lactate sediment and anaerobic granular PCE —> Ethane sludge absence of methanogenesis. This observation was inconsistent with previous work, which indicated that methanogens participated in PCE dechlorination Authors concluded that other organisms, potentially acetogens, played a role in PCE dechlorination Lactate- and ethanol-fed cultures Gibson and dechlorinated PCE (5 mg/L) with smaller Sewell lag times than did cultures fed butyrate, (1992) crotonate, or propionate. Dechlorination activity was not enriched using acetate, isopropanol, or methanol. Unlike the other substrates tested, methanol and acetate usually do not produce large amormts of hydrogen during their anaerobic metabolism. Therefore, the authors concluded that hydrogen, generated through alcohol and fatty acid metabolism, was the final electron donor in PCE dechlorination. Hydrogen has been proposed as the direct DiStefano, electron donor, derived from the Gossett, and fermentation or breakdown of other more Zinder complex compounds. Hydrogen was found (1992) to sustain dechlorination in an anaerobic enrichment culture for a period of 14 to 40 days. Dechlorination beyond 40 days could not be sustained without the addition if nutritional factors from a culture containing the same inoculum but fed methanol. Acetogenesis was inhibited in both the methanol- and hydrogen-fed cultures by vancomycin, and eubacterial inhibitor of cell wall synthesis. In the presence of vancomycin, dechlorination was inhibited in only the methanol-fed cultures. Authors concluded that dechlorination in the methanol-fed cultures was sustained by hydrogen produced during acetogenesis of methanol. BES was found to inhibit dechlorination in cultures fed methanol and hydrogen, suggesting that hydrogen- utilizing methanogens, not acetogens, were possibly the dechlorination microorganism. PCE (1.5 ppm) was dechlorinate to ethene DeBruin er and ethane in a fixed-bed column. When al. (1992) BES inhibited methanogenesis, ethene was not reduced to ethane. Both river sediment and sludge were need for complete dechlorination 28 PER-K23 (Dehalobacter restrictus) Aquifer solids from Traverse City, MI Enrichment culture seeded with D. tiedjei DCB-l Laboratory culture used by Freedman and Gossett Anaerobic digested sewage sludge Hydrogen and forrnate PCE -—> Ethane (?) Mixtures of lactate, acetate, and propionate PCE —> DCE 3-chlorobenzoate PCE —> DCE Methanol PCE —> VC Glucose, yeast extract, propionate, hydrogen, methanol and acetate A gram-negative bacterium (PER-K23) was insolated from the inoculum from DeBruin et a1 (1992). First isolate to couple the dechlorination of PCE to c-DCE to growth. Growth on PCE or TCE could only be sustained using hydrogen or formate. Biomass and dechlorination were found account for all of the electrons generated from hydrogen and forrnate utilization Hollinger, Schraa, Stams, and Zehnder (1993) A mixture of fatty acids, at three different concentrations, was fed in conjunction with 5 mg/L, PCE to microcosms containing aquifer solids. The amount of PE dechlorinated was similar regardless of fatty acid concentration, although higher substrate concentrations led to shorter lag periods. A zero order rate constant of 0.3 uM-day“ was calculated for PCB and TCE dechlorination TCE and the DCE isomers were the reduced end products formed. The authors concluded that butyrate oxidation supported dechlorination activity since it preceded the onset of CE reduction to TCE. Gibson, Robertson, Russell, and Sewell (1994) F athepure and Tiedje (1994) Biofilm reactor packed with an enrichment culture containing Desulfomonile tiedjei DCB-l was shown to dechlorinate PCE to cis- and trans-DCE. Acetate, methanol, glucose, and benzoate could not replace 3- chlorobenzoate (3-CB) as electron donor. Dependence of the culture on 3-CB indicated that strain DCB-l of a similar microorganism was involved in dechlorination. The maximum PCE dechlorination rte was 10.3 umol-L“.hr". Tandoi et al. (1994) Anaerobic enrichment culture degraded 53 mg/L PCE to VC within 20 hours. The PCE degradation rate was calculated to be 4.6 i 0.4 mol PCEmg VSS'loday'l. VC dechlorination was inhibited by the presence of PCE. Zero order kinetics were used to describe PCE, TCE, cis-DCE and 1,1-DCE reductions to VC. Trans-DCE conversion to VC was modeled using first order kinetics. Cis-DCE dechlorination in anaerobic Komatsu er culture was supported by glucose, yeast al. extract, propionate, and to a lesser extent by ( 1994) hydrogen The rate of cis- dechlorination was slower in methanol-fed cultures than in 29 Dehalaspirillum multivorans D. multivarans Laboratory culture used by Freedman and Gossett Butyrate enrichment culture seeded from Distefano et al. (1991) Strain MS-l c-DCE—a ethene, ethane Pyruvate, hydrogen, and forrnate PCE —> c-DCE Pyruvate and PCB PCE —> DCE TCE —> DCE Hydrogen PCE —+ VC, ethene Hydrogen and formate PCE -> VC Glucose, pyruvate, formate, lactate, acetate, yeast extract, amino acids cultures in which no electron donor was added. This implied that methanol might have had an inhibitory effect on dechlorination Dehalaspirillum multivorans, an anaerobe capable of utilizing PCE as its terminal electron acceptor, dechlorinated PCE to cis- DCE. Pyruvate, formate, and hydrogen could serve as electron donors. Furnarate was found to be an alternative terminal electron acceptor. The enzyme PCE dehalogenase was recovered in cell free extracts. D. multivarans grew in defined medium with PCB and H2 as sole energy sources and acetate as carbon source. Alternatively to PCE, fumarate and nitrate could serve as electron acceptors: sulfate could not. The organism utilized a variety of electron donors for dechlorination (pyruvate, lactate, ethanol, formate, glycerol, H2). Using the anaerobic enrichment culture developed by (Freedman and Gossett), a hydrogen-utilizing culture capable of growing on PCE was isolated The hydrogen-PCE culture was capable of dechlorinating PCE to VC and ethene, but required supplements of vitamin 812, supematant from an anaerobic digester sludge, and acetate as a source of carbon. Methanol and acetate could not replace hydrogen as an electron donor, and the culture did not produce methane or acetate. This suggests that the hydrogen-PCE culture did not contain methanogens or acetogens. Dechlorinating organisms are able to utilize lower concentrations of hydrogen than competing bacteria such as methanogens. A facultative aerobe capable of dechlorinating PCE at 0.5 umol PCE.hr' l.mg (dry weight) cell“1 was isolated from a site in Victoria, TX, contaminated with PCE. PCE was dechlorinated to cis-DCE. 3O Neumarm er al. (1994) Scholz- Muramatsu- Heidrun et al. ( 1 995) Maymo- Gatell, Tandoi, Gossett, and Zinder (1995) Smatlak, Gossett, and Zinder (1996) Sharma and McCarty (1996) Strain TF4B Laboratory methanogenic consortium Laboratory culture Contaminated soil sample PCE —-> c-DCE Acetate or pyruvate PCE —> c-DCE Lactate, acetate, hydrogen, and propionate PCE —> TCE VC, ethene Butyric acid, ethanol, lactic acid, propionic acid PCE —> DCE Citrate, pyruvate, succinate, formate, acetate, and acetate with hydrogen PCE —> DCE The presence of oxygen and nitrate inhibited dechlorination, suggesting that strain MS-l would only dechlorinate when more thermodynamically favorable terminal electron acceptors were absent. Numerous electron donors could sustain dechlorination, although high concentrations of fermentable compounds were inhibitory. Characteristics of strain MS-l were found to be very similar to the Enterobacteriaceae family. Enterobacter agglomerans was also shown to dechlorinate PCE to c-DCE. Strain TT4B was isolated from stream Krumholz, sediments contaminated with TCE and sharp, and toluene. The isolate dechlorinated PCE to Fishbain c-DCE, and was found to use other terminal (1996) electron acceptors (TCE, fumarate, and ferric nitroacetate). Higher dechlorination rates were observed Ballapragada to correspond to higher hydrogen partial et al. pressures. Dechlorinators were found to (1997) have an advantage in competing with methanogens for hydrogen (K values for hydrogen uptake by dechlorinators were reported to be 12-28 ppm). Four electron donors were compared for F ennell er their ability to stimulate and sustain al. (1997) dechlorination in short and long-term studies. It was found that for butyrate and propionate, which are fermented at low hydrogen partial pressures of 10-3.5 and 10- 4.4 at respectively and ethanol and lactic acid fermented at partial pressures orders of magnitude higher, long term differences were negligible. In all cases, increased electron donor levels resulted in greater ethene production Demonstrated that dechlorinating bacteria Lee at al. are relatively abundant in nature by (1997) showing that cultures from both contaminated and uncontaminated sites could dechlorinate PCE (nominal concentration of 10 mg/L) within two weeks. PCE dechlorinating activity increased with increasing PCE concentration up to 150 mg/L nominal concentration (58 rug/L in aqueous phase). 31 Sediments taken from several contaminated sites Dehalococcoides ethenogenes strain 195 Laboratory enrichment culture seeded from an anaerobic upflow sludge blanket; sediments from a contaminated site in Texas Laboratory culture Methanol, lactate, acetate, and sucrose PCE —> TCE Hydrogen PCE —+ ethene Methanol, lactate, and hydrogen PCE -> VC, ethene Butyric acid, ethanol, lactic acid, and propionic acid PCE —> c-DCE, VC The optimum values for pH and temperature were 7.0 and 30°C, respectively. Yeast extract was required to maintain activity, and all electron donors tested supported dechlorination. Dechlorinating rate was determined to be 0.4 umol PCE-mg VSS'I-hr'1 Tested the ability of several electron donors to enrich PCE dechlorination in contaminated sediments. Long lag-times were observed before PCE dechlorination commenced in all systems. Lactate was the only electron donor that was able to support dechlorination in more than one of the sediments tested. Lactate also supported greater extents of dechlorination. Gao, Skeen, Hooker, and Quesenberry (1997) An anaerobic bacterium was isolated that had the capability to dechlorinate PCE to ethene. Strain 195 coupled the reduction of PCB with growth on hydrogen, and required the addition of anaerobic digester sludge supernatant for growth. This is the first report of an isolate with the capability of reducing PCE completely to ethene. Maymo- Gatell et al. (1997) Carr and Hughes (1998) A high-rate PCE dechlorinating culture was enriched from a culture with no previous exposure to chlorinated ethenes. The ability to enrich and sustain dechlorination activity with various electron donors was tested with this culture either alone or with a (1:1) mixture containing a second culture derived from contaminated aquifer sediments. Over extended periods of time (approx. 430 days), it was demonstrated that similar rates and extents of PCB dechlorination could be achieved regardless of electron donor fed or inoculum used. Later studies demonstrated that PCE dechlorination could be sustained at high hydrogen partial pressures despite the presence of an actively methanogenic community. Fennell and Gossett (1998) Using data obtained from a laboratory, a model was developed to predict the formation of hydrogen from the fermentation of various substrates coupled with the competition for hydrogen among methanogens and dechlorinating bacteria Model simulations suggested that 32 Aquifer material Benzoate and from a PCE- propionate contaminated groundwater site in c-DCE —> ETH Victoria, TX Aquifer material Benzoate and from a PCE- propionate contaminated grormdwater site in Victoria, TX TCE-contaminated Glucose aquifer PCE —> VC compensating for competition of electron donor by adding excess donor eventually led to failure of dechlorination and the development of a predominantly methanogenic population. Demonstrated that in batch studies Yang and containing benzoate and c-DCE, McCarty dechlorinators maintained hydrogen (1998) concentrations at levels below that which support methanogenesis. Cultures fed propionate, which is fermented to hydrogen at slower rates than benzoate, had 100% conversion to ethene rather than 73% in benzoate-fed cultures. Dechlorinators had higher hydrogen utilization efficiencies in continuous flow, completely mixed systems than in batch reactors, indicating that different approaches may be used to favor dechlorination over other competing microbial processes. The maximum degradation rates and half- Haston and velocity coefficients for PCE, TCE, c-DCE, McCarty and VC dechlorination were determined in (1999) batch cultures. Degradation rates were highest with PCB (77 uM/day), and degradation rates for c-DCE and VC were similar (14 uM/day, 13 uM/day, respectively). Half-velocity coefficients for PCE, TCE, c-DCE, and VC were 0.11, 1.4, 3.3, and 2.6 uM, respectively. Authors concluded that the common observation of slow or incomplete dechlorination to the level of c-DCE and/or VC could be partially explained by kinetics. Rates of PCE and TCE dechlorination were Nielsen and shown to increase with increasing PCB and Keasing TCE concentrations, including (1999) concentrations near the solubility limit. Under saturating conditions (i.e., DNAPL present), PCB and TCE dechlorination did not result in the buildup of VC, potentially due to less competition for electron an electron donor. PCE dechlorination was shown to follow Monod kinetics, and TCE and VC dechlorination were first-order. 33 (b) Electron Donors and source of inoculums in sulfate-reducing mixed and pure cultures. Source of Electron Donor (5) Summary Reference Inoculum Soil sediments Acetate Trace halogenated aliphatics were Bouwer transformed under denitrifying, sulfate— and Wright reducrng, and methanogemc COIldlthIlS. (1 98 8) Laboratory culture Lactate PCE dechlorinated to TCE and c-DCE Bagley and under sulfate-reducing conditions. Extent Gossett PCE —+ DCE and rate of PCB dechlorination was less (1990) TCE _) DCE than that observed in mixed methanogenic cultures. Neither lactate, acetate, methanol, isobutyric acid, valeric acid, isovaleric acid, hexanoic acid, succinic acid, nor hydrogen appeared directly to support tetrachloroethene dechlorination, although lactate-fed inocula demonstrated longer- term dechlorinating capability. Enrichments from Sodium lactate and TCE reductively dechlorinated to c-DCE by Pavlostathis contaminated acetate sulfate-reducing. The highest observed & Zhuang subsurface soil transformation rate of TCE was 213 (1991) TCE —> c-DCE micromoles 1'1 per day at 35°C. No further dechlorination of c-DCE. A decrease in the rate and extent of TCE transformation was observed with an increase in the concentration of bromoethanesulfonate up to 50 mM. Desulfitabacreriu Lactate, pyruvate, Anaerobic bacterium was isolated from a Genitse er m sp. Strain PCEl butyrate, formate, PCE dechlorinating enrichment culture a]. (1996) succinate, and containing sulfate reducers and acetogens. ethanol Desulfitobacterium could utilize PCE, 2- chlorophenol, 2, 4, 6-trichlorophenol, PCE —> DCE fumarate, sulfite, thiosulfate, and 3-chloro- 4-hydroxy-phenylactetate as terminal electron acceptors. PCE was dechlorinated to TCE, and c-DCE and t-DCE. Hydrogen was found to inhibit dechlorination. Laboratory Acetate (1 mM) The rate of transformation and the by- DeBest et Enriclunent products depended on the concentrations of al. (1997) 1,1,1-TCA —9 TCA, acetate and sulfate. Both packed-bed DC A and reactor studies and batch experiments with DC A _) unknown BEA and molybdate demonstrated the products involvement of methanogens and sulfate- reducing bacteria in transformation of TCA. 34 Subsurface soils PCE-enrichment PCE contaminated soil Methanol, lactate, acetate and sucrose PCE —> TCE Yeast extract, sodium lactate PCE —> TCE PCE —-> c-DCE, TCE Various levels of sulfate-reducing, acetogenic, ferrnentative, and methanogenic activity were observed in all sediments. PCE dechlorination was detected in all microcosms, but the amount of dehalogenation varied by several orders of magnitude. Lactate-amended microcosms showed large amounts of dehalogenation but elevated levels of dehalogenation were not consistently associated with any observable anaerobic metabolisms. A methanogenic and sulfate-reducing consortium showed that dehalogenation was due to the direct activity of methanogens. In the presence of sulfate, methanogenesis and dechlorination decreased because of interspecific competition, probably between the H2-oxidizing methanogenic and sulfate- reducing bacteria in batch conditions. PCE was transformed to c-DCE and TCE under sulfate-reducing conditions using acetate, lactate, and methanol as electron donors. Biotransformation of PCB was dependent upon the type of electron donor used. Gao- J ianwei et al(1997) Cabnoler al(1998) Ndon et a]. (2000) 35 (c) Summary of dechlorination by iron-reducing or iron related abiotic mechanism. Source Dechlorination Summary Reference Anaerobic aquifer Vinyl chloride First demonstration of vinyl chloride can be Bradley sediments added as oxidized to CO2 under iron (ED-reducing and substrate conditions. Transformation was dependent Chapelle upon the bioavailability of Fe (III), which (1996) was added as Fe-EDTA. VC —> C02 Creek bed c-DCE and VC c-DCE and VC mineralization were observed Bradley sediments obtained added as under both methanogenic and Fe (III)- and near discharge of substrates reducing conditions. In the methanogenic Chapelle contaminated microcosms, 5% to 44% of 1“VC and 4% to (1997) groundwater C-DCE, VC_> 14% of 14DCE were recovered as 1"C02. C02 Under iron reducing conditions, the recovery of 14co2 from labeled vc was twice that of the methanogenic microcosms, and similar for labeled c-DCE. The kinetics of DCE and VC mineralization varied between the two compounds: DCE was modeled using first order kinetics, and VC was modeled with Michaelis-Menten kinetics. 1.1 M Na2S added PCE —> TCE and PCB were transformed by 10 g/L Butler and to 0.57 M F€C12 in acetylene F eS in aqueous solution at pH 8.3. Hayes anaerobic glove TCE _) Acetylene was the major reaction product for (1999) bag acetylene both TCE and PCB. 1,1-DCE was persistent over 120 days. Zero-valent iron Reductive B-elimination accounts for 87 % William of PCB, 97% of TCE, 94 % of c-DCE, and and 99 % of t-DCE. Reaction of 1,1-DCE gives Roberts rise to ethylene, consistent with a reductive (2000) u-elimination pathway. Pyrite and Abiotic reductive dechlorination of Lee and magnetite chlorinated ethenes by pyrite and magnetite Batchelor was monitored in batch system. Target (2002) organics were mainly transformed to acetylene and small amount of chlorinated intermediates, which suggests that B- elimination, was the main dechlorination pathway. 36 3.8. Literature cited Maymo-Gatell, X.,Anguish, T. and S. H. Zinder. 1999. Reductive dechlorination of chlorinated ethenes and 1,2-dichloroethane by Dehalococcoides ethenogens 195. App]. Environ. Microbiol. 65(7): 3108-3113. Sawyer, C. N., McCarty, PL, and G. F. Parkin. 1994. Chemistry for environmental engineering, McGraw-Hill, Inc., New York Yaws, C. L. 1999. Chemical properties handbook: physical, thermodynamic, environmental, transport, safety, and health related properties for organic and inorganic chemicals, McGraw-Hill, Inc., New York. Gossett, J. M. 1987. Measurement of Henry’s Law Constants for C1 and C2 chlorinated hydrocarbons. Environ. Sci. T echnal. 21(2): 202-208. Haston, Z.C., and PL. McCarty, 1999. Chlorinated ethene half-velocity coefficients (Ks) for reductive dechlorination. Environ. Sci. T echnol. 33(2): 223-226. Vogel, T.M., C.S. Criddle, and PL. McCarty. 1987. Transformation of halogenated aliphatic compounds, Environ. Sci. Techno]. 21; 722-736. Freedman, D. L. and J. M. Gossett, 1989. Biological reductive dechlorination of Tetrachloroethylene and Trichloroethylene to ethylene under methanogenic conditions. Appl. Environ. Microbiol. 5 5(9): 2144-215 1. Fetzner, S., and F. Lingens. 1994. Bacterial dehalogenation; Biochemistry, genetics, and biotechnological applications. Microbiol. Rev. 58: 641-685 Mohn, W.W., and J .M. Tiedje. 1992. Microbial reductive dehalogenation. Microbiol. Rev. 56: 482-507 McCarty, Perry. 1997. Breathing with Chlorinated Solvents. Science. 6(276) 1521-1522 Holliger, C., G. Schraa, Stams Alfons, J. M., Zehnder J. B. Alexander. 1993. A highly purified enrichment culture couples the reductive dechlorination of tetrachloroethene to growth. Appl. Environ. Microbiol 59(9): 2991-2997. Neumann, Anke., Scholz, Muramatsu Heidrun., Gabriele Diekert. 1994. Tetrachloroethene metabolism of Dehalaspirillum multivorans. Archives of Microbiology, 162(4): 295-301 . Krumholz, Lee 1L Sharp, Richard, F ishbain, S. Susan. 1996. A freshwater anaerobe coupling acetate oxidation to tetrachloroethylene dehalogenation, Appl. Environ. Microbiol. 62(11): 4108-4113 37 Maymo-Gatell, X., Chien, Y. T., Gossett, J. M., and S. H. Zinder, 1997. Isolation of a bacterium that reductively dechlorinates tetrachloroethene to ethene. Science 276(5318): 1568-1571. Distefano, T. D., Gossett, J. M., and S. H. Zinder, 1992. Hydrogen as an electron-donor for dechlorination of Tetrachloroethene by an anaerobic mixed culture. Appl. Environ. Microbiol 58(11): 3622-3629. Fennell, D. E., Gossett, J. M., and S. H. Zinder, 1997. Comparison of butyric acid, ethanol, lactic acid, and propionic acid as hydrogen donors for the reductive dechlorination of tetrachloroethene. Environ. Sci. T echnol. 31(3): 918-926. Ballapragada, B. S., Stensel, H. D., Puhakka, J. A., and J. F. Ferguson, 1997. Effect of hydrogen on reductive dechlorination of chlorinated ethenes. Environ. Sci. Techno]. 31(6): 1728-1734. Yang, Y. R. and P. L. McCarty, 1998. Competition for hydrogen within a chlorinated solvent dehalogenating anaerobic mixed culture. Environ. Sci. T echnol . 32(22): 3591- 3597. Fatherpure. B. Z., Nengu, J. P., S. A. Boyd. 1987. Anaerobic bacteria that dechlorinate perchloroethene. Appl. Environ. Microbial. Washington: American Society for Microbiology. Nov 1987. v. 53 (1 1): 2671-2674. Fathepure, B. Z. and S. A. Boyd, 1988. Reductive dechlorination of perchloroethylene and the role of methanogens. FEMS Microbiology Letters 49(2): 149-156. Bouwer, El. and PL. McCarty. 1983. Transformations of halogenated organic compounds under denitirifying conditions. Appl. Environ. Microbiol. 45(4): 1295-1299 Jafvert, CT. and N. L. Wolfe. 1987. Degradation of selected halogenated ethanes in anoxic sediment-water systems. Environ. T axical. Chem. 6;827-83 Egli, C., R.Scholtz, AM. Cook and T. Leisinger. 1987. Anaerobic dechlorination of tetrachloromethane and 1, 2-dichloroethane to degradable products by pure cultures of Desulfobacterium sp. and Methanobacterium sp. FEMS Microbial. Lett. 43: 257 —261. Bagley, D. M. and J. M. Gossett. 1990. Tetrachloroethene transformation to trichloroethene and cis-1,2-dichloroethene by sulfate-reducing enrichment cultures. Appl. Environ. Microbiol. 56(8): 251 1-2516. Kohring, G., X. Zhang and J. Wiegel. 1989. Anaerobic dechlorination of 2,4- dichlrophenol in freshwater sediments in the presence of sulfate. Appl. Environ. Microbiol. 55(10): 2735-2737 38 Stevens, T.S., T.G. Linkfield and J. M. Tiedje. 1988. Physiological characterization of strain DCB-l, a unique dehalogenating sulfidogenic bacterium. Appl. Environ. Microbiol. 54; 293 8-2943. Bouwer, E]. and J. P. Wright. 1988. Transformations of trace halogenated aliphatics in anoxic biofilm columns. .1. Contam. Hydro]. 2; 155-169. Gibson, SA. and J. M. Suflita. 1986. Extrapolation of biodegradation results to groundwater aquifers; reductive dehalogenation of aromatic compounds Appl. Environ. Microbiol. 52(4): 681-688 Debruin, W. P., Kotterman, M. J. J., Posthumus, M. A., Schraa, G., and Zehnder, A. J. B. 1992. Complete biological reductive transformation tetrachloroethene to ethane. Appl. Environ. Microbiol. 58(6): 1996-2000. DiStefano, T. D., J. M. Gossett and S. H. Zinder. 1991. Reductive dechlorination of high concentrations of tetrachloroethene to ethene by an anaerobic enrichment culture in the absence of methanogenesis. Appl. Environ. Microbiol. 57: 2287-2292. Pavlostathis, 8G. and P. Zhuang. 1991. Transformation of trichloroethylene by sulfate- reducing cultures enriched from a contaminated subsurface soil. Appl. Microbial. Biotech. 36 (3): 416-420. Komatsu, T, Momonoi, K., Matsuo, T., K. Hanaki. 1994. Biotransformation of c-DCE to ethylene and ethane under anaerobic conditions. Water Science and Technology. Kidlington, Oxford, UK; Elsevier Science Ltd. v. 30 (7): 75-84. Gibson, S.A., Roberson, D.S., Russell, H.H., and G.W. Sewell. 1994. Effects of three concentrations of mixed fatty acids on dechlorination of tetrachloroethene in aquifer microcosms. Environmental Toxicology and Chemistry, 13 (3): 453-460. Tandoi, V., Distefano, T. D., Bowser, P. A., Gossett, J. M., and S. H. Zinder. 1994. Reductive dehalogenation of chlorinated ethenes and halogenated ethanes by a high-rate anaerobic enrichment culture. Environ. Sci. T echnol. 28(5): 973-979. Wild, A. P., Winkelbauer, W., and T. Leisinger. 1995. Anaerobic dechlorination of trichloroethene, tetrachloroethene and 1,2-dichloroethane by an acetogenic mixed culture in a fixed-bed reactor. Biadegradatian 6(4): 309-318. Scholz. Muramatsu Heidrun, Neumann. Anke, Messmer. Michael, Moore. Edward, and Diekert. Gabriele. 1995. Isolation and characterization of Dehalaspirillum multivorans gen. nov., sp. nov., a tetrachloroethene-utilizing, strictly anaerobic bacterium. Archives ofMicrobialogy; 163 (1): 48-56. 39 Sharma. P. K. and PL. McCarty. 1996. Isolation and characterization of a facultatively aerobic bacterium that reductively dehalogenates tetrachloroethene to cis—l,2- dichloroethene. App]. Environ. Microbial. 62 (3): 761-765. Lee, T. H., Yoshimi, M., Ike, M., and M. Fujita. 1997. Characterization of an anaerobic soil enrichment capable of dechlorinating high concentrations of tetrachloroethylene. Water Science and Technology. 36(6/7): 117-124. Luijten. M.L.G.C., S.W.M. Kengen, G. Schraa, and AIM. Stams. 1997. Complete reductive dehloagnation of chlorinated ethenes. In: H.Verachtert and W.Verstraete (Ed), Intern. Symposium. Environ. Biotechnol, pp. 24-25. Shelton, D.R., Boyd, SA. and J .M. Tiedje. 1984. Anaerobic biodegradation of phthalic acid esters in sludge. Environ. Sci. T echnol. 18 (2) p. 93-97. Rickard, D. T. 1969. The microbiological formation of iron sulphides, Stockholm Cantr, Geol. (26): 49-6 Freney, J. R. 1979. Sulfur Transformations, in R. W. Fairbridge, and C. W. Fink], Jr., Eds, The Encyclopedia of Soil Science Part 1, in Encyclopedia of Earth Science Volume XII, Dowden, Hutchison and Ross, Inc., Stroudsburg, PA, pp. 536-544. Sivavec, TM. and D. P. Horney. 1997. Reduction of chlorinated solvents by Fe(II) minerals. Preprints of papers presented at the 213 th ACS National Meeting, April 13-17, 1997, San Francisco, CA, V0137 (1), pp. 115-117 McCormick, M.L., Kim, H.S., Bouwer, El, and P. Adriaens. 1998. Abiotic transformation of chlorinated solvents as a consequence of microbial iron reduction: An investigation of the roles of biogenic magnetite in mediating reductive dechlorination, Proceedings of the 13 th Mid-Atlantic Industrial and Hazardous Waste Conference, Suri, R.P.S. and Christensen, G.L., eds, Villanova University, Philadelphia, PA, July 12-15, 1998, pp. 339-348. Kriegman-King, M. R. and M. Reinhard. 1991. Reduction of hexachloroethane and carbon tetrachloride at surfaces of Biotite, Vermiculite, Pyrite, and Marcasite, in R Baker, Ed., Organic Substances and Sediments in Water, Vol. 2., Processes and Analytical, Lewis Publishers, Inc. Chelsea, Michigan, pp. 349-3 64. Kriegman-King, M. R. and M. Reinhard. 1994. Transformation of carbon tetrachloride by Pyrite in aqueous solution, Environ. Sci. T echnol (28), 692-700. Butler, E., and Kim F. Hayes. 1999. Kinetics of the transformation of Trichloroethylene and Tetrachloroethylene by iron sulfide. Environ. Sci. Technol. 33(12): 2021-2027 40 Lee, W., and B. Batchelor. 2002. Abiotic reductive dechlorination of chlorinated ethylenes by iron-bearing soil minerals. 1. Pyrite and Magnetite, Environ. Sci. T echnol, 36(23): 5147-5154 Bouwer, E. J., B. E. Rittmann, and P. L. McCarty 1981. Anaerobic degradation of halogenated 1- and 2-carbon organic compounds. Environ. Sci. Techno]. 15: 40-49. Kleopfer, R.D., Easley. D. M. , Haas, B.B., Deihl, T.G., Jackson, DE, and OJ. Wurrey. (1985). Anaerobic degradation of trichloroethylene in soil. Environ. Sci. T echnol, 19(3): 277-280. Vogel 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(5): 1080-1083. Fathepure, B. Z. and S. A. Boyd. 1988. Dependence of tetrachloroethylene dechlorination on methanogenic substrate consumption by Methanosarcina Sp. Strain Dcm. Appl. Environ. Microbial. 54(12): 2976-2980. Sewell, G. W., and S. A. Gibson. 1991. Stimulation of the reductive dechlorination of tetrachloroethene in anaerobic aquifer microcosms by the addition of toluene. Environ. Sci. T echnol., 25(5): 982-984 Gibson, SA, and G.W. Sewell. 1992. Stimulation of reductive dechlorination of tetrachloroethene in anaerobic aquifer microcosms by addition of short-chain organic acids or alcohols. Appl. Environ. Microbial, 58 (4): 1392-1393. Fathepure, El. and J.M. Tiedje. 1994. Reductive dechlorination of tetrachloroethylene by a chlorobenzoate enriched biofilm reactor. Environ. Sci. Technology. 28:746-752. Maymo-Gatell, X., Tandoi, V., Gossett, J. M., and S. H. Zinder. 1995. Characterization of an H2-utilizing enrichment culture that reductively dechlorinates tetrachloroethene to vinyl chloride and ethene in the absence of methanogenesis and acetogenesis. Appl. Environ. Microbiol. 61(11): 3928-3 933. Smatlak, C. R., Gossett, J. M., and S. H. Zinder. 1996. Comparative kinetics of hydrogen utilization for reductive dechlorination of tetrachloroethene and methanogenesis in an anaerobic enrichment culture. Environ. Sci. T echnol. 30(9): 2850-2858. Gao, J. W., Skeen, R.S., Hooker, B. S., and R. D. Quesenberry. 1997. Effects of several electron donors on tetrachloroethylene dechlorination in anaerobic soil microcosms. Water Research. 3 1(10): 2479-2486. Carr, C. S. and J. B. Hughes. 1998. Enrichment of high rate PCE dechlorination and comparative study of lactate, methanol, and hydrogen as electron dopers to sustain 41 activity. Environ. Sci. Technol. 32(12): 1817-1824. Nielsen, R B. and J. D. Keasing 1999. Reductive dechlorination of chlorinated ethene DNAPLs by a culture enriched from contaminated groundwater. Biotechnol. Bioeng. 62 (2): 160-165. Gerritse, J., Renard, V., Games, T. M. P., Lawson, P. A., Collins, M. D., and J. C. Gottschal. 1996. Desulfitabacterium sp. strain PCE], an anaerobic bacterium that can grow by reductive dechlorination of tetrachloroethene or ortho-chlorinated phenols. Archives of Microbiology 165(2): 132-140. DeBest]. H., Jongema. H, Weijling. A, Doddema. H.J., Janssen. D. B., and W. Harder. 1997. Transformation of 1,1,1-trichloroethane in an anaerobic packed-bed reactor at various concentrations of 1,1,1-trichloroethane, acetate and sulfate. Applied Microbiolog/ and Biotechnology. 48 (3): 417-423. Gao, J. W., Skeen, R.S., Hooker, B. S., and R. D. Quesenberry. 1997. Effects of several electron donors on tetrachloroethylene dechlorination in anaerobic soil microcosms. Water Research. 31(10): 2479—2486. Cabirol. Nathalie, Jacob. Francois, Perrier. Joseph, Fouillet. Bruno, and P. Chambon. 1998. Interaction between methanogenic and sulfate-reducing microorganisms during dechlorination of a high concentration of tetrachloroethylene. Journal of General and Applied Microbiology. 44 (4): 297-301. Ndon, U. J ., Randall, A. A., and T. Z. Khouri. 2000. Reductive dehalogenation of tetrachloroethylene by soil sulfate reducing microbes under various electron donor conditions. Journal of Environmental Monitoring and Assessment, 60: 329-3 36. Bradley, P. M., and F. H. Chapelle. 1996. Anaerobic mineralization of vinyl chloride in Fe (RD-reducing, aquifer sediments. Environ. Sci. T echno]. 30 (6): 2084 -2086 Bradley, P.M., and Chapelle, F .H. 1997. Kinetics of DCE and VC mineralization under methanogenic and Fe(III)-reducing conditions. Environ. Sci. T echnal. 31 :2692-2696. Arnold, W. A.and A. L. Roberts. 2000. Pathways and kinetics of chlorinated ethylene and chlorinated acetylene reaction with F e(0) particles. Environ. Sci. T echnol, 34(9): 1794-1805. 42 Chapter Four: MATERIALS AND METHODS 4.1. Chemicals and Reagents. All chemicals used in this thesis were ACS reagent grade or higher. The chlorinated compounds used as substrates in microcosms studies were obtained in neat liquid forms from Sigma-Aldrich (Milwaukee, WI): PCE (HPLC grade, 99.9+%); TCE (certified ACS grade); c-DCE (certified ACS grade); 1,1, l,-TCA, trans-1,2- dichloroethene (t-DCE), 1,1-dichloroethene (1,1-DCE), 1,1-dichloroethane (1,1-DCA). Analytical standards such as PCE, TCE, c-DCE, t-DEC, 1,1-DCE, 1,1,1-TCA, 1,1-DCA, 1,2-DCA, VC, and CA (Chloroethane) were purchased from Supelco (Bellefonete, PA) in methanol-dissolved forms. Gas standards including vinyl chloride (1000 ppm), methane (100%), hydrogen (100%), ethene (100 ppm), acetylene (l %), carbon dioxide (100 %), and ethane (100%) were purchased from Scott Specialty Gases. 4.2. Description of Site and Sediments. The study site is an aquifer located approximately 16 km south of Kalamazoo, MI (Figure 4.1). This aquifer (designated by Michigan Department of Environmental Quality as Plume G) has been contaminated with mixed volatile organic compounds (VOCs) and metal compounds due to PCE use and improper disposal of wastes by a plastic and rubber manufacturer, which impacts 1.7x107 cubic yards of sediments (Mayotte, 1994). Immediately down gradient of the Plume G source, groundwater and sediments are contaminated with chromium and arsenic from a lumber preserving 43 operation (Plume F). Migration of these two plumes to the southeast has resulted in a zone of significant hexavalent chromium/chlorinated organic contamination. Data collected from 1988 on has revealed evidence of intrinsic reductive dechlorination that the daughter products (i.e.TCE and c-DCE), not the parent compound (PCE) prevail. The background chlorinated compounds and their levels in plume G site are summarized in Table 4.1. Table 4.1. Typical Contaminants found from The Study Area“ Contaminants Concentration (pg/L) TCE 1015 1, l, l-TCA 232 c-DCE 1518 1,1-DCA 249 VC 19 Chromium 7O Arsenic 152b aVOC concentrations fiom location MPS-S. No PCE was detected in MPS-5 vicinity. bMaximum concentration detected in MDEQ monitoring wells. No significant arsenic is present in MPS-S vicinity The Plume G aquifer is comprised of relatively homogeneous glacial outwash sands. This glacial outwash consisted of mostly coarse to medium sand with two gravel and cobble layers located just above the confining clay layer (depth 80-83 ft). Sulfate, nitrate, dissolved iron concentrations and total organic carbon analysis indicated that the 44 site groundwater was predominantly anaerobic and low in electron donors (Table 4.2). The mean annual temperature of this plume is about 12°C. The sediment used in this study was collected from the mixed waste impacted section of the plume G site, depth between 50 ft and 75 it, which were mainly composed of fine to medium sand. Batch experiments were designed to imitate field environment (temperature, electron acceptor levels, and chloroethene concentrations) as much as possible. Table 4.2. Typical Chemical Composition of G plume site“: Parameter Value Parameter Value Dissolved iron (aqueous) 12-20 mg/L Nitrate 45 mg/L Total iron 2.5-8.2 mg/g Sulfate 38 mg/L pH 7 .2 Total Organic Carbon < 0.04 %b aAnion concentrations from groundwater (MW-MSUI). Iron concentrations fi'om core sediment samples. ”Provided by Xianda Zhao. 4.3. Microcosms All experiments were conducted either in 250—ml serum bottles containing 200 ml of groundwater or 40-ml vials containing 35 ml of anaerobically collected groundwater from the same aquifer location (MW-MSUI) at 65 it below groundwater surface (bgs). Bottles were sealed with Mini-inert® (teflon) valves or sealed with screw cap containing 22-mm teflon-lined butyl rubber septa. Groundwater was supplemented to create a sulfate reducing media as described in Atlas (2). The detailed description of the amended media is shown in Table 4.3. The groundwater medium was kept under 02 free N2-CO2 (80:20) prior to addition of NaHC03 and vitamin solutions. The vitamin solutions were filter 45 sterilized. The pH of the medium was adjusted to 7.3 to 7.5 with a sterile HCl or Na2C03 solution. Approximately 20g of soil with large particles removed by sieving with 3.35 mm screen were added to 250-ml bottle. Abiotic (pasteurized) controls were included in all experiments. For abiotic controls, the medium and sediments were pasteurized for 8 hours in 70°C water bath and cooled to room temperature under 02 free N2-CO2 (80:20) prior to addition of NaHCO3 and vitamin solutions. Sorption controls were added to account for potential partitioning associated with large-chain hydrophobic carbon sources such as CMB and palmitate. The pH of cultures was routinely checked with a pH meter and probe (Orion). A sub-sample of the culture was removed and pH was determined both before and after adjustment additions. The pH probe was calibrated before each use (Fisher Scientific). Table 4.3. Basal Medium for Microcosm" Component Trace metal solution Vitamin solution NI-14C1 0.16 g CoCl2 061-120 012 g p-Aminobenzoic acid KH2P04 23 m8 MnCl2 -4H20 0.1 g 5 m8 Resazurin 10 mg 21102 0.07 g Vitamin Bu 5 mg NaHCO; 5.0 g H3B03 0.06 g Biotin 1 mg Trace metal solution 1 mL Na2M004 02H2O 0025 g Groundwater 100 mL Vitannn solution 1 mL NiCl2 06H2O 0.025 g Groundwater 1000 mL CuCl2- 2H2O 0.015 g HCl (25 %) 6.5 mL Groundwater 1000 mL a Modified from Atlas (1993) 46 Avoom .owaueeméfisfimv 80¢ 85.893 .83 >95 Co 89304 a 5. 2am WEBB: Z— NA 47 4.4. Analysis of Volatile Organics 4.4.1. Acetylene, Methane, Ethane, Ethene, and C02 Concentrations of acetylene, ethane, ethene, and methane were routinely determined by gas chromatographic (GC) analysis utilizing PerkinElmer (Autosystem) equipped with a flame-ionization detector and a l % SP-1000 on 60/80 Carbopack-B (3.2 mm x 2.44 mm) column from Supelco. Standards and identification were accomplished by matching the retention times of certified gas standards with peaks detected in headspace samples. The oven program was as follows: 35 °C, hold 2 min, 35 °C/min to 180 °C, no hold, flow rate was at 20 ml/min. The detector and injection port temperatures were 200°C. Standards were prepared in serum bottles closed with aluminum crimp cats and teflon lined butyl rubber stoppers. The detection limit of ethene was 65 nmol/L. In case of high concentrations of methane and CO2, these were quantified by direct injection of headspace samples (500 pL) into GC equipped with a thermal conductivity detector (TCD) and a molecular sieve 5A column (15 ft x 1/8 in. stainless steel 60/80 Carboxen-lOOO: Supelco) using helium (30 ml/min) as carrier gas. The GC oven was operated isothermally at 200°C. Standards were prepared by volumetric dilutions of certified gas standards. 4.4.2. Chlorinated compound Liquid samples (0.5 mL) were collected with a gas-tight syringe and immediately dispensed to 40 mL EPA/VOA vials (0.125” TFE/Silicone Liner: I-Chem), which contains 4.5 mL of 2 % sodium bisulfate as a preservative. Vials were sealed immediately with 48 Teflon lined septa and kept in the refrigerator until final analysis. All samples were analyzed within a 14-day holding period. Identification and quantification of PCE, TCE, c-DCE, t-DCE, 1,1-DCE, and VC was made by gas chromatography/mass spectrometry (Agilent 5973 Mass Selective Detector) combined with a purge-and-trap system (Tekmar/Dohrmann Precept II and Tekmar/Dohrmann 3100 Sample Concentrator). GC calibration was performed by adding known masses of PCB, TCE, c-DCE, trans-DCE, 1,1-DCE, and VC directly to a 40-mL EPANOA vial containing 5-ml chilled distilled deionized water. Quantification of chlorinated ethenes and ethene was corrected for partitioning between the aqueous and gas phases. Dimensionless Henry’s Gas Law constants were obtained from Gossett (1987) except for gases. These values are as follows: TCE, 0.392; c-DCE, 0.154; VC, 1.077; ethane, 20.4 (Schwarzenbach et al. 1993); ethylene, 8.5 (Y aws, 1999); methane, 33.1 (Freedman & Gossett, 1989), and acetylene, 0.887 (Butler & Kim, 1999). Throughout this thesis, all reported aqueous and gas-phase concentrations have been corrected for partitioning between aqueous and gas phase. 4.5. Analysis of Volatile Fatty Acids Liquid samples were dispensed into Eppendolfs and clarified by centrifirge at 18, 000 rpm for 5 minutes. The supernatant was the acidified with 50 pl of 1.25 N H2SO4, and dispensed into a 1.5 ml glass vial (Sun Microsystems). The samples were kept frozen until analysis. Acetic, propionic, and lactic acids were measured by high performance liquid chromatography using a Supelco Discovery C8® column (ultraviolet absorbance at 210 nm) and the flow rate was 0.6 mL/min. Acetonitrile was used to regenerate the column between 49 samples and 3.4 g/L of KH2P04 was used as an eluent. Chromatograms were recorded and data integrated using Turbochrom R 4 software (Perkin Elmer Corp). 4.6. Anion Analysis Anions were assayed by ion chromatography with suppressed conductivity detection on a Dionex model 2000i/ SP ion chromatography equipped with a Dionex AS4A IonPac column and utilizing a 1.8 mM biocarbonate—1.7 mM carbonate mobile phase (3mL/min). Chromatograms were recorded and data integrated using Turbochrom R 4 software (Perkin Elmer Corp.) Five—point calibration curves were prepared by diluting primary anion standards into secondary deionized water standards. 1 mL liquid samples were taken from bottles and filtered through a 0.22 um nylon filter (Scientific Resources Inc.) 120 microliter of filtered samples were diluted with 480 pl of deionized water, and dispensed into a polypropylene sample vial (Alcott Chromatography). Samples were stored at -20 °C until analysis was performed. Each sample was analyzed for acetate, bromide, chloride, nitrate, nitrite, phosphate, and sulfate. 4.7. Hydrogen Analysis Hydrogen levels were monitored by injecting 0.5 mL of headspace gas into a mercury vapor reduction gas analyzer (RGA2 Reduction Gas Analyzer;Trace Analytical, Menlo Park, CA). Data was collected with Turbochrom R4 software. Hydrogen partial pressures were measured in the headspaces of bottles at room temperature. The detection limit for hydrogen in headspace was 10 ppb (0.4 nM). 50 Standards were prepared in serum bottles purged with N2 and capped with butyl septa prior to adding H2. Hydrogen values are expressed in parts per million by volume (1 ppmv = 0.1 pa = 10*5 atm) according to the convention adapted by Conrad (1996). Since most of the values published in other studies are expressed as hydrogen dissolved in aqueous, dissolved H2 concentrations were calculated by using the following equations: LP H dissolved = — 4.1 .( ) RT ( ) Where, H2 (dissolved) is the aqueous concentration in males per liter; L is the Ostwald coefficient for H2 solubility, which is 0.01913 at 25°C (Wilhelm et a]. 1977); R is the universal gas constant (0.0821 liter atm K'lmol'l); P is the absolute pressure (in atmospheres); and T is the Kelvin temperature (Lovley et a]. 1994). Therefore, 1 ppmv = ca. 0.78 nM at 25 °C. 4.8. Hydrogen Sulfide Analysis Hydrogen sulfide was colormetrically measured using sulfide kit (Chemtrics, Inc.) based on Methylene blue method. 4.9. Data Reduction Transformation of chlorinated compounds was decided using first-order kinetics. First-order rate constant k can be obtained by plotting natural logarithm of measured aqueous concentrations of chlorinated compound as a function of time. Evidence for fit of 51 the data to such a rate law includes the good agreement between experimental data and rate laws. First order reaction rate (r) in the liquid phase: r = (dCaq /dt) = —k,,,,sCaq (4.2) Due to the existence of headspace in batch systems, a fraction of the chlorinated ethenes partitioned to the headspace and was effectively withdrawn from the reaction phase (aqueous phase). As a result, the measured rates of transformation can be impeded. The inherent rates of transformation occurring in aqueous phase can be determined fiom measured rates using the volumes of the gaseous and aqueous phases, assuming equilibrium partitioning occurring between these phases at all times. The inherent rate is the product of the measured rate, corrected by factor f (Burris eta]. 1996) V f=1+HV—g (4.3) Where H is the dimensionless Henry’s law constant (molL"gas / mol L'1 alqmam), and V3 and Va,l are the gas and aqueous volume respectively. Values of k were corrected to account for the effects of partitioning of the reactant between the aqueous and gas phases (kobs =f*k)- 52 4.10. DNA Analysis 4.10.1. DNA Extraction and RTm-PCR Microcosm DNA samples were extracted from vigorously mixed suspended sediment slurries (3 ml). After collection, sediment slurries were centrifuged (10,000 rpm) for 6 minutes to remove supernatant and DNA was extracted from sediment pellets using DNA extraction kit (MO BIO Laboratories. Inc. Solana Beach, CA.) following manufacturer’s guidelines. Total DNA quantification was accomplished by detecting and quantifying double- stranded DNA (dsDNA) in sediment using the PicoGreen® method (Molecular Probes co. Eugene, OR). PicoGreen® quantification reagent is an ultra sensitive fluorescent nucleic acid stain for quantitating double-stranded DNA (dsDNA) in solutions. Quantification assays were performed using 1-10 % dilutions of sample extracts resuspended in 50 11L of 10 mM TRIS-HCl [pH 7.5]. The samples were excited at 485 nm and the fluoresecence emission intensity was measured at 538 nm using a spectroflurometer. Fluoresecence emission intensity was then plotted versus DNA concentrations. Sample quantitations were accomplished by using MFX Microliter Plate Flourometer and integration software. Real-time PCR (TaqMan-PCR) was used to quantify Dehalococcoides populations and conducted by following the outlines by Gruntzig, et al. (Gruntzvig eta]. 2001). PCR amplification was performed in 50 )1] reaction volumes. The reaction mixture for real-time PCR consisted of IX TagMan Universal Master Mix (containing AmpliTaq Gold DNA polymerase, AmpErase Uracil-N-glycosylase, which degrades PCR carryover products from previous reactions, deoxynucleoside triphosphates with 53 dUTP, a passive reference [6 carboxy-X-rhodamine], and optimized buffer components) (PE Applied Biosystems), 50 nM forward primer and 50 nM reverse primer. MicroAmp optical caps and tubes were used for the final reactions. The 168 rRNA gene forward and reverse primers were 5’ CTGGAGCTAATCCCCAAAGCT 3’; 5’ CAACTTCATGCAGGCGGG 3 ’(He et a]. 2002)’. PCR conditions were as follows: 2 minutes at 50 °C, 10 minutes at 95 °C, then 40 cycles of 15 seconds at 95 °C and 1 minutes at 60 °C. Negative controls with no template DNA or no probe were run in each reaction. The increase in fluorescence emission, due to the degradation of the probe by the DNA polymerase in each elongation step, was monitored during PCR amplification using the 7700 Sequence Detector (PE Applied Biosystems). The fluorescence signal was normalized by dividing the emission of the reported dye (6-carboxyflourscene) by the emission of the passive reference dye, 6-carboxy-X-rhodamine. The parameter threshold cycle (Tc) is the fractional cycle number at which the fluorescence emission crosses an arbitrarily defined threshold within the logarithmic increase phase (0.1 in our reactions). The higher the amount of initial template DNA, the earlier the fluorescence will cross the threshold and the smaller will be Tc. The TC values obtained for each sample were compared with a standard curve to determine the initial copy number of the target gene. 4.10.2. T-RFLP analysis T-RFLP analysis was performed as previously described by Liu, et al. (1997). The amplification was performed using the pair of universal bacterial primers, 8F [5’ AGAGTTTGATCCTGGCTCAG 3 ’] labeled with hexachlorofluorescene (Hex) at the 5’ 54 end (synthesized by Operon Technologies, Inc., Alameda, CA). PCR mixtures included Taq DNA Polymerase (Gibco BRL, Gaithersburg, MD) and were done according to manufacturer’s recommendations with the addition of 0.2 mg/ml of bovine serum albumin (Sigma Chemical Co., St. Louis, MO). PCR amplification was conducted in a GeneAmp PCR system 9600 thermal cycler (Perkin-Elmer Cetus, Norwalk, CT). The protocol consisted of an initial denaturnation step at 94 °C for 30 sec, annulling at 57 °C for 45 sec, elongation at 72 °C for 1 min 30 sec and one extension cycle at 72 °C for 7 min to finalize the PCR. Negative controls included tubes that received no template DNA, as well as positive controls containing pure culture genomic DNA. Aliquots (10 pl) of the PCR products were separated by electrophoresis in a 1.5 % agarose gel using 1 X TAE buffer. The gel was stained with ethidium bromide (500 ng/ 1.11) and visualized by ultraviolet excitation (Helton, 2000). Amplified PCR products were purified using the Wizard PCR purification kit (Promega, Madison, WI) and separately digested with the restriction endonucleases HhaI overnight at 37 °C. Resulting fragments were resolved on an ABI 373A sequencer in a 6 % urea-containing polyacrylamide gel (PE Applied Biosystems Sequencer), running in gene scan mode. The resulting electropherograms were analyzed for similarities using GeneScan software version 3.1 (PE Applied Biosystems). 55 4.11. Literature cited Mayotte, Timothy. J. 1994. Field Trips Guidebook. The geological society of America. North Central Section. Ronald, M. Atlas. 1993. Handbook of Microbiological Media. CRC Press, Florida Jaime A. Graulau-Santiago. 2003. Development and Application of a Methodology to Evaluate Natural Attenuation for Chlorinated Solvents using Conceptual and Numerical Models. A Doctoral Thesis. Civil and Environmental Engineering. Michigan State University. Gossett, J. M. 1987. Measurement of Henrys Law Constants For C1 And C2 Chlorinated Hydrocarbons. Environ. Sci. T echnol. 21(2): 202-208. Schwarzenbach R.P., P.M. Gschwend, and D.M. Imboden. 1993. Environmental Organic Chemistry. Wiley and Sons, Inc. Yaws, CL. 1999. Chemical Properties Handbook. McGraw-Hill. Freedman, D. L. and J. M. Gossett. 1989. Biological reductive dechlorination of tetrachloroethylene and trichloroethylene to ethylene under methanogenic conditions. . Appl. Environ. Microbiol. 5 5(9): 2144-21 5 l . Butler, E., and Kim F. Hayes. 1999. Kinetics of the transformation of trichloroethene and tetrachloroethene by Iron Sulfide. Environ. Sci. T echnal. 33: 2021-2027 Conrad, R 1996. Soil microorganisms as controllers of atmospheric trace gases (H2, CO2, CI-Ia, OCS, N20, and NO). Microbiol. Rev. 60: 609-640 Wilhelm, E., R. Battino, and R]. Wilcock. 1977. Low pressure solubility of gases in liquid water. Chem. Rev. 77:219-262 Lovley, D.R., F.H.Chapelle, and J .C. Woodward. 1994. Use of dissolved H2 concentrations to determine distribution of microbially catalyzed redox reactions in anaerobic groundwater. Environ. Sci. T echnol. 28: 1205-1210 Burris, D.R., C.A. Delcomyn, M.H. Smith, and AL. Roberts. 1996. Reductive dechlorination of tetrachloroethylene and trichloroethylene catalyzed by vitamin B-12 in homogeneous and heterogeneous systems. Environ. Sci. Technol. 30:3047-3 052. Gruntzvig, V., Nold, S.C., Zhou, J., and J .M. Tiedje. 2001. Pseudomonas stutzeri Nitrite Reductase gene abundance in environmental samples measured by RT-PCR App]. Environ. Microbial. 67, 760-768. 56 He, J., Sung, Y., Dollhopf, ME, Fathepure, B.Z., Tiedje, J. M., and FE. Loffler. 2002. Acetate versus hydrogen as direct electron donors to stimulate the microbial reductive dechlorination process at chloroethene-contaminated sites. Environ. Sci. T echnol. 36, 3945-3952. Liu, W., T.L. Marsh, H. Cheng, and LI. Fomey. 1997. Characterization of microbial diversity by determining Terminal Restriction Fragment Length Polymorphisms of genes encoding 16S rRNA. Appl. Environ. Microbiol. 63:4516-4522 Helton, Rebekah R. 2000. Assessment of the microbial communities derived from two tetrachloroethene contaminated aquifers. MS. Thesis. Michigan State University. 57 Chapter Five: EVALUATION OF ELECTRON DONOR EFFECTS ON THE TRANSFORMATION OF TCE AND c-DCE USING ANAEROBIC SEDIMENTS 5.1. Abstract Anaerobic microcosms prepared with sediment and groundwater from an aquifer contaminated with mixed chlorinated compounds (perchloroethene (PCE), 1,1,1- trichloethane (1,1,1-TCA), and daughter products (c-DCE and vinyl chloride», and chromium were used to investigate the effects of electron donors on dechlorination. Screened substrates were ethanol, methanol, lactate, palmitate, propionate, and commercial food grade palmitate (Crisco®) delivered via Microporous Beads (CMB). Substrates were added at levels calculated to stoichiometrically reduce all available anions in the system. Hydrogen levels were ranged from 0.3 to 1.1 mM in biotic bottles. But no clear relationship was observed between extent of dechlorination and hydrogen levels. In microcosms spiked with TCE, the greatest transformation was observed with CMB (44 %) or lactate (36 %) as carbon sources. For all electron donors tested, the predominant biotransformation daughter product was c-DCE, with little further dechlorination beyond c-DCE over a 240-day period. Productions of acetylene have indicated that abiotic degradation by iron sulfide also participated in dechlorination of TCE. 58 5.2. Introduction Chlorinated volatile organic compounds (VOCs) are the most frequently occurring type of contaminant in groundwater at Superfund sites in US (EPA, 1994). It has been estimated by EPA that cleanup of these sites will cost more than $45 billion (EPA, 2000). Biological remediation of these compounds offers the potential for economical and destructive removal of these hazardous compounds from contaminated aquifers. Biological treatments may be passive (natural attenuation) or engineered. Naturally occurring biological activity often leads to incomplete or slow attenuation or incomplete degradation of contaminants. This is generally due to a deficiency in an electron donor or acceptor, other nutrients required for the dechlorinating microbial population, and/or absence of appropriate dechlorinating microbial populations. In these cases, biostimulation or bioaugmentation may be successful. Biostimulation can be a less complicated and costly alternative compared to bioaugmentation, if the appropriate population(s) are present. Many substrates have been shown to stimulate dechlorination of chlorinated solvents (Gibson et a]. 1994; Gibson and Sewell, 1992; Gao et a]. 1997). These are generally organic acids and alcohols. Four fatty acids (lactate, palmitate, propionate, and CMB) and two alcohols (methanol and ethanol) were used as electron donors to support dechlorination in this study. Palmitate and CMB were evaluated as potential slow degrading/persistent electron donors for dechlorination. Propionate is a common fermentation intermediate that could be produced during anaerobic metabolism. Alcohols such as ethanol and methanol were used as easily degradable substrates, and 59 methanol has been often used to support methanogens (F ennell et a]. 1997; Carr and Hughes, 1998; Fathepure and Boyd, 1988 (a)(b)). Furthermore, fermentation of these substrates is known to produce different levels of hydrogen (F ennell er a]. 1997; Carr and Hughes, 1998). Ethanol and lactate are known to generate relatively high hydrogen levels, whereas propionate is known to generate lower hydrogen levels. Since differences in dechlorination might be due to the possible competition between dechlorinators and other hydrogen-utilizing microorganisms for hydrogen, hydrogen production was monitored. Much of the work on biodegradation of VOCs has evaluated degradation under methanogenic conditions (Fathepure and Boyd, 1988 (a)(b); Freedman and Gossett, 1989). Fewer studies have been conducted on the transformation of chlorinated compounds under sulfate reducing conditions (Pavlostathis and Zhuang, 1991) in spite of the dechlorinating potential of sulfate-reducing bacteria and the relative abundance of sulfate in subsurface environments. FeS has been identified as a soil precipitate in anaerobic environments where it is formed through the biologically mediated reduction of sulfate to sulfide and subsequent reaction of sulfide with available iron species (Richard, 1969; Freney, 1979). Several forms of ferrous sulfide can be produced; amorphous iron sulfide, mackinawite (F e29”- 1_o23S), greigite (F e384), and pyrrohotitie (F eSLl). Formation of amorphous iron sulfide predominates at low pH, while pyrite (F eS2) can be developed by further reaction with elemental sulfur. Butler and Hayes (1999) have demonstrated that ferric sulfides and ferric disulfrdes such as mackinwite and pyrite can promote the abiotic dechlorination of 60 Chloroethenes. Trichloroethene was transformed to 65% acetylene, 6% c-DCE, and 9% residual TCE (at pH 8.3) after 120 days. Lee and Batchelor (2002) reported that pyrite degraded TCE to 3.3% c-DCE, 43% acetylene, 2.2% ethene, and 50% residual TCE after 32 days. However, reports by Hassan et a]. (1995) and Doong and Wu (1992) indicated that FeS is unreactive in the transformation of chlorinated compounds. Discrepancies in experimental conditions among these different studies may be responsible for these disparate results. The purpose of the studies reported here were to screen for the biostimulation potential of the native site microbes when augmented with different electron donors. I sought to demonstrate that by adjusting the levels and types of substrates, I could poise the systems reducing potential above methanogenesis and thus selectively influence which indigenous populations were enriched. I also attempted to evaluate the production and reactivity of FeS since it is naturally occurring in sulfate-reducing environments and known to be involved in dechlorination of chlorinated compounds (Butler and Hayes, 1999; Lee and Batchelor, 2002; Hassan et a]. 1995; Doong and Wu, 1992). Here I report on comparative studies in TCE/c-DCE degradation by both biostimulation and biogenic iron sulfide. 61 5.3. Materials and Methods All experiments were conducted in 250—ml serum bottles containing 200 ml of groundwater anaerobically collected from the same aquifer location (MW -MSU1) at 65 it below groundwater. The bottles were sealed with mini-inert® teflon valves, and incubated in the dark, at aquifer temperature (15°C) and inverted to minimize volatile loss. The bottles were shaken periodically (twice a week). Reagent-grade chemicals were used in nutrient medium preparation. Groundwater was supplemented to create a sulfate reducing media as described in previously in Chapter 4. In order to study the effect of reductant (N a2S) on dechlorination of TCE and c-DCE, some bottles were spiked with 0.15 or 1.5 mM of sodium sulfide. Not only abiotic (pasteurized) controls, but also sorption controls were added to account for potential partitioning associated with hydrophobic carbon sources in all experiments. To prevent the formation of siderite (F eC03), and mimic in-situ conditions, the pH was maintained between 7 .0-7 .5 by addition of NaOH. The concentrations of TCE and c-DCE in the bottles were 15 11M and 20 11M, respectively. To study the effect of different substrates in dechlorination, soluble nutrients including ethanol, 2 mM; lactate, 2 mM; methanol, 4.25 mM; palmitate, 0.25 mM; propionate, 1.5 mM; and slow release nutrients (CMB, 400 mg/L each) were added, respectively. CMB carbon beads were prepared by loading molten Crisco® into clay hydrophobic plant growth media. Substrate concentrations were calculated using McCarty’s stoichiometric approach at standard conditions to poise the reducing potential to sulfate reducing environments to prevent methanogensis due to excess carbon (See Appendix A). The equations for the heterotrophic sulfate reducing of various electron 62 donors were calculated using cell efficiency of energy transfer as 0.6 and ammonia as N- source (McCarty, 1969). The identification of iron sulfide’s crystalline structure produced in the microcosms was made by powder x-ray diffraction (XRD) analysis (McCormick et al. 2002). Figure 5.1 shows a X-ray diffraction pattern for the experimentally-prepared FeS match with the mackinawite (The mackinawite power diffraction pattern is from the Powder Diffraction File, International Center for Diffraction Data, Swarthmore, PA). Poorly-crystalline mackinawite and amorphose FeS are found in natural systems (Butler, 1998). X—ray diffraction analyses were conducted using a Rigaku Rotaflex rotating anode XRD apparatus (Cu K01 radiation, 40 kV, 100 mA). XRD samples were prepared by freeze-drying and back-filling under nitrogen to potential oxidation. Samples were mounted on glass slides in the anaerobic glovebox and then sealed under tape to prevent sample oxidation during analysis. Headspace samples (acetylene, ethane, ethene, and methane) were determined by gas chromatographic (GC) analysis (PerkinElmer) as described in Chapter 4. 63 Figure 5.1. Powder X-ray diffraction pattern by experimentally prepared FeS. (The mackinawite power diffraction pattern is from the Powder Diffraction File, International Center for Diffraction Data, Swarthmore, PA). "“l Intensity PI'I'I‘I'flSTTI I" I'I'TYT J'T'T'T‘ T‘T‘Trl I'I'T‘ ' r'r'r‘r' I'Ivr'r TYT'IrI I'I'I‘I' 20 64 5.4. Results 5.4.1. Effect of carbon sources on transformation of c-DCE and TCE 1. TCE transformation In order to determine which substrates support enhanced dechlorination activity, multiple carbon sources were screened, including both alcohols and fatty acids. Long- term incubation cultures were provided with carbon sources including methanol, ethanol, lactate, palmitate, CMB, and propionate at the beginning of incubation only. Within a week of nutrient addition, nitrate presented in the site groundwater (average concentration of 35 ppm) was consumed. All biotic bottles blackened within two weeks and the presence of sulfide was confrrmed by Chemetrics kit (methylene blue method). Anion levels, turbidity, and color remained unchanged in pasteurized abiotic controls throughout the 240-day period. All bioactive (non-control) microcosms exhibited some dechlorination of TCE over a 240-day period. The activity generally initiated after approximately 60 days of incubation, however the onset of degradation and 240 day removal amounts varied among replicates. Lactate, CMB, and palmitate were effective in stimulating TCE degradation (Figure 5.2). Propionate, methanol, and ethanol stimulated slight dechlorination of TCE. The dominant byproduct of TCE degradation was c-DCE, but small amount of acetylene was detected in most replicates. The percent removal of TCE by fatty acids and their sorption controls are shown in Figure 5.3. Sorption controls are included for large molecular weight (C-16) carbon sources to accounts for partitions. After 240 days of incubation, CMB and lactate showed the most significant removal (about 44 % and 36 % removal respectively, corrected for partition). An approximately 15 % decrease in TCE 65 .998 ovm macaw meadow sebum—o 8 26 mega—Emcee“ MOB Eng 3v .m> 32m 3 a 80:80me .N.m eBwE Beam 3 32m 8 3.550 m5 ESE—am IT 8393 IT 2:538.— Eiucm 111 .8282 II... 353nm 19... are a: 2.3 a: a: a: a: 8 c as a: as 8 e . _ P _ Q — p p p o r In r G ‘— (wt) °ou03 snoanbv . . . . .V“\\\\\\\>{\§s\\)\s\\:0‘r .. . .. (r. E 66 level was observed in the CMB specific abiotic control bottles, most likely due to partitioning to the hydrophobic Crisco® in the CMB beads (Figure 5.2). This indicated the removal in the CMB treated microcosms were not due to partitioning. However, without additional amendments, c-DCE appeared to resist firrther dechlorination under the conditions of this study. 80 - 60 - "<3 40 — O 8 d) 04 °\° 20 - O I l ' Coiuol Lactate Palmitate Plamitate CMB CMB control control -20 - Electron Donors Figure 5.3. The differences in dechlorination of TCE after 240 days expressed as percentage removal of abiotic soil control. This experiment was conducted in the presence of Na2S (1.5 mM). Sorption controls are included for large molecular weight (C-16) carbon sources. Values are averages of 4 replicates and the error bars represent the standard deviation. 67 2. Cis-DCE transformation Unlike TCE, only slight degradation of c-DCE was observed over a 240-day period. Palmitate and CMB amended microcosms demonstrated some degradation of c-DCE (Figure 5.4). For c-DCE-amended microcosms, minimal degradation daughter products were detected. As can be seen on Table 5.2, about 25 % and 12 % degradation of the initial c-DCE were observed in the CMB and CMB specific abiotic control bottles, respectively, after 240 days of incubation. This again indicated partitioning did not account for c-DCE removal in the CMB fed microcosms. 80- 60- % Removal Cool Lactate Palmitate Palmitate CMB CMB control control '20 I Electron Donors Figure 5.4. The differences in dechlorination of c-DCE after 240 days expressed as percent removal of abiotic soil control. This experiment was conducted in the presence of Na2S (1.5 mM). Sorption controls are included for large molecular weight (C-16) carbon sources. Values are averages of 4 replicates and the error bars represent the standard deviation. 68 5.4.2. Oxidation-Reducing Potential and pH The pH of the bottles dropped slowly during the course of incubation and remained stable around 6.8, while the oxidation-reduction potential (ORP) changed rapidly at first (from an initial value of -102 mV to about -230 mV on day 30), then remained relatively constant. The transformation of Chloroethenes, pH/ORP final values after 240 days of incubation, and hydrogen production after 25 days of incubation at 15°C was summarized in Table 5.1. Table 5.1. Effect of substrates on transformation of chloroethenesa after 240 days of incubation in sulfate-reducing conditionsb. Electron H dro en Chlorinate ethene Reactant donors p HIGH) 3(1th Reduction (as %)° TCE Controls 7.2/-138 O 8.6 i 2.6 No nutrient 7.2/~189 0 12.8 i 2.6 CMB 6.8/—206 0.4 i 0.03 43.7 i 4.7 Ethanol 7.2/-168 0.4 i 0.12 14.2 :t 3.3 Lactate 6.9/-192 0.3 i 0.07 35.7 i 4.1 Methanol 7.1/-211 1.1i0.18 17,0;th Palmitate 6.9/-182 0.3 i 0.09 32, 5 i 29 Propionate 7 .0/-202 0.5 i 0.03 14_1 i 3.6 c-DCE Controls 7. l/-123 0 12.1 i 2.7 No nutrients 7.2/-147 0 11.6 i 2.7 CMB 6.8/-248 0.5 i: 0.08 25.3 i 5.6 Ethanol 7.1/-180 0.6 i 0.12 9.1 i 2.8 Lactate 6.9/-217 0.4 i 0.09 11.8 i 3.8 Methanol 7.0/-199 1.0 i 0.28 13.0 i 4.4 Palmitate 7 .0/-192 0.4 i 0.09 10.9 i 32 Propionate 7.0/-185 0.5 i 0.01 13.2 i 3.5 “ Values represent average of triplicate. i indicates standard deviation bAll experiments are done in the presence of reductant NaZS of 1.5 mM; Methane levels were below 0.1 uM at the end of the incubation; At end of incubation H2 levels were below 0.01 mM. °Based on initial moles of Chloroethenes individually. 69 5.4.3. Anion Utilization Within a week of carbon addition, complete removal of available nitrate/nitrite was observed. When Nags was not added as a reductant, all sulfates present in the groundwater (65 to 92 ppm) were removed within 12 days (Figure 5.5). In order to see the effect of sulfate on TCE transformation, a similar amount of sulfate was respiked on llO-th day for CMB fed microcosms (Figure 5.6). Even though complete removal of sulfate occurred again, no significant enhancement on TCE transformation was observed during this period. 90 -I— Sulfate ~6- Nitrite 65 : "air-“Nitrate Conc. (ppm) A O p... kit - 10 5 10 1 5 Time (days) Figure 5.5. Anion degradation in CMB fed microcosm. Values represent average of triplicate. The error bars indicate stande deviation 70 100 - —~ 16 Sulfate respiked at 65 ppm 75 ~ J + 12 A .- T r 8 g was . m 3 50 - “L .. i —- s 9 5 ‘2‘? fi v .2 25 __ 4 c: < 0 2.29““. 0 0 45 90 135 180 Time (days) +Nitrate %Nitrite + Sulfate %TCE Figure 5.6. Effect of sulfate on transformation of TCE in CMB fed microcosm. 5.4.4. Hydrogen production High levels of hydrogen were produced by fermentation of substrates at first (Table 5.2). Ethanol fed microcosms exhibited the highest levels of hydrogen immediately following the electron donor addition. But no clear relationship was observed between extent of dechlorination and hydrogen levels. Methane production was minimum in all biotic bottles. All these observations indicate that amount of added substrate was enough to sustain sulfate reducers but not for methanogens. 71 5.4.5. Effect of sodium sulfide on TCE transformation To determine the effect of reductant on transformation of TCE, additional microcosms were prepared with 2 different levels of sodium sulfide (0.15 and 1.5 mM). Approximately 32 % of TCE were transformed with and without reductant, respectively (lactate as substrate). Addition of reductant did not lead to a significant increased in TCE transformation (Table 5.2). Sterile controls amended with 1.5 mM of sodium sulfide showed no significant degradation of TCE (i.e. < 7 %), and no c-DCE or other daughter products were detected. Microcosms with and without added Na2S, demonstrated similar ORP values at the same substrate levels (ranging from -182 to -248). Thus, not only microcosms with reductant, but also those without reductant sustained desirable reducing conditions, indicating the biological processes, as well as added reductants were capable of generating the drop in reducing potential. Therefore, carbon loading (not use of reductant) may be sufficient to control ORP in the field. Table 5.2 shows transformation of TCE in the presence and absence of reductant using CMB and lactate as substrates. Table 5.2. Reduction of TCE depending on presence of reductant after 180 days of incubation at 15°C. Substrate Reductants (mM) Reduction (as %)a ORP CMB 0 37.8 i 5.8 -187 i 7.3 0.15 40.0 i 4.4 -209 i' 5.9 1.5 41.4:72 -211¢4.5 Lactate 0 32.4 i 1.9 -201 i- 5.7 0.15 3l.4i2.1 -l96i4.5 1.5 32.7:t3.7 -213 123.1 “ Values represent average of triplicate. i indicates standard deviation. 72 5.4.6. Byproduct Analysis Table 5.3 shows the byproduct investigation after the dechlorination of TCE after 180-day period. Productions of trace amount of ethene and acetylene initiated approximately 90 days after incubation, however the onset of degradation and 180-day removal amounts varied among replicates. The principal reaction byproduct for TCE transformation using CMB as substrate was c-DCE and acetylene; c-DCE accounted 65 % of the transformed TCE, while acetylene accounted for 27 %. Even though trace amounts of ethene were detected in some bottles, neither t-DCE nor 1,1-DCE was observed above the method detection limit. Unlike TCE, c-DCE treated microcosms exhibited little degradation of c-DCE. However, trace amount of acetylene was observed above the method detection limit (acetylene: 50 nM) using CMB as substrate. No 1,1-DCE or VC was observed above the method detection limit. Mass balance of c-DCE shown in Table 5.3 shows more than 17 percent discrepancy, which may have resulted from VOCs loss from handling and sampling. In order to see the effect of sulfate on TCE transformation, a similar amount of sulfate was respiked on 110-th day for CMB fed microcosms. Even though complete removal of sulfate was observed again, no significant enhancement in acetylene production was observed during this period. The crystalline structure of microcosms is illustrated in Figure 5.7. The figures show wide peaks and relatively small intensities. These diffraction patterns by the microcosm sediment shows poor degree of crystallinity and thus, indicating that the iron sulfide produced was amorphous. A crystal structure of microcosm treated with added 73 sulfate can be seen in Appendix B (no difference in crystal structure was observed due to added sulfate). Table 5.3. The byproducts and mass recoveries for transformation of c-DCE and TCE after 180 days“. Reactant Amendment ijroduct Mass recovery (%)b TCE CMB Acetylene 7.3 i 4.4 Ethene 2.2 i 3.7 eDCE 175i49 TCE remaining 62.2 3: 5.8 Tm“ 892:68 Ethanol Acetylene 5.3 i 2.0 C-DCE 1.8 i 0.5 TCE remaining 87.7 i 7.3 Tad 948:83 Lactate Acetylene 8.2 i 3.7 Ethene 2.1 i 1.3 eDCE 152:24 TCE remaining 67.6 i- 1.9 Tad 931:89 Methanol Acetylene 4.7 i 1.2 04303 1.5 i 0.9 TCE remaining 84.9 i- 5.8 Tmfl 911:6? Palmitate Acetylene 6.3 i 2.1 Ethene 2.2 i 0.7 eDCE 15:39 TCE remaining 68.2 i 9.3 de 842:128 Propionate Acetylene 1.9 i 1.4 c-DCE 2.0 d: 1.1 TCE remaining 88.2 i 5.3 TNfl 922:68 c-DCE CMB Acetylene 7.3 i 1.0 c-DCE remaining 755 i 49 Tmm 828:58 “The pH of microcosms was maintained and nearly constant between 7 -7.5. Reductant (N a28) was not added. Only trace amounts of VC were detected. Values represent average of triplicates. i indicates standard deviation bBased on initial moles of chloroethenes individually. 74 0.250 0.200 - 0.150 d 0.100 . CPS norm 0.050 ~ 1 0a» . . . L 11.4.1.1 1...; -l. 1. link... .‘J 11‘.“ A . .LL L. 10 20 30 40 50 60 70 80 9C -0.050 29 Figure 5.7 . Powder x-ray diffraction pattern by iron sulfide produced in microcosm. 75 5.4.7. Field Estimation The extrapolation of laboratory result to field-scale processes is an uncertain endeavor, particularly in environmental microbiology (Madsen, 1996). However, studies of several sites contaminated with chlorinated ethenes suggest that rates of reductive dechlorination estimated from laboratory microcosms are similar to those estimated from field studies (Wilson et al. 1996). Half-life prediction of chloroethene in the groundwater was made using first order rate equation to obtain treatment estimation. As can be seen in Table 5.4, remediation of TCE and especially c-DCE in groundwater will results in prolonged operation time and high cost without electron donor addition Natural attenuation occurring in the aquifer seems not only incomplete, but also insufficient. The effects of sulfate, reductant (N azs) and nutrient and their interactions on the dechlorinations are summarized on Appendix C. 76 Table 5.4. Estimated transformation rate and half-time of TCE and c-DCE biodegradation in Plume G Chloroethene Electron donor k (l/month)a 2311:3133 TCE CMB 0.13 5.6 Ethanol 0.09 8.0 Lactate 0.11 6.5 Methanol 0.08 8.3 Palmitate 0.08 9.0 Propionate 0. 1 6.6 No nutrient 0.006 108 c-DCE CMB 0.03 20.3 Ethanol 0.01 62.4 Lactate 0.03 24.5 Methanol 0.02 37.5 Palmitate 0.01 47.1 Propionate 0.01 53.7 No nutrient 0.003 256 “ Based on first-order rate equation C=Coe‘k‘. b t1/2 = 111 Z/k. 77 5.5. Conclusion and Discussion Several researchers have demonstrated the stimulation of reductive dechlorination of chlorinated ethene by addition of organic supplements or hydrogen (Gibson et al. 1994; Fathepure and Boyd, 1988(a); Freedman and Gossett, 1989). These studies were conducted in soils that were rich in organic material (Gibson et a1. 1994; Freedman and Gossett, 1989) or a pure culture derived from anaerobic sludge (Fathepure and Boyd, l988(a)). In contrast, our study used anaerobic materials from an oligotrophic aquifer (maximum total organic carbon < 0.04 %) impacted with mixed chlorinated compounds (PCE, 1,1,1-TCA, and daughter products), and chromium to investigate the effects of nutrients on dechlorination. TCE degradation was achieved by selectively stimulating indigenous microflora using fatty acids such as CMB and lactate. The predominant transformation daughter product was c-DCE and acetylene. Previous studies on the transformation of chloroethenes under sulfate-reducing conditions have also resulted in the production and accumulation of c-DCE (Pavlostathis and Zhuang, 1991; Bagley and Gossett, 1990). Even though TCE degradation continued afier depletion of sulfate, it occurred at a much lower rate in our study. This suggests these conditions led to enrichment of sulfate reducers that the degradation observed might have been linked to sulfate reducing populations such as Desulfomonile tieay'ei or Desulfovibrio fiuctosivorans (Drzyzga and Gottschal, 2002; Mohn and Tiedje, 1991). Hydrogen is an important intermediate in the anaerobic degradation of organic matter (Conrad, 1999). Microbial competition for H2 plays an important role in the natural attenuation of chloroethenes, and recent studies have focused on the role of H2 as a key electron donor for the reductive transformation of these compounds. In our study, we 78 designed a sulfate-reducing environment where sulfate is depleted with minimal methane production. Even though high H2 production was observed in some cases, no clear relationship with transformation rate was observed. Even though the predominant transformation byproduct was c-DCE, small amounts of acetylene, known as an abiotic dechlorination indicator (Butler and Hayes, 1999; Lee and Batchelor, 2002), were also detected. This observation suggested that the abiotic degradation by iron sulfide have participated in dechlorination of TCE in microcosm. A plausible scheme for the reduction of the TCE to acetylene by abiotic degradation (hydrogenolysis (replacement of a halogen by hydrogen), reductive elimination (dihaloelimination), and hydrogenation (reduction of multiple bonds)) has been proposed by Lee et al (2002). Because acetylene was one of the main transformation products, a reductive elimination pathway would appear occur (Figure 5.8). The removal of TCE without observing stoichiometric production of a chlorinated intermediate suggests that the acetylene formed during the TCE dechlorination is produced via an unstable intermediate (chloroacetylene) that quickly decays. The diffraction pattern by XRD indicated poor-crystalline (amorphous) formation of FeS in microcosm. According to Bemer (1964), the poorly-crystalline mackinawite and amorphous FeS are often found in natural systems and their crystallinity is similar to the materials precipitated from sodium sulfide and ferrous sulfate. Additional sulfate, which would have provided added source of the reductive power in microcosm, did not coincide with rapid enhanced dechlorination. No immediate increase on the TCE degradation by the added sulfate indicated that the dechlorinations 79 by indigenous sulfate reducers and iron sulfide are very slow. This suggests that the pulse feeding strategy is required to achieve sufficient dechlorinating activity. 80 Figure 5.8. Possible pathways for the reductive dechlorination of chlorinated ethenes by iron sulfide (modified from Lee et al. (2002)). Chemical compounds in italic characters were detected in this experiment C1 C1 / Cl H \ C1 C1 TCE H—m— Cl H H Chloroa et 1 cis-DCE \ _ c y we Reductive l B-elrmmatron l Hydrogenolysis H>==< H H Ethene l Hydrogenation RH» H H Ethane 81 5.6. Literature cited EPA. 1994. “Common Chemicals found at Superfund Sites”, Office of Emergency and Remedial Response, Washington, DC. EPA. 2000. Engineered Approaches to In Situ Bioremediation of Chlorinated Solvents: Fundamentals and Field Applications. Gibson, S.A., Roberson, D.S., Russell, H.H., and G.W. Sewell. 1994. Effects of three concentrations of mixed fatty acids on dechlorination of tetrachloroethene in aquifer microcosms. Environmental Toxicology and Chemistry, 13 (3): 453-460. Gibson, SA, and G.W. Sewell. 1992. Stimulation of reductive dechlorination of tetrachloroethene in anaerobic aquifer microcosms by addition of short-chain organic acids or alcohols. App] Environ. Microbial. 58 (4): 1392-1393. Gao, J. W., Skeen, R.S., Hooker, B. S., and R. D. Quesenberry. 1997. Effects of several electron donors on tetrachloroethylene dechlorination in anaerobic soil microcosms. Water Research, 3 1(10): 2479-2486. Fennell, D. E., Gossett, J. M., and S. H.Zinder. 1997. Comparison of butyric acid, ethanol, lactic acid, and propionic acid as hydrogen donors for the reductive dechlorination of tetrachloroethene. Environ. Sci. T echnol., 31(3): 918-926. Carr, C. S. and J. B. Hughes. 1998. Enrichment of high rate PCE dechlorination and comparative study of lactate, methanol, and hydrogen as electron donors to sustain activity. Environ. Sci. T echnol., 32(12): 1817-1824. (a) Fathepure, B. Z. and S. A. Boyd. 1988. Dependence of tetrachloroethylene dechlorination on methanogenic substrate consumption by Methanosarcina sp. Strain Dcm.. Appl. Environ. Microbiol. 54(12): 2976-2980. (b) Fathepure, B. Z. and S. A. Boyd. 1988. Reductive dechlorination of perchloroethylene and the role of methanogens. FEMS Microbiology Letters 49(2): 149- 156. Freedman, D. L. and J. M. Gossett. 1989. Biological reductive dechlorination of tetrachloroethylene and trichloroethylene to ethylene under methanogenic conditions. App]. Environ. Microbiol. 55(9): 2144-215 1. Pavlostathis, S. G. and Ping Zhuang. 1991. Transformation of trichloroethylene by sufate-reducing cultures enriched from a contaminated subsurface soil. Appl. Microbial. Biotech, 36(3): 416-420. 82 Rickard, D. T. 1969. The microbiological formation of iron sulphides, Stockholm Contr, Geo]. (26): 49-6 Freney, J. R 1979. Sulfur Transformations, in R. W. Fairbridge, and C. W. Fink], Jr., Eds, The Encyclopedia of Soil Science Part 1, in Encyclopedia of Earth Science Volume XII, Dowden, Hutchison and Ross, Inc., Stroudsburg, PA, pp. 536-544. Butler, E., and Kim F. Hayes. 1999. Kinetics of the Transformation of Trichloroethene and Tetrachloroethene by Iron Sulfide. Environ. Sci. Techno]. 33: 2021-2027 Lee, W., and B. Batchelor. 2002. Abiotic reductive dechlorination of chlorinated ethylenes by iron-bearing soil minerals. 1. Pyrite and Magnetite, Environ. Sci. T echno], 36(23): 5147-5154 Hassan, S. M., Wolfe, N.L., and MG. Cippolone. 1995. Reaction mechanisms involved in the reduction of halogenated hydrocarbons with sulfated iron, “209'” ACS National Meeting, April 2-7, Anaheim, CA, Vol. 35 (1), pp.73-737. Doong, R. A. and S. C. Wu. 1992. Reductive dechlorination of chlorinated hydrocarbons in aqueous solutions containing ferrous and sulfide ions, Chemasphere 24, 1063-107 5. McCarty, P. L. 1969. Energetics and Bacterial Growth. 5'” Rudolf Research Conf. Rutgers. McCormick, M.L. E.J. Bouwer, and P. Adriaens. 2002. Carbon tetrachloride transformation in a model iron-reducing culture: relative kinetics of biotic and abiotic reactions. Environ. Sci. T echnol. 36(3); 403-410 Butler, Elizabeth Constance. 1998. Transformation of haolgenated organic compounds by iron sulfide. A Ph.D. Thesis. pp. 28. Univeristy of Michigan. Madsen, E.L. In Soil Biochemistry. Vol.9; Stotzky, G., Bollag, J ., M. Eds; Marcel Dekker: New York, 1996. pp.287-370 Wilson, J .T.; Kampbell, DH. and J .W. Weaver. In US. Environmental Protection Agency Symposium in natural attenuation of chlorinated organics in ground waters. Sept. 11-13, 1996; Dallas, TX; EPA/540/R-96/509; 1996;pp.124-127 Bagley D. M. and J. M. Gossett. 1990. Tetrachloroethene transformation to trichloroethene and cis-1,2- dichloroethene by sulfate-reducing enrichment cultures. Appl. Environ. Micro]. 56(8): 251 1-2516. 83 Oliver Drzyzga and Jan C. Gottschal. 2002. Tetrachloroethene dehalorespiration and growth of Desulfitobacteriumfrappieri TCE] in strict dependence on the activity of Desulfavibrio fructosivorans. Appl. Environ. Microbiol, p. 642-649, Vol. 68, No. 2 Mohn, W.W., and J .M. Tiedje. 1991. Evidence for chemiosmotic coupling of reductive dechlorination and ATP synthesis in Desulfomonile tieay'ei. Arch. Microbiol. 157:1-6. Conrad, R. 1999. Contribution of hydrogen to methane production and control of hydrogen concentrations in methanogenic soils and sediments. FEMS Microbiol. Ecol. 28, 193-202. Bemer, R A. 1964. Iron sulfides formed from aqueous solution at low temperatures and atmospheric pressure. .1. Geol. 72, 293-306. 84 Chapter Six: STINIULATION OF HIGH-RATE COMPLETE DECHLORINATION ACTIVITY BY PULSE FEEDIN GS OF CHLOROETHEN ES 6.1. Abstract Once-fed microcosms that demonstrated initial positive dechlorinating activity were further characterized for chloroethene/ethane transformation potential by pulse- feedings of VOCs, lactate, and native levels of sulfate using plume G groundwater. Experiments were performed in batch microcosms receiving two levels of lactate; low (0.5 mM) or high (12.5 mM). PCE was fed approximately 15 umol/ 2 month for the first year followed by 15-umol/ month in the second year. High lactate enriched bottles exhibited limited transformation of PCB with TCE or c-DCE as byproduct (0.03 6/day). Low lactate enrichments demonstrated high rate of degradation; PCE, TCE, c-DCE, t- DCE, 1,1-DCE, and VC to ethenes (rates ranged from 0.19 to 0.82/day). Further reduction to ethane was observed in many cases. Highest rate of PCE transformation was observed at 25 °C in low lactate/PCE enrichments. 16S rRNA gene-based RTm-PCR identified that the pre—enrichments contain no quantifiable amounts of Dehalococcoides populations. Significant levels of Dehalococcoides were detected in all of the post- enrichment samples. The low lactate/PCE microcosms contained Dehalococcoides at least two orders of magnitude higher than high lactate/PCE enriched microcosms. This demonstrated that pulse feedings of chlorinated ethenes, and low levels of electron donor effectively stimulated Dehalococcoides populations. 85 6.2. Introduction Methanogens are known to dechlorinate PCE to TCE and c-DCE (F athepure and Boyd, 1988(a)(b); Freedman and Gossett, 1989). However, the dechlorination by methanogens is known to be incomplete and proceed at low rates. So far only Dehalococcoides ethenogenes is known to fully dechlorinate PCE to ethene. It uses H2 as a sole electron donor, grows slowly in pure culture, and is fastidious in its growth conditions (Maymo-Gatell et a]. 1997). In pure culture, Dehalococcoides ethenogenes requires the extract of an anaerobic sludge community for sustained growth, suggesting that it relies on biochemical collaboration with other microorganisms and that it benefits significantly by growing in a mixed community. Several molecular approaches are currently available as means of exploring microbial communities including T-RFLP and RTm—PCR (Hendrickson et a]. 2002; Liu et a]. 1997). This research discusses the characterization of highly enriched mixed cultures that reductively dechlorinates PCE to ethene/ethane with lactate as an electron donor using two molecular techniques. Effects on the community structure and activity on the original community by prolonged exposure to high versus low lactate/PCE feedings were studied. Terminal restriction length polymorphism (T-RFLP) analyses were employed to identify community structure in the mixed cultures and to track population dynamics after enrichment by PCE. Additionally, specific primers for detecting Dehalococcoides species were used to determine the presence and enumeration of such Dehalococcoides populations. 86 Studies presented here focus on the enrichment of a high-rate PCE dechlorinating culture through pulse feedings of PCB, lactate, and native levels of groundwater sulfate over 2-year period. Microcosms which demonstrated initial positive dechlorinating activities, were further enriched using 2 different ratios of lactate and pulse feedings of chlorinated ethenes to evaluate the effect of excess electron donor on stimulating dechlorinating activity. I have evaluated the effect of lowering the ratio between substrate to PCE, along with pulse feedings on the substantial enrichment of Dehalococcoides. Characterization of these enrichments and enumeration of Dehalococcoides are presented later in this chapter. Finally, I conducted daughter product analysis to evaluate the completeness of PCB transformation by rapidly dechlorinating enrichments. 87 6.3. Materials and Methods Enrichment Procedure Anaerobic microcosms assembled from plume G sediments initially exhibited neither rapid nor complete transformation of parent compounds (TCE and c-DCE). In order to enhance dechlorinating activity, enrichments that demonstrated the highest dechlorination activity were subcultured and incubated with three different concentrations of lactate (0, 0.5 mM, and 12.5 mM) and equal amount of PCE (15 umol/bottle). Table 6.1 shows the additions of chlorinated compounds and lactate corresponding to the appropriate experiment. Chlorinated compounds were added neat using SuL microsyringe except for VC, which was added as a gas using airtight syringe. The Roman numeral corresponds to the chronological order in which the experiments were performed. Experiment I indicates two conditions that favor one or the other process based on electron donor/acceptor ratio. Experiment 11 was designed to screen for the substrate specificity of PCE over TCE, and c-DCE. Experiment [I was set up using c-DCE long-term microcosms from chapter 5. Only low levels of lactate (see Table 6.1 for exact concentration) were fed once every 2 months for 6 months and once a month thereafter for c-DCE and TCE. TCA microcosms were spiked once every 2 months, as the TCA microcosms never exhibited high rates of dechlorination. Therefore, all chloroethene enrichments consumed similar amounts (12 cycles of 15 umol) before the rate experiments except TCA. TCA enrichments had only 6 cycles of 15 umol over 360 days of enrichment. The additions and consumptions of chlorinated compounds for experiment II are depicted in 88 Appendix D. The overview of enrichments and experiments performed according to their time sequence is described in Figure 6.1. Sampling and Analysis Over the initial period of enrichment, complete transformation of PCB was not observed; however, periodic respiking with PCE was conducted. During enrichment, one- quarter of the bottle (liquid volume, ZOO-mL), contents were replaced with fresh groundwater media (typically once/2 month), to replenish vitamin, phosphate, and lactate (0.5 or 12.5 mM). Groundwater media was made by adding phosphate (KH2PO4, 0.2g/L), sodium bicarbonate (NaHCOg, 5g/L), and vitamin supplements (see Table 4.3) to plume G groundwater, which contained native amounts of sulfate (range between 35 to 60 ppm) and nitrate (range between 25 to 40 ppm). After lactate addition, PCE was added at 15 umol/bottle. Throughout these studies, bottles were purged thoroughly afier sampling but prior to each PCE addition to remove VC and ethene, and thus to prevent potential toxicity (Schink, 1985). The pH was monitored and maintained between 6.8-7.5 by addition of small amounts of 1 M NaHC03 solution. Chlorinated compounds, headspace, anion, and fatty acid analyses were done as described previously in Chapter 4. Terminal restriction length polymorphism (T-RFLP) and Real time PCR (RTm-PCR) analyses were performed as described in Chapter 4. Effect of Temperature All bottles were kept in a constant room temperature during 2-year enrichment period. Nine low lactate/PCE enriched bottles were incubated at three different 89 temperatures (15, 25, and 35 °C) for 3 weeks to evaluate temperature effects on transformation of PCE. Table 6.1. Experiment arrangements for respective conditions. Microcosms were exposed to pulse feedings of PCB over 1.5 years for Experiment I and pulse feedings of TCE, c-DCE, and TCA over 1 year for Experiment 11“. Chloroethene added Lactate/PCE uequiv of umol of Conditions (umol) (equiv/ equiv) lactate lactate Experiment I: Cometabolic reductive dechlorination favoring versus chlororespiration favoring conditions (using TCE exposed bottles from Chapter 5) PCE 1 15 0/0 0 0 PCE 2 15 10/1 1200 100 PCE 3 15 250/ 1 30000 2500 Experiment 11: Chlororespiration favoring conditions (using c-DCE exposed bottles from Chapter 5) TCE 15 10/1 900 75 c-DCE 15 10/1 600 50 TCA 15 10/1 600 50 aThe electron donor ratio was calculated assuming: PCE 8 equiv/mol; TCE 6 equiv/mol; c-DCE 4 equiv/mol; and TCA 4 equiv/mol. Lactate was assumed to provide 12 equiv/mol (on CO2 basis). 90 0.5 mM Lactate, 12 mM Lactate, 15 mol PCE/2 mo. 15 mol PCE/2 mo. f """""" '1 ¢ : 1 st year 5 0.5 mM Lactate, 12 mM Lactate, 15 umol PCE/l mo 15 umol PCE/l mo. Figure 6.1. The overview and sequence of the experiments performed during enrichments. 91 6.4. Results 6.4.1. High lactate conditions 6.4.1.1. Transformation of PCE After one year of repeated feedings, transformation rates became faster (once a month feeding of PCE) and c-DCE was produced as byproduct for some bottles. Figure 6.2. depicts repeated additions and consumptions of PCB at the early stages of enrichment period. However, transformation beyond TCE was limited, with only slight c-DCE degradation during 2 years. Figure 6.3 illustrates PCE transformation by high lactate/PCE enrichments and the first order rate achieved by high and low lactate/PCE enrichments. O I T T 1 I Y 0 60 120 180 240 300 360 420 480 540 Time (days) Figure 6.2. PCB additions and consumption by high lactate/PCE treatment over 1.5 years. 12 cycles of PCB was added. 92 60 20 40 Time (days) + VC + c-DCE +TCE + PCE Figure 6.3. Effects of prolonged exposure of high lactate (12.5 mM) on enhancing transformation activity (after 12 month of incubation). Table 6.2. Effects of prolonged exposure of high versus low lactate on enhancing transformation activity (after 12 month of incubation). Transformation rate Half life Final products (l/day) a (day) b Low lactate/PCE 0.71 0.98 ethene and ethane High lactate/PCE 0.04 19.4 c-DCE and TCE 3Based on first-order rate equation C=Coe'1“. b rm = ln 2/k. 93 6.4.1.2. Fermentation balance An electron balance was determined after 10 days of fermentation. To accomplish this, fatty acid transformation, methane, and hydrogen were monitored 10 days after the spiking of lactate and PCE. Using lactate and PCE as starting compounds, microequivalents of lactate consumed were balanced with the sum of the microequivalents of reduced products formed. The electron balance for one of the six replicate bottles is shown in Table 6.3. It is apparent that nearly 76.5 % of the reducing equivalents from lactate were used to produce acetate and propionate and the other 10 % being used in methane production. Dechlorination and hydrogen production were insignificant sinks (< 1.5 %) for electrons on 10th day electron balance. The other five replicate bottles exhibited similar results. The fraction of lactate equivalents used in dechlorination varied from 0.7-1.5 % and averaged 1.2 % among 5 bottles. The electron equivalents for dehalogenation were calculated assuming complete dechlorination from PCE to ethene requires 8 equiv/mol and lactate breaks down to propionate and acetate. Biomass increment was ignored. However, CO2 change during this 10 day was not negligible. The missing 8 % of reducing equivalents produced from lactate consumption may have used in nitrate or sulfate reduction. Negative and positive values represent consumption and production during 10 days. 94 Table 6.3. Electron balance for a single, PCE degrading microcosm (15 umol PCE /bottle, 12.5 mM (2500 umol) lactate; data obtained at day 10) Amount Reducing Electron Electron Percenta present Equivalents equiv. equiv. (%) Compound (umol) (eq/mol) consumed formed (ueq) (ueq) Lactate -476 4 - 1 904 Propionate 52 6 3 12 16.4 Acetate 129 8 1032 54.2 CH4 23 8 184 9.7 H2 0.27 2 0. 54 0 . 03 PCE - l 3 0 TCE l 3 2 26 l .4 DCE isomers 0 4 VC 0 6 Ethene 0 8 aPercent calculated by dividing reducing equivalents produced over total reducing equivalents consumed. 95 6.4.2. Low lactate conditions 6.4.2.1. Enrichment of high rate activity Low lactate/PCE bottles were respiked with PCB and replenished with fi'esh groundwater media just like high lactate/PCE bottles. However, lactate was supplied 25 times less than high lactate/PCE bottles (100 mol). Reducing potential for low lactate/PCE reached approximately -200 and pH was monitored and maintained between 6.8-7 .5. After 4 cycles of repeated feedings, transformation rates became faster (once a month feeding of PCE) and more complete. For example, primarily VC was produced as byproduct. By 9 cycles of repeated feedings (12 month), bottles treated with low lactate ones exhibited complete transformation of PCE to ethene and 20 times faster than high lactate treated ones (Figure 6.4). Since complete transformation of PCB was only observed under the low lactate conditions, the ranges of VOCs, which could be degraded under these conditions, were determined. As parent products, TCE, c-DCE, and TCA were evaluated since these compounds already present in plume G. TCE and c-DCE were completely transformed to ethene after 6 month of repeated feedings. However, limited number of TCA fed microcosms (5 out of 12) transformed to Chloroethane (CA). Ethane production from CA was not observed over 60 days. 96 0 l 1 0 90 180 270 360 450 Time (days) Figure 6.4. PCE additions and consumption by low lactate/PCE treatment during 450 days (12 cycles). 97 6.4.2.2. Fermentation balance An electron balance was calculated after 10 days of fermentation for low lactate treated microcosms. To accomplish this, fatty acid transformation, methane, and hydrogen were monitored 10 days after the spiking of lactate and PCE. Using lactate and PCE as starting compounds, microequivalents of lactate consumed were balanced with the sum of the microequivalents of reduced products formed. The electron balance for one of the six replicate bottles is shown in Table 6.4. It is apparent that about nearly half of the reducing equivalents from lactate were used to produce acetate (43%) and the other 20 percent being used in reductive dechlorination. Methanogensis and hydrogen production were insignificant sinks (4 %) for electrons on 10th day electron balance. The other five replicate bottles exhibited similar results. The fraction of lactate equivalents used in dechlorination varied from 12-28 % and averaged 24 % among 5 bottles. The electron equivalents for dehalogenation were calculated assuming complete dechlorination from PCE to ethene requires 8 equiv/mol and lactate breaks down to propionate and acetate. Biomass increment was ignored and CO2 change during this 10 day was negligible. Negative and positive values represent consumption and production during 10 days. 98 Table 6.4. Electron balance for a single, PCE degrading microcosm (15.2 mol PCE /bottle, 0.5 mM, lactate; data obtained at day 10) Amount Reducing Electron Electron present Equivalents equiv. equiv. Percenta Compound (umol) (eq/mol) consumed formed (%) (ueq) (ueq) Lactate -102 4 -408 Propionate 13 6 78 19 Acetate 22 8 17 6 43 CH4 0.2 8 1.6 0.4 H2 < 0.01 2 0.02 < 0.005 PCE ~14 0 TCE 0 2 DCE isomers 3.2 4 12.8 3 VC 11.5 6 69 17 Ethene 0 8 aPercent calculated by dividing reducing equivalents produced over total reducing equivalents consumed. 99 6.4.2.3 . Utilization of Chlorinated Ethenes Low lactate/PCE enrichments which showed complete degradation of PCB were employed to determine the ability to degrade other chlorinated compounds. In addition, degradation rates were compared with bottles enriched with TCE and c-DCE. The transformation rates observed in these experiments are summarized in Table 6.4. The enrichments usually required 5-15 days to completely degrade PCE depending on the bottles and concentration of parent products. There was stoichiometric conversion of PCE to VC, and VC dechlorination occurred only when all other chlorinated intermediates were nearly consumed. Figure 6.5 shows the sequence and duration of intermediates produced by high concentration of PCE dechlorination. The productions of TCE and c-DCE were observed as PCE degraded. Cis-dichloroethene was completely dechlorinated to VC in approximately 20 days. When high concentrations of PCE or TCE were received, some bottles produced t-DCE as intermediate, which persisted longer than c-DCE. When t-DCE alone was added in the bottles, its rate of degradation was again slower than c-DCE. VC production from t-DCE was slow and transformation of VC was substantially slower than any other chloroethenes tested. Utilization of 1,1-DCE by the culture was similar to c-DCE in that conversion from VC to ethene only occurred after depletion of 1,1-DCE (Figure 6.6). When VC was added alone, it was converted to ethene with no lag period. All these tests were done in the bottles that were previously enriched with repeated PCE feedings over a 2-year period. However, the disappearance of TCE and c- DCE from TCE, c-DCE incubated bottles did not show any increase in rate of transformation compared to PCE enriched bottles. For example, the rates of 100 transformation of c-DCE in PCE enriched bottles were not significantly different from that of c-DCE enriched bottles (Table 6.5). Figure 6.5. Transformation of PCB and intermediate productions from lactate/PCE enrichments received high dosage of PCE (80 umol). 100 e 80 — 604 amol/bottle A O 20- Time (days) + PCE -u- TCE -------- c-DCE -I- t-DCE ~>K~ VC ~0— Ethene 101 Figure 6.6. Conversion of chlorinated ethenes to ethene by low lactate/PCE enrichment. This experiment was done with the microcosms showing highest activity. Approximately 15 umol of chlorinated ethenes and 0.5 mM of lactate were used: (a) PCE; (b) TCE; (c) c-DCE; (d) t-DCE; (e) 1,1-DCE; and (0 VC. Error bars represent the standard deviation of triplicates. (a) PCE d) E 0 fl 0 E 3; -5 Tune (days) -0- PCE mi»— TCE + c-DCEwir/iw VC 916- Ethene (b) TCE 0 7 14 21 Time (days) + TCE W“: c-DCE '1'- VC ”We”: Ethene 102 (c) c-DCE 0 6 12 l 8 '5 ‘ Time (days) + c-DCE -~I-- VC Ethene (d) t-DCE “mol/bottle 0 1 5 3 0 45 Time (days) +t-DCE +VC - Ethene 103 (e) 1,1-DCE 20 ‘ +1,1-DCE wan-vc 15 - dwEthene 0 7 14 21 Time (days) 20 (f) VC 1 15 0 g 1 Q10 — E t :r --o— VC 5 _ -I-- Ethene 0 r - 1 3—9 0 6 12 18 Time (days) 104 Table 6.5. Summary of dechlorination rate observed by low lactate/PCE, low lactate/T CE, and low lactate/c-DCE enrichments (/day)“’b Compound PCE Enriched TCE Enriched c-DCE Enriched PCE 0.82 a: 0.07 N/Ac N/A° TCE 0.69 a 0.08 0.76 a: 0.02 N/A° c-DCE 0.66 2: 0.04 0.70 a. 0.04d 0.67 :2 0.03 t-DCE 0.19 i 0.06 N/P° N/Ac 1,1-DCE 0.45 :t 0.04 N/P° N/Ac VC 0.29 d: 0.03 0.292 0.03 d 0.31 :t 0.04‘1 “ Based on first-order rate equation C=Coe"“. Microcosms had 15 cycles of respking. b Experiments were conducted in triplicates (i indicates standard deviation from triplicates). cN/A = Not Available. dSome compounds like c-DCE and VC were produced as daughter products. °Some compounds are NP (Not Present or Not Produced). 6.4.2.4. Metabolism of chlorinated ethanes. The enrichments (pulse fed TCA for 12 month) also transformed 1,1-DCA and TCA in a rather different way. Some of the bottles received TCA did not show any signs of degradation. Some bottles showed rapid transformation after a l-month lag phase. Only bottles that showed TCA transformation degraded 1,1-DCA. 1,1-DCA had been found to be the main product of TCA biotransformation. Chloroethane was the final product and no ethane was observed over 2-month period (Table 6.6). 105 Figure 6.7. Conversion of chlorinated ethanes by low lactate/T CA enrichment. Individual serum bottles have received 15 pmol of chlorinated ethanes and lactate of 0.5 mM: (a) TCA and (b) 1,1-DCA. Error bars represent the standard deviation of triplicates. (a) TCA 20 +TCA + 1,1-DCA 15 +CA O 8 x- %10 - Fax“ 8 N»- a 5 — f. /rr"’/ 0 K I I F r . 0 9 18 27 36 45 Time (days) (b) 1,1-DCA 20 ~ -+— 1,1-DCA 15 if .4 .243- --------- a) _ flflflflflflfl %fl,... .1310 — ”””” £1 % / 8 :3. 5 — ,. 0 r“ 1 . . . Y 0 9 18 27 36 45 106 Table 6.6. Effects of prolonged exposure of TCA on stimulation of transformation activity (after 12 month of incubation). Transformation rate Half life Final product (l/day)ab (daY)° TCA 0.23 i 0.03 3.0 1,1-DCA and CA 1,1-DCA 0.12 at 0.02 3.7 CA “ Based on first-order rate equation C=Coe'k‘. bExperiments were conducted in quadruplicates (i indicates standard deviation from quadruplicates). c11/2 = ln Z/k. 6.4.2.5. Effect of temperature The microcosms spiked with 6 umol/bottle of PCE were kept in stationary and dark place at different temperatures (15 °C, 25 °C, and 35 °C). The effects of temperature on PCE transformation were monitored over 3 -week periods. As can be seen from Figure 6.8, all microcosms exhibited similar rates of PCB transformation. However, the byproduct disappearance was difference among treatments. The ethene production from 25 °C started at day 10, when others started at day 14. In additions, cis-DCE persisted longer at low temperature (15 °C) than others. 107 Figure 6.8. Temperature effects on PCE transformation on low lactate/PCE enrichment. Individual serum bottles have received 6 umol of PCB and lactate of 0.5 mM. Error bars represent standard deviation of triplicates. 15 °C + PCE "I" TCE + c-DCE ”)6“ VC --)1(--- Ethene umollbottle 25 °C + PCE «aw TCE ~A-~ c-DCE mam VC -)K- Ethene 35 °C + PCE -I— TCE ~*- c-DCE --\-- VC --)K.- Ethene 108 6.4.2.6. Did it go all the way to ethane? Reduction of ethene to ethane has been previously demonstrated under methanogenic condition by enrichment (deBruin et a]. 1992). In order to examine whether the enriched mixed community can produce ethane from ethene, 8 bottles were injected with PCB and sampled every 10 days. Ethane production was observed from ethene after 60 days. Utilization of ethene by the culture was similar to the rest of the chloroethenes, in that the conversion from ethene to ethane only occurred after depletion of all chlorinated parent and intermediate products (Figure 6.9). However, some bottles exhibited very slow reduction of ethene to ethane. As can be seen from high standard deviation of total ethane produced, the degree of reduction from ethene to ethane varied between microcosms. Mass balance of PCE to ethane is shown in Table 6.7. Table 6.7 . Mass balance determined for PCE transformation to ethane after 60 days Total PCE added Total Ethene produced Total Ethane produced Conversion (umol i SD)a (umol i SD) (umol :1: SD) (%) 13.15:l:1.15 0.69i1.24 10.10i1.36 82% “ SD — standard deviation from 6 out of 8 microcosms tested. The rest 2 bottles had very slow ethane production that they were not included in the calculations. 109 r l '.. ‘F‘ Figure 6.9. Final PCE transformation byproducts by low lactate/PCE enrichments. 6 out of 8 microcosms tested. The rest 2 bottles had very slow ethane production that they were not included in this figure. Individual serum bottles have received 15 umol of PCB and lactate of 0.5 mM. “mol/bottle 110 6.4.3. Microbial Community Analysis After extracting DNA from 3—ml samples of sediment groundwater slurry of pre- enriched, low, and high lactate enrichment, PCR is amplified for T-RFLP analysis and Real time PCR (RTm-PCR) for the quantification of dechlorinating populations. T-RF LP was used to determine if the microbial community had been changed over prolonged exposure to chlorinated compound loadings and lactate. Community shifts were indicated by changing peak patterns, i.e. the terminal restriction fragments (TRFs), on the electropherograms. Since it is possible for different organisms to share a common restriction site in the 168 genes, only a single dominant peak on the electropherogram was considered for that fragment. Using digestions enzyme HhaI, pre-enrichment and enrichments showed differences in fragments numbers. The patterns show that the numbers of terminal restriction fragments (TRFs) detected in the enriched ones have significantly increased from those in the pre—enrichment samples. All low lactate chloroethene enrichment (PCE, TCE, and c-DCE) had similar fragments and overall bacterial community structure, indicating that the dominant indigenous populations were similar to each other (Figure 6.10). The TCA and high lactate enrichments had different patterns from low chloroethene enrichments in terms of terminal restriction fragments (TRFs) sizes indicating change in overall bacterial community structure. T-RFLP profile fail to show a single fragment that could represent as Dehalococcoides specific fragment (List of the T-RFLP fragments and peak areas using HhaI are shown on Appendix D). 111 Figure 6.10. T-RFLP profiles generated using pre—enrichment, low lactate/PCE, and high lactate/PCE enrichments using HhaI. Each fragment size indicates percent fragment area: (a) pre-enrichment; (b) high lactate/PCE; (c) low lactate/PCE; (d) low lactate/c- DCE; (e) low lactate/TCE; and (0 low lactate/1‘ CA (a) Pre-enrichment 25- % Total Area 207 650 700 900 950 999 Fragment Size (bp) (b) High Lactate/PCE Enrichment % Total Area -.- :1 - ->u5 W5 a 450 475 500 550 600 650 700 750 801 850 899 950 1000 Fragment size (bp) 112 (0) Low Lactate/PCE Enrichment 228 400 450 475 500 550 600 651 700 751 800 850 900 950 Fragment size (bp) ((1) Low Lactate/T CE Enrichment 12- % Total Area 126 219 250 351 450 475 501 551 Fragment size (bp) 113 (e) Low Lactate/c-DCE Enrichment % Total Area 650 701 750 800 901 950 1000 475 551 600 Fragment size (bp) (f) Low Lactate/T CA Enrichment 12- % Total Area 369 450 475 500 550 600 648 650 700 751 850 900 951 Fragment size (bp) 114 Compared to pre-enrichment bottles which exhibited non-quantifiable amounts of Dehalococcoides, greater numbers of Dehalococcoides were observed in enriched bottles. As described in Table 6.7, the enrichment by low lactate/PCE treatment contained at least 160 times higher numbers of Dehalococcoides populations than the high lactate/PCE enrichments. The differences in total DNA for low chlorinated compounds were small. However, the differences in Dehalococcoides concentration were substantial. TCA enrichment did not have any detectable Dehalococcoides. The quantitative estimates on total DNA and Dehalococcoides 16S rRNA gene copies are summarized in Table 6.8. Table 6.8. Quantitative Estimation on Total DNA and Dehalococcoides populations in pre—enrichment, Low/PCE, High/PCE, c-DCE, TCE, and TCA enriched bottles using PicoGreen and RTm—PCRa Total DNA Dehalococcoides l6S rRNA gene (ng/ml) (pg/ml) (copies/ml) b Pre—enrichment 85.33 at 16.42 NQ° NRd High lactate/PCE 186.28 i 64.25 17° [1.0 a 0.78] x 103 Low lactate/PCE 129.60 :1: 77.41 256.8 [1.6 a 0.11] x 105 Low lactate/c-DCE 130.57 at 47.33 195.3 [1.2 a 0.57] x 105 Low lactate/TCE 78.76 a 4.25 118.1 [7.2 :1: 0.40] x 104 Low lactate/T CA 90.59 i 15.70 NDf NRd aTriplicate samples of slurry extracted DNA were used except for the high lactate/PCE enrichment. Three PCR reactions were run for each sample point. "16 s rRNA gene copies/ml = DNA (ug/ml)x6.023x1023/(1.5x106x660x106) from He et al. (16) “N Q = detectable but Not Quantifiable by RTm-PCR. NR = Not Reported due to the absence of data c4 Out of 8 results were used in calculated due to the high variation of DNA between samples. ‘ND = Not Detectable by RTm-PCR. 115 6.5. Conclusion and Discussion After 2 years of pulse feedings, once-fed microcosms (microcosms from chapter 5), which initially exhibited slow degradation, showed much higher rates of PCB dechlorination. The high lactate/PCE treatments exhibited improved rates of PCE dechlorination and produced TCE and c-DCE as final byproducts. The low lactate/PCE treatments were capable of sequentially dechlorinating chlorinated ethenes to ethene, and then ethane as final byproduct. These results provide evidence that the ratio between electron acceptor and donor is important. Even though 16S rRNA gene targeted primers revealed the presence of PCE-dechlorinating Dehalococcoides in high lactate enrichments, none of the laboratory microcosms completely reduced PCE to ethene. Possible reasons for incomplete dechlorination could be that the high nutrient environment resulted in elevated H2 levels, which stimulated methanogenic populations, and resulted in limited growth of halorespirating populations. In this study, 100-umol lactate (electron donor) and 15 umol of PCE were used. This is 10 times the minimum reducing equivalents necessary for a complete reduction of the 15 umol PCE provided to ethene, and is similar to the amounts supplied in other studies (Freedman and Gossett, 1989; Bagley and Gossett, 1990; Vogel and McCarty, 1985). It is not clear from our study whether a smaller amount of lactate or other electron donors would have been a better candidate for even faster and complete dechlorination. The highest rate of transformation was 0.82/day using PCE as parent product. VC, 1,1-DCE, and even t-DCE were dechlorinated by the PCB enrichment, but degradation rates were slow compared to PCE. Transformation of chlorinated ethenes 116 did not show specificity over what compound it was enrich with. TCE dechlorination by TCE enrichment have similar rate as TCE dechlorination by PCE enrichments. Lowering the ratio between substrate to PCE, along with pulse feedings has resulted in substantial enrichment of Dehalococcoides. Dehalococcoides populations were detected in all of the enriched samples by RTm-PCR On the contrary, the once fed long-term microcosms (from chapter 5) and TCA enrichments did not show detectable amounts of Dehalococcoides. Distinct differences in Dehalococcoides DNA levels were observed in low lactate and high lactate enrichments. The low lactate/PCE contained at least 160 times higher Dehalococcoides numbers than high lactate/PCE bottles, supporting the hypothesis that the Dehalococcoides cells compete better in a low rather than high nutrient environment. T-RFLP was used to generate profiles of the bacterial communities in chlorinated compound enrichments and pre-enrichment. The number of fragments was consistently higher after enrichment. The detection of greater number of terminal fragments in chlorinated enrichment is an indication that a relatively higher bacterial diversity exists after enrichment on lactate/PCE. The profiles generated by T-RF LP can vary in two ways. First, there can be variation in the number and size (in basepairs) of terminal restriction fragments (T-RFs) present in profile. Secondly, variation can be found in the height (and consequently the area) of any particular peak (Osborn et a]. 2000). However, T-RFLP does not resolve closely related organisms or organisms that happen to share the same restriction sites and hence yielding the same TRF. Therefore, a single peak in the electropherogram may r epr esent more than one organism. 117 A single common fragment present in all of the enriched microcosms (not found in pre-enrichment) by T-RF LP was not found. This may have resulted since these populations are a small percentage of the total community and only small sample amount (3 ml) was used to extract the DNA that it is not sufficient to represent whole microbial system in the bottle. Even though CO2 and methane have also been observed as dechlorination products under methanogenic conditions (V ogel and McCarty, 1985; Bradley and Chapelle, 1999), most biotransformation results in the accumulation of partially dechlorinated ethenes such as c-DCE or VC. Complete dechlorinations of chloroethene are known to produce ethene as a frnal transformation product (Maymo-Gatell, 1995). Despite the reports of ethene persistence in the subsurface (Oremland, 1981; Schink, 1985), ethene reduction was occurred predominant in low lactate enriched conditions. DeBruin (1992) also reported that ethane could be the final product of dechlorination in mixed community using lactate as an electron donor. These often-observed diverse transformations byproducts emphasize the need for better understanding of the underlying microbiological process involved. Low temperature (15 °C) had no significant effects on transformation of PCE. Transformation of PCE was highest at room temperature instead of 35 °C. This might have resulted since the bottles have been enriched in room temperature over 2 years. The high and complete transformation of chlorinated compound, and the absence of chlorinated end products like VC, makes bioremediation an attractive method for the removal of chlorinated compounds at the plume G site. 118 It is not known which organisms are responsible for the reduction of ethene to ethane in our study. It is also not clear whether that same organisms are also responsible for TCA transformation since ethane production from CA was not observed. Since they were pulse fed with a different compound over an extended period, we suspect both communities are very different. More research will be required in the fiJture to backup this assumption. 119 6.6. Literature cited (a) Fathepure, B. Z. and Boyd, S. A. 1988. Dependence of tetrachloroethylene dechlorination on methanogenic substrate consumption by Methanosarcina Sp. Strain Dem. Appl. Environ. Microbiol. 54(12): 2976-2980. (b) Fathepure, B. Z. and Boyd, S. A. 1988. Reductive dechlorination of perchloroethylene and the role of methanogens. FEMS Microbiology Letters. 49(2): 149- 156. Freedman, D. L. and J. M. Gossett. 1989. Biological reductive dechlorination of tetrachloroethylene and trichloroethylene to ethylene under methanogenic conditions. Appl. Environ. Microbiol. 55(9): 2144-215 1. Maymo-Gatell, X., Y-T, Chien, J.Gossett, and SH. Zinder. 1997. Isolation of a bacterium that reductively dechlorinates tetrachloroethene to ethene. Science 276;]568-1571. Hendrickson, E.R. et a]. 2002. Molecular analysis of Dehalococcoides 16S Ribosomal DNA from chloroethene-contaminated sites throughout North America and Europe. Appl Environ. Microbiol. 68 (2) 485-495. Liu, Wen-Tso, Marsh T.L., Hans Cheng, and Larry J. Forney. 1997. Characterization of microbial diversity by determining Terminal Restriction Fragment Length Polymorphisms of genes encoding l6S rRNA. Appl. Environ. A/Iicrobiol. 63 (l 1) 4516- 4522. Schink, B. 1985. Inhibition of methanogensis by ethylene and other unsaturated hydrocarbons. FEMS Microbiol. Ecol. 31:63-69 Debruin, W. P., Kotterman, M. J. J ., Posthumus, M. A., Schraa, G., and A. J. B. Zehnder, 1992. Complete biological reductive transformation tetrachloroethene to ethane. Appl Environ. Microbiol. 58(6): 1996-2000. Bagley, D. M. and J. M. Gossett. 1990. Tetrachloroethene transformation to trichloroethene and cis-l,2- dichloroethene by sulfate-reducing enrichment cultures. Appl. Environ. Micro]. 56(8): 251 1-2516. Vogel 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(5): 1080-1083. Osborn, A. M., E. R. B. Moore and K. N. Timmis. 2000. An evaluation of terminal- restriction fragment length polymorphism (T-RF LP) analysis for the study of microbial community structure and dynamics. Environm. Microbiol. 2: 39-50. 120 Bradley, P. M. and F. H. Chapelle. 1999. Methane as a product of chloroethene biodegradation under methanogenic conditions. Environ. Sci. T echno]. 33(4): 653-656 Maymo-Gatell, X., Tandoi, V., Gossett, J. M., and, S. H. Zinder. 1995. Characterization of an H2 utilizing enrichment culture that reductively dechlorinates tetrachloroethene to vinyl chloride and ethene in the absence of methanogenesis and acetogenesis. Appl. Environ. Microbiol 61(1 1): 3 92 8-3 933 . Oremland, RS. 1981. Microbial formation of ethane in anoxic estuarine sediments. Appl. Environ. Microbiol 42: 122-129. He, Jianzhong, Kirsti M. Ritalahti, Michael R Aiello, and Frank Loffler. 2003. Complete detoxicification of vinyl chloride by an anaerobic enrichment culture and identification of the reductively dechlorinating population as a Dehalococcoides species, Appl. Environ. Microbiol 69(2): 996- 1003 . 121 Chapter Seven: CHARACTERIZATION OF DECHLORINATIN G COWUNITY BY METABOLIC INHIBITORS 7.1. Abstract Metabolic inhibitors were used to characterize the anaerobic enrichment, which dechlorinates chloroethene to ethene/ethane using lactate as an electron donor. The effects of BES, molybdate, vancomycin, and sulfate on these enrichments were evaluated over 18 days. While the inhibitory effects depended on experimental concentrations, dechlorination was not hindered by molybdate (inhibitor of sulfate-reducer), but was inhibited by high levels of sulfate. High levels of sulfate stimulated PCE to c-DCE reduction, but not the complete and rapid dechlorination. Therefore, sulfate reducers appear to be not involved in the observed rapid and complete dechlorination. Low levels of BES (5 mM; methanogen inhibitor) specifically inhibited methane production, but the cultures retained dechlorination activity, indicating that the responsible microorganism(s) is (are) non-methanogenic. High concentration of BES (50 mM) inhibited the on-set of dechlorination and dechlorination stopped at c-DCE. The decrease in Dehalococcoides DNA in the presence of high levels of molybdate and BES suggests non-specific toxic effects. High levels of vancomycin (100 mg/L; acetogens inhibitors) successfully inhibited acetate production from lactate. The reduction in acetate production also inhibited dechlorination, further suggesting that acetate-utilizing dechlorinating microbes are responsible for the observed dechlorination. 122 7.2. Introduction One of the main topics under investigation with regard to the process of reductive dechlorination is the interaction between microorganisms (methanogens, sulfate reducers, and acetogens, etc.) inhabiting contaminated groundwater and aquifer solids. To clarify the: se interactions and the roles of specific microorganisms in dechlorination, specific inhibitors were used to block the activity of potentially degradative bacteria. Molybdate is often used as an inhibitor of sulfate reduction (Smith and Klug, 19 81) and vancomycin is used to inhibit acetogenic bacteria since it is an inhibitor of cell wall synthesis in gram-positive eubacteria (Distefano et al 1992). The 2- bro moethanesulfonic acid (BES) is frequently used to inhibit of methanogenesis (D istefano et al 1992). Unlike vancomycin, molybdate, and BES, sulfate is not a sp ecific inhibitor of any microbial physiological group. Sulfate addition usually stimulates grth of sulfate-reducing bacteria and thus inhibits metabolic activity of methanogenic bacteria. The effects of molybdate, BES, vancomycin, and sulfate on PCE dechlorinating community enriched from mixed chlorinated solvents contaminated sediments was investigated. The main objective of this study was to characterize the anaerobic dechlorination activity observed in low lactate/PCE enrichment stimulated by pulse feedings of chlorinated compounds and lactate over 2 years. High rates of dechlorination rndic at ed that the transformation observed in the lactate/PCE enrichment was occurring V13 h'C‘I-lorespiration. 123 “‘0‘ MO (a) BES (b) Molybdate HO me i 0 0°“ . O 0 9 F0... 0 H 5042. 0% 0&0 o :11, 1...; .. $.47 .a’p ; 0 We" "1’ m 0" (c) Sulfate (d) Vancomycin Figure 7.1. Chemical Structures of inhibitors used in this study. Vancomycin structure obtained from Doug et a]. (2002). Two molecular studies were conducted to characterize the effect of inhibitors on enrichment. Terminal restriction length polymorphism (T-RFLP) analyses were employ ed to identify community shifts due to the addition of inhibitors. Specific primers for dete cting Dehalococcoides species were used to determine the effects of inhibitors on SUCh P O pulations. 124 7.3- Materials and Methods Cl: etnicals Bromoethanesfulfonic acid (sodium salt, 98 %), vancomycine hydrochloride (4 % water content), and molybdate (sodium salt, 99+%) were purchased from Sigma-Aldrich (1V1 ilwaukee, WI). In hibitors Studies Inhibitors were applied to low lactate/PCE vials to investigate the roles of ac etogens, methanogens, and sulfate-reducers on the onset of dechlorination. The itlhibitors were fed only once at the beginning of the experiment and their effects on transformation were monitored for 18 days. Approximately 4 g (wet weight) of sediments were anaerobically transferred to 40-ml serum vials each and sealed with screw cap containing 22-mm teflon—lined butyl rubber septa. Only low lactate/PCE rni crocosms were used in this study. The 30-mL of growth medium was amended with lactate (200 umol), vitamin solutions, phosphate, and groundwater, as described previously (Chapter 4). After transferred to their new 40-ml vials, vials were spiked with 3 umol of PCE/month for 3 month. Thirty-three vials out of 45, which showed similar tr 3118f O rmations rates to each other, were selected for this experiment in order to reduce the variation of PCB transformation between vials. A total of 33 vials were employed: 3 were prepared with no inhibitor, 9 had 3 different levels of BES (0.5, 5, 50 mM), and 6 had two different levels of vancomycin (25 mg and 100 mg), 9 had 3 different levels of 125 molybdate (0.5, 2, 6 mM), and 6 had 2 levels of sulfate (0.6 mM, 2 mM). The inhibitor concentrations were similar to other reported studies using dechlorinating enrichments for comparisons (Distefano et a]. 1992; Pavlostathis and Zhuang, 1991;,DeBest et al 1997; Bagely and Gossett, 1990; dean et al. 1992). Table 7.] depicts the list of inhibitors and their concentrations used in this chapter. Prior to each electron donor and PCE addition, the vials were thoroughly purged with 80% N2- 20% C02. The intent of purging was to prevent accumulations of VC and/or ethene, which is a known inhibitor of methanogenesis (Schink. 1985). All vials had 200 umol (5.72 mM) lactate as electron donor before 3 umol (85 11M) PCE addition. Vancomycin was added directly after purging the vials. Therest of inhibitors (BES, no lybdate, and sulfate) were bubbled with N2, autoclaved, and then used for inhibition experiments. Vials were kept in a constant room temperature throughout the course of th i 5 experiment. 126 Tab le 7.1. The list of inhibitors and their concentrations used in this chapter Samples Inhibitor Concentrations Number of (mM)a bottles Controlb 0 3 BES 0.5 3 5 3 50 3 Molybdate 0.5 3 2 3 8 3 Sulfate 0.6 3 2 3 Vancomycin 25 mg/L 3 100 mg/L 3 aConcentrations were in mM otherwise noted. bControl had no inhibitor addition. 127 7.4- Results 7.4- 1. Effects of BES The effects of three different levels of BES on the dechlorination of PCE by low lactate/PCE enrichments was examined. Earlier microcosm results (Chapter 6) indicated that methane production was low, suggesting that methanogens do not play significant role in dechlorination. However, this observation does not completely rule out a role for methanogens in PCE dechlorination. The effects of three levels of BES on PCE transformation during 18 days are depicted in Figure 7.3. Low levels of BES (0.5 mM) amendment showed no immediate eff‘ects on dechlorination, but 5 mM amendments exhibited delayed and partially inhibited dechlorination that DCEs and VC accumulated. Even though, 50 mM BES inhibited PCE dechlorination significantly, it did not stop PCE degradation completely. P CE transformation was delayed by 50 mM BES and did cause the production and accumulation of c-DCEs. This indicates that repeated dose of BES may be need to completely step PCE dechlorination. Ethene production was observed only by 0.5 mM BES treated vials during this 18-day period and both 0.5 mM BES treated vials and control vials reduced PCE in 6 days that VC and ethene were final byproduct after 18 days (Figure 7 .2 and 7.3). This result suggests that small dose of BES had no effect on dechlorination. Methane formation was inhibited by all doses of BES (Table 7 .3). Methane produCtions from 0.5 mM and 5 mM BES were only 43 and 18 percent of the methane pr Oduced by control microcosms. This result shows that the residual PCE were probably 110i su‘ITT‘icient to suppress methanogenesis and BES was the one, which inhibited 128 methanogenesis. Hydrogen sulfide, lactate, acetate, and propionate concentrations were measured for all vials, where possible (Table 7 .3). Even though BES inhibited PCE transformation, no apparent negative or positive effects on acetogenesis was observed (Table 7 .3). Methane production in the BES 50 mM was less than 1 % of total equivalents were measured as methane, however non-specific effects were observed at this concentration. Sulfide production was not measured in the BES inhibited mi crocosms due to interference by BES. + Ethene 100 — WVC -—*-- cDCE ween TCE % PCE 0 6 1 2 1 8 Time (days) Figure 7 .2. The PCE transformation by control. 129 .bao 08w Son 3 mmm mo 28 cm 98 .28 m .28 md we Began 0.53 358832 .mmm he 28 on e5 .28 m 28 no 3 8888382 mofie 38% By .2. 63mm Ea 8 ma 9 Es m mmm 3v Ea 3 ma 3 a A33 baa. E63 bag a eg new M: S e o 2 S e o M: S e o mop. {kmvmtza a. . m5 mph man—to 1‘1 89.2.4.1. to f U> + mUQIO E I OO~ m8 Ofiufiwm + f COM o>1..-T t o2 U>i§t 130 7.4- 2. Effects of Molybdate and Sulfate Two levels of sulfate and three level of molybdate were added to observe their potential effects on PCE transformation. The effects of molybdate and sulfate on PCE dechlorination during 18-day period are shown on Figure 7.4 and Figure 7 .5. Dechlorination was neither inhibited nor improved by low levels of molybdate ad ditions (0.5 and 2 mM). Both molybdate treated vials and control vials reduced PCE in 3 — 6 days and VC and ethene were final byproducts after 18 days (Table 7.2). High levels of molybdate (6 mM) amendments showed delayed and partially inhibited dechlorination that DCEs and VC accumulated. This suggests non-specific toxic effect of molydate on d e chlorination. The PCE dechlorination was neither delayed nor inhibited by 0.6 mM sulfate, but high levels of sulfate (2 mM) significantly reduced dechlorination of PCE that c—DCE were major byproduct after 18 days. Molybdate did inhibit sulfate reduction (Figure 7 .6). After 18 days of incubation, the residual sulfate from groundwater persisted in the microcosms amended with molydate. In sulfate-amended vials, low levels of sulfate (0.6 mM) were all reduced within 7 days, but high levels of sulfate (2 mM) persisted over 18 days. Hydrogen sulfide, lactate, acetate, and propionate concentrations were measured for all Vials, where possible (Table 7.3). Essentially stoichiometric sulfide production was observed in the 0.6 mM and 2 mM sulfate treated vials. Methane production was very low, accounting for only 1 % of the electron equivalents measured. 13] Both high sulfate and high molybdate inhibited methane production. Lesser amount of methane was observed by sulfate (0.6 mM and 2 mM) amendments compare to the molybdates and controls. However, high levels of molybdate (6 mM) did exhibit inhibitory effect on methanogenesis. As shown in Table 7 .3, low levels of molybdate (O -5 and 2 mM) did not inhibit methane production while high level inhibited tn ethanogensis up to 40 percent. 132 .380 08% Son 3 339308 mo 28 o 98 .28 N .28 md as 3258a 803 m8moo€oflz .3828: we Ea e .8 .E8 N Ea no 3 8886.88 mom .8 38% 65 .I. 28E E8 e eceBreE 3 Ea N eaBbeE 3V Ee no eeBEoE 3 88V 98. Ages «a: was a: M: S o o M: NM o o E S o o mom lyml. . 881T - . m8. . t . mom o 111 mon if: r U> 1 OS U> o3 mooé ldl o3 285m 101 985m Iol o>lal 133 Figure 7.5. The effects of PCE transformation by 0.6 mM and 2 mM of sulfate. Microcosms were amended with sulfate at time zero only. 100 a -O— Ethene -*- c-DCE mew» TCE (a) 0.6 mM Sulfate 0 6 12 18 Time (days) 100 — "*"VC +c-DCE we» TCE +PCE (b) 2 mM Sulfate 0 6 12 18 134 300 - + 2 mM 200 sulfate 100 . \] (It Sulfate (ppm) O l T 7‘ [TR Tune (days) +6mMmolybdate will~-~2mMmolybdate - ~ 0.5 mM molybdate ->(- 0.6 mM sulfate Figure 7 .6. Sulfate reduction and effect of molybdate on residual sulfate in microcosms. The reduction of high sulfate (2 mM) is shown on the bigger scale. 135 7.4.3. Effect of Vancomycin Two levels of vancomycin were added to lactate/PCE enrichments to observe its potential effects on PCE transformation. On the basis of the electron balance from Chapter 6, acetogenesis from lactate was the predominant pathway, and the role of acetogens in the PCB dechlorination must be considered. Results for vancomycin (acetogenesis inhibitor)-amended bottles are shown in Figure 7.6. Figure 7.7 indicates that 25 mg/L of vancomycin had no significant effect on transformation while 100 mg/L of vancomycin delayed the transformation of PCE that VC and c-DCE was observed as final byproduct during 18 days. Table 7 .3. shows only low acetate production, demonstrating the inhibitory effect of vancomycin on acetogens. Methane productions from vancomycin treated vials were also smaller than controls. The decline in ethene production and the detection of VC and DCEs by higher concentration of vancomycin indicated that acetogens are indirectly involved in PCE dechlorination. Lactate metabolism and the resultant products are shown in Table 7 .3. Acetogenesis accounted for almost 50 percent of electron donor used for controls, and reduced products from PCE dechlorination represented nearly 10 percent. 136 100 - +Ethene w“WVC (a) 25 mg/L + c-DCE 0 6 12 18 Time (days) 100 - + VC «qt-- c-DCE "Mic“ TCE 75 2 "7'16— PCE (b) 100 mg/L > 0 6 12 18 Figure 7.7. The effects of PCE transformation by 25 mg/L and 100 mg/L of vancomycin. Microcosms were amended with vancomycin at time zero only. 137 Table 7.2. Effect of inhibitors on reductive dechlorination of PCE after 18 days3 (Initial PCE load was between 2.33 and 2.63 umol/bottle). Amt of product formed (pmol) Inh’b‘tor PCE TCE c-DCE vc Ethene BES 0.5 mM 0 0 0 0.7 1.8 5 mM 0 0 0.5 1.9 0 50 mM 0 0 2.2 0.3 0 Sulfate 0.6 mM 0 0 0 0.9 1.4 2 mM 0 0 2.3 0.2 0 Vancomycin 25 mg/L 0 0 0 0.1 2.4 100 mg/L 0 0 0.5 1.8 0 Molybdate 0.5 mM 0 0 0 0.6 1.6 2 mM 0 0 0 1.0 1.5 6 mM 0 0 0 0.5 2.0 No inhibitors 0 0 0 1. 1 l .2 aLactate (200 mol) was supplied as electron donor. The data are means obtained for triplicates samples after 18 days. The standard deviation was < i 0.2. 138 CHE 88 .2 .3 38890 a 80: “008308 8038800800 8800a 880.55. 033308 .3 008088888 9 030 2:056 33 083308 28 o 28 28 m 88m 0508 watae0§e 808880088 08.28 08308 8 5:38: 8 03. 3:808 88 #26 80808308 08.23. HEB 00:0o0f88 mmm 9 020 0385308 8: .22.. .w ..mm 88 am emu mm .8808 6 6380808 we .882: éo83>833 3850mm Bum—:28 0803 38033200 w803e0m 6088 838080 4838033360: 03 38: =< .8820 8500—0 mm @2395 33 9081 88 8825.. 2 a. 2N- em mm mm wN 3.88 2 t. :N- S o: 0 ON Eae : 3 weN- wN :: o E Ea N E 3 MN. _N 3 0 mm Ea no 8683362 2 2 NE. a 2 mm : ewa OS t 2 m2- a N R M: Ema mN 83888.8( _ 2 Se- CM 2 maN m Ea N _ 2 SN. Ne on E m Ea no Saw—Sm _ Nm SN- 3 an N Ea ow m 3 MN. 8 g m Ea m a we NeN- mm 9: .32 NE Ea no mmm g g OHSOMEU o~30wdo~wuoo< ofifiowd OuwfiomQOunm Dawn—004% .mm £0 HOHEEGH 580: “0002005 e330 N. San 00:23 80800—0 no 88358 cc 8803?: .2. 033. 139 7.4.4. Microbial Community Analysis After extracting DNA from 3-ml sediment slurry of microcosms treated with inhibitors for 18 days, PCR is amplified for T-RFLP analysis and Real time PCR was conducted for quantification of dechlorinating populations. T-RFLP was used to determine if the microbial community had been changed by the exposure to inhibitors. Changes in peak patterns and number of fragments can indicate community shifts. Since it is possible for different organisms to share a common restriction site in the 16S genes, only a single dominant peak on the electropherogram was considered for that fragment. Not all T-RFLP produced usable patterns for inhibitor study. Many of them were over digested or not digested enough to represent the effect of inhibitors on microbial communities. Using digestions with the enzyme Hhal, inhibitors produced different patterns compare with controls. Each inhibit treated microcosms had different patterns from control in terms of terminal restriction fragments (TRFs) sizes indicating change in overall bacterial community structure (Figure 7.8). List of the T-RFLP fragments and peak areas using HhaI are shown on Appendix F. 140 Figure 7 .8. T-RFLP profiles generated by using dechlorinating enrichments treated with inhibitors. Each fragment size indicates percent fragment area: (a) BES (0.5 mM); (b) molybdate (2 mM); (c) sulfate (0.6 mg/L); and (d) vancomycin (25 mg/L). (a) BES ._- N N Ur 0 tn 1 J % Total Area 8 176 268 348 393 475 476 600 651 685 950 Fragment size (bp) (b) Molybdate % Total Area 334 476 500 600 651 700 751 850 900 950 999 Fragment size (bp) 00 15 141 (c) Sulfate 16 - % Total Area 00 ii? A 185 217 464 475 500 600 649 850 900 951 Fragment size (bp) ((1) Vancomycin 12] oo 1 % Total Area 207 450 476 492 501 550 601 650 700 750 800 850 900 950 Fragment size (bp) 142 Microcosms amended with inhibitors such as BES showed significant decay in the total populations and Dehalococcoides populations at high levels (Table 7 .4). These high levels (50 mM BES) exhibited non-specific toxic effect, which led to a 60 % decrease in total DNA. The 16S rRNA demonstrated a decrease in Dehalococcoides populations due to the high levels of BES. High levels of molybdate (6 mM) also exhibited non-specific toxic effects in that both total DNA and Dehalococcoides levels declined. The low levels of sulfate and vancomycin did not significantly alter the total DNA levels and Dehalococcoides populations. While 25 mg/L vancomycin exhibited no effects on both DNA and Dehalococcoides yield, less total DNA and Dehalococcoides were observed at high levels (100 mg/L). 143 .808 020:3 8000883 205 was Rama 8:888 £20.... 8052:. £053 8 m8: mom .8.“ vow—0%: 0S 8098:: .300 0806 .A8m0080m8 8003009 805283 38388 3 0:3 820 880808 1:63:08. mm 8327. was 820 Mom.8._.d 325.3 a 6: ace Aeoaoeexeoim. Qe2meo83aa5 <78 n 88286 86» <28 may .88 088.8 :08 com 8: 0803 803008 Mom 0083. E02. 0203 <75 “008.0820 bus—m .«o 0,038.8 88:28.? 22 a 2 02 a ow 2. a 02 mm a n: 3860 .20.? .20.; 2 «EN : «NMN Eae 0.2 a? .20.; $32 :32 EaN .2 e. we .2 a an 8 a on. E a aNN Ea no .8886: .2 x3 .2 a: £32 EnSN Ana 2: .2 x 3 N2 0. S 2 2. NM mm a SN .28 N 803800§> .2 0.3 m2 xNe Rae: aNnEN EN .2 x 3 .2 03% Nate 2 nme Eaed Beam .2 f: .2 x3 ern 2 NNafiN anw .2 x m.» .2 a 2. mm in em 2. a 8N Ea m .2 SN .2032 $32. an”: $82 28 come 883 Ste 000.22. aoaaeaeev eaem <28 2: cane <78 86H ROAAER 88 8005005 mean 8200808 @8088 80280053 23 .0323 033308 .mmm 2: 8 82338 8.800000230Q 23 <75 :38. .«0 aosmgmm 03238830 .1. 03a... 144 7.5. Conclusion and Discussion Characterization of the dechlorination activities in an undefined microbial community is an important step in the identification of the responsible microbial populations. Information from these studies can provide a basis for further enrichment, and may facilitate the identification and isolation of microorganisms capable of dechlorination. The microbial group specific inhibitors can sometimes be used to distinguish which group is responsible for a specific activity (Oremland and Capone, 1988) To investigate the role of methanogenic, acetogenic and sulfate-reducing microorganisms on the transformation of PCB, 4 inhibitors; BES, molybdate, sulfate, and vancomycin were used. Although methane production was low in lactate/PCE enrichments, this fact alone does not rule out a role for methanogens in PCE reduction. BES, considered a selective inhibitor of methyl coenzyme M reductase, the enzyme that catalyzes the final step in methanogensis (Gunsalus et a1. 1978), was therefore used to inhibit methanogenesis in the enrichments. Adding BES eliminated methanogenic activity, which plays an important role in anaerobic microbial processes (Bhatnagar et al. 1991; Zeikus, 1977), but a partial dechlorination activity was still remained. Repeated additions of BES might have been required in order to completely stop dechlorination, possibly because microbes can degrade BES. Schink (1985) has noted this is a draw back to the use of BES. Many puzzling (diverse) results have been published regarding the use of BES. Previous works have demonstrated that methanogens can dechlorinate PCE (F athepure et al. 1987 ; Fathepure and Boyd, 1988). However, when methanogenesis stops, PCE 145 dechlorination by methanogens also stops. In previous research conducted by Freedman and Gossett (Freedman and Gossett, 1989), 5 mM BES was found to be sufficient for the complete cessation of methane production and TCE dechlorination by mixed methanogenic cultures. In the case of mixed sulfate-reducing cultures spiked with 50 mM BES, methane was not produced and PCB dechlorination declined by approximately 25 % as compared to non-inhibited cultures (Bagley and Gossett, 1990). Lofiler et al. (1997) showed that BES (2 mM) inhibited dechlorination in the absence of methanogens, indicating BES results should be interpreted with caution. From our study, BES inhibited dechlorination beyond DCEs when added at 50 mM. Methane production in the presence of 50 mM BES was less than 1 % of the total measured electron equivalents, indicating very little methanogenic activity. However, PCE dechlorination was not completely stopped. These observations deviate substantially from the pattern of PCE dechlorination observed with methanogenic systems. Therefore, it appears probable that dechlorinators are responsible for the observed dechlorination. Lactate/PCE enrichment also had high acetate production that a majority of reducing equivalents was accounted for acetate production with the remainder being accounted for by PCE reduction (refer to Chapter 6). Vancomycin effectively inhibited acetogenesis in lactate fed bottles in every concentration. Vancomycin is a eubacterial peptidoglycan synthesis inhibitor, (is bacteriocidal,) and is more effective against gram- positive than gram-negative eubacteria (Bock and Kandler, 1985; J oklik et al. 1984). It is unlikely that vancomycin directly inhibited acetogenesis in the microcosms but rather the 146 - I" .—m-—_ inhibition was probably due to the bacteriocidal effects of vancomycin on acetogens, most of which are gram positive (Ljungdahl, 1986). In contrast to 100 mg/L—amended bottles, 25 mg/L amended bottles continued to dechlorinate well after the inhibition of acetogenesis. The inhibition of PCB dechlorination by the high levels of vancomycin (100 mg/L) suggests that PCE dechlorinator could not use lactate directly but rather required lactate metabolism, most likely acetate (or H2) as the electron donor for reductive dechlorination. Figure 7.9 shows a hypothetical model of the lactate metabolism and PCB transformation by the enrichment, which includes the hypothesis that acetate and H2 as the actual electron donor for PCE dechlorination. Control (no inhibitors) performed similarly with respect to electron donor use; acetate production accounted for a majority of reducing equivalents, with the remainder being accounted for by PCE reduction and methane production. Molybdate is generally regarded as a specific inhibitor of sulfate-reducing bacteria (SRB) (Taylor and Oremland, 1979) and thus widely used in microbiology studies due to its specificity (Oremland and Capone, 1988). Molybdate is an analog that compete with sulfate for the active site of ATP sulfurylase, resulting in formation of an unstable analog-AMP complex, which readily hydrolyzes to AMP and the sulfate analog; the latter is then available to again react with ATP sulfierlase (Wilson and Bandurski, 195 8). Repetition of these events depletes intracellular ATP, thereby halting growth of the bacteria and, as a result, inhibiting sulfate reduction. Even though sulfate reduction was efficiently prevented by small amounts of molybdate (0.5 and 2 mM), no inhibition on PCE dechlorination was observed. The 147 inhibition of PCB dechlorination by high molybdate concentrations might have been resulted by a toxic effect since it can bind free sulfide ions to form a molybdosulfide complex (Tonsager and Averill, 1980), and this could influence microorganisms requiring sulfide ions (Oremland and Capone, 1988) e. g. methanogens (Bhatnagar et al. 1991; Zeikus, 1977). The methane production from 6 mM molybdate treated vials was less than 0.5 mM or 2 mM, which confirms the toxic effects on methanogens. T-RF LP was used to discern the microbial communities’s response to the inhibitor treatments. The restriction enzyme digest using the restriction enzymes Hhal was used to obtain a T-RF LP community fingerprint from the extracted DNA of the microcosm sample. Profiles of the bacterial communities showed extensive shifis due to inhibitors in terms of terminal restriction fiagments (TRFs) numbers/sizes. Figure 7 .9. Hypothetical model based on DiStefano et al. (1992), for carbon and electron flow in a lactate-PCE enrichment Lactate Acetogens Dechlorinating organisms 4 HCl Acetate ,1 / \ /,/ PCE Ethene /'/ o/. H2 / ’ CH4 Methano gens 148 7.6. Literature cited Smith, LR. and M.J. Klug. 1981. Electron donors utilized by sulfate-reducing bacteria in eutrophic lake sediment. Appl. Environ. Microbial. 42:116-121 Distefano, T. D., Gossett, J. M., and S. H. Zinder. 1992. Hydrogen as an electron donor for dechlorination of tetrachloroethene by an anaerobic mixed culture. Appl. Environ. Microbiol. 58(11): 3622-3629. Steven D. Dong, Markus Oberthur, Heather C. Losey, John W. Anderson, Ulrike S. Eggert, Mark W. Peczuh, Christopher T. Walsh, and Daniel Kahne. 2002. The structural basis for induction of vanB resistance. Am. Chem. Soc., 124 (31), 9064 -9065 Pavlostathis, S.G. and P. Zhuang. 1991. Transformation of trichloroethylene by sulfate- reducing cultures enriched from a contaminated subsurface soil. Appl. Microbial. Biotech. 36 (3): 416-420. DeBest, J. H., Jongema, H.,Weij1ing. A., Doddema, H. J., Janssen.,D. B. and W. Harder. 1997. Transformation of 1,1,1-trichloroethane in an anaerobic packed-bed reactor at various concentrations of 1,1,1-trichloroethane, acetate and sulfate. Applied Microbiology and Biotechnology. 48 (3 ): 417-423 . Bagley, D. M. and J. M. Gossett. 1990. Tetrachloroethene transformation to trichloroethene and cis-l,2-dichloroethene by sulfate-reducing enrichment cultures. Appl. Environ. Microbial. 56(8): 2511-2516. DeBruin, W. P., Kotterman, M.J. J, Posthumus, M.A., Schraa, G., and A.J.B. Zehnder 1992. Complete biological reductive transformation of tetrachloroethene to ethane. Appl. Environ. Microbiol. 58(6) 1996-2000. Schink, B. 1985. Inhibiton of methanogensis by ethylene and other unsaturated hydrocarbons. FEMS Microbiol. Ecol. 31:63-69 Ballapragada, B. S., Stensel, H. D., Puhakka, J. A., and J. F. Ferguson. 1997. Effect of hydrogen on reductive dechlorination of chlorinated ethenes. Environ. Sci. T echnal. 31(6): 1728-1734. Fennell, D. E., Gossett, J. M., and, S. H. Zinder. 1997. Comparison of butyric acid, ethanol, lactic acid, and propionic acid as hydrogen donors for the reductive dechlorination of tetrachloroethene. Environ. Sci. T echnal. 31(3): 918-926. He, Jianzhong, Kirsti M. Ritalahti, Michael R. Aiello, and Frank Loffler. 2003. Complete detoxicification of vinyl chloride by an anaerobic enrichment culture and identification of the reductively dechlorinating population as a Dehalococcoides species, Appl. Environ. Microbiol. 69(2): 996- 1 003 . 149 Oremland, RS, and D.G. Capone. 1988. Use of specific inhibitors in biogeochemistry and microbial ecology. In K.C.Marshall (ed.), Advances in microbial ecology. Vol. 10. Plenum Publishing Co. Gunsalus, R. P., J. A. Romesser, and R. S. Wolfe. 1978. Preparation of coenzyme M analogues and their activity in the methyl coenzyme M reductase system of Methanabacterium thermotrophicum. Biochemistry 17:23 74-23 7 6 Bhatnagar, L., M.K.Jain, and J .G.Zeikus. 1991. p. 251-270. Methanogenic bacteria. In variations in autotrophic life. Academic Press, Orlando, FL. Zeikus, J .G. 1977. The biology of methanogenic bacteria. Bacterial. Rev. 411514-541 Fathepure, B., Nengu, J. P. and SA. Boyd. 1987 . Anaerobic bacteria that dechlorinate perchloroethene. Appl. Environ. Microbiol. 53(11): 2671-2674. Fathepure, B. Z. and S. A. Boyd. 1988. Reductive dechlorination of perchloroethylene and the role of methanogens. FEMS Microbiology Letters. 49(2): 149-156. Freedman, D. L. and J. M. Gossett. 1989. Biological reductive dechlorination of tetrachloroethylene and trichloroethylene to ethylene under methanogenic conditions. Appl. Environ. Microbiol. 5 5 (9): 2144-21 5 1 . Loffler, Frank E., K. M. Ritalahtiand James M. Tiedje. 1997. Dechlorination of chlorethenes is inhibited by 2 —Bromoethanesulfonate in the absence of methanogens. Appl. Environ. Microbiol. 63(12): 4982-4985 Back, A. and O. Kandler. 1985. Antibiotic sensitivity of Archaebacteria, p. 525-544. In C. R Woese and RS. Wolfe (ed.), The bacteria: a treatise on structure and function, vol. VIII. Archaebacteria. Academic Press, Inc., New York. Joklik, W.K., H.P. Willett, and D. B Amos. 1984. Zinsser microbiology. Appleton- Century-Crofts, East Norwalk, Conn. Ljungdahl, LG. 1986. The autotrophic pathwa of acetate synthesis in acetogenic bacteria. Annu. Rev. Microbiol. 40: 415-450. Taylor, B. F., and R. S. Oremland. 1979. Depletion of adenosine triphosphate in Desulfovibrio by oxyanionsof group VI elements. Curr. Microbiol. 3: 101-103 Wilson, L. G., and R. Bandurski. 1958. Enzymatic reactions involving sulfate, sulfite, selenate, and molybdate. J. Biol. Chem. 233:975-981 Tonsager, SR, and B. A. Averill. 1980. Difficulties in the analysis of acid-labile sulfide in Mo-S and Mo-Fe-S systems, Anal. Biochem. 102213-15. 150 Chapter Eight: ENGINEERING APPLICATIONS 8.1. Enrichment of Rapid and Complete Dechlorinating Activity The research presented in this thesis provides insights into the practical application of selective biostimulation: how to enrich microbial populations that are capable of complete and rapid degradation of a contaminant. For oligotrophic aquifers such as Plume G, the native dechlorinating microflora are present at such small numbers that selective stimulation of the degradative indigenous soil microorganisms is required. However, the success of stimulating dechlorinating activity may be severely limited by biological factors such as competition among dechlorinating and non- dechlorinating microorganisms, and the nutritional requirements of dechlorinating bacteria. Although the complete transformation of PCE to ethene have been observed (DiStefano et al. 1991; Freedman and Gossett, 1989), the PCE transformation often stopped at TCE (F athepure and Boyd, 1988), or c-DCE (Sewell and Gibson, 1991) or VC (Vogel and McCarty, 1985). These incomplete reductive dechlorination reactions pose environmental risks because the intermediates c-DCE and VC are themselves hazardous environmental pollutants. One of the most significant findings from this research has been the achievement of complete transformation of chlorinated ethenes by mixed cultures, which originally exhibited only limited and incomplete dechlorination. By pulse feedings of lactate and chlorinated compounds over extended periods, rapid and complete dechlorinating activity has been enriched. However, when enrichments were stimulated in high nutrient 151 conditions, slow and incomplete dechlorinating activity was expressed. The time sequences of the enrichments are summarized in Table 8.1. After 1 year of enrichment, 36 out of 36 bottles from low lactate/PCE enrichment showed complete activity compared to 0 out of 28 bottles from high lactate/PCE enrichments (Table 8.2). In addition, the transformation rate of PCE became 15 times faster in low lactate/PCE enrichments than high lactate/PCE enrichment after approximately same pulse feedings (12 cycles). Table 8.1. Enhancement of first order rate k (1/day) on disappearance of parent products by pulse feedings overtime. Number of High Low Low Low Pulse feedings Lactate/PCE Lactate/PCE Lactate/TCE Lactate/c-DCE 3 0.008 0.012 0.01 0.01 6 0.04 0.38 0.42 0.33 12 0.06 0.85 0.72 0.62 Higher levels of Dehalococcoides populations were present in low lactate enrichments compare to high lactate enrichments. This demonstrated that pulse feedings strategy of chlorinated compounds and low electron donor could be a preferred method for the stimulation of halorespiring microbes. The mechanisms by which nutrient pulse feeding stimulates Dehalococcoides has two main components: (1) pulse feeding allows sustained Hz/acetate production and cofactors generation by fermentators, and (2) the relative low nutrient/high chlorinated ethene levels favor microbes that can exploit the energy (AG°') present in the oxidized Cl-C bonds. Another examples of cofactors known to affect bioremediation are iron and copper for Pseudomonas KC (Tatara et al. 1993; 152 Dybas et al. 1995; Kim, 1998; Dybas et al. 1998). Schematics showing these relationships are included in Figure 8.1. Hz and Acetate Lactate Fermentators Dehalococcoides J Cofactors and B 1 2 Figure 8.1. A schematic of the relationship between fermentators and Dehalococcoides. 153 Table 8.2. Comparisons of maximum rate between the pulse fed and once-fed enrichments in enhancing dechlorination activity observed from each condition. First order rates are obtained after the end of the enrichment experiment. Method Chloroethene Electron donor Maxrmun: k Success RatesCl (l/day) Once-fed TCE" High Lactate 0.004 0/220 Pulse-fed PCEb High Lactate 0.03 0/28 Pulse-fed PCE” Low Lactate 0.82 36/36 Pulse-fed TCEb Low Lactate 0.73 36/36 Pulse-fed cDCE" Low Lactate 0.66 36/36 Pulse-fed TCAc Low Lactate 0.25 12/36 ‘ Based on first-order rate equation C=Coe"“. b Final byproduct to ethene or ethane. “ Final byproduct to Chloroethane. d Occurrence of complete and rapid degradation. 8.2 Field Applications The remediation of groundwater containing chlorinated compounds is challenging. Traditional approaches for groundwater remediation have relied an extraction, followed by a physical/chemical process (e. g., air stripping and carbon adsorption). This approach has some major disadvantages such as high cost, inefficiencies in removing contaminants absorbed to the aquifer material, and requirement of pumping, treating, and ultimately disposal of large volumes of water. In situ bioremediation is an alternative approaches to these more traditional methods due to the discovery of bacterial populations that grow on chlorinated compounds, thus efficiently reducing and detoxifying these compounds. Major 154 advantages of employing such a method are potentially lower operation costs and minimum ecological disturbances. In addition, in-situ bioremedation does not generate waste solids for disposal. Biostimulation refers to enhancing the metabolic activity of indigenous microflora to transform target compounds and usually requires amendments (electron donors and/or acceptors) to enhance the microbial populations already present. Factors such as the delicate nutritional requirements of dehalogenating microbes and electron donor competition between dechlorinating and non-dechlorinating microorganisms have presented challenges to implementation of this promising technology in the field. Much of the work presented in this thesis has been directed toward feasibility of biostimulation in Schoolcrafi Plume G site. For oligotrophic aquifers such as Plume G, natural attenuation may be insufficient, and electron donor addition is often needed to accelerate reductive dechlorination or stimulate cometabolic activities. Although bioaugmentation offers the promise of increased control over transformation of a specific compound, competition between indigenous microflora and the introduced organisms usually presents a challenge. Laboratory and field investigations were used to develop cost effective methods of biologically generated reducing conditions to support reductive dehalogenation of PCE/TCE/T CA mixtures. We have evaluated the two levels of enhanced remediation in the Plume G: biostimulation and bioaugmentation: and compared their results over three- year period. Biostimulation could be part of a remediation strategy for a VOC plume (PCE and daughter products, 1,1,1-TCA and daughter products) co-mingled with a chromium and arsenic plume. 155 A series of periodic lactate additions were conducted in the field to test the L. . hypothesis that pulse feedings of nutrient influence the stimulation of dechlorinating activity in the study region. Figure 8.1 depicts the delivery setups employed in the Schoolcrafi Plume G. Groundwater was pumped from the flux control well to a mixing tank where lactate was added once a week using an in-tank recirculation system. Solution from the mixing tank was injected into the ground in delivery wells 1 and 2. The screened interval for these wells is between 18.9 and 25.0m bgs. Biostimulation by weekly nutrient feeding resulted in TCE, c-DCE, and VC degradation (Figure 8.2). TCE and daughter products were degraded with a different motif than occurred under bioaugmentation conditions, with slower initial TCE removal, but more rapid c-DCE and VC removal. 1,1,1-TCA was more rapidly degraded in the biostimulation than the bioaugmentation test cell (data not shown). Bioaugmentation with halorespiring microbes in a flow through system can be accomplished with a one-time inoculation of Bachman Road halorespiring consortium and weekly nutrient feeding. No pre-reducing of the aquifer was required. Degradation of TCE, c-DCE and VC has been observed, with aqueous phase reductions of up to 99% for TCE within 60 days. Degradation of TCE and daughter products has continued for over 15 months following inoculation. Both laboratory results and field evaluations have successfully demonstrated the importance of periodic feedings in stimulation of indigenous dechlorinating populations. Observed enhanced dechlorination and enumeration of Dehalococcoides populations are summarized in the Table 8.3. Compared to a test grid inoculated with Bachman Consortium (0.2 %), biostimulation shows similar levels of Dehalococcoides. 156 Overall, in situ biostimulation using native flora to remove chlorinated compounds shows promise for cleaning up contaminated groundwater and sediments. The levels of Dehalococcoides stimulated in the biostimulation grid was comparable to the bioaugmentation system, suggesting that stimulation using indigenous flora shows promise for cleaning up chlorinated compounds contaminated aquifers in a cost-effective manner. Table 8.3. Levels of Dehalococcoides DNA in the biostimuation and bioaugmentation test sites. Dehalococcoide DNA (%) Days 36 Days 235 Bioaugmentation 2.08 0.024 Biostimulation 2.86 0.025 157 floumuqemtsflzflw .< 082: 80c 0000008 80026 020002: 20000 05 m0 3033 Na. 0503 0.20.52 1 N 0 av G 2522 ”<0: O 2.432 225 6 200% 20a oEésZm 303 >202me a .503 092.200 05.: ® :02 0228202 O QZmOmA 158 Figure 8.3. Comparison on the Observed Reduction in Chlorinated Compounds in the biostimuation and bioaugmentation test sites. (a) Biostimulation System 1500 pt N c Q l 900 _ c-DCE VC Chlorinated Ethenes (ppb) Weeks from Initial Feed (b) Bioaugmentation System 1500 1200 - 900— g TCE It: i / X 600- g ‘ : , .. .' Chlorinated Ethenes (ppb) 0 10 20 30 40 50 60 70 Weeks from Inoculation 159 8.3. Literature cited DiStefano, T. D., J. M. Gossett and S. H. Zinder. 1991. Reductive dechlorination of high concentrations of Tetrachloroethene to ethene by an anaerobic enrichment culture in the absence of methanogenesis. Appl. Environ. Microbiol. 57: 2287-2292. Freedman, D. L. and J. M. Gossett. 1989. Biological reductive dechlorination of Tetrachloroethylene and Trichloroethylene to ethylene under methanogenic conditions. Appl. Environ. Microbiol. 5 5(9): 2144-215 1. F athepure, B. Z. and S. A. Boyd. 1988. Reductive dechlorination of perchloroethylene and the role of methanogens. FEMS Microbiology Letters 49(2): 149-156. Sewell, G. W., and S. A. Gibson. 1991. Stimulation of the reductive dechlorination of tetrachloroethene in anaerobic aquifer microcosms by the addition of toluene. Environ. Sci. T echno]., 25(5): 982-984 Vogel 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(5): 1080-1083. Tatara, G. M., Dybas, M. J ., and C. S. Criddle. 1993. Effects of medium and trace metals on kinetics of carbon tetrachloride transformation by Pseudomonas sp. strain KC. Appl. Environ. Microbiol. 59(7), 2126-2131. Dybas, M. J ., Tatara, G. M., and C. S. Criddle. 1995. Localization and characterization of the carbon tetrachloride transformation activity of Pseudomonas sp. Strain KC. Appl. Environ. Microbiol. 61(2): 758-762. Kim, HaeKyung. 1998. The role of trace copper in the transformation of carbon tetrachloride by Pseudomonas stutzeri KC. A MS Thesis, Department of Civil and Environmental Engineering, Michigans State University. Dybas, M. J ., Barcelona, M., Bezborodnikov, 8, Davies, S., Forney, L., Heuer, H., Kawka, O., Mayotte, T., Sepulveda-Torres, L., Smalla, K., Sneathen, M., Tiedje, J., Voice, T., Wiggert, D. C., Witt, M. E., and C. S. Criddle. 1998. Pilot scale evaluation of Bioaugmentation for in-situ remediation of a carbon tetrachloride-contaminated aquifer. Environ. Sci. T echnal. 32(22): 3598-3611. Jaime A. Graulau-Santiago. 2003. Development and Application of a Methodology to Evaluate Natural Attenuation for Chlorinated Solvents using Conceptual and Numerical Models. A Doctoral Thesis. Civil and Environmental Engineering. Michigan State University. 160 Chapter Nine: CONCLUSIONS AND FUTURE STUDY 9.1. Conclusions The following specific and general conclusions have been drawn regarding the development and characterization of dechlorinating microorganisms based on research performed using Plume G sediment and its enrichment: o Incomplete and slow dechlorination is stimulated in Plume G by microcosms by addition of nutrients such as CMB, lactate, and palmitate. o Cometabolic dechlorination was dominant in high lactate and PCE pulse fed conditions 0 Halorespiration was dominant in low lactate and PCB pulse fed conditions 0 Rates and extent of dechlorination increased with repeated feeding of chlorinated compounds 0 High hydrogen concentration did not favor dechlorination over methanogenesis. Enhanced dechlorination activity under these conditions was not observed, which agrees with predictions by other researchers based on half-velocity constants with respect to hydrogen. 0 Dehalococcoides 16S rRNA gene targeted real—time PCR confirmed that two order of magnitude higher amounts of Dehalococcoides DNA present in low lactate enrichments than high lactate enrichment. 0 Metabolic inhibitors such as molybdate did not impact the low PCE/lactate enrichments, but high levels of sulfate inhibited dechlorination. This suggested that the sulfate-reducers are not involved in the observed dechlorination. 161 0 Low levels of BES specifically inhibited methane production, but retained dechlorinating activity, indicating that the responsible microorganism(s) is (are) non-methanogenic. o The reduction of acetate production was observed in vancomycin treated enrichment and inhibited dechlorination. This suggests that acetate-utilizing microbes are responsible for the observed dechlorination. 9.2. Future Study Results from research presented in the previous chapters indicated that fithher research is required in specific areas. m, no attempts have been made to isolate PCE, TCE, and c-DCE dechlorinating microorganisms. The isolation of these dechlorinators from the enrichment appears to be necessary to better understand its nutritional dependencies on other organisms present in the culture. A better understanding of the growth requirements of TCA dechlorinating microorganisms can also be obtained through research using highly purified or pure cultures. Sew—11d, abiotic dechlorination observed in the microcosms has to be explored further due to the prevalence of high sulfate area (such as CA and CO) contaminated with chlorinated compounds. I‘m, the organism(s) responsible for the reduction of ethene to ethane has to be explored. This can be achieved again by purifying or isolating microorganisms from the enrichment. Egyfih, the role of sulfate reducing dechlorinators such as Desulfomonile tieay'ei in enrichment has to be addressed clearly. Enumeration of such microorganisms using specific primers would further dissect the enrichment community and their interactions with each other. 162 APPENDICIES 163 APPENDIX A. Electron Donor Calculations (Adapted from McCarty McCarty, Perry L. 1969. Energetics and Bacterial Growth. The Fifth Rudolf Research Conference, Rutgers, New Brunswick, New Jersey) Simple Models of Electron Donor Requirement Calculations. Microorganisms obtain energy for growth and maintenance by removing electrons from electron donors and transferring them via macromolecules to electron —deficient compounds, such as nitrate or sulfate, the terminal electron acceptors. A fraction of the electrons removed from the electron donor may also be used to reduce oxidized forms of carbon and nitrogen in the creation of new biomass. These two pathways for the consumption of reducing power are summarized in this equation where fs is the fraction of electrons diverted for synthesis and the fraction used for energy generation is termed fe. R = Rd + feRa + stc Where: Rd = half reaction for the oxidation of an electron donor normalized by the moles of electrons removed fi'om the donor, Ra = half reaction for the reduction of an electron acceptor used for energy normalized by males of electrons added to the acceptor, Rc = half reaction for the reduction of an electron accept used for synthesis normalized by the males of electrons added to the acceptor, and fs + fe = 1 AG°' Electron Half reaction (Kl/mol) donor Methanol 0.167 methanol + 0.167 water => 0.167 C0; + H+ + e' -37.53 Lactate 1/ 12 lactate + 1/3 water => 1/6 C02 +1/12HC03'+ l-F + e' -32.96 Ethanol l/12 ethanol + 3/12 water => 2/12 C02 +1/12HCO3‘+ IF +e' -31.13 Propionate 1/14 propionate + 5/14 H20 => 1/7 C02 +1/14HC03‘+ H)“ + e' -27.9 Acetate 1/8acetate + 3/8 water => 1/8C02 + 1/8HC03‘+ I-F + e' -27.98 Palmitate 1/92 C16HglOz' + 31/92 water => 15/92 C02 + 1/92HC03'+ H* + e' -27.87 164 E0 I AGO I Electron Half reaction acceptor (V) (KJ/mol) Nitrate 1/5 N03" + 6/5H+ + e’ => 1/10N2+2/5 H20 0.74 71.71 reduction 1/8 SO42' + 5/4l-F + e' => 1/8 HS’ + 1/2 H20 -0.22 20.79 Sulfate reduction Cell % C0; + 1/20 NH; + H+ + 6 => synthesis “20 C5H702N + 2/5 H20 The appropriate value for fs depends upon the type of microorganism or enrichment involved, its electron donors and acceptors, and the amount of decay a culture experiences. -kAGr (film. = —kAGr+188+A0n+ AGp k k’" where: AGs = free energy per e- mole of the electron donor used for the creation of biomass = AGp + 18.8 + AGn AGr = free energy per e- mole released by oxidation of the electron donor AGp =free energy per e- mole of the electron donor— energy need for pyruvate AGn = free energy need for N assimilation k = the efficiency which energy is converted into new chemical m=-1 if AGp<0 m=1ifAGp>0 In our system: Residual amounts of nitrate and sulfate in the groundwater; 65 ppm and 80 ppm each Assume ammonia is N-source K is 0.6 165 28 m3. 8 53.5. n 5&8 x 35.0va .0 Eu . n 5%. x N22 023 208 32 3932.0 m mm 8 @0820 m; u e .20 n e 2 no u a .366 n c 35502 022.232 .o $00023 Ea N cc 208 02 32% n 2.28.. $8380 5232 n .828 x 88286 3.0 n a .20 u a 83 u a .200 u a 200 0:00: 2520.008; 02$onde 5%. HEQMN X em 3 6288.0 memo n a .5 .o u a $8022 EQNm. “Ema. x a 8 6228.0 98 u a do u e 35ch 28 2. co 208 a; ea: 8.22:. EQQM. ”Seem x m 8 33mm; 8.0 u a .30 u 0 8232223 Em . HERE x cm om me 8323 was t a .33 u a 200 0500< cocon— aoboflm 230,—. 32am 800:2 .8099? 000 2. 0003002 00:00 005000 m0 200080 019 166 28 2 8 008 N2 $3.2“ u 5% x 0220,2020 sq . u 5% x 8600.200 00 5080 2 we 8 522 0 28 200 u a .0000 u a 0.20 n a 030 u a eaoaea 26 2.0 5 .208 cm 5.00 n 3&0 x 02200002 0 E . u. 3% x «280.200 0 $300200 0 2 8 $300200 28 200 u a .500 n a 5.0 u a 030 u e 20.58 167 APPENDIX B. Effect of Added Sulfate on the Powder X-ray Diffraction Pattern by Iron Sulfide produced in Microcosm. 0.020 0.015 - 0.010 - E Q g.” 0 0.005 - 0000 . ll l 30 40 50 60 70 80 9c 0005 29 168 APPENDIX C. Design of Experiment A statistical design of experiment was implemented to evaluate the effects of sulfate, reductant (Na2S), and nutrient in the stimulation of dechlorination (Table 1). Replications were performed to gain a sufficient level of confidence in the estimated parameters. The main and interaction effects were analyzed by AN OVA and presented in Table 2. The two main effects such as sulfate and Na2S exhibited as insignificant components on dechlorination. The nutrient alone had significant effect. According to the P-value, the two selected factors sulfate and reductant were insignificant at the 0.01 levels. The interactions between each factor were insignificant. In statistical analyses a P value of <0.01 was considered the criterion for significance. Table 1. Experimental Design Matrix for Dechlorination Run Sulfate Reductant Nutrient Run label (A) (B) (C) 1 - - - (1) 2 + - - 3 - + - b 4 + + - ab 5 - - + c 6 + - + ac 7 - + + bc 8 + + + abc 169 2300 u :80 5-2 28.8 n 5.0 822.0 M m anamomu mm 130,—. S .2 345— 2 SEN mod cod cod cod _ 0.3.... mod cod v0.0 vod _ 0.... hmd cwd ova: ova: ~ 0...a $0.0 mmd No.0 No.0 ~ new 2:. cwduu cmdhbm cmdhbu — 0 2 .o VmN mwdm mwdm _ a E .0 mm; ooém ooém _ a a— m w: mm mm 00.50m .m .8.— 00=0€a> ._0 20%—0:0. — .c N 605... 0 ~ .9 N .005.— a a .c N 105.— a 0035» £0.04 09¢. .8000..— 0 .0 .0 «0000.» 2 25624.. 0.0.0. QNM fiom wAm ”NM wd— 3.0 0.0 NS» fiwm fiwm WON NS 2: md 50 QNm mm wdm NNM 92 52 3.0 wd 2.08 .882 .N 3 $620. .0 03a... UOOOOv—Iv—dv—iv—roooot—ar—tt—tt—roooOv—tu—tv—nv—t noo~~oo~~oo~~oo~~oo~r-toov—H— «GHQ—'Or—rO—‘Ov—O—rOv—Ov—ro—rov—rOv—Ov— 170 Main Effects Ptot (data mews) for y a b 3%} ~ 25 -— M “MA 21% W W” ,1’5 * $~ “5 13} , . . r: 1 fl 1 Q a z: E f, 38 - x”, 2S - / 23 ~ / 15 ~ 18 ( 1 Interactinlot (data mewslforv e 1 e 3 f, a ff ‘ 30 -~§- o a 7” ‘““' * W Ii’a‘" ’ 33 / _ f A; / .10 1: b ,/ ' 3° -O~ o z/ a» z B f”? - as x} :f/ .w G 171 APPENDIX D. The Additions and Consumptions of Chlorinated Compounds for Experiment II umol TCE A ‘7'— 300 360 172 mono 683 120 180 240 300 360 60 Time (days) <3 68: p 120 180 240 300 360 Tune (days) 60 173 1. EARLY microcosms after 1 yr of enrichment % Total Area s—A N \O 1 l (bp) 350.54 400.64 450.82 475.3 500.6 550.63 600.76 650 700.61 750.26 300.7 350.15 900 950.7 351 401 451 475 501 551 601 650 701 750 801 Fragment size (bp) Terminal fragment 174 Peak area 1335 1234 1355 1690 1532 1462 1845 1820 1784 1890 1558 1599 1932 2088 APPENDIX E. T-RFLP Patterns from Chapter 6 850 900 951 2. T-RFLP PATTERS SHOWN IN CHAPTER 6 (Figure 6.7). (a) Pre-enrichment (b) High Lactate/PCE (c) Low Lactate/PCE Terminal Peak area Terminal Peak area Terminal Peak area fragment (bp) fragment (bp) fragment (bp) 206.92 518 449.68 1168 228.09 1115 650.26 1144 474.55 1352 400.3 1211 700 1102 499.58 1257 450.34 1243 900.19 1055 550.23 1170 475.43 1427 949.8 1341 600.12 1348 500.44 1301 999.33 1416 650.13 1455 550 1473 699.86 1358 600.48 1775 750.29 1272 650.64 1654 800.62 1292 700.4 1515 849.83 1308 750.58 1664 899.46 1515 800.31 1321 949.8 1503 850.33 1512 999.56 1956 899.82 1602 949.6 1853 (d) Low Lactate/TCE (e) Low lactate/cDCE (1) Low lactate/T CA Terminal Peak area Terminal Peak area Terminal Peak area fragment (bp) fragment (bp) fragment (bp) 126.02 475 475.29 1379 368.61 252 218.56 385 550.8 1208 450.11 1045 250.25 765 600.31 1495 475.3 1227 350.92 649 650.33 1625 500.41 1128 450.21 1223 700.92 683 550 1207 475.2 1532 750.49 1441 600.44 1457 500.72 1388 799.74 1221 648.47 610 550.53 1579 900.6 1476 650 1561 600.33 1670 950.32 1851 699.87 1414 650 1764 1000 1970 750.67 1403 700.25 1713 850.15 1251 750.4 1691 900 1517 800 1446 950.72 1675 850.45 1735 900.33 1752 950.36 2084 175 oo0_ m~m~ w5- mv__ 00N~ env— wm0_ mum— mm- 00m— omv w¢m~ awofi v~a_ no.8 018; ES mmv EQAEOQ§> A3 memo «0.3” 50.3% m ficow ”0.35 8.85 2 .30 5.000 _ _.omm 05.80 Vm. 50 $050. demv $.08 35 05:92.0 35:55 no.8 xmom mwvfl _5m~ ~_N_ maa va~ wm__ mwmfl vow ova Nm5 A28 0.8 mmdmm ovdca 2.03 3.540 94.000 3.0% _.m5v M5.m0v 3.5g 5v.mw_ €00 Eon—wee..— EEEBH mos xmom Sufism on vm0_ Nwm_ _5N_ w0o~ Nm~_ V5~_ m_m_ mmm_ om~_ mw__ Nw5~ m0v 55:35 3.03 w 0 .25 0mm 00.005 0N.¢¢5 mdm0 0v.oo0 8.000 Nm.m5v vw.mmm 05.52 300 EoEwEm amm— vcm mm: «5: NE wwm wmm 5.: 30 :m 2.03 3.30 $.80 mm.oo0 _m.m5v _.m5v 3.3m 5.5g ao.w0~ and: ae Eofiwfim Ecmgoh. No.8 xwom Ecmgor—t A26 NV 22506: 30 :25 08 5mm 3 a. 3.926 .80 “E636.— mqmmé .m 02:25.? A: 63.6 5 $9.25 E 2305 $395 Sun: A 176 \ .0 ion in u “—0? 000_ 000 000 500— 000— N0__ 000— N00_ 000_ 000 _50_ no; 030 0...: 8: 2658:; 00 05.000 00.550 00.050 0.000 00.000 5_.000 005 55.000 50.000 00.000 00.050 300 32:08:“ 358.5... N000 0050 000_ _0N_ N_~_ 000— N_m_ 0000 000_ N050 _00_ 0_0 000 N00 000 __0m 0000 020 030 ES 0 00.000 00. _ 00 0_.000 00.000 00.000 00.005 00.005 00.000 $000 000 0 00.050 000 0.050 00.50— _0.00_ 05 00.05 300 058000 358.5... «80 080 80.200 A00 500 _00 002 000 0000 00m— _0_ 000 02.: 8 £58 8.35 688 :33 028 ”ES ”28 $.08 30 aggro: 30 0__0 000_ 000 0000 0.0— 000— 050— 000— 500— 0_5_ 000— 005— 000 500— 000_ 000_ 0___ N00 _ 058080 —m=m§0r—t no.5 Jack 3):: 00 05.000 0.000 00.00 00.30 00.000 00.000 _0.000 00.005 0 _ .005 00.000 00.000 000 00.500 2.000 _N.050 000 .000 00.000 300 288000 REES... fig E 00300 0:80.06 80 8.803 0573-10 .0 177 (a) BESA (5 mM) 000— SN 20* 0‘ 6 1 n _ 2 8 82 565 .x. (b) Molybdate (6 mM) _ 5 2 . O 2 5 0 82 use .x. Fragment size (bp) 178 (c) Sulfate (2 mM) 25 ] I l 1 20 4 1 % Total Area 75 75 164 197 280 350 475 560 600 651 700 750 801 850 900 951 1000 Fragment size (bp) .fi_'_.. (d) Vancomycin (100 mg) .3 E 2 § 1— 550 600 651 750 850 901 950 974 978 1000 Fragment s'me (bp) 179 000— 000— 0_0_ 000_ 000— 0500 0000 0000 000— 05__ 500_ 000— 000— 020 Run 2003 E00E02R> :0 00.000 0.000 00.000 00.000 00.000 00.005 00.000 00 .000 00 .000 00.00 0 00 .000 0 0 .050 00.000 05 8202 REESE 8R Rom 000_ 000— 500_ 000— 500_ 000— 000_ 0000 500— 050— 000_ _000 00__ 000_ 000_ 0.000 00.000 00.000 0_.000 50.005 50.005 0_.000 00.000 00.000 0 0.000 00.050 000 5.000 00.00 0 _0.500 05 “55020 355.5%. 00.3 Java 2800 800:0 30 0000 0000 000— 000. 000— 550_ 0000 500_ 050— 000— 000_ 000— 20000 8000032 30 05.000 0.000 00.000 000 000 005 50.000 00.000 000 00.000 00.000 00.050 300 EoEwR0 REESE SR 030 0000 000— 050— 0000 550 000— 050_ 000— 000_ 00__ 00__ 05__ 05.000 50.000 0.000 3.000 00.005 2.005 00.000 00.000 00.000 _0.000 050 00.00 0 800 020E080 REESE 20.60